Southern Asia, Australia, and the Search for Human Origins 9781107017856, 1107017858

This is the first book to focus on the role of Southern Asia and Australia in our understanding of modern human origins

182 49 19MB

English Pages 348 [350] Year 2014

Report DMCA / Copyright

DOWNLOAD PDF FILE

Table of contents :
Cover......Page 1
Southern Asia, Australia and the Search for Human Origins......Page 3
Title......Page 5
Copyright......Page 6
Contents......Page 7
Illustrations......Page 9
Tables......Page 11
List of Contributors......Page 13
Introduction......Page 19
Issues and Challenges East of Africa......Page 20
When Did Homo sapiens First Appear between Africa and Australia?......Page 22
Multi-Regional Evolution or Replacement?......Page 23
How Ancient Is “Modern”?......Page 24
Conclusion......Page 25
Introduction......Page 26
The East as Ancient......Page 27
Helmut de Terra......Page 28
Hallam Movius......Page 30
G. H. R. von Koenigswald......Page 31
Assessment......Page 32
Handaxes: Why the Big Deal?......Page 33
Wider Perspectives......Page 34
The World View: Near, Middle and Far East......Page 35
Western Perceptions of the Far East......Page 36
Discussion: The Movius Line and the Advent of Homo sapiens......Page 37
Encountering the “Primitive”......Page 39
Skullology......Page 40
Studies of Australian and Tasmanian Skulls......Page 41
Australia: Fossil and Archaeological Records......Page 43
Fossils at Last......Page 46
Acknowledgments......Page 49
The Levant......Page 51
South Asia......Page 52
Mainland Southeast Asia (Malaysian Peninsula, Thailand, Cambodia, Laos, Vietnam)......Page 57
Laos......Page 58
Vietnam......Page 59
Ngandong......Page 60
The Philippines......Page 61
Tongtianyan Cave, Liujiang......Page 62
Zhirendong, Mulanshan (Mulan Mountain)......Page 63
Australia......Page 64
Discussion......Page 65
Interpretative Frameworks: Where Paradigms Clash......Page 66
Towards the Future......Page 67
Introduction......Page 69
Hominin Fossils......Page 72
Genetics......Page 74
Archaeology......Page 75
Developing an Interdisciplinary Perspective......Page 79
Evaluating the Models......Page 80
Acknowledgments......Page 81
Models for Dispersal of Homo sapiens into South Asia......Page 82
Current Models and the Human Skeletal Record......Page 85
Routes of Dispersal......Page 86
Current Models and the Timing and Routes of Dispersals......Page 87
Genetic Evidence......Page 88
Current Models and Genetic Coalescence Data......Page 89
Archaeological Evidence......Page 90
Evaluation of Models for Modern Human Dispersals to South Asia......Page 91
Acknowledgments......Page 93
Introduction......Page 94
Africa......Page 95
The Levant......Page 97
Arabia......Page 98
South Asia......Page 99
Australia and Melanesia......Page 103
Diminishing Diversity East of Africa......Page 105
Conclusion......Page 106
Acknowledgments......Page 107
Introduction......Page 108
The Discovery of the Deep Skull......Page 110
The West Mouth Sedimentary Sequence and the Location of the Deep Skull......Page 114
The Sediments within and around the Deep Skull......Page 117
The Niah Deep Skull and Cultural Modernity in Southeast Asia......Page 122
Acknowledgments......Page 124
Introduction......Page 126
Faunal Dispersals......Page 127
Island Faunal Sequences and Evolution......Page 128
Early Hominin Dispersals......Page 130
Austronesian Colonisation of ISEA......Page 132
Conclusions......Page 133
Acknowledgments......Page 135
Introduction......Page 136
Palaeoenvironment of Late Pleistocene Southeast Asia......Page 137
Foraging Strategies in the Late Pleistocene of Southeast Asia......Page 139
Discussion......Page 150
Acknowledgments......Page 152
Introduction......Page 153
Early Human Colonization of the Philippines......Page 155
The Paleoenvironmental Setting of the Callao Hominin......Page 158
Microwear Analysis and Modern Human Behaviour......Page 160
Discussion......Page 164
Acknowledgments......Page 165
Introduction......Page 166
Demographic and Ecological Changes......Page 167
A Social and Symbolic Framework......Page 170
Bonding and Open Social Networks......Page 171
Bounding and Closed Social Networks......Page 173
Europe......Page 178
Africa......Page 179
Conclusion......Page 180
Introduction......Page 182
Historical Views of Tool Use by the First Colonisers of Sahul and Island Southeast Asia......Page 183
Sahul......Page 185
The Southeast Asian Islands......Page 189
Discussion......Page 191
Acknowledgments......Page 192
Introduction......Page 193
Tasmanian Palaeoenvironment......Page 195
Human Skeletal Evidence......Page 197
Technology......Page 198
Subsistence and Land Use......Page 200
Stencil Art......Page 202
The Neanderthal Case......Page 203
Conclusion......Page 206
The Importance of Australia......Page 207
Thermal Physiology and Clothing......Page 208
Simple and Complex Clothing......Page 209
Clothing and Behavioural Modernity......Page 210
Modernity in Pleistocene Australia......Page 211
Standardized Lithics......Page 212
The Tasmanian Challenge......Page 213
Adaptive Modernity......Page 214
A Eurocentric Perspective......Page 215
Acknowledgments......Page 216
Introduction......Page 218
Previous Interpretations of Sahul......Page 219
The Impact of Taphonomy and Sampling on Sahul’s Record of Pleistocene Complex Cultural Behaviour......Page 220
Pleistocene Sahul as Reminiscent of Middle Palaeolithic Eurasia?......Page 226
Discussion......Page 228
Conclusion......Page 229
Acknowledgments......Page 230
Colonisation of Sahul: Entry from the West......Page 231
Timing the Initial Occupation of New Guinea......Page 232
Coastal Adaptations......Page 236
Interior Adaptations......Page 239
Developments after 30,000 Years Ago......Page 242
Acknowledgments......Page 244
Introduction......Page 246
Regional Setting......Page 247
Phylogeography......Page 248
Interbreeding with Neanderthals......Page 249
Interbreeding with Denisovans in Oceania......Page 250
Interbreeding with Denisovans in China?......Page 253
Single versus Multiple Recent Exit Models for AMH......Page 254
A Last Interglacial Exit into Southwest Asia?......Page 255
Summary of Genetic Arguments against a Genetically Successful Exit during MIS 5e......Page 256
India and Toba......Page 257
Archaeology, Radiocarbon and Luminescence Dating......Page 258
Genetics......Page 259
Conclusions......Page 260
Introduction......Page 261
Theoretical Questions......Page 262
Physical Anthropology......Page 265
Stone Tool Studies......Page 266
What Is to Be Explained?......Page 268
Material Culture......Page 269
Subsistence......Page 270
Population Ups and Downs......Page 271
Demography and Modern Human Behaviour......Page 272
The Importance of Australia for Understanding Human Evolution......Page 273
Acknowledgments......Page 274
Introduction......Page 275
Historical Legacies......Page 276
Human Revolutions: Open and Disguised......Page 278
Modern Humans and Adaptationist Explanations......Page 280
Conclusion: Absolute or Relative Capacity......Page 281
References......Page 283
Index......Page 337
Recommend Papers

Southern Asia, Australia, and the Search for Human Origins
 9781107017856, 1107017858

  • 0 0 0
  • Like this paper and download? You can publish your own PDF file online for free in a few minutes! Sign Up
File loading please wait...
Citation preview

Southern Asia, Australia and the Search for Human Origins This is the first book to focus on the role of southern Asia and Australia in our understanding of modern human origins and the expansion of Homo sapiens between East Africa and Australia before 30,000 years ago. With contributions from leading experts that take into account the latest archaeological evidence from India and Southeast Asia, this volume critically reviews current models of the timing and character of the spread of modern humans out of Africa. It also demonstrates that the evidence from Australasia should receive much wider and more serious consideration in its own right if we want to understand how our species achieved its global distribution. Critically examining the “Out of Africa” model, this book emphasises the context and variability of the global evidence in the search for human origins. Robin Dennell is professor of archaeology, University of Exeter. The recipient of a Leverhulme Research Fellowship (1989–1992) and a British Academy Research Professorship (2003–2006), Dennell has conducted extensive fieldwork in Bulgaria, Iran, Pakistan (where he was Field Director of the British Archaeological Mission) and China. He is the author of The Palaeolithic Settlement of Asia (Cambridge University Press, 2009) and European Economic Prehistory: A New Approach (1983), among other books. He is also a Fellow of the British Academy. Martin Porr is associate professor of archaeology and a member of the Centre for Rock Art Research and Management at the University of Western Australia. He has published widely on issues related to Palaeolithic art and archaeology. He is the editor of Ethno-Analogy and the Reconstruction of Prehistoric Artefact Use and Production (1999, with Linda Owen) and The Hominid Individual in Context: Archaeological Investigations of Lower and Middle Palaeolithic Landscapes, Locales and Artefacts (2005, with Clive Gamble). He is currently engaged in research projects on the Pleistocene settlement of the Philippines; the indigenous art of the Kimberley, northwest Australia; and the Early Upper Palaeolithic art of Central Europe.

Southern Asia, Australia and the Search for Human Origins Edited by

Robin Dennell University of Exeter Martin Porr University of Western Australia

32 Avenue of the Americas, New York, NY 10013-2473, USA Cambridge University Press is part of the University of Cambridge. It furthers the University’s mission by disseminating knowledge in the pursuit of education, learning and research at the highest international levels of excellence. www.cambridge.org Information on this title: www.cambridge.org/9781107017856 © Cambridge University Press 2014 This publication is in copyright. Subject to statutory exception and to the provisions of relevant collective licensing agreements, no reproduction of any part may take place without the written permission of Cambridge University Press. First published 2014 Printed in the United States of America A catalog record for this publication is available from the British Library. Library of Congress Cataloging in Publication data Southern Asia, Australia and the search for human origins / [edited by] Robin Dennell, Martin Porr. pages  cm. Includes bibliographical references and index. ISBN 978-1-107-01785-6 (hardback) 1.  Human evolution – South Asia.  2.  Human evolution – Australia.  3.  Human beings – South Asia – Origin.  4.  Human beings – Australia – Origin.  I.  Dennell, Robin.  II.  Porr, Martin. GN 281.S 685  2013 599.93′80954–dc23    2013027296 ISBN

978-1-107-01785-6 Hardback

Cambridge University Press has no responsibility for the persistence or accuracy of url s for external or third-party Internet Web sites referred to in this publication and does not guarantee that any content on such Web sites is, or will remain, accurate or appropriate.

Contents

List of Illustrations List of Tables List of Contributors One.  The Past and Present of Human Origins in Southern Asia and Australia Robin Dennell and Martin Porr Two. East Asia and Human Evolution: From Cradle of Mankind to Cul-De-Sac Robin Dennell Three. “Rattling the Bones”: The Changing Contribution of the Australian Archaeological Record to Ideas about Human Evolution Sandra Bowdler Four. Smoke and Mirrors: The Fossil Record for Homo sapiens between Arabia and Australia Robin Dennell Five. An Arabian Perspective on the Dispersal of Homo sapiens Out of Africa Huw S. Groucutt and Michael D. Petraglia Six. Assessing Models for the Dispersal of Modern Humans to South Asia James Blinkhorn and Michael D. Petraglia Seven. East of Eden: Founder Effects and the Archaeological Signature of Modern Human Dispersal Christopher Clarkson Eight. Missing Links, Cultural Modernity and the Dead: Anatomically Modern Humans in the Great Cave of Niah (Sarawak, Borneo) Chris Hunt and Graeme Barker Nine. Faunal Biogeography in Island Southeast Asia: Implications for Early Hominin and Modern Human Dispersals M. J. Morwood† Ten. Late Pleistocene Subsistence Strategies in Island Southeast Asia and Their Implications for Understanding the Development of Modern Human Behaviour Philip J. Piper and Ryan J. Rabett Eleven. Modern Humans in the Philippines: Colonization, Subsistence and New Insights into Behavioural Complexity Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares Twelve.Views from Across the Ocean: A Demographic, Social and Symbolic Framework for the Appearance of Modern Human Behaviour Phillip J. Habgood and Natalie R. Franklin

page vii ix xi 1 8 21 33 51 64 76 90 108 118 135 148

v

Contents Thirteen. Early Modern Humans in Island Southeast Asia and Sahul: Adaptive and Creative Societies with Simple Lithic Industries Jane Balme and Sue O’Connor Fourteen. Tasmanian Archaeology and Reflections on Modern Human Behaviour Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks Fifteen. Clothing and Modern Human Behaviour: The Challenge from Tasmania Ian Gilligan Sixteen. Patterns of Modernity: Taphonomy, Sampling and the Pleistocene Archaeological Record of Sahul Michelle C. Langley Seventeen. Late Pleistocene Colonisation and Adaptation in New Guinea: Implications for Modelling Modern Human Behaviour Glenn R. Summerhayes and Anne Ford Eighteen. Modern Humans Spread from Aden to the Antipodes: With Passengers and When? Stephen Oppenheimer Nineteen. It’s the Thought that Counts: Unpacking the Package of Behaviour of the First People of Australia and Its Adjacent Islands Iain Davidson Twenty. Essential Questions: Modern Humans and the Capacity for Modernity Martin Porr References Index

vi

164 175 189 200 213 228 243 257 265 319

Illustrations

2.1 Movius’s interpretation of the early Palaeolithic world 2.2 Terra’s long-distance geological correlations for dating Early Palaeolithic artefacts and climatic periods 5.1 The correlation of paleoenvironmental change and key archaeological sequences in Arabia 6.1 Map of sites: Site 55, 16R Dune, Pushkar Valley, Singhbum and Wadri Atri, Bamburi and Patpara,Visadi, Hathnora, Patne, Inamgaon, Jwalapuram, Fa Hien, Batadomba Lena 6.2 Synthesis of recent environmental, genetic and archaeological evidence relating to the dispersal of Homo sapiens into South Asia, and the timing of arrival of H. sapiens as predicted by the MIS 4–3 and MIS 5 models 7.1 Examples of key artefact types from East African MSA sites, 45–100 kya: Ethiopian sites, Taramsa and Sodmein Cave 7.2 Examples of Arabian Middle Palaeolithic assemblages: Jebel Faya, United Arab Emirates; Jebel Qattar, Saudi Arabia; and Aybut Al Auwal, Oman 7.3 Examples of artefacts from South Asia: Toba Ash in the Jurreru Valley, central India; and Site 55, Riwat, Pakistan 7.4 Hypothetical correlation between phases of colonisation and technological change from MSA/Middle Palaeolithic to microlithic in South Asia 7.5 Examples of stone artefacts from Southeast Asia and Australia: Lang Rongrien, northwest Thailand; Nauwalabila, Lake Mungo and Jerimalai, Australia 7.6 Diversity of technological elements found in assemblages from 57 sites associated with presumed earliest modern humans in each region 8.1 Borneo, showing sites mentioned in the text 8.2 The West Mouth of Niah Great Cave, looking west 8.3 Tom Harrisson overseeing excavations in the Hell Trench of the West Mouth, Niah Great Cave 8.4 The discovery of the Deep Skull with reference to the location of the charcoal found previously 8.5 The Deep Skull after restoration and a human skull of recent age 8.6 Schematic representation of the Late Quaternary lithostratigraphy of the northern part of the West Mouth of Niah Great Cave 8.7 The principal Pleistocene sediments (Lithofacies 2, 2C, 3, 4) identified in the northwest corner of the West Mouth of Niah Great Cave, their inferred ages, the age of the Deep Skull, and the correlations with the climatic (isotopic) sequence in the GRIP Greenland ice core 8.8 Palynology of the sediments in the Deep Skull and in the H/6, 107" sample

page 9 12 58 66 74 79 81 84 84 86 88 91 91 93 94 95 97

98 103 vii

List of Illustrations 9.1 The distribution of land mammals from Sunda to Sahul 9.2 Predominant ocean currents in ISEA from the Pacific to the Indian Ocean (the Indonesian Throughflow) 9.3 The Flores faunal sequence over the past 1.1 million years, characterised by very few species, long-term phylogenetic continuity and two faunal turnovers 9.4 Japanese troop movements in Southeast Asia and Oceania between December 1941 and May 1942 10.1 Map of Mainland and Island Southeast Asia with the sites discussed in the text 10.2 The cave mouth at Hang Boi,Vietnam, showing the archaeological compound 10.3 A section of the west-facing profile of the Hang Boi excavation, illustrating the stratified land-snail midden 11.1 Map of the Philippine Islands with archaeological and palaeontological sites mentioned in the text 11.2 A cut mark on an animal bone from the 67 ka layer in Callao Cave 11.3 Stone tools from the Late Pleistocene layer of Ille Cave, Palawan, with corresponding microwear traces and residues 12.1 Zones of innovation in Pleistocene Sahul 12.2 Phases for behavioural innovations in Pleistocene Sahul 12.3 Stencils of Baler shell pendants (che-ka-ra), Carnarvon Gorge, central Queensland 12.4 Panaramitee tradition rock engravings at the Kybra site, Western Australia 12.5 Panaramitee tradition rock engravings at N’Dhala Gorge, central Australia 12.6 Simple figurative engraving of a "shark", Sydney region 13.1 Map of Sahul and Island Southeast Asia showing sites and locations mentioned in the text 14.1 Distribution of western and central Tasmanian Late Pleistocene archaeological sites 14.2 A small pressure flaked, denticulate quartzite flake from Bone Cave recovered from below the level dated to 23,130 ± 460 BP 14.3 Five red ochre hand stencils located on a limestone lintel in Judds Cavern, south-central Tasmania 15.1. Archaeological evidence for behavioural modernity in Late Pleistocene Tasmania and location of sites mentioned in the text 16.1 Distribution of 223 Pleistocene sites identified in Sahul 16.2 Comparison of the archaeological record of complex behaviours in Pleistocene Sahul and Middle Palaeolithic Eurasia 17.1 Possible corridors of entry into Sahul 17.2 Late Pleistocene sites of New Guinea and the Bismarck Archipelago 17.3 Archaeological sites of the Ivane Valley 17.4 Pleistocene stone tools from the Ivane Valley: South Kov, Layer 4; Joe’s Garden, Layer 3b; Airstrip Mound, Layer 4; and South Kov, Layer 4 18.1 Narrative map of modern human dispersals, as reconstructed in this review 18.2 Illustrative gene tree based on the first mtDNA complete sequence data available in 2000 (from 52 individuals randomly selected around the world) 18.3 Map showing a single southern route out of Africa and beachcomber arc route from the Red Sea along the Indo-Pacific coast to Australia, including likely extensions to China, Japan and New Guinea, from MIS 4 18.4 Map with archaeological dating of early AMH evidence in ISEA and Sahul as mentioned in the text

viii

110 111 113 116 122 127 128 136 141 145 149 151 155 156 157 159 165 178 181 185 195 203 205 214 216 217 223 229 230 234 241

Tables

4.1 4.2 4.3 5.1 6.1 8.1 8.2

Dates for Levantine hominin skeletal remains The fossil hominin evidence from southern Asia, 250–30 ka Suggested dates of Australian Pleistocene skeletal remains Key models for the dispersal of Homo sapiens out of Africa Summary of model predictions Lithofacies adjacent to the findspot of the Deep Skull in the Great Cave ICPMS geochemistry of the material from the Deep Skull and from the H/6, 107" sediment sample 8.3 Major element geochemistry of the material from the Deep Skull by X-ray fluorescence 8.4 Magnetic susceptibility of material from the Deep Skull 8.5 Geochemistry by X-ray fluorescence of the top half metre of the Hell Trench sequence in the West Mouth of Niah Great Cave 10.1 Number of identified specimens of each reptile, bird and mammal taxa (excluding the chiroptera and other smaller mammals such as murid rodents) recovered from the earliest deposits recorded at Niah Cave (ca. 45–30 ka) 10.2 Number of identified specimens of each reptile and mammal taxon recovered from the earliest deposits recorded at Ille Cave (ca. 14–13 ka) 10.3 Minimum number of individuals and number of identified specimens of invertebrates recorded within a sample of Early Holocene deposits from Hang Boi, northern Vietnam 10.4 Number of identified specimens recorded for the Late Pleistocene, Pleistocene–Holocene boundary and Early Holocene from Hang Boi, northern Vietnam, during the 2008 field season 14.1 Behavioural attributes of Sahul (Australia–New Guinea–Tasmania) compared with those of the European Middle and Upper Palaeolithic 14.2 Biological, behavioural and cultural comparisons between the late Middle Palaeolithic and the Upper Palaeolithic in Europe 15.1 Features distinguishing simple and complex clothing 15.2 Archaeological markers of behavioural modernity and the suggested strength of their association with the development of clothing 17.1 Dates of archaeological sites mentioned in text 19.1 Problems solved in Australian colonisation 19.2 Examples of early indications of emerging complex technology 19.3 Evidence of early voyages around the edge of Sahul 19.4 Separate traditions of bifacially flaked points 19.5 Some counts of languages in Sahul

page 35 36 47 52 68 98 100 101 101 102 123 125 129 130 176 187 191 193 215 245 246 246 250 251 ix

List of Tables 19.6 Selection of material culture items that show variation across Australia 19.7 Studies of the subtlety of subsistence of fisher-gatherer-hunters in Australia 19.8 Early sites in Sahul and their environments

x

252 253 254

List of Contributors

Jane Balme is a professor of archaeology in the School of Social Sciences at the University of Western Australia. Most of her research and publications are on the archaeology of Indigenous Australia, ­especially the Pleistocene period, and particularly about the technology, subsistence, social ­organisation and symbolic behaviour associated with the colonisation of Australia. She has also ­published on the development of archaeologists’ approaches to understanding gender in huntergatherer societies and on archaeology education. She has worked with Indigenous groups on archaeological projects in different parts of the Australian continent, most recently the Kimberley region. Graeme Barker is Disney Professor of Archaeology and Director of the McDonald Institute for Archaeological Research at the University of Cambridge, where he is also a Professorial Fellow at St. John’s College. His overarching research focus has been the mutual interactions between people and landscape, an interest that has involved him in field investigations of Mediterranean, desertic and tropical landscapes (Italy, Libya, Jordan, Sarawak). He has written extensively on the beginnings of farming and is currently leading a team investigating the expansion of modern humans into North Africa, through excavations at the Haua Fteah Cave in Libya. James Blinkhorn is a Fyssen Foundation Post-Doctoral Researcher based at PACEA, Université Bordeaux 1, having recently submitted his doctoral thesis (“The Palaeolithic occupation of the Thar Desert”) at the Research Laboratory for Archaeology and the History of Art, University of Oxford. His current research is focussed upon lithic technology in western India and southeastern Pakistan and integrating this within a framework of palaeoenvironmental change during the Upper Pleistocene. He has undertaken a range of excavations and survey work in India since 2003, including studies of Palaeolithic and later prehistoric sites, rock art and palaeoenvironmental landscape studies. He is a recent guest editor of a special Issue of Quaternary International titled “The Middle Palaeolithic in the desert”. Sandra Bowdler is Emeritus Professor and Senior Honorary Research Fellow in Archaeology (and Music) at the University of Western Australia. Her most recent publications include “Gatherers and grannies: further thoughts on the origins of gender” (with Jane Balme), in Australian Feminist Studies, and “The empty coast: conditions for human occupation in southeast Australia during the late Pleistocene”, in the volume Altered ecologies: fire, climate and human influence on terrestrial landscapes (ANU E Press, 2010). Christopher Clarkson is associate professor and QEII Fellow in the School of Social Science at the University of Queensland. He has published two books on lithic analysis, Lithics down under (British Archaeological Reports, 2005) and Lithics in the land of the lightning brothers (ANU E Press, 2007), and more than 50 articles and book chapters on Palaeolithic technology, human dispersals and Australian and Indian archaeology, as well as general theoretical and methodological aspects of archaeological and lithics research. He is currently engaged in projects excavating Middle and Upper Palaeolithic sites in India as well as re-excavating the site of Malakunanja II in northern Australia. xi

Contributors Richard Cosgrove is a reader in the Archaeology Program at La Trobe University, Melbourne. He has carried out fieldwork and research in most Australian states and in France, China, Jordan and England over the past 30  years. His research and teaching experience have been in Late Pleistocene human behavioural ecology, rock art studies, palaeoecology, zooarchaeology and hunter-gatherer archaeology. He has recently co-authored a book chapter, “Aboriginal exploitation of toxic nuts as a late-Holocene subsistence strategy in Australia’s tropical rainforests” (with A. Ferrier), in Peopled landscapes: archaeological and biogeographic approaches to landscapes (Terra Australis 34, ANU Press, 2012), and co-authored a paper, “Ages for hominid occupation in Lushi Basin, in catchments of South Luohe River, central China”, in Journal of Human Evolution. Iain Davidson is Emeritus Professor of Archaeology at the University of New England, Australia, and holds honorary positions also at Flinders University, the University of Queensland, Harvard University and the University of Colorado. He has published on Spanish and Australian archaeohistories, prehistoric economy, rock art, working with Aboriginal people, language origins and the evolution of cognition. He has recently published two edited volumes: Stone tools and the evolution of human cognition (with April Nowell; University Press of Colorado, 2010) and “People colonizing New Worlds” (Quaternary International, 2012). His two particular interests are the narrative of the peopling of Sahul and the role of rock and cave art in generating diversity between modern peoples. He is a Fellow of the Australian Academy of the Humanities and held the chair of Australian Studies at Harvard in 2008–2009. Robin Dennell is professor of archaeology at the University of Exeter. He is the author of European economic prehistory: a new approach (Academic Press, 1983) and The Palaeolithic settlement of Asia (Cambridge University Press, 2009) and has published widely on the early history of farming (1972–1983) and, thereafter, on the Palaeolithic of Eurasia. His main fieldwork has been in Bulgaria (1970–1972), Iran (1975–1978) and Pakistan (1981–1999), and he has recently participated in Palaeolithic research programmes in North and Central China. Robin has also played an advisory role to UNESCO in the consideration of human evolution sites for world heritage status. Anne Ford is a lecturer in the Department of Anthropology and Archaeology at the University of Otago. She recently completed her PhD modelling colonisation and occupation strategies of modern humans using lithic sources and stone tool technology in Late Pleistocene Papua New Guinea. She has guest-edited for Archaeology in Oceania on Papua New Guinean archaeology, and recent publications include “Human adaptation and plant use in Highland New Guinea 49,000 to 44,000 years ago,” in Science (with G.R. Summerhayes et al.), and “Learning the lithic landscape: using raw material sources to investigate Pleistocene colonisation in the Ivane Valley, Papua New Guinea,” in Archaeology in Oceania. Previously, she has researched and published on stone tool technology and regional economic systems in Neolithic and Early Bronze Age China. Natalie R. Franklin is an adjunct senior lecturer at the University of Queensland and an internationally renowned rock art specialist. She has published widely in academic journals and edited or authored a number of books. She has also been a regular contributor to the Rock art news of the world series. Natalie has experience in archaeological fieldwork encompassing rock art recording across Australia, and excavation both in Australia and overseas. She has extensive experience within the cultural heritage field, including developing government policies for the assessment of archaeological significance and management plans for rock art sites, as well as compliance guidelines for cultural heritage sites.Together with P. J. Habgood, she has co-authored “Modern human behaviour and Pleistocene Sahul in review” (Australian Archaeology, 2007), and she is author of the monograph Explorations of variability in Australian prehistoric rock engravings (British Archaeological Reports, 2004). Ian Gilligan is a postdoctoral research Fellow in the School of Archaeology and Anthropology, Australian National University. His primary research interest is the prehistoric development of clothing and its wider relevance to the emergence of human modernity. He graduated in xii

Contributors psychology and medicine and, before gaining his doctorate in palaeoanthropology in 2010, he completed a master’s thesis at the University of Sydney on the use of clothing in Aboriginal Australia, focusing on Tasmania. He has been involved in archaeological excavations in the Russian Far East and Spain, and his doctoral research involved examining human skeletal material from around the world relating to body shape and climate, at museums in Thailand, Japan, the United States, England and France. Related interests include the interaction between biological and behavioural cold adaptations during the Late Pleistocene (including the role of clothing technology and rapid climate change in Neanderthal extinction), together with the advent of warm-climate textile clothing during the early Holocene and the significance of fibre production (as opposed to food production) as an initial impetus for the transition to agriculture. Huw S. Groucutt is a postdoctoral researcher at the Research Laboratory for Archaeology and the History of Art, University of Oxford, where he recently submitted his doctoral thesis (“Hominin dispersals and the Middle Palaeolithic of Arabia”). His research focusses on Middle Palaeolithic lithic technology in Arabia and surrounding areas in the context of hominin population dispersals and adaptation. He has recently conducted excavations in Saudi Arabia and is undertaking a comparative analysis of lithic technology in Africa and southwest Asia. He is a recent guest editor of a special issue of Quaternary International titled “The Middle Palaeolithic in the desert”, and other recent publications include “The prehistory of the Arabian Peninsula: deserts, dispersals and demography” with Michael Petraglia in Evolutionary Anthropology. Phillip J. Habgood is an Honorary Research Fellow at the University of Queensland with extensive international experience in archaeological excavation, cultural heritage management and academic research. He has been involved in major international field research programmes including Middle Palaeolithic sites in Spain, and Neolithic, Chalcolithic and Bronze Age sites in Jordan, Syria, Bahrain and Turkey. He has also studied many of the key skeletal specimens from the Middle and Upper Pleistocene as part of his ongoing research interest into the origins of modern humans. In Australia Phillip has project-managed heritage and archaeological impact assessments and major Aboriginal surveys and mitigation and excavation programmes and has been field director on excavations of Aboriginal and non-Aboriginal (historical) archaeological sites. He works as a technical advisor for a number of Aboriginal Traditional Owner groups across Queensland. Together with N. Franklin he has co-authored “The revolution that didn’t arrive: a review of Pleistocene Sahul” (  Journal of Human Evolution, 2008) and “Explanations for patterning in the ‘package of traits’ of modern human behaviour within Sahul” (Bulletin of the Indo-Pacific Prehistory Association, 2010). Chris Hunt is reader in palaeoecology and director of research in environmental change at the Queen’s University of Belfast. His main interests are in past human-environment interactions, and he has carried out field investigations in Europe, North Africa and Western and Southeast Asia. He has worked in Borneo for the past 12  years, first as a member of the AHRB Niah Cave Project, later as co-investigator of the AHRC Cultured Rainforest Project and leader of the British Academy Loagan Bunut Project. Recent publications include “A 50,000-year record of Late Pleistocene tropical vegetation and human impact in lowland Borneo”, in Quaternary Science Reviews; “Early Holocene vegetation, human activity and climate from Sarawak, Malaysian Borneo”, in Quaternary International; and “A 2300  year record of sago and rice use from the Southern Kelabit Highlands of Sarawak, Malaysian Borneo”, in The Holocene. He is an editor of Journal of Archaeological Science. Michelle C. Langley is a PhD candidate at the Institute of Archaeology, University of Oxford, where her research focuses on identifying cultural trends in the maintenance of osseous projectile points (barbed points and sagaies) by Magdalenian populations. Previous research has centred on exploring the evidence for advanced and symbolic cognition in the Pleistocene archaeological records of Eurasia and Sahul (Greater Australia) and identifying the impact of taphonomic processes on these records. Issues of the development and use of symbolic behaviour and xiii

Contributors social technologies within Pleistocene Neanderthal and Modern Human populations remain the underlying focus of her research. She is co-author of “From small holes to grand narratives: the impact of taphonomy and sample size on the modernity debate in Australia and New Guinea Langley” (with C. Clarkson and S. Ulm; Journal of Human Evolution, 2011) and “Behavioural complexity in Eurasian Neanderthal populations: a chronological examination of the archaeological evidence” (with C. Clarkson and S. Ulm; Cambridge Archaeological Journal, 2008). Armand Salvador B. Mijares is an associate professor and currently the director of the Archaeological Studies Program, University of the Philippines. His research interest focusses on understanding early human migration in Island Southeast Asia, in particular reconstructing hunter-gatherer subsistence strategy during the Late Pleistocene and Early Holocene periods. In order to address this, he is currently working in Northern Luzon, especially in the Callao Cave Complex, where the earliest human fossil bone in the Philippines (67 kya) was found. He is also the research program head for the Mindoro Island Archaeological Research. His field of specialisations includes lithic use-wear analysis and geoarchaeology (soil micromorphology). He is the lead author for the recent article “New evidence for a 67,000-year-old human presence at Callao Cave, Luzon, Philippines” (  Journal of Human Evolution, 2010) and is author of the monograph Unearthing prehistory: the archaeology of northeastern Luzon, Philippine Islands (British Archaeological Reports, International Series, 2007). M. J. Morwood† was professor in Archaeology in the Centre for Archaeological Science, University of Wollongong. He undertook regional research projects across northern Australia – as outlined in his book Visions from the past: the archaeology of Australian Aboriginal art – and in Indonesia, where he co-directed excavations at Liang Bua on the island of Flores that yielded evidence for a tiny, endemic hominin species, Homo floresiensis. Recent publications included guest editing a special issue of Journal of Human Evolution on Liang Bua and a monograph, The discovery of the hobbit. His latest research projects focussed on the Kimberley, northwest Australia, and on the Indonesian islands of Flores and Sulawesi. Sadly, he died in July 2013 and will be immensely missed. Sue O’Connor is an ARC Laureate Professor and Head of Archaeology in the College of Asia and the Pacific, The Australian National University. Her research focuses on all aspects of initial colonisation in the Asia Pacific region as well as on theoretical issues surrounding human migration and settlement. She has current field projects in Australia, Indonesia and East Timor. Her books and co-edited volumes include 30,000 years of aboriginal occupation, Kimberley, northwest Australia (1999); East of Wallace’s Line: studies of past and present maritime cultures of the Indo-Pacific region (2000); The archaeology of the Aru Islands, Eastern Indonesia (2005); and New directions in archaeological science (2009). Recent papers in Science, the Journal of Archaeological Science and Antiquity cover topics as diverse as Pleistocene fishing strategies and coastal subsistence, the dating of Southeast Asian rock art and contemporary cave use in East Timor. Stephen Oppenheimer’s first career as an academic specialising in clinical medicine resulted in the publication of more than 100 papers on nutrition, infections and tropical diseases. He is better known to academics and a wider public for his second career as a multi-disciplinary author who writes about prehistoric migrations. In addition to numerous related original papers in scientific journals, he is the author of Eden in the east: drowned continent of Southeast Asia (1999), which examined the demographic consequences of rising sea levels at the end of the last Ice Age in Southeast Asia; Out of Eden: the peopling of the world (2004), which covered the dispersal of our species from Africa; and recently, The origins of the British: a genetic detective story (2007). Alfred F. Pawlik is associate professor and coordinator for research at the University of the Philippines, Archaeological Studies Program. His area of research includes the archaeology of hunters and gatherers in Asia and Europe and the reconstruction of prehistoric technology and human behaviour, development of stone tool technology and identification of tool use and function with xiv

Contributors optical and electron microscopy and X-ray microprobes. He has worked in Southeast Asia since 1995 and is currently researching the migration of hominins into the Philippine Archipelago with field projects in Northern Luzon and on Mindoro and Ilin Island. Maintaining an interaction between the prehistory of Southeast Asia and Europe, he is participating in several research projects on the European Palaeolithic, Mesolithic and Neolithic.This includes the investigation of the Neanderthal type locality Feldhofer Grotto and the Eemian open-site of Inden-Altdorf, which featured the earliest evidence for housing structures and the manufacture of hafting adhesives in Central Europe. Recently, he has co-authored with Jürgen Thissen “Hafted armatures and multi-component tool design at the Micoquian site of Inden-Altdorf, Germany” (  Journal of Archaeological Science, 2011); with Miriam Haidle “The earliest settlement of Germany: is there anything out there?” (Quaternary International, 2010) and “Pleistocene modernity: an exclusively Afro-European issue?” (Bulletin of the Indo-Pacific Prehistory Association, 2011); and, with Eusebio Dizon “The Lower Palaeolithic record in the Philippines” (Quaternary International, 2010). Michael D. Petraglia is professor of human evolution and prehistory, Senior Research Fellow and Co-Director of the Centre for Asian Archaeology, Art & Culture, School of Archaeology, University of Oxford. He is also a Senior Research Fellow, Linacre College (Oxford), and a member of the Human Origins Program, Smithsonian Institution. He is co-editor of The evolution and history of human populations in South Asia: inter-disciplinary studies in archaeology, biological anthropology, linguistics and genetics (with B. Allchin; Springer, 2007) and The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics (with J. Rose; Springer, 2009). Anne Pike-Tay is professor of anthropology and zooarchaeologist atVassar College, Poughkeepsie, New York. She has conducted and published seasonality studies of Middle and Upper Palaeolithic sites in Europe, Late Pleistocene sites in Tasmania (with Richard Cosgrove) and Neolithic faunas from Europe and China. Her books and edited volumes include Red deer hunting in the Upper Paleolithic of Southwest France: a study in seasonality (Tempus Reparatum, 1991); Innovations in assessing season of capture, age and sex of archaeofaunas (editor, ArchaeoZoologia XI, La Pensée Sauvage, 2001); Before Lascaux: the complex record of the Early Upper Paleolithic (co-editor with H. Knecht and R. White, CRC Press, 1993) and Mingei (  Japanese folk art) from the Brooklyn Museum (as co-editor with R. Moes and author of Ainu section, Universe Books, 1985). Philip J. Piper is an ARC Future Fellow at the School of Archaeology and Anthropology, Australian National University, and adjunct faculty member of the University of the Philippines Archaeological Studies Program. His research has focussed on the reconstruction of Late Pleistocene and Holocene human subsistence strategies, migration, colonisation and behaviour in the island environments of Southeast Asia. He has published numerous journal articles together with Ryan Rabett, including a recent paper entitled “The emergence of bone technologies at the end of the Pleistocene in Southeast Asia: regional and evolutionary implications”, in the Cambridge Archaeological Journal, and in 2009 they jointly edited a special issue of the International Journal of Osteoarchaeology entitled “New approaches to Southeast Asian zooarchaeology: papers on the vertebrate fauna at Niah Caves, Sarawak, Borneo”. Martin Porr is associate professor in archaeology at the University of Western Australia. He has published widely on Palaeolithic art and archaeology as well as on general theoretical aspects of archaeological research. With C. Gamble he edited the volume The hominid individual in context: archaeological investigations of Lower and Middle Palaeolithic landscapes, locales and artefacts (Routledge, 2005). He is currently engaged in research projects into the Pleistocene settlement of the Philippines; the Indigenous art of the Kimberley, northwest Australia; and the Early Upper Palaeolithic art of Central Europe. Ryan J. Rabett is a Research Fellow at the McDonald Institute for Archaeological Research, University of Cambridge. He has worked in Southeast Asia since 1998, and since 2007 has been xv

Contributors the director of an international archaeological project in northern Vietnam. He is the author of more than 40 articles, focusing primarily on prehistoric subsistence and technological strategies in Asia, and on early human colonization and behavioural evolution. His recent publications have included the 2012 monograph Human adaptation in the Asian Palaeolithic, published by Cambridge University Press. Wil Roebroeks is professor in Palaeolithic Archaeology at Leiden University, the Netherlands. He works on various aspects of the Pleistocene occupation history of Europe and western Asia, in collaborative and interdisciplinary projects. These projects usually entail fieldwork, for which he has carried out numerous excavations at key Palaeolithic sites in Europe.The focus of his research is the archaeological record of Neanderthals. Wil is one of the founding members (and acting vice-president) of the recently (2011) established European Society for the Study of Human Evolution (ESHE). Glenn R. Summerhayes is professor in the Department of Anthropology and Archaeology, University of Otago, where he was head of the department from 2005 to 2010. Before that, he was based at the Australian National University, as head of archaeology and natural history, RSPAS, in 2004. Glenn specialises in the archaeology of Papua New Guinea and has worked extensively across that country. His projects have been numerous, ranging from understanding the earliest occupation from 50,000 years ago to Lapita expansions from 3,300 to 2,000 years ago, and the development of present-day socio-economic configurations.

xvi

Chapter 1 The Past and Present of Human Origins in Southern Asia and Australia

Robin Dennell and Martin Porr

Introduction Debates about modern human origins remain among the most controversial in the fields of archaeology and anthropology and have generated both specialist literature and a high level of public interest.The debate not only addresses a topic relevant to each individual human being but also deals with possibly the greatest intellectual challenge altogether, trying to understand what makes us human and how humans became what they are today. Over the past decades, the scientific examination of these questions has developed into a remarkable interdisciplinary endeavour involving several fields, such as developmental psychology, socio-cultural anthropology, palaeoanthropology, molecular biology and palaeolithic archaeology – to mention just a few. Thorough discussion of the integration of these types of evidence with their respective strengths and weaknesses is beyond the scope of this volume. While information from human fossil remains as well as ancient and recent genetic material continues to have a large impact on the reconstruction of the history of Homo sapiens, this volume – edited by two palaeolithic archaeologists – ­concentrates on the role of archaeological evidence during the Upper Pleistocene (ca. 125–10 ka) in the vast geographical region that lies east of Africa. As with the editors, most of the contributors are palaeolithic archaeologists, and the volume is aimed primarily at the palaeolithic community. Two chapters have been included that provide links to two disciplines that are deeply involved in researching modern human origins: one by Oppenheimer (Chapter 18) on the genetic evidence from living populations, and one by Dennell (Chapter 4) that summarises the human skeletal evidence between Arabia and Australia from 125 to 30 ka. Given the inevitable constraints of an edited volume, we have had to choose between coherence and diversity: whether to include several chapters from one discipline or a few papers from many. In order to produce a coherent volume, we opted for the former. Large parts of discussions about modern human origins and how they might be detectable through material culture are framed in rhetorics emphasising centres of origins and subsequent dispersals. The reasons for this are complex and have as much to do with the nature of the evidence as with the politics, intellectual history and foundation of Palaeolithic archaeology and palaeoanthropology (see, e.g., Gamble & Gittins 2004; Landau 1993). The past hundred years 1

Robin Dennell and Martin Porr have seen the supposed centre of modern human origins shifting from Europe to Asia and, most recently, to Africa (Dennell 2001 and Chapter 2, this volume). During this time, an increasing emphasis on cognitive abilities that are detectable through either changes in the sophistication of material culture or the presence of symbolism has occurred, which leads to a conceptual decoupling of the supposed development of anatomical and behavioural modern features in human evolution. Most authors probably support a view that both developments originated in Africa, as seen by recent evidence for artistic behaviours at Blombos Cave in South Africa (Henshilwood et al. 2011) and the earliest anatomically modern humans from Herto and Omo Kibish, Ethiopia (White et al. 2003; McDougall et al. 2005). These discoveries are clearly important pieces of evidence for understanding the global history of H. sapiens, but it needs to be stressed that it is far from clear how these finds are related to the cognitive, anatomical and behavioural foundations of all past and present human beings – if such shared foundations indeed exist. Consequently, all questions dealing with so-called modern human origins also relate to the conceptual treatment of the causes and character of human variability, in both morphology and behaviour.These issues naturally become more relevant if more evidence of greater spatio-temporal depth and extent is included.

Issues and Challenges East of Africa This collection of papers attempts to redress an imbalance in discussions about the early history of our species outside Africa by focusing on the southern rim of Asia, from the Arabian Peninsula through India to Southeast Asia and then onwards to the Australian landmass that includes New Guinea and Tasmania. As most readers are aware, there is a mountain of literature in journals, research monographs and popular accounts of the evidence from the Levant, which contains the earliest evidence yet found of our species outside Africa as well as evidence of Neanderthals, and an even bigger mountain of literature from Europe, where H. sapiens displaced or replaced Neanderthals between 30 and 40 ka. As this was accompanied by the replacement of Middle Palaeolithic, Mousterian industries of the Neanderthals by Upper Palaeolithic ones marked by blade assemblages and shaped tools of bone, antler and ivory, great emphasis has been placed on the significance of this Upper Palaeolithic revolution (see, e.g., Mellars & Stringer 1989). While the evidence from the Levant and Europe understandably attracts so much attention, it inevitably detracts from the intrinsic interest and significance of other less well-researched regions. Many of these lie across the southern rim of Asia, between Arabia and Australia. As an example, the Arabian Peninsula is as large as Western Europe and lies immediately opposite northeast Africa across the Red Sea. Although it is an obvious dispersal route to areas further east, it has no Pleistocene human skeletal data, and only the barest outlines of a dated Palaeolithic sequence. Groucutt and Petraglia summarise what is currently known in their contribution and show why Arabia should be an essential component of the story of our species outside Africa. Given its size, it should have a complex Palaeolithic record in its own right, and not just one showing the footprints of those who crossed it from Africa. Hopefully we shall soon know much more from their current, ongoing research. India is another region that should have far greater prominence in human evolutionary studies: it is larger and more diverse than the European Union but is treated as little more than a corridor between the Levant and Australia. Blinkhorn and Petraglia discuss various theories over when and how H. sapiens first appeared in South Asia, and highlight two major recent discoveries. The first is that the stone tools before the Toba eruption of 74 ka may be similar to later ones that were probably made by H. sapiens, which raises the possibility that it was present before 74 ka, long before the common estimated date of arrival (derived from genetic studies of extant populations) of circa 60 ka. The second is the recent discovery of microliths dated to 35 ka in Central India. This discovery raises a central argument over whether such innovations have to be African-derived (see Mellars 2006b) or could instead be indigenous developments (see Clarkson et al. 2009). 2

The Past and Present of Human Origins Moving further east towards mainland and island Southeast Asia, we cannot escape two realities: one is that their Palaeolithic record is extremely poor, particularly before 40 ka, and the other is the long shadow cast by the pronouncements of Hallam Movius (1948) more than 60 years ago that the Palaeolithic inhabitants of Southeast Asia were primitive, backward and conservative. Such negative views also helped denigrate perceptions of Australian aborigines, as these were likely derived from populations in Southeast Asia. Two papers explore the history of research in Southeast Asia and Australia. Dennell discusses Western perceptions of Asia as ancient, exotic but backward and argues that Movius helped reinforce such prejudices on the basis of poorly dated surface assemblages of stone artefacts, and Bowdler shows how negatively Aborigines have been portrayed. As both argue, it is time to move on to a less prejudiced perspective. The greater part of the volume is taken up with considerations of Sunda – the great landmass of Southeast Asia that during interglacials (as now) becomes an archipelago of more than 7,000 islands – and Sahul, the continental landmass that at times of low sea level unites Australia with New Guinea and Tasmania. Sunda and Sahul are fascinating for a number of reasons. One is that despite the distance of Australia from East Africa, humans arrived here long before they entered Western Europe between 30 and 40 ka. Although the first definite evidence in Sunda for our species dates from circa 35 and 37 14C ka (40 and 44 cal ka BP) at Niah Cave, Borneo (see Hunt and Barker, this volume), humans were already venturing into the highlands of New Guinea circa 49 ka BP (see Summerhayes and Ford, this volume), and (depending upon which dates are preferred) were in Australia between 45 and 60 ka ago.The second source of fascination is that the colonisers of Sahul could have arrived only by using watercraft that were navigable by paddle or sail, as the sea currents rule out the chances of arrival by accidental drifting (see Morwood, this volume). To place this evidence in wider perspective, by 30 ka humans in Southeast Asia were routinely making round trips across 100 km or more of open sea to islands not visible from the coast; in Europe, there is no similar evidence until the Holocene. Thirdly, the Pleistocene inhabitants of Sunda and Sahul were simultaneously using a remarkably simple lithic technology (by European standards) to exploit extremely complex environments, some of which (in the case of Australia and New Guinea) were ones that had never been occupied before. As Habgood and Franklin (2008) have already pointed out, the first Australians did not arrive with an “African package” of “modern” traits such as blades, ground stone and stylised artefacts but developed these piecemeal over several millennia according to local circumstance in a “revolution that wasn’t”.These themes are explored by several contributors: Davidson, and Balme and O’Connor examine seafaring and speed of colonization as evidence for complex behaviours with simple technologies, and different forms of evidence of complex behaviours – such as maritime technologies, organic technologies, movement of plant and animal species, burial practices, detoxification, pelagic fishing and hafting – are examined in the contributions by Hunt and Barker for Borneo, Summerhayes and Ford for New Guinea, Piper and Rabett for Island Southeast Asia, and Pawlik et al. for the Philippines. For Sahul, Habgood and Franklin show in Chapter 12 how the appearance and development of art and symbolism in Aboriginal Australia should be seen as part of a set of responses of environmental instability and demographic pressure. They further suggest these factors could be usefully explored in studies of the appearance of symbolic behaviour in Middle Stone Age sub-Saharan Africa and the Middle to Upper Palaeolithic transition in Europe. A rather different approach to the Australian evidence is taken by Langley (Chapter 16), who examines the importance of taphonomic factors when considering evidence for the early manifestations of art and symbolism in Africa, Europe and Australia. She rejects comparisons between Australia and Eurasia, and argues instead that the Australian evidence shows the remarkable adaptability and flexibility of the Pleistocene colonists of Sahul. For Europeans, the fascination of the evidence from Sunda and Sahul must surely be that it inverts so many perceptions of how modern behaviour might be recognised. Here, Australia reverts to type, as it seemed an inverted world to European observers in the1830s. As J. Martin complained in the 1830s, “trees retained their leaves and shed their bark instead, the swans were 3

Robin Dennell and Martin Porr black, the eagles white, the bees were stingless, some mammals had pockets, others laid eggs, it was warmest on the hills and coolest in the valleys, [and] even the blackberries were red” (cited by Dunlap 1997). In Palaeolithic terms, the indigenous inhabitants of Sahul routinely navigated across open seas and survived in complex and often harsh environments with an astonishingly simple technology and without an Upper Palaeolithic revolution. Nonetheless, the inhabitants clearly had the potential to act in an Upper Palaeolithic manner if it was in their interests. This is borne out in Gilligan’s fascinating study of Tasmania, which on contact with Europeans was inhabited by people with the simplest technology then recorded. He shows that they had developed in the late Pleistocene some hallmarks of the European Upper Palaeolithic – sewing needles – but later discarded them and relates that to the need for sewn clothing in the severest parts of the Pleistocene. In milder times, sewn clothing was seen as unnecessary and needles were thus discarded – so they entered, and left, the “advanced” Palaeolithic depending on ambient temperature. To develop this further, the emphasis in the European Upper Palaeolithic on scrapers and needles may simply reflect the need for sewn clothing and, by extension, the need to emphasise social status by adornment of clothing with beads and jewellery rather than through using the body for colouring, tattoos or scarification – none of which would of course be preserved.

Three Major Debates about Homo Sapiens and the Southern Rim of Asia Synthetic discussions about the first appearance and subsequent development of Homo sapiens in southern Asia tend to focus on three inter-related themes: When did our species first appear between Arabia and Australia? Did it appear through local evolution or through dispersal from Africa? When can it be regarded as modern?

When Did Homo sapiens First Appear between Africa and Australia? As is well-known, the first unambiguous skeletal evidence for our species outside Africa dates from circa 125 ka and comes from the cave of Skhu¯l in northern Israel. Slightly later evidence comes from the cave of Qafzeh, also in Israel, and dated to circa 100–80 ka. After this date, H. sapiens appears to become extinct in the Levant and was replaced by Neanderthals between circa 70 and 50 ka, after which they in turn were replaced by H. sapiens. Some researchers (e.g., Shea 2008) regard this sequence as representing a failed dispersal event by H. sapiens, as it did not persist, nor did it appear to have ventured beyond the Levant. Opinions vary enormously over when our species dispersed across southern Asia and entered Australia. One argument is that as the Arabian Peninsula is much closer to East Africa – seen by many as the most likely place where our species originated – H. sapiens could have entered it as early as in the Levant (i.e., during MIS5 or late MIS6), and then dispersed eastwards (see, e.g., Dennell & Petraglia 2012). Most researchers prefer a later date: Field and Lahr (2006) suggest that it spread eastwards during MIS 4 (ca. 80–70 ka); Petraglia et al. (2007) and Clarkson et al. (2012) have argued on the basis of similarities in stone core reduction that H. sapiens was already in India before the super-eruption of the Toba volcano in Sumatra circa 74 ka; Oppenheimer (2009; 2012a; this volume) and the majority of researchers who infer population histories from genetic studies of modern populations prefer a dispersal date of < 70 ka, and likely only circa 60 ka (but see Scally & Durbin 2012); Mellars (2006c) argues on archaeological as well as genetic grounds for a dispersal date of circa 60 ka, and Klein (2009) favours a later dispersal date of circa 40–60 ka. The crucial evidence here of course is skeletal, as that alone can indicate when H. sapiens is first evidenced in Arabia, South and mainland Southeast Asia, and Sunda, as well as the identity of its predecessors in those regions, and the date of their extinction. Unfortunately, as Dennell shows in Chapter 2, this is currently impossible, because the only clear landmark in southern Asia east of 4

The Past and Present of Human Origins the Levant and East Africa is the cranium from Niah Cave, Borneo, dated to circa 40–44 cal ka BP. There is no relevant skeletal evidence older than 30 ka from South Asia, or the Holocene in the Arabian Peninsula. Although there are hints that H. sapiens is present in Sunda and South China by 100 ka, these claims are beset by uncertainties of dating, stratigraphic context, or identification as unambiguously H. sapiens. Nevertheless, one of the most unexpected recent surprises has been the discovery of a human metatarsal at Callao Cave, Luzon, in the Central Philippines dated to 67 ka, as described by Pawlik et al. in this volume. Although the metatarsal has been described as similar to that of H. habilis, H. floresiensis and H. sapiens, the last-mentioned seems the most likely as Luzon could have been reached only by boat (and H. habilis can surely be discounted as a potential coloniser!). While the evidence from Callao Cave might therefore also be indicative of further so-far-undetected hominin species in island Southeast Asia, it certainly raises the possibility that the colonisation of Australia was part of a wider process of maritime colonisation in Southeast Asia in the early Upper Pleistocene – when Neanderthals still inhabited the Levant. If so, the timing of dispersal of H. sapiens from Africa may have been seriously underestimated, as well as its capability. Because the current human skeletal record from Arabia, South and Southeast Asia, and Sunda is so inadequate, we are forced to rely on proxy indicators, notably genetic studies of regional population histories derived from modern populations, and archaeological studies of lithic assemblages. There is no doubt that population geneticists have made enormous progress in elucidating population histories across Asia for the past 60 ka, and this may well show that successful dispersals – in the sense of having extant descendants – across Asia occurred only after 60 ka. However, this leaves open the possibility of earlier dispersals that left no surviving genetic legacy in extant populations. As example, the earliest populations of H. sapiens in the Levant date from 125 ka and have left no genetic imprint on the present. Lithic evidence is also highly problematic as a source of evidence about when H. sapiens first appears in Southern Asia. In the Levant, both the earliest groups of H. sapiens and the Neanderthals who replaced them used the same type of Middle Palaeolithic assemblage, and in East Africa, the earliest groups of H. sapiens continued to use a Middle Stone Age technology. Conversely, changes in lithic technology need not indicate a change of the species that used it. As an example, Mellars (2006c) has argued the similarities between microliths 35 ka old in India and older ones in the Howiesons Poort industry of southern Asia are such that it is an “impossible coincidence” that the Indian ones were developed indigenously. Yet microliths developed indigenously in Australia and are (to date) absent from the Arabian Peninsula until the terminal Pleistocene.These issues are also addressed by Clarkson in Chapter 7, who examines the composition of artefact assemblages across southern Asia, argues that these undergo a reduction in diversity with distance from Africa and develops an argument for “for a non-microlithic dispersal of anatomically modern humans before 60 kya”. As he notes, better (and, in particular, skeletal) evidence is needed to confirm or refute these proposals, but much could still be learnt by focusing archaeological attention upon southern Asia between 50 and 60 ka.

Multi-Regional Evolution or Replacement? Arguably the longest-running debate in studies of human origins is whether our species originated in one “centre” or whether it arose indigenously from local populations. At risk of generalisation, most researchers currently favour the former model, and envisage H. sapiens as originating in Africa and then dispersing initially to the Levant, and then later throughout Eurasia and ultimately to Australia, the Pacific and the Americas. For those favouring a multi-regional model, China and Southeast Asia have often been put forward as the regions with the most compelling evidence outside Africa for the local evolution of our species from an indigenous background. If one relies solely on human skeletal evidence, neither model can at present be completely discounted. As concluded in Chapter 4, those favouring a replacement model for the appearance 5

Robin Dennell and Martin Porr of H. sapiens between Arabia and Australia cannot yet satisfactorily demonstrate when it first appeared, how often it may have dispersed or whom it replaced. Multi-regionalists face three major problems: first, in Sunda, it is not currently possible to demonstrate population continuity between the population from Ngandong – now re-dated to the Middle Pleistocene – and that at Niah Cave which is at least 100 ka younger. Secondly, the multi-regional model depends upon the existence of genetic networks across Asia, for which there is currently no indication in mainland Southeast Asia between the late Middle Pleistocene (ca. 150 ka) and the Upper Pleistocene. Additionally, as Lahr (1996) has pointed out, it is very hard to demonstrate anatomical traits that are unique to the hominin sample from China, Southeast Asia and Australia. Finally, the fossil record from South China has so many problems over dating, stratigraphic context and identification that it impedes any firm conclusions about ancestor-descendant lineages. Clearly, we need not only more fossil evidence – particularly from Arabia, India and mainland Southeast Asia – but more evidence that is not dogged by issues over its age, context or identification to species level.

How Ancient Is “Modern”? An integral but often ill-fitting part of discussions over when Homo sapiens first appeared in the regions between East Africa and Australia is the issue of modernity: at what point did the archaic populations of H. sapiens evidenced in the Levant and East Africa before 100 ka become “modern” in the sense of having the same cognitive and behavioural capabilities as ourselves? As these capabilities are regarded as cultural or behavioural and not skeletal, identification of modern humans thus shifts from the physical anthropologist to the archaeologist. The complex issues involved in these discussions are explored by Porr in the final chapter of this volume. Despite some considerable advances in the past few decades, current discussions continue to reflect issues that Darwin and Wallace struggled with in the 19th century.These involved “the higher faculties” of human thinking for which the former assumed a gradual development and perfection during human evolution and history, while the latter favoured revolutionary origins (as a product of divine intervention) (see, e.g., Porr 2010; Ingold 2004). To a certain extent, these differences are contained in recent discussions about which traits might be used to identify modernity. For example, McBrearty and Brooks (2000) argue that the origins of “modern” behaviour were acquired incrementally and have roots deep in the African Middle Stone Age; at the other extreme, Klein (2008) argues that the shift to modernity occurred circa 50 ka and involved a small but critical number of genetic changes, akin to switching on a light, that resulted in modern capabilities in language, abstract thought and symbolism. While both approaches continue to use the European Upper Palaeolithic record as a benchmark to measure behavioural complexity, they each would answer the question differently if certain material expressions, such as evidence for symbolism (e.g., ochre, shell beads and ornaments) or behavioural flexibility (Shea 2008) can be regarded as critical indicators for modernity. Both approaches consequently provide different answers to the question whether particular expressions are unique to H. sapiens or might be found in other hominin species; put another way, would evidence of art, symbolism or behavioural flexibility in Arabia, India or Sunda necessarily indicate the presence of H. sapiens? As yet, there are no traits that can be regarded as universal but exclusive to Homo sapiens. One solution to this dilemma is to invoke the evidence from Australia, which we know was colonised only by H. sapiens. As these colonists had the ability to build navigable watercraft, navigate open seas and colonise an environment wholly different from those west of the Wallace Line, we can reasonably assume that they were “modern”, and presumably acquired or developed that modernity by the time that they appeared in Southeast Asia. Australia thus provides, as Davidson points out in this volume, a baseline for when modern behaviour can be safely assumed. In a comparable manner, Cosgrove, Pike-Tay and Roeboeks discuss the claimed contrasts between “archaic” and “modern” behaviour by reference to the Tasmanian evidence. European writers such as Sollas (1911) depicted the 6

The Past and Present of Human Origins indigenous inhabitants of Tasmania as comparable to Neanderthals, even though they were clearly humans like ourselves. They argue that “models used to describe human groups as either ‘archaic’ or ‘modern’ are faulty and are clearly unhelpful in explaining issues such as the Middle to Upper Palaeolithic transition in Europe”: instead, they suggest, we need to understand cultural variability in terms of responses to solving social and environmental problems.

Conclusion We regard this volume as a work in progress: we are far from understanding how and when Homo sapiens first appeared and subsequently developed between Arabia and Australia – and offer no definite set of conclusions on these topics. There are inevitably regrettable omissions. In Arabia, Anne Delagnes and her group have conducted exemplary fieldwork in Yemen, and thankfully their work at Shi’bat Dihya is now published (Delagnes et  al. in press). Sri Lanka is another unfortunate omission and the recent publication of Perera et al. (2011) shows the rich occupation record dating back to 36 ka in that island. Nevertheless, we hope this volume shows the extraordinary diversity and complexity of the evidence so far obtained on the Pleistocene inhabitants of Sunda and Sahul and the potential richness of poorly documented regions such as Arabia and India.The southern rim of Asia, from Arabia to Australia, deserves to be treated as more than simply a corridor that humans had to traverse on their way to Australia, and as potentially every bit as fascinating in its own right as Europe or the Levant.

7

Chapter 2 East Asia and Human Evolution From Cradle of Mankind to Cul-De-Sac

Robin Dennell

Introduction Our thinking about the Early Palaeolithic of the “Far East” – the region containing China, Korea, Japan and Southeast Asia  – is still framed in terms of the “Movius Line”, whereby the Early Palaeolithic inhabitants of East and Southeast Asia supposedly retained a simple, Mode 1 flake and core lithic technology until the Late Pleistocene and in some regions, even the Early Holocene. In contrast, those hominins west of the Movius Line, in Africa, most of Europe, Southwest Asia and India, developed an Acheulean Mode 2 lithic industry in which handaxes were prominent, along with cleavers in Africa, Southwest Asia and India (Movius 1948). Thereafter, these same inhabitants later developed prepared-core Middle Palaeolithic or Middle Stone Age (Mode 3) assemblages, and later still, Upper Palaeolithic or Late Stone Age (Mode 4) blade-based assemblages. Because of these differences, early Palaeolithic societies west of the Movius Line have often been portrayed as “dynamic”, unlike those to the east, which have been envisaged as unchanging in their technology and deeply conservative. Since Movius published his synthesis of the Old World early Palaeolithic in 1948, a substantial literature has accumulated on the Movius Line. Some authors have argued that the presence of (a few) bifaces in China and Korea invalidate the concept entirely (Yi & Clark 1983; Gamble & Marshall 2001); others have argued that their occasional presence in East Asia indicates the need to envisage a “Movius Line sensu lato” (Lycett & Bae 2010; Norton et al. 2006), and several have proposed reasons why a bifacial, Acheulean technology was rarely utilised in East Asia (Pope & Keates 1994). Suggestions have ranged from a reliance on bamboo (Pope 1989; Watanabe 1985) to demographic factors (Schick 1994; Lycett & Norton 2010). It is hard to name any other paper that has been discussed and cited so frequently in Palaeolithic archaeology more than 60 years after publication as Movius’s paper of 1948. Likewise, probably no figure has been reproduced in textbooks and articles on Asian palaeoanthropology as often as Movius’s (1948) map showing the demarcation of bifaces and non-biface assemblages in the Lower Palaeolithic (Figure 2.1). At the risk of generalisation, most who have written on the “Movius Line” have tended to regard it as a concept that is basically sound, and then attempted to explain or modify it. Here, I suggest it is useful to examine the origins of the concept, and assess whether the ideas underlying 8

East Asia and Human Evolution

Figure  2.1.  Movius’s interpretation of the early Palaeolithic world (Movius 1948, map 4; reproduced with kind permission from the American Philosophical Society).

it can be regarded as valid. In order to do this, we need to examine three basic ideas: that humanity in the “Far East” was very ancient, that it was thereafter very conservative, and that it was “primitive” relative to contemporaneous developments further west. I suggest it is also useful to assess the origins of the Movius Line in relation to wider perceptions of the “Far East” by Western investigators prior to World War II.

The East as Ancient Although Darwin (1871) tentatively suggested that humans originated in Africa, most physical anthropologists by the end of the 1930s thought East Asia was a more likely place of origin. This was because almost all the relevant fossil hominin specimens came from Asia: from Trinil (found in 1891), Modjokerto (1935) and Sangiran (1937–1941) in Java, and in China, Choukoutien (now Zhoukoudian), where numerous fossils attributed to Sinanthropus pekinensis (now H. erectus) and artefacts were found from 1925 to 1937. Other important specimens came from the European peninsula in western Eurasia, notably Mauer (1907) and Steinheim (1933) in Germany, and in 9

Robin Dennell Britain, Piltdown (1912), which was not unmasked as a hoax until 1953, and Swanscombe (1935– 1936). Africa in 1939 had no fossil evidence for the earliest phases of human evolution, with the obvious exceptions of Dart’s discovery of Australopithecus africanus in South Africa in 1925, and Broom’s discoveries of A. transvaalensis at Sterkfontein (1936–1939) and Paranthropus robustus at Kromdraii (1938); at the time, most researchers (particularly in Britain) regarded these as ape-like, and not directly or even remotely relevant to human evolution. The remains from Broken Hill (Kabwe), found in 1921, were regarded as an African Neanderthal or early H. sapiens, and Leakey’s discovery of hominin specimens at Kanam and Kanjera (Kenya) in 1932 elicited a damning degree of scepticism over their age and provenance (Boswell 1935; Kent 1942). A second point favouring an Asian origin for humanity was the strong appeal of biogeography, particularly as developed by Matthew (1915) and Black (1925), who argued that the cradle of mammalian and human evolution lay in the invigorating realm of East Asia because of the uplift of the Tibetan Plateau: as it became higher, drier, less forested and more seasonal, primitive forms of animals either became extinct or dispersed to marginal areas in Southeast Asia and Africa, and more successful types took their place. Finally, racial prejudice played its part in favouring Asia over Africa, as eminent British palaeoanthropologists such as Sir Arthur Keith and Sir Eliot Grafton Smith refused to countenance the idea of a black ancestry for Europeans and preferred instead a Eurasian one (see Dennell 2001). Despite the overwhelming amount of fossil skeletal evidence favouring an Asian origin for humankind prior to 1939, some argued on archaeological grounds that humanity was more ancient in Africa than Asia. According to the chronological frameworks of the time, which were based on correlations of geological deposits to four major glaciations in Europe and a similar parallel sequence of pluvials in Africa, the oldest artefacts were thought to be the Kafuan industry from Uganda. This was considered to be older than the earliest Oldowan industry, which was then dated to a warm period before the first Alpine (Günz) glaciation, along with the oldest “preChellean” artefacts from Europe and those from Choukoutien, China (see, e.g., Leakey 1934, 126). However, the evidence that the Kafuan was earlier than the Oldowan was weak, and in 1939 it was still possible to argue that tool making was no older in Africa than in Europe or Asia. (The Kafuan industry was effectively shown to be an assemblage of geofacts by Desmond Clark [1958], by which time the primacy of Africa was well established.)

The East as Conservative The notion that the Palaeolithic inhabitants of the “Far East” were ancient but thereafter extremely conservative in their lithic technology (and perhaps other aspects of their behaviour) stems primarily from the shared fieldwork experiences of four researchers in Java in the summer of 1938. These were the German American geologist Helmut de Terra (1900–1981), the French palaeontologist and Jesuit Teilhard de Chardin (1881–1955), the palaeontologist G. H. R. von Koenigswald (1902–1982) and the American archaeologist Hallam Movius (1907–1987). All four met in Java following the expedition of Terra, Chardin and Movius to the Upper Irrawaddy Valley in autumn 1937 and spring 1938. Between them that summer, they shaped discussion and debate about the early Palaeolithic of East Asia for the following 70 years. In terms of its intellectual influence on subsequent generations, the fieldwork in Burma was the most significant piece of Palaeolithic research in East Asia before World War II. The role of each should be briefly summarised.

Helmut de Terra Helmut de Terra was one of the giants of Pleistocene studies of Asia in the 1930s. In 1927–1928, he had been on an expedition to Central Asia and later studied the geology of Chinese Turkestan (now Xinjiang), Tibet and the eastern Himalayas. He surveyed in Kashmir in 1932 with the 10

East Asia and Human Evolution British archaeologists Jaquetta and Christopher Hawkes (Hawkes et al. 1934). In 1935 he directed an enormously influential survey in the Soan Valley, Kashmir and the Narmada Valley of (then) British India, with the English archaeologist Thomas Paterson (1909–1994) (Terra 1937; Terra & Paterson 1939) and, for part of that survey, Teilhard de Chardin (Terra & Chardin 1936), whom he first met in 1933 and was a long-term colleague and close friend. In 1937–1938,Terra, Chardin and Movius conducted a four-month survey of the Irrawaddy Valley, Burma (now Myanmar) (Terra & Movius 1943), which led directly to Movius’s (1948) synthesis of the East Asian early Palaeolithic. After World War II, he moved to Mexico, where he made many notable discoveries before retiring to Switzerland. Terra’s geological fieldwork is a prime example of the broad-brush approach that characterised much Pleistocene fieldwork in the interwar years. In the years before palaeomagnetism and absolute dating techniques, the main geological window into the Pleistocene were large river systems, since these hopefully contained terrace sequences that could be arranged in a relative order and correlated with each other on the supposition that each terrace responded to a worldwide change in sea level. In areas adjacent to mountains, moraines and outwash deposits from glaciers could also (in theory) be tied to glacial phases, as in the European Alps. Terra’s work in the Soan Valley in British India and the Irrawaddy in Burma could be matched by comparable investigations along the Thames and Somme in Northern Europe (Breuil & Kozlowski 1931), the Nile in Egypt (Sandford & Arkell 1933), the Vaal in South Africa (Söhnge et al. 1937) and the Kagera in Uganda (Wayland 1934). To this can be added Leakey’s pre–World War II work at Olduvai (Leakey 1951 [but completed in 1938]) in which he attempted to correlate the Olduvai sequence to African pluvial phases that could in turn be linked to European glaciations; and Caton-Thompson and Gardner’s (1934) survey in the Fayyum Basin, Egypt, which also attempted to identify major dry and moist periods that could be correlated with the Alpine sequence. The bedrock of all this work was Penck and Bruckner’s (1909) and Eberl’s (1930) classic studies of the river sequences of the northern Alps and southern Germany that led to the recognition of four major glacial phases of Günz, Mindel, Riss and Würm, in order of decreasing age. In Terra’s fieldwork in North India and Burma, a major horizon that appeared to link many of these sequences comprised large gravel deposits (the so-called Boulder Conglomerate) that were linked to the second Himalayan glaciation, which Terra correlated with the Mindel glaciation of the Alpine sequence. This formed the baseline for his regional stratigraphic frameworks, and for his correlations with the Alps. He argued that hominins first appeared in India during the second glacial phase, as rolled, derived “pre-Soan” tools were found in the gravels of the Boulder Conglomerate. What he thought were younger terraces contained similar artefacts that were less rolled, and often had more retouch: the degree of abrasion and rolling of artefacts could thus be used as an indicator of relative age. Paterson (Terra & Paterson 1939) showed that the lithic sequence in the Soan Valley showed little change from Pre- through Early to Late Soanian, just like the Clactonian sequence he studied at Barnham St. Gregory, England (Paterson 1937; see Dennell and Hurcombe 1993). In contrast, handaxes from river terrace deposits in northern France appeared to show “progressive evolution” in size, proportions and amount of retouch: Breuil and Kozlowski (1931) demonstrated 11 stages in the Somme sequence, and Leakey (1951) produced the same 11-stage sequence for Olduvai Gorge (Dennell 1990). In the 1937–1938 expedition to Burma, Terra (1938; 1943a) based his approach on his previous work in India and established a similar Pleistocene framework for the Irrawaddy that was also based on the supposed succession of four river terraces (Terra 1943a, 330). These were linked to comparable sequences in the Soan Valley but also to sections as far removed as the Pyrenees and Java (see Figure 2.2). According to Terra, the oldest terrace along the Irrawaddy was formed in the second Himalayan glaciation, which he correlated with the Boulder Conglomerate of the Soan Valley and the Mindel glaciation in the Alps.The supposed presence of artefacts in deposits in and below what may have been the oldest Irrawaddy terrace indicated the presence of hominins at a very early date, as early as in India. As similar Palaeolithic tools were found in (but usually on the 11

Robin Dennell

Figure  2.2. Terra’s long-distance geological correlations for dating Early Palaeolithic artefacts and climatic periods (Terra 1943a, Fig. 54). Note the four-fold Alpine sequence of glaciations, and the massive Middle Pleistocene erosion layer that formed his main marker horizon from the Pyrenees to China (reproduced with kind permission from the American Philosophical Society). surfaces of) younger terraces along the Irrawaddy, it was concluded that the inhabitants had been very conservative in their tool making.

Teilhard de Chardin In 1938 Chardin was in his late 50s and at the height of his status as an authority on the palaeontology and the Palaeolithic of East Asia. Before World War I, he had worked at Piltdown in 1913, which most British experts prior to 1953 regarded as the oldest hominin fossil known. In 1923– 1924 he and the French missionary Émile Licent explored the Nihewan Basin and Ordos Desert in North China and discovered Shuidonggou, which is still the most famous Upper Palaeolithic site in North China. In 1928–1929, he joined a French expedition to Somalia and Djibouti and then participated in expeditions to the Gobi Desert in 1931 and to Chinese Turkestan in 1932 (Aczel 2007). In the1930s, he was heavily involved in the excavations at Choukoutien but also collaborated with Terra in Kashmir in 1935, and later that year he also visited Java, where he met Koenigswald. He was in regular correspondence with Henri Breuil (who visited him in China in 1935), the leading Palaeolithic expert in the first half of the 20th century, and had a wide range of contacts with other Palaeolithic researchers. By 1939, his knowledge of the Pleistocene and Palaeolithic of East, South and Southeast Asia was unrivalled. It is entirely fitting that Chardin is still honoured in China (along with Licent, Pei Wen-Zhong and Jia Lan-po) as one of the founding figures of studies of the Chinese Palaeolithic. As the senior-most participant in the 1937–1938 expedition to Burma and the subsequent trip to Java, Chardin’s opinions on the Asian Palaeolithic would have carried enormous weight with Movius even though Chardin was not listed as an author of the ensuing monograph.

Hallam Movius In 1937–1938, Movius was the youngest and most junior of the expedition to Burma, and at the start of his post-doctoral career. Although he is probably best known for the “Movius Line”, it was his only early Palaeolithic fieldwork. His first experience of European prehistory was during a summer school in 1931. In 1932 he excavated at the cave of Mugharet es-Skhu¯ l in (then) 12

East Asia and Human Evolution British Palestine and then participated in Harvard’s Irish Survey from 1932 to 1936, during which he excavated six mainly Mesolithic sites, which became the basis of his PhD, awarded in 1937. In 1937 he was a research associate at the Peabody Museum and was delegated to join de Terra’s expedition as Yale had also provided some of the finances. He and his wife spent a maximum of 10 weeks in the Irrawaddy Valley surveying for Palaeolithic material and 4 weeks in a different area looking for Neolithic evidence. Movius’s participation in the 1937–1938 expedition to Burma was also the last occasion when he was the junior-most partner. After World War II, his energies were directed towards the French Upper Palaeolithic, first at La Columbière in 1948, and then at the Abri Pataud, where he directed an enormously successful campaign from 1958 to 1964 that is still one of the anchors of the French Upper Palaeolithic sequence (Bricker 2007). The 1937–1938 expedition to the Irrawaddy resulted in the discovery of around 600 pieces that were regarded as artefacts from 16 locations. This material was grouped into a “culture” or “technocomplex” known as Anyathian, after a local term for Upper Burma. One peculiarity of the Irrawaddy material is that many artefacts were made from fossilised wood, which was locally abundant; the remainder were made from local stone. Almost all artefacts were found on the surfaces of exposures of fluvial gravels, which were in turn linked to Terra’s sequence of river terraces. As in the case of Paterson in the Soan Valley, the degree of rolling and abrasion was taken as an indicator of age, so the earliest claimed tools were invariably rolled, and the youngest were fresh. On the basis of this assumption, Movius (1943) recognised a five-stage sequence of a threestage Early Anyathian, and a two-stage Late Anyathian (for reasons unexplained, there was no “Middle Anyathian”). However, the assumption that abrasion was a reliable indicator of relative age was already questionable, as Oakley and (Mary) Leakey (1937) had discussed how the degree of abrasion of artefacts in fluvial deposits at Clacton, England, could simply indicate the distance they had been transported, rather than their age relative to fresher-looking artefacts. As Movius was well aware, the most striking result of the Irrawaddy survey was what was not found: unlike in, for example,Western Europe, peninsular India and Palestine, handaxes, prepared cores and Upper Palaeolithic blades were absent. The same observation seemed true of the Soan Valley, where Paterson had identified a “Soanian” flake and core assemblage that passed through four stages, each marked by greater “refinement”, that is, a gradual reduction in size and increased amount of retouch. (Handaxes were occasionally found in the Soan Valley sequence but explained as the result of incursions by handaxe users into an otherwise undisturbed cultural milieu of flake and core usage [Terra and Paterson 1939].)

G. H. R. von Koenigswald Koenigswald was appointed in 1930 as a stratigrapher and palaeontologist to the Geological Service of Netherlands India (present-day Indonesia). In 1931 he participated in the excavation of Ngandong on the Solo River and was also involved in the discovery of the infant cranium at Mojokerto in 1935. Additionally, he recognised the giant ape Gigantopithecus from teeth in Chinese pharmacies (as fossil bones and teeth were supposed to have great healing powers). In 1935 he met both Chardin and Terra in Java and showed them the main fossil localities. In 1937, on the recommendation of Chardin, he became a research associate of the Carnegie Foundation and started work at Sangiran in Central Java. The following year, Weidenreich visited him in Java to examine Koenigswald’s fossil specimens and, in 1939, invited him to study the Zhoukoudian Sinanthropus specimens in Beijing (Wolpoff & Caspari 1997a, 188), after which they published a joint paper (Koenigswald et al. 1939) on the Javanese and Chinese hominins. By the time of his internment by the Japanese in 1942, Koenigswald had become one of the most important figures in East Asian palaeoanthropology (see Frenzen 1984). His main contribution to the archaeological discussions that took place with Chardin and Movius in 1938 was over the artefacts he had collected in 1935, when he (Koenigswald 1936) collected flaked stones from the hilltop surface of a 13

Robin Dennell conglomerate at Kampong Ngebung, in the Sangiran area, and attributed them to Pithecanthropus (see Koenigswald & Ghosh 1973). Later that year, he also collected artefacts along a dry water course near Pajitan on the south coast of Java, as well as from a boulder conglomerate in the riverbank (Koenigswald 1936). As the conglomerate was tilted, he thought it had to be at least as old as the strata at Trinil. The artefacts he collected included some he classed as Chellean handaxes; because of their crude appearance, he argued that the conglomerate from which they came must also have been very ancient. (He was not the only one at the time to employ that type of circular argument: Terra did exactly the same in both the Punjab and the Irrawaddy Valley.) Terra (1943b, 456–457) attributed the Sangiran artefacts to the Upper Pleistocene, as he thought them too advanced for Pithecanthropus, and the Pacitan artefacts to the late Middle Pleistocene.

Assessment In Movius’s synthesis of 1948, the basic structure of the Palaeolithic in East Asia seemed clear. According to Terra’s chronological framework, there were simple flake and core assemblages in the Soan Valley of North India and the Irrawaddy Valley of Burma dating from the Mindel Glaciation of the Middle Pleistocene. At Choukoutien broadly similar ones were associated with a Middle Pleistocene fauna, and in Java the earliest artefacts from Pacitan were dated to the later Middle Pleistocene. Although not known at first hand, Collings (1938) had reported a simple flake and core assemblage at Kota Tampan in the Malay Peninsula, although there was no clear evidence that it was Middle Pleistocene in age (Movius 1948, 403). Finally, a few cores and flakes had been collected (under unbelievably harsh conditions) near Bhan-Kao in northern Thailand by Heekeren (1948) whilst a Japanese prisoner of war on the Burma railway; these came from what may have been Pleistocene terrace gravels and were elevated by Heekeren into a “Fingnoian” variant of the chopper-chopping tool complex of East Asia. All this evidence from India, Burma, Malaya, Thailand, Java and North China was incorporated into Movius’s (1948) scheme of East Asian “chopper-chopping tool” industries. The over-riding weakness of this model was the lack of evidence that any of the cited artefacts from India, Burma and mainland and island Southeast Asia were Middle Pleistocene in age (and thus contemporary with Acheulean assemblages further west). This is particularly true of Terra’s chronological framework for the Soan Valley of India, the Irrawaddy Valley of Burma and the Pacitan River of Java. In the Soan Valley, all that can be said about the dating of most of the localities examined by Terra is that they overlie folded late Miocene strata and are under late Pleistocene loess (Rendell et al. 1989). In the Irrawaddy Valley, the alleged “terrace” deposits are younger than tilted Irrawaddy Beds containing a Siwalik fauna (Colbert 1943), but there is no demonstrable correlation of the highest terrace that forms the basis of his sequence with the second (or any other) Himalayan glaciation. In neither the Soan nor the Irrawaddy Valley was there any demonstration that the artefacts were associated with specific Middle Pleistocene terraces. Elsewhere in mainland Southeast Asia, there is still virtually no evidence of Middle Pleistocene occupation. In northern Thailand, there are many undated or poorly dated archaeological sites, but the only dated archaeological evidence are three flakes from under a basalt dated to 0.73 Ma (Pope et al. 1986) at Mae Tha, Lampang Province in northern Thailand, that are questionable on stratigraphic grounds (Marwick 2009, 54). There are also some cranial fragments from Had Pu Dai, Thailand (Subhavan 2009), that may be considerably more or less than the claimed date of 500 ka (see Dennell, Chapter 4, this volume), and one tooth classified as Homo sp. from the late Middle Pleistocene cave of Thum Wiman Nakin (Tougard et al. 1998), but otherwise there is no clear evidence of occupation before the Upper Pleistocene. From northern Vietnam, evidence is limited to perhaps one hominin tooth (Schwartz et  al. 1994; 1995) from the cave of Tham Khuyen, circa 475 ± 125 ka (Ciochon et al. 1996) and attributed to H. erectus, and two teeth of “archaic” Homo sp. from the cave of Ma U’Oi (Demeter et al. 2004; 2005) that are late Middle to 14

East Asia and Human Evolution Late Pleistocene in age. In Malaysia, Kota Tampan is now regarded as Upper Pleistocene, with a likely age of circa 30 ka (Zuraina Majid & Tjia 1998; Reynolds 1993), and the site of Bukit Jawa is dated to 50–100 ka (Zuraina Majid 1997). Likewise in Java there is no evidence that the artefacts collected by Koenigswald from Pacitan and Sangiran in 1936–1938 are Middle Pleistocene in age. Several decades later, Bartstra (1983; 1985) argued that the these were probably derived from much later contexts. He also showed that the only true stone tools in the Sangiran area came from very recent gravel veneers capping the hills (Bartstra 1985); as for the artefacts from Pacitan, they were probably Late Pleistocene, and very like Hoabinhian assemblages on mainland Southeast Asia (Bartstra 1983, 429; 1984). Even today, remarkably little archaeological evidence from Java can be safely assigned to the Middle Pleistocene. The most convincing is a small assemblage of circa 20 pieces (two cores, three blades and fifteen flakes) from a Middle Pleistocene sequence of five stratigraphically distinct sets at Ngebung (Sémah et al. 1992; Simanjuntak & Sémah 1996), although the assemblage is too small to allow a confident assertion that handaxes were never used in Middle Pleistocene Java. Two artefacts were also found in context at Sambungmachan in a layer variously dated to the Middle or Late Pleistocene (see Keates 1998, 183). Sixty years after Movius’s 1948 publication, the only well-dated Early Palaeolithic assemblages from Southeast Asia are those from Flores (which has its own unique record of H. floresiensis) (Morwood et al. 2004) and the late Middle Pleistocene material from Panxian Dadong, southern China (Miller-Antonio et al. 2004).

The East as “Primitive” and “Conservative” The notion that early humanity in the “East” was more primitive, retarded and conservative than its counterparts in “the West” stems from the writings of Chardin and Movius in the 1940s. Chardin (1941, 60) suggested that “in contrast with the already ‘steaming’West, Early Pleistocene Eastern Asia seems to have represented . . . a quiet and conservative corner amidst the fast advancing human world”, by which he meant regions where handaxes were used. This view is developed in Movius’s (1948, 411) conclusions on the Early Palaeolithic of southern and eastern Asia: it “cannot be considered in any sense as ‘progressive’ from a cultural point of view”; the tools are “relatively monotonous and unimaginative assemblages of choppers, chopping tools, and handadzes . . . as early as Lower Palaeolithic times, southern and eastern Asia was a region of cultural retardation . . . it seems very unlikely that this vast area could ever have played a vital and dynamic role in early human evolution. . . . Very primitive forms of Early Man apparently persisted there long after types at a comparable stage of physical evolution had become extinct elsewhere”. Why did Chardin and Movius equate backwardness or the primitive with an absence of handaxes? Put another way, why did the use of handaxes make an early Palaeolithic population “progressive”, “steaming” or “fast advancing”? To answer these, we need to return to the knowledge base of Palaeolithic studies before World War II.

Handaxes: Why the Big Deal? One reason why hominins with Acheulean assemblages were seen by Western researchers as more “advanced” than those who did not use them may have been that they were first identified in Western Europe (then the most technologically advanced part of the world) and later recognised in areas ruled by Britain or France in Africa, India and Southwest Asia. Another likely reason is that Acheulean handaxes are the most distinctive and best-known artefact type in the entire Palaeolithic, and their importance was perhaps exaggerated because they were found preferentially by early collectors relative to other simpler tools. Palaeolithic researchers also value them because they are unambiguous examples of hominin agency, hence their importance in arguments in favour of the antiquity of humankind by John Frere at Hoxne (1797), Boucher de Perthes and 15

Robin Dennell Prestwich, Falconer and Evans in the Somme Valley in 1859 and Bruce Foote in India in 1863. Handaxes are also aesthetically pleasing in their symmetry, and many are also expertly made. The best ones are indeed examples of how craftsmanship and art can co-exist in one piece. These reasons may explain some of the attraction that Acheulean bifaces have had for archaeologists, but do not explain why they have been seen as evidence of dynamism or progress. One reason may be that they had an obvious appeal to an industrialised society that mass-produced goods, many of which were complex and required high levels of skill. Handaxes could be envisaged as the Palaeolithic example of skilled mass production and contrasted with the simpler, expedient type of flakes that were used in areas such as the “Far East”, whose inhabitants were therefore less “dynamic” than those in Western Europe. Their absence in the Early Palaeolithic of East Asia thus reinforced the impression that the modern weaknesses of East and Southeast Asian societies were deep-rooted: “By the end of the century, European commentators were increasingly inclined to see the ‘stagnation’ of non-European societies as a hereditary condition. Cultural differences, whatever their origin, became ‘racial’ differences, and cultural habits the product of racial ‘instinct’” (Darwin 2007, 342). In a post-imperial context, matters are less clear-cut. Firstly, there is the issue of double standards: the absence of handaxes in East and Southeast Asia was seen as evidence of backwardness, but the same accusation was not levelled at Lower Palaeolithic societies that lacked them in Eastern Europe or Britain. There, the absence of handaxes was seen as more likely a response to environmental factors (see, e.g., McBurney 1950; Mithen 1993). Secondly, it is debateable whether hominins that used handaxes were more intelligent than those who did not, as these may have had the ability to make them but decided not to. Even among communities designated as “Acheulean”, some groups used them rarely, and quite possibly, many individuals in “Acheulean” groups never made one. Thirdly, one should question how a lack of handaxes rendered an early Palaeolithic society primitive or retarded. After all, the hunters at Schöningen (Germany) made superlative javelins, were able to kill large numbers of horses, and indeed, provide perhaps the best example of Middle Pleistocene group hunting of large animals, yet never used handaxes (see Thieme 1997). As another example, Australia was colonised by people with a remarkably simple stone technology (see Balme & Connor, this volume). Rather than being “stagnant”, the nonhandaxe using societies of East Asia (whether Middle Pleistocene or later) might simply have been highly efficient in using a simple stone technology in a wide range of environments (see Dennell 2009, 436–437). To summarise, there is no justification  – contra Chardin and Movius  – for supposing that early Palaeolithic societies were “primitive”, “conservative” or examples of “cultural retardation” because they did not use handaxes.

Wider Perspectives The development of Palaeolithic archaeology and palaeoanthropology cannot be divorced from its social and political context (see, e.g., Dennell 1990, 2001; Díaz-Andreu 2007). First and foremost, before 1960 these subjects developed during an imperial era. Thus, when Chardin, Movius, Paterson, Koenigswald, Terra and others conducted their fieldwork in South and East Asia, they did so in regions that were either parts of empires ruled by the British (India, Burma and Malaya), the French (Indo-China) or the Dutch (Netherlands East Indies), or in China, in which Western powers exercised extra-territorial rights in 92 treaty ports (Holcombe 2011). We can also include here Japan’s pre-1945 colonies of Formosa, the Korean peninsula, Shandong in North China and Chinese Manchuria. Although we need to be careful when ascribing all pre-World War II research to the stereotype of a “grand imperial biogeographical tradition” as the practitioners were not automatons, often disagreed with each other and were rarely overtly political, we equally cannot envisage them as wholly impervious to the ideas of their time. 16

East Asia and Human Evolution

East and West: The Development of a World View One striking aspect of the development of the “advanced” European powers, notably Britain and France, in the 18th and 19th centuries was the rapid and profound application of science, initially within but increasingly beyond Europe. One important manifestation of this development was the systematic recording of the natural world, and particularly the cataloguing of life, whether plants, birds, animals, shells or insects. Another important aspect was the mapping of countries and coastlines, using the most advanced instruments and trigonometric techniques. Particular attention was paid to those parts deemed commercially, politically or strategically relevant to the nation-state. In the British case, the development of Ordnance Survey maps of Britain (1803), the Great Triangulation Survey of India (1850s onwards) and the mapping of the world’s coastlines by James Cook and others were notable achievements. The past was also mapped. In Europe, the prehistoric past was charted in terms of the Three-Age system of stone, bronze and iron, to which the further categories of Palaeolithic and Mesolithic were added after 1870, and the familiar categories of Lower, Middle and Upper Palaeolithic were in place by 1914. The historic past of Europe was also seen within a similar Three-Age system of ancient, medieval and modern history. Both schemas were seen as applicable to non-European societies. In a real sense, Europeans in their imperial eras created a global past, but one that was created in European terms: it served as a basis for their world view of how they saw themselves relative to the rest of the world.

The World View: Near, Middle and Far East Europeans also created a geography of Asia, with its perceived components of a Near or Middle East, and a Far East. These terms stem from the mid-19th century. The Near and Middle East were (and still are) often regarded as synonymous, but initially the Near East referred to the Ottoman lands of the East Mediterranean, and the Middle East, the Islamic lands further inland in Iraq, Iran and Central Asia. The “ancient Near East” was a convenient device for demarcating the land of and adjoining the Fertile Crescent of the Tigris and Euphrates, including the lands of interest to biblical Old Testament scholars (Palestine, the “Holy Land”). The “Far East” referred to the regions furthest in Asia from Western Europe and thus denoted China (including Formosa [Taiwan]), Japan, the Korean Peninsula and mainland Southeast Asia, comprising Malaya (now the main part of Malaysia), Siam (Thailand) and Indo-China (Vietnam, Cambodia, Laos), a term dating from 1810. It usually included the Philippines (then a Spanish and, after 1898, an American colony) and the Dutch East Indies (modern Indonesia), and sometimes the Russian Far East, that is, Siberia east of Lake Baikal. As a maritime power, the British tended to think of the Far East as the lands east of the Indian Ocean and west of the Pacific Ocean – thus the Far East comprised Southeast Asia, China, Japan and Korea – but excluded New Guinea and (white-dominated) lands such as Australia and New Zealand, even though these lay still further east. As defined, the “Far East” was a massive area with a latitudinal range from 50o N in Chinese Manchuria to10o S in Indonesia, equivalent to Edinburgh to Nairobi, or Vancouver to Lima. Today, the Eurocentric term “Far East” is usually replaced by East Asia (generally denoting China, Korea, Japan and the Philippines), and Southeast Asia (former Indo-China, Thailand, Myanmar and Indonesia). The “Far East” was not only distant from Europe but exotic. Even now, East and Southeast Asia seem exotic to European eyes, especially for the transient, first-time traveller who usually arrives by air, and can now confidently find a Starbucks or McDonalds in most large Asian cities, and see that most “natives” wear Western-style clothes, often speak (some) English, drive familiar makes of car, and inhabit familiar types of villas or high-rises. In the world before World War II, East Asia was incomparably more alien and exotic, with hardly any familiar cultural landmarks, and reached 17

Robin Dennell after several weeks of travelling (usually by ship) from Western Europe. How did Europeans view the Far East relative to their own European culture?

Western Perceptions of the Far East Far Eastern societies were usually regarded negatively, as exotic but inferior, backward and stagnant: these were societies that may have had a distinguished past but had let it slip: “It was all too easy to explain their cultures in terms of a stagnant past from which only Europe (or its developed regions) had escaped” (Darwin 2007, 314). Europeans “alone had broken out of the cycle of growth and decay to which all other civilisations were subject. They alone had discovered the secret of the wealth of nations.They had achieved an unequalled technological mastery.They had broken through the old barriers of superstition and myth to found their intellectual life on the rigorous collation of empirical knowledge” (Darwin 2007, 339–340). Thus, the perceived “backwardness” of societies in the “Far East” and others (e.g., Indian, African) helped reinforce the sense of superiority that Europeans felt about themselves: “It has been claimed that the invention of an oriental ‘other’, sunk in the quagmires of moral and intellectual ‘backwardness’, was essential to the European self-image of progress. Only by insisting on the failings of the ‘Orient’ (in practice all non-western peoples) could the Europeans be sure of their own progressive identity” (Darwin 2007, 340; see also Said 1978). Although Darwin (2007, 340) adds that “this almost certainly exaggerates the intellectual interest that Europeans took in other parts of the world”, it nevertheless must have pervaded the interpretations of those who did bother. One who may have bothered was Chardin. By 1937, he had spent most of the previous 15 years in China during the long exile that was imposed upon him by the Vatican (Aczel 2000). During that time, he had travelled extensively throughout China, including its remotest, poorest and often most dangerous regions. He was also present during one of the worst periods in its recent history: after the Qing Empire collapsed in 1912, the first republic had faltered and the central government had lost effective control over much of the country by 1916. In the 1920s, huge swathes were ruled by war-lords, usually arbitrarily and often at war with each other. This situation was partly overcome by the Northern Expedition of 1928 when Chang Kai-Shek brought the eastern seaboard under control, but this move was complicated by the outbreak of civil war in 1929. Conditions in China worsened considerably after 1931 because of Japanese aggression in Manchuria and Japan’s invasion of northern China in 1937. There is little evidence that Chardin cared greatly about present-day conditions in China, but two of his comments indicate his general perception.The first is that Early Palaeolithic China was “a quiet and conservative corner” “on account of its marginal geographical position” (Chardin 1941, 60). China’s position may have been “marginal” to a homesick French Jesuit yearning to return to the “Far West” but was hardly so to the hundreds of millions who lived there. (The modern term for China, Zhongguo, can be translated as “central country”.) Nor is it easy to argue that its position made it “quiet and conservative”, given the preceding two millennia of its history. The second is his assertion that “East Asia gives the impression of having acted (just as historical China and in sharp contrast with the Mediterranean world) as an isolated and self-sufficient area, closed to any major human migratory wave” (1941, 86, 88; emphasis added). This European perception of China, as isolated, self-sufficient and sealed behind its Great Wall dates from the writings in the 16th to 17th century of the Jesuit Matteo Ricci (1552–1610), who was one of the first westerners allowed to reside in Beijing (Lovell 2006). It betrays, however, an alarming degree of ignorance about China’s dynastic history, which includes several periods when it was as cosmopolitan and expansionist as many Mediterranean states before 1800: “The idea of a changelessly static East Asia, at any rate, is a fantasy, sustained only by a lack of historical knowledge” (Holcombe 2011, 160). In seeing China as geographically marginal and historically static, Chardin can be accused of allowing his Eurocentric views of geography and history to prejudice his interpretations of the early Palaeolithic of East Asia. 18

East Asia and Human Evolution

Discussion: The Movius Line and the Advent of Homo sapiens What is remarkable about the research of both Terra and Movius is the longevity of their influence. However, this is not because their research was superior to comparable fieldwork in Africa or Europe in the 1930s, but because early Palaeolithic fieldwork in South and East Asia virtually ceased for geopolitical reasons after World War II. In Pakistan, Terra and Paterson’s fieldwork in 1935 was not re-examined until the 1980s (Rendell et al. 1989; Dennell & Rendell 1991; Dennell 1995); no significant fieldwork has taken place in Myanmar since 1938; in Java, systematic investigations were not resumed until the 1980s (e.g.,Vos 1985; Leinders et al. 1985; Sondaar 1984); and although research in China continued after 1949, it was isolated from the “Far West” until the 1970s. Although Terra and Movius were unsuccessful in demonstrating that the “monotonous and unimaginative assemblages” of Southeast Asia were contemporary with Middle Pleistocene Acheulean assemblages in India, western Asia, Europe and Africa, their conclusions continue to influence discussion of how the inhabitants of Southeast Asia eventually became behaviourally modern. It logically followed from Movius’s (1948, 411) assessment that Southeast Asia was an area of “cultural retardation” occupied by primitive types of humanity that modern behaviour had to be intrusive – just as, in recent times, the process of modernisation in the “Far East” was a direct result of contact with the West (Holcombe 2011, 160).The Movius Line thus provided the necessary background for a replacement model to explain the appearance of Homo sapiens in Southeast Asia in the Upper Pleistocene: H. sapiens replaced H. erectus in Southeast Asia, in the same way that it replaced Neanderthals in Europe. However, a different perspective is obtained if we disregard the basis on which the Movius Line was created for Southeast Asia. Because none of the assemblages from India, Burma, Malaysia, Thailand and Indonesia considered by Movius are demonstrably Middle Pleistocene in age, it follows that there is no “long persistence” of a simple lithic technology in Southeast Asia. Instead, almost all the Palaeolithic assemblages from Southeast Asia (including Indonesia) are Upper Pleistocene in age, and most are less than 50 ka old. On current evidence, most of Southeast Asia was uninhabited before this time, apart perhaps from northern Thailand (see Dennell, Chapter 4, this volume). Ciochon (2009) has suggested that the rainforest regions of mainland Southeast Asia were occupied by a Stegodon-Ailuropda fauna in which Homo erectus was absent, and Gigantopithecus and perhaps another “mystery ape” were the largest primates. If this was the case, H. sapiens in Southeast Asia may have colonised a vacant niche rather than replacing its indigenous (and primitive) inhabitants. Genetic studies of modern populations suggest that this process of colonisation may have begun circa 50–60 ka (Liu et al. 2006; Su et al. 1999; Oppenheimer, this volume), although earlier dispersals cannot be excluded (Dennell & Petraglia 2012). Some populations of H. erectus may also have dispersed southwards from China into Southeast Asia. Schepartz et al. (2000), for example, have suggested that upland regions of southern China were colonised in the late Middle Pleistocene, and this process may have continued into the Upper Pleistocene and extended into parts of mainland Southeast Asia, such as northern Thailand. Some H. erectus populations might even have interbred with H. sapiens: one interpretation of the morphology of the Zhirendong mandible from South China, perhaps dated to circa 125 ka, is that it may indicate interbreeding between the two species (Liu et al. 2010b). (The situation in Java seems different: because the deposits containing the Ngandong H. erectus assemblage now appear to be Middle Pleistocene in age [143–546 ka] [Indriati et al. 2011], and not Upper Pleistocene, as previously indicated [Swisher et al. 1996;Yokoyama et al. 2008], H. erectus was probably locally extinct before the arrival of H. sapiens). To summarise, the population history of Southeast Asia was likely much more complex than envisaged by Movius. If we disregard his synthesis of 1948, we can also disregard his notions that Southeast Asia was culturally retarded and inhabited by populations that were incapable of innovation or progress. Instead, we can better assess its later Palaeolithic, Upper Pleistocene, record 19

Robin Dennell as indicating a process of colonisation by groups that were among the first to adapt to tropical rainforests. Far from being marginal, the appearance of populations after 50 ka in Southeast Asia may mark a behavioural threshold in hominin adaptive behaviour. In the 21st century, we need to break away from the rationale that underpinned the Movius Line: it lacks legitimacy, has no credible foundations, and impedes our thinking both about the Middle Pleistocene settlement of East and Southeast Asia and about how Homo sapiens first appeared in these regions.

20

Chapter 3 “Rattling the Bones” The Changing Contribution of the Australian Archaeological Record to Ideas about Human Evolution

Sandra Bowdler

On the whole they appear to me to stand some few degrees higher in the scale of civilization than the Fuegians. Charles Darwin, The Voyage of the Beagle (1845)

Encountering the “Primitive” Europeans of an intellectually curious bent first encountered the Indigenous people of the Pacific and Indian Oceans in the 18th century, the Age of Enlightenment. During this period, explanations of the observable world moved from the biblical to the irreligious empirical – what was considered “scientific”. By the middle of the 19th century such ideas dominated in the academic institutions of Western Europe. One of the major events in the history of biological science was Darwin’s theory of evolution by natural selection, published by him in 1859 in On the origin of species. While the idea of evolution in itself – the idea that species were not immutable – was hardly new, Darwin’s formulation was accepted as a revelation and landmark in the history of biology. The first explorers in the Pacific and Indian Oceans, driven by scientific curiosity rather than, or as well as, dreams of riches or empire, brought intellectual currents of the day with them. One of the challenges of encountering the Indigenous peoples of the antipodes was how to explain them; why were they physically different from Europeans, why was their technology so apparently simple? In biblical terms, they could be rationalised as benighted heathens descended from the unfortunates driven from Eden and going in the wrong direction from the Land of Nod, or fleeing after the Flood, or the collapse of the Tower of Babel. For scholars of the Enlightenment, some other explanation was needed. The widespread 18th-century concept of the “noble savage” enabled Europeans to perceive Indigenous peoples through a more benign lens but did little to explain their difference.The first step in understanding was accurate recording, and to this end most voyages of exploration carried natural historians and dedicated draughtsmen (Smith 1985, 7, 9); thus, aristocratic scientist Joseph 21

Sandra Bowdler Banks with assistants and artists accompanied Cook on the voyage of the Endeavour in 1769–1771. This desire to capture what was considered to be objective information about little-known, “other”, humans was carried on until well into the 20th century. Back in Europe, anatomists began collecting human skeletal material, particularly cranial, just as their wandering colleagues were collecting data from living specimens. A vast collection of anatomical objects from humans and other animals was amassed by Scottish anatomist John Hunter (1728–1793), who tried to discern patterns amongst his specimens.1 By 1779, he had turned from the notion of immutable species and considered that similar species had descended from common ancestors, changes being brought about by environmental conditions (Moore 2005, 491), a very short step from Darwin’s theory of natural selection. Impressed by Hunter’s collection and his written work, German anatomist Johann Friedrich Blumenbach (1752–1840) did not advance the cause of evolution, but he did propose a model for human variation that was to have far-reaching consequences, the concept of “race”. In three editions of his MD thesis (first published in 1775), he proposed classifications of the human species (and he was of no doubt that humans constituted a distinct species) that, in the third edition, were distilled down to five races, Caucasians,2 Mongolians, Ethiopians, Americans and Malays. While this has had unwelcome ramifications to the present day, Bhopal (2007) defends Blumenbach from being himself a racist. By the end of the 18th century, intellectual scaffolding was in place for assessing human physical and cultural variation in a context of social evolutionism, and a tradition of empirical data gathering by observation and recording had been established. To this point such observations were not generally quantified, and assessments of primitivism could hardly be considered objective.There was no doubt however, despite some lingering ideas of the noble savage, that Australian Aborigines, and Tasmanian Aborigines in particular, were considered “primitive” and inferior to Europeans. All that was now needed was the social Darwinism engendered by the publication of On the origin of species to set these concepts within an acceptable theoretical framework.

Skullology During the 19th century a whole branch of enquiry sprang up which concentrated on establishing which of the varieties of humans were the most advanced, and which the most primitive. This whole concept is ahistorical, in that it assumes that certain varieties, or races, of humans were somehow arrested at an early stage of development, while others (inevitably European) somehow advanced to their full potential. While it was already clear to westerners who was advanced and who was primitive, the science of the 19th century demanded some form of objective quantifiable approach to ranking. What better than to measure skulls, the locus of intelligence? Craniometry became a dominant branch of 19th-century biological science, beginning well before Darwin, and including some major studies by collectors such as Samuel Morton (1799–1851) in the United States. In Europe, after 1859, skull measuring was refined and applied in the pursuit of a number of different aims, most of which are now considered to rest on dubious premises, to say the least. Events in 1859 included not only the publication of Darwin’s theory of evolution but also the establishment of the Anthropological Society of Paris by Paul Broca (1824–1880) (Gould 1996, 115). This led to skull measuring on a grand scale, with further elaboration of measurements, but was still designed to show the same thing: the evolutionary superiority of white races over darker and especially black ones. Broca spawned a considerable school, his best-known disciple being Paul Topinard (1830–1911). The popularity of craniometry in Europe and North America led to the craze for skull and bone collecting which resulted in the substantial collections from the Antipodes in the museums of the Northern Hemisphere as well as Australia, and in such episodes as the stealing of William Lanné’s head in Hobart in 1868 (Petrow 1997). 22

Contribution of the Australian Archaeological Record The observations of 18th-century explorers provided 19th-century evolutionists with people to put even lower on the evolutionary ladder than Africans. Not only were they more primitive in appearance (i.e., less attractive to Europeans); they also had a dismally inferior material culture to anyone else on earth except Tierra del Fuegians and Andaman Islanders.The Tasmanians with their pitiful few possessions were at the very bottom of the heap. During the 19th century, Indigenous skulls travelled from Australia into the collections of Western Europe to reinforce these views.

Studies of Australian and Tasmanian Skulls As the 19th century drew on, Darwinian evolution tended to hold the floor, and with respect to the human species it was accepted that “primitive races” like those of Australia and Tasmania provided insights into human ancestors. One of the first published studies of Australian and Tasmanian skulls was in Richard Owen’s (1804–1892) Descriptive catalogue of the osteological series contained in the Museum of the Royal College of Surgeons, published in 1853. In this massive work – and here we are concerned only with volume II, Mammalia placentalia  – Owen described the remains of species from acouchis to zebras, including everything from musk voles to domestic cats, culminating in Man [sic]. He covers remains from around the globe, finally tailing off into foetuses, children, teeth, oddments and oddities. It is clear that the aim is to describe, not to analyse or draw conclusions, but such may be inferred. Gorilla and chimpanzee skeletons are described in particular detail so as to facilitate comparison with humans (Owen 1853, 802) – some anyway, as the only comparison is actually with a “male negro”, referred to as a “low race” (Owen 1853, 785–786).The overall conclusion is that humans are a unique species with no closer relations with the “brute-kind” than those common to all placental mammals (Owen 1853, 802). In discussing Australian and Tasmanian specimens, he has no qualms in spelling out his views. Having made the observation above about the unique nature of Man, an Australian female skeleton is described as a member of the lowest race of the Melanian [dark-brown or black] variety: such, e.g., as the well-known modifications of the pelvis and pelvic extremities for maintaining the erect posture, and the specific distinctions of the skull, as described in the comparison of that of the Gorilla with the skull of the Negro.The inferior characters of the Melanian as compared with other races of Mankind are illustrated in the present skeleton by the narrowness of the cranium, the prominence of the alveolar parts of the jaws, the flatness of the nose-bones, and the recession of the chin. The squamosal unites with the frontal on both sides of the cranium, as in the Chimpanzee. (Owen 1853, 805) As for the skull of a male Australian, in the narrow form of the cranium, the low receding forehead, the prominent obtuse borders of the orbits, the prominence of the jaws, and advanced position of the canines of the lower jaw, . . . presents, irrespective of any artificial distortion, the lowest character of any Human skull in the Museum; but in all the essentials it adheres to the Human type and departs from that of the Gorilla and Chimpanzee. (Owen 1853, 823) Owen did not at this time believe in evolution (Owen 1853, 802), yet while he considered humans to be unique, some varieties of humans were considered to be “lower” than others, particularly of the dark-brown to black variety, and one Aboriginal skull had the “lowest character” in the collection. A few years later, a similar study by surgeon George Williamson (1857) was carried out on crania in the Museum of the Army Medical Department in Fort Pitt, Chatham (Kent, UK). This 23

Sandra Bowdler collection included ten Australian and two Tasmanian skulls. Owen described his specimens anatomically, but no metric data were used in his descriptions.Williamson did provide some metrical information of post-cranial material, but with respect to skulls, he simply enumerated characteristics (Williamson 1857, 84, 78). The real number crunching began after Darwin and tended to coincide with the decline of the Indigenous populations under study in terms of population and agency in the face of increased government control.3 Broca and particularly Topinard led the way with elaborate studies of skulls including examples from Australia and Tasmania. Gould has argued that Broca used his measurements to support his preconceptions that black races were inferior to whites: he “did not fudge numbers; he merely selected among them or interpreted his way around them to favoured conclusions” (Gould 1996, 110). Topinard however was not quite so determined to cling to preconceived ideas and, in his study of eight Tasmanian Aboriginal skulls, concluded that, rather than being at the bottom of the evolutionary ladder, they did in fact fall between Polynesians and Aboriginal Australians (Topinard 1868, 325). He based his comparison on measurements that illustrated “très-dolichocéphalique” (very long-headed) and “très-prognathe” (very protruding lower face) characteristics along with non-continuous variables, some of which have been used to characterise Tasmanian skulls to this day, such as rounded cranium, keeled vault and parietal bossing (Topinard 1868, 825). Topinard (1872, esp. 326) also incorporated non-skeletal data in his studies of Tasmanians and Australians, such as observations about hair and skin colour – the very characteristics that led to their being classed as “primitive” to start with. It is hard to credit a totally objective interpretation when he considers Tasmanian skulls to be “farouche et sinistre” (fierce and sinister) (Topinard 1868, 314). In England, physician Joseph Barnard Davis (1801–1881) amassed his own collection of skulls, the largest in the world at the time: 1,474 specimens by 1867 (MacDonald 2005, 96). He was not a proponent of Darwinian evolution; rather he favoured the principle of polygenism, whereby all human races are distinct and of distinct origins. Separate does not, in his case, mean equal. In his 1874 paper, he compared Tasmanians to Australians, not limiting himself to skeletal characteristics but, like Topinard, including soft tissue and cultural items in his analysis. In one respect he differentiates himself from his European colleagues: he finds Tasmanians to have had a larger cranial capacity than Australians, and he argues that Topinard’s figures show the same, even if “given in different terms” (Barnard Davis 1874, 15). This puzzles him greatly but is not to be denied. He is puzzled by the Tasmanians’ poverty of material culture compared to the Australians but cannot avoid concluding that their cranial capacity reflects the “intellectual and moral superiority of the Tasmanian” (Barnard Davis 1874, 15). He concludes nonetheless that they were “a peculiar and distinct race of people dwelling in their own island and different from all others”. Barnard Davis’s contemporary William Flower (1831–1899), conservator of the Hunterian Museum from 1861 to 1884 and then director of the Natural History Museum, took a different position. He was perfectly and unusually explicit about his preconceptions. One of his reasons for discussing the Australian “race” was that a “comparison of its characteristics with those of the best known race (that to which we ourselves belong and which is commonly taken as the standard in works of human anatomy) will afford a good idea both of the kind and degree of variation to be met with between one of the lowest and one of the highest groups of mankind” (Flower 1878, 603). Flower was far more a lumper than a splitter, and classified the Tasmanians with Australians and Papuans as “Melanesian” (Flower 1878, 629). A flow of similar studies continued, and increasingly more so in Australia, with measurers not only raiding the collections of museums for candidates for the callipers but also receiving specimens dug from graves or fresh from the bodies of Aborigines murdered by settlers or native police (Turnbull 1991; 1994, 12–13). The level of excitement when new specimens turned up, particularly from Tasmania, can be gauged by the introduction to Berry and Robertson’s paper for the Royal Society of Victoria. 24

Contribution of the Australian Archaeological Record In the whole annals of the scientific Tasmanian literature there has never yet been recorded in a single communication such a large number of Tasmanian crania as we have the privilege to lay before the Royal Society of Victoria tonight. (Berry & Robertson 1909, 47) Many of these papers provide a plethora of measurements with little interpretation at all beyond what was gleaned from the non-continuous and often non-skeletal characteristics (e.g., Duckworth 1894, 1902;Turner 1908, 1910; Büchner 1912a, 1912b, 1914; Hrdlic˘ka 1928; Krogman 1932; Morant 1927, 1939; Wunderly 1939). Work concentrating on non-metric characteristics led to no conclusions at all (Fenner 1939; Kellock & Parsons 1970). Papers of nothing but Tasmanian skull measurements were published explicitly to provide data for “our masters, the savants of the Old World” (Harper & Clarke 1898, 97). Arguments arose at the turn of the century as to which measurements were “useless” (Berry & Robertson 1909, 53). Berry and Robertson (1910, 41–42) quote Cunningham, fulminating, “What is the value of this glabellocerebral index of Schwalbe? Can we rely upon it giving a true and proper idea of the relative extent and degree of projection of the pars glabellaris of the cranium?” Should we care? We are not told. In fact, there is little to no discussion of actual methodologies in most of these cases, in terms of discussing the measurements used and what they are meant to convey, and that continued well into the 20th century. Berry and Robertson are something of an exception, both as to making some attempt to discuss their methodology (e.g., 1910, 43) and in maintaining some sense of scientific method, with hypotheses evaluated against their results. It is never quite clear, however, as to the significance of the measurements in terms of the aims of the exercise, and clearly this was also not always clear to the scientists involved (e.g., Berry & Robertson 1910, 47). In the latter case, the title of their two-part paper indicates its aims: “The place in nature of the Tasmanian Aboriginal as deduced from a study of his calvarium. Pt. I – His relations to the anthropoid apes, Pithecanthropus, Homo primigenius, Homo fossilis, and Homo sapiens; Pt. II – His Relation to the Australian Aboriginal” (Berry & Robertson 1910; 1914) – that is, to put Tasmanian Aborigines somewhere on the evolutionary scale on the basis of head measurements. “Of recent man the Tasmanian stands nearest to Homo fossilis [Cro-Magnon, anatomically modern humans from the later Ice Age] but morphologically has progressed a very long way from Homo primigenius [Neanderthal and Pithecanthropus] and the anthropoid apes – very much farther than most writers would seem to believe” (Berry & Robertson 1910, 67). In Australian scientific circles considerable discussion was centred on the differences or otherwise between mainland Aboriginal people and those of Tasmania, a question not resolved to everyone’s satisfaction until the latter part of the 20th century, despite a frenzy of cranial measurements. Berry and Robertson (1914) conclude that Tasmanian and mainland Aborigines were descended from common stock but that the Australians are a “hybrid” race (a view shared by Topinard [1872]); perhaps surprisingly, they also conclude that “the modern-day Australian aboriginal stands rather nearer the anthropoid ape, or the common ancestor, than did the Tasmanian. Isolated individuals of the Australian race have, on the other hand, surpassed the Tasmanian” (Berry & Robertson 1914, 186) – rather a puzzle for the evolutionists, one would think. The notion of hybridity, or “race purity”, seems odd given our modern understanding of human variation, but it certainly exercised the minds of evolutionists in the early 20th century (Berry et al. 1910; Büchner 1912a), as it was thought to bear on the issue of whence Aboriginal ancestors had come.

Australia: Fossil and Archaeological Records Globally, the first fossil of an extinct hominin recognised was the skull from the valley of the River Neander, or Neanderthal, discovered in 1856 but not officially recognised as a potential human precursor until its naming in1863. Not until the end of the 19th century did a coherent 25

Sandra Bowdler story begin to emerge from the fossil record, alongside a new science, archaeology, which offered the means to construct an equally coherent story about the development of human culture. The discovery of the fossils we now call Homo erectus (Pithecanthropus, Java man) in Java by Eugene Dubois (1858–1940) in 1891 led to the real beginning of a meaningful hominin fossil record, with subsequent finds of Homo erectus (Sinanthropus, Peking Man) at Zhoukoudian (Choukoutien) in China between 1927 and 1929, and the Australopithecines of South Africa first identified by Raymond Dart in 1924. Given those finds, it was possible to recognise a logical progression from small-brained but bipedal species with increasing manual skills towards a larger-brained fully cultured one, in a succession of Australopithecines through Homo erectus to Homo sapiens. Australia was slow in producing a fossil and archaeological record similar to that of Europe at the turn of the 19th century. The usual prejudices contributed to this state of affairs. When antiquarians turned their attention to stone artefacts, the aspect of Aboriginal culture with the potential for archaeological survival, they were unimpressed. Where were the carefully worked and easily classified fossiles directeurs of the Old World Palaeolithic sequences? In the last decade of the 19th century, eminent anthropologist Edward Burnet Tylor (1832–1917) turned his attention to stone artefacts from Tasmania and concluded that “these belong, notwithstanding their modern date, to the order of the very ancient ‘paleolithic’ implements of the Drift and Cave Periods [Upper Pleistocene]” (Tylor 1894, 147). This idea found full expression in Sollas’s (1911) book, Ancient hunters and their modern representatives which, just as the title promised, compared modern hunter gatherers with their ancient analogues (or vice versa). Tasmanians were depicted as exemplars of Lower Palaeolithic “folk”, and the somewhat more culturally advanced mainland Aborigines were illustrative of life in the Middle Palaeolithic (Mousterian) (Sollas 1911, 69, 169). What point was there in looking for a coherent archaeological sequence to document a history of these “unchanging people living in an unchanging environment” (Pulleine 1929, 310)? Etheridge had asked as early as 1890, “Has man a geological history in Australia?” It was many years before anyone could provide an answer. Enter anthropologist and archaeologist (and erstwhile entomologist) Norman B.Tindale (1900– 1993) of the South Australian Museum. Paying no heed to Pulleine’s dictum, Tindale set about looking for archaeological sites, excavating them and creating sequences. He was not blind to the potential for discovering skeletal remains in this way, either. In 1927 he excavated a small rock shelter on the Murray River in South Australia named Devon Downs. In so doing, he demonstrated that Australia did have an archaeological record, showing cultural change over time (Hale and Tindale 1930). One of the problems at that time was that there was no independent way of dating archaeological sites in a text-free context, and not until the 1950s did radiocarbon dating become available. Tindale, however, was able to establish a “geological antiquity”: he identified stone tools on Kangaroo Island, which had been unoccupied when first visited by a European, Matthew Flinders, in 1802. Tindale argued that the site must have been occupied before it was severed from the mainland by post-glacial sea level rise, that is, during the Pleistocene (Tindale 1937; 1957). He constructed an overall sequence with the Kartan culture of Kangaroo Island as the oldest, dating to the late Pleistocene, followed by a Tartanga culture, a Pirrian culture, a Mudukian culture and a Murundian culture. While this made Australia suddenly look a lot more like the rest of the world in terms of cultural change over time, it must be said that most of these “cultures” had few clear criteria and were conceptually underdeveloped. Tindale (1957) did however show that there was a definite regional change, with the larger Kartan cobble tools being oldest, and with a range of smaller tools including microliths appearing at a later time.This basic scheme is still of relevance to Australian archaeology, if not quite as Tindale had conceived it (e.g., Bowdler 1993). At some point in an American sojourn (1936–1937), Tindale chummed up with physical anthropologist Joseph Birdsell (1908–1994), for what became a lifetime professional partnership. Together they embarked on a series of trips around Australia to various Aboriginal communities under the joint aegis of the universities of Harvard and Adelaide, making particularly significant 26

Contribution of the Australian Archaeological Record visits to the Bass Strait islands at one end of the country and the Queensland rainforest on the other. Emerging from this was Birdsell’s “trihybrid” theory of Aboriginal origins. This concept was instantly attractive to scientists who had to this time struggled with making sense of the quantities of data that had been, and continued to be, generated without any kind of conclusions being reached, other than the primitiveness of the subjects. Birdsell (1949; 1967) posited three waves of Aboriginal people colonising Australia, an earliest “negrito” group named Barrinean, which survived in pockets of the Queensland rainforest and in a mixed form in Tasmania; a “Murrayan” group, which came and swamped the negritos and became the more or less norm of Australian Aboriginal people; and a late “Carpentarian” group, evident in northern Australia. In many ways, the whole package is a throwback to the early 20th century (and to be fair the idea was developed by Birdsell between 1937 and 1940), with its concern for racial “purity” and belief that measuring contemporary crania would be informative about the evolutionary background of the subjects. The idea of the Tasmanians as “negritos” (or partly so) manages to put them back at the bottom of the evolutionary ladder. Birdsell’s primary scientific opponent was Andrew Arthur Abbie, who argued that the Aboriginal population was essentially homogeneous; his work was equally based on studies of contemporary peoples (Abbie 1951; 1969). If Australia was to make any contribution to the wider story of human evolution other than the constant arguments about “racial purity” and endless measuring of modern skulls, it was going to have to come up with some fossils, that is, human remains of some consequential time depth. In the rest of the world, the early 20th century was a time of significant discoveries of human ancestral forms not only in Europe and Africa but, closer to home, in Java and China. Where had humans originated? Could Australia contribute to this unfolding story? The Chinese Homo erectus fossils from Zhoukoudian, initially called Sinanthropus pekinensis, were subjected to detailed study by German anatomist Franz Weidenreich (1873–1948), who had been based in Beijing up to 1941, when the Japanese occupation made his situation, and that of the fossils, precarious. He oversaw the excavation of some of these specimens in the 1930s and also had access to the Javanese Homo erectus (Pithecanthropus erectus) fossils for comparative study. In 1943, he published the substantial monograph, The skull of Sinanthropus pekinensis. Weidenreich was a great and meticulous scientist, and it is unfortunate that his views were misrepresented and distorted into a form of scientific racism that persisted almost to the present day, and which in some cases continues. Weidenreich argued that Sinanthropus and Pithecanthropus were at the same general stage of evolution (Weidenreich 1943, 248), which occupied “the lowest place of all the hominid forms so far known” (Weidenreich 1943, 214).4 He also examined another group of remains from Java, generally known as Solo Man, or the Ngandong fossils. These he named Homo soloensis, and considered them to be related to but more phylogenetically evolved than Pithecanthropus (Weidenreich 1943, 274). He went on to discuss the implications of the different hominins, as follows: Homo sapiens is a collective name for numerous racial groups which have certain general features in common. The possession of these features proves that all these groups belong to the same advanced stage of human evolution but does not mean that all came from one form with “Neanderthalian” qualities. As Pithecanthropus and Sinanthropus represent two groups of the same early stage, and as the Neanderthal and Late Palaeolithic mankind consists of similarly differentiated groups, so, it can be assumed, the various racial groups of modern mankind took their origin from ancestors already differentiated in the same manner. The Australian aborigines of today, regarded by some authors as relics of the “European” late Palaeolithic man, driven from Europe to far-off Australia, are, in one sense, rather autochthons of South East Asia. There is an almost continuous line leading from Pithecanthropus through Homo soloensis and fossil Australian forms5 to certain modern primitive Australian types. (Weidenreich 1943, 276) 27

Sandra Bowdler In the same way, Weidenreich (1943, 277–276) saw “clear evidences” for Sinanthropus as “direct ancestor of Homo sapiens with closer relation to certain Mongolian groups [i.e., East Asians] than to any other races”. Thus the different races of modern humans derived at a very early date from identifiable groups of ancestral hominins in different places. Again emphasising his point about Australian Aborigines, he comments that “the Australian aborigine as an offspring of the Pithcanthropus line is not a relic of European Late Palaeolithics, marooned on the Australian continent, but is a late branch of the human stem passing now through the same evolutionary stage as that through which the Europeans have passed during the Late Pleistocene” (Weidenreich 1943, 277). Not a relic, then, but a very very late developer. There were, and still are, considerable difficulties with this proposition. If Weidenreich believed that Sinanthropus and Pithecanthropus were different species, and he continued referring to them under those rubrics until his death, it is biologically incorrect to have a single species evolve from two different ones. The subsuming of both under the single species name of Homo erectus did not occur until 1951. If the two types are in fact a single species, it is still problematic to have another species arise in two different places. We shall return to this point. Weidenreich’s multilineal model for human evolution might have remained safely quarantined in the scientific literature, but it was exported to the world at large by anthropologist Carleton Coon (1904–1981) in a volume The origin of races (1962) which was hugely popular, if reviled by Coon’s peers even as it was published. He took Weidenreich’s work and reinterpreted it to show that there were five human races (as per Blumenbach), which had each evolved in five different places, with some being more “successful” than others. Not surprisingly, the most successful were Europeans, ahead of Africans, and least of all the Australian Aborigines (Coon 1962). By the 1960s, retarded or not, a decent archaeological time depth had been demonstrated for Aboriginal Australia, with radiocarbon dating clearly demonstrating a Pleistocene antiquity (Mulvaney 1964; Mulvaney & Joyce 1965). Mulvaney had also confirmed Tindale’s findings that there was cultural change in time in the Australian archaeological record, with smaller stone tools appearing after the end of the Pleistocene. Rhys Jones’s (1966; 1968; 1971) work in Tasmania showed time depth there also, and subtle changes in the archaeological record; there were no small stone tools of the later mainland kind, however.The lack of developed formal patterning in the earlier strata of stone tools found in early deposits of Australian sites, and until recent times in Tasmania, may have shown that Australia had an interesting archaeological record but did little to dispel notions of primitiveness.

Fossils at Last During the 1960s, skull measuring continued in Australia (Freedman 1964; Larnach & Macintosh 1966, 1970; Macintosh & Barker 1965). No clear-cut conclusions emerged, but Sydney anatomist Norman Macintosh at least recognised the need for properly dated human remains of some antiquity, especially for Tasmania (Macintosh & Barker 1965, 55). By 1943, three skulls able to be called “fossils” (i.e., with some claim to antiquity) were known. Talgai and Cohuna had been known for some time and were incorporated in Weidenreich’s 1943 publication. The third, Keilor, was announced and described in a special edition of the Memoirs of the National Museum (Melbourne) that same year (Mahony 1943a, 1943b;Wunderly 1943). None was without problems of one sort or another. The Talgai cranium had been found in Queensland in a flood channel in 1884, somewhat early for any proper archaeologist to be on the job (Smith 1918). Not only is its provenance shaky; the skull itself had been broken, was thickly encrusted with calcified deposit and was of a juvenile to boot. That did not stop its being invoked in the debate about Piltdown, making it the first Australian fossil on the international stage (Hammond 1988; Allen 2010). Cohuna was in better condition, but its provenance was not much better: ploughed up by a farmer in the Murray River basin in 1925 (Macintosh 1965, 50). 28

Contribution of the Australian Archaeological Record Keilor similarly had a pick driven into it by a quarry worker on a river terrace on the outskirts of Melbourne in 1940 (Macintosh 1965, 41). While Cohuna and Talgai had already received notice by Weidenreich and others as being of a primitive nature, the publication of the 1943 volume prompted a review by Ashley Montagu (1944), who saw none of the Australian fossils as being outside the range of modern humans. Macintosh in 1965 was able to add a few more examples, offer some dating suggestions, and draw some tentative conclusions; the influence of Weidenreich is evident. With regard to the three examples discussed, the dating of Keilor (the best-provenanced specimen) seemed to be anything from 3,000 to 18,000 bp (Macintosh 1965, 44–45). Talgai is suggested to be 11,000 years old, and the nearest best guess for Cohuna is mid-Holocene (Macintosh 1965, 50–51). The additional specimens are the Mossgiel skeleton, excavated by spade by graziers from a claypan in Western New South Wales (Macintosh 1965, 51); the Tartanga remains, found originally in 1928 but incorporating some excavated by Tindale from Devon Downs (Macintosh 1965, 52); and some skull fragments from Aitape in the west Sepik area of Papua New Guinea (Macintosh 1965, 55). Of these, only the Tartangan remains have some form of associated dating, between circa 6,000 and 4,000 bp. Reviewing these remains, and extending the caveat that all needed better morphological description and examination, Macintosh was nevertheless moved to remark that “the mark of ancient Java is on all of them”, going on to add that that “can be seen in modern Aboriginal crania too” (Macintosh 1965, 59). Just two years later, in the wake of a further discovery, Macintosh was able to refine his ideas. The Green Gully remains were found not far from the Keilor find spot in 1965, in the usual way, by a contractor, this time with a front-end loader. While some damage was sustained, the site was at least able to be examined expeditiously by more or less professional archaeologists, and the remains were in reasonable condition for examination. A radiocarbon date of circa 6,500 bp was obtained. Macintosh, comparing these remains with the fossils previously described, was able to recognise in this admittedly small series two morphological types. On the one hand, he grouped together Cohuna, Talgai and Mossgiel and, on the other, Green Gully, Keilor and Tartanga. The first group was characterised by overall robusticity and a larger size than the other group, with a markedly receding frontal vault, marked post-orbital constriction, obvious prognathism and other distinguishing characteristics of similar nature. The second group had a full curved frontal bone, minor post-orbital constriction and orthognathism and was generally smaller and less robust.With acknowledgment to Weidenreich, Macintosh suggested that the first group “exhibits individual and collective traits which, . . . subjectively are reminiscent of characters in Homo erectus and Solo man” (Macintosh 1967, 97). Rather than see Australians as “surviving Homo erectus or Solo Man” however, he suggested that Homo erectus should be brought nearer to reclassification as Homo sapiens. In 1970 geomorphologist Jim Bowler came across human remains exposed in a lunette dune at Lake Mungo in western New South Wales, not far from Mossgiel, and further investigations were carried out with archaeologists Harry Allen and Rhys Jones and physical anthropologist Alan Thorne. In due course, two Pleistocene burials, a cremation and an interment, were uncovered with next to impeccable stratigraphic and dating contexts. The Lake Mungo site with its archaeological as well as fossil record was originally dated to between 26,000 and 32,000 bp; recent research has suggested that Lake Mungo III (the interment) dates to circa 40,000 bp (Bowler et al. 1970; Bowler & Thorne 1976; Bowler et al. 2003). In the early 1970s, following astute detective work by Alan Thorne, the location of the Cohuna cranium in the Murray River area was revisited, and the extensive burial ground of Kow Swamp was found and excavated. Some 22 individuals were uncovered (Thorne 1975, 95–96), although not all were suitable for detailed analysis. Thorne concentrated on nine skulls, to produce data for comparison with Lake Mungo and the other known fossils. In interpreting his data, he followed Macintosh in distinguishing two groups: Kow Swamp aligned with Cohuna, Mossgiel and Talgai to form an “archaic” robust group, while the Mungo remains fell in with the Keilor–Green 29

Sandra Bowdler Gully gracile group (Thorne 1975, 269). Before dates were available, it was expected by many that the Kow Swamp remains might be of a considerable antiquity, and definitely older than those from Lake Mungo. Indeed the reverse was the case; they are dated between 13,000 and 9,000 bp (Thorne 1976, 96). Here was a problem: Why was an “archaic” group younger than a “modern” one? Thorne proposed several models to account for what he perceived to be two very distinct morphologies: either they represented two ends of a range or they were two evolutionarily distinct forms. Like Macintosh, he turned to Weidenreich, and indeed Coon, to explain the difference. Lake Mungo and the gracile group were on the Mongoloid line, Kow Swamp and the robust group were on the Australoid line. Aware of the taxonomic and phylogenetic difficulties, Thorne (1977a, 200) suggested that Homo erectus from Asia at least be referred to Homo sapiens. This still left much unexplained: Why would Mongoloids be turning up in Australia, presumably from China, earlier than Australoids from Java? Is the robust group a survivor from an earlier time somehow marooned in the Murray Valley, not exactly an isolated place? Shortly thereafter, Thorne threw in his lot with American physical anthropologist Milford Wolpoff, and other scholars, notably Wu Xinzhe from China, in a series of papers proclaiming the multiregional or multilineal theory of human evolution: Homo erectus in different parts of the world evolved into different human races. Chinese Homo erectus evolved into Mongoloids, Javanese Homo erectus into Australoids, European Homo erectus through Neanderthal into Caucasoids, African Homo erectus (or ergaster) into negroids. They got round the small difficulty of a species evolving into another species in different places by invoking gene flow between the ancestral groups (e.g., Thorne & Wolpoff 1981), a rather problematic concept given the sparse hominin populations usually imagined for the Pleistocene. There is still a problem here: Are modern Australoids purely descended from Javanese forbears? If so, what happened to the Mongoloids at Lake Mungo? If not, does that mean modern Australian Aborigines are a mix of Mongoloids and Australoids? Doesn’t that rather dilute the model? While one would not want to impute a less than purely scientific spirit of inquiry to this model, it nevertheless leaves the impression that Aborigines were somehow descended from a more primitive group than other races. The mark of ancient Java was still weighing heavily; Australia’s contribution to an understanding of human evolution seemed, once again, to serve as an illustration of the primitive. Alternative views came from a variety of quarters. Habgood (1986) used cranial data from early anatomically modern humans from various parts of the world, but Australian data from both the gracile and the robust groups, for a multivariate statistical analysis. He concluded that the early Australian fossils fell within a morphological continuum ranging from the gracile to the more robust crania. Within this great morphological variation the crania do display an “Australianness” which is unique to them. That is, the “gracile” and “robust” groups are more similar to each other, overall, than they are to any other early anatomically modern Homo sapiens crania from around the world. (Habgood 1986, 136) Perhaps the most startling, and at the time controversial, was DNA evidence, suggesting a quite different story from that of the multiregionalists. Cann et al. (1987) announced that their research on mitochondrial DNA from living populations suggested that modern humans had a single origin, in Africa about 200,000 years ago from Homo ergaster, and spread out to establish themselves in the rest of the world, replacing any other H. erectus–derived populations that were already there. This was seen by many archaeologists and human palaeontologists to fit with their interpretation of the archaeological and fossil records (e.g., Stringer & Andrews 1988). In Australia, the research of physical anthropologist Peter Brown also provided alternative explanations. He argued that the unusual appearance of the frontal bones, and especially their very low recession index, could be explained by artificial cranial deformation such as some kind 30

Contribution of the Australian Archaeological Record of headbinding (Brown 1981); this continues to be a strongly fought argument (e.g., Curnoe & Thorne 2006). Brown however, working from a collection of skulls from Coobool Creek, undated but from a similar location to Kow Swamp, does find a significant difference between terminal Pleistocene populations from the Murray valley and other, more gracile populations (Brown 1989, 175). Hence he sees the robust morphology as being regionally restricted and looks for adaptive rather than wider human evolutionary explanations. Stephen Webb (1989) similarly studied a collection of human remains from the Willandra Lakes, which include Lake Mungo, some of which were dated. He also notes the range of morphologies in Pleistocene human populations, in this case seeing an “extraordinarily gracile morphology” amongst his study sample (Webb 1989, 80). While research has come a long way, we still to some extent seem to be stuck with head measuring leading to uncertain conclusions; but at least the measured heads are now largely specimens assumed to have some antiquity. More recent researchers such as Darren Curnoe are seeking alternative techniques and approaches to explain Australia’s past human morphologies. He suggests ways to look for explanations that are not phylogenetic, or even adaptive, but depend on epigenetic and ontogenetic explanations, suggesting that robusticity might result from “phenotypic plasticity alone” (Curnoe 2011). In other words, sometimes genes do not entirely dictate what individuals look like (an example might be Russell [1985] on brow ridges). This is an encouraging approach to understanding the past of Aboriginal people. If we accept the out-of-Africa theory, and the DNA evidence continues in the main to do so, we should consider firstly that Australia and Tasmania were settled by modern humans, Homo sapiens sapiens, and the observed variation all falls within that species. No serious scholar has suggested that any Australian or Tasmanian human fossil represents any other species.With respect to the Tasmanians in particular, Pardoe (1991b) carried out a study of non-metric characteristics on a collection of Tasmanian skulls and a collection of skulls from southeast Australia. He expressed surprise that, after at least 10,000 years of separation, they were so similar. We know Australians arrived some 50,000 years ago, and of course there is no reason why there was not continuous immigration after that. There is no reason not to suspect that a range of morphologies was represented, but all within the species Homo sapiens sapiens. Surely the time has come to stop the endless skull measuring in search of evolutionary evidence. This is not to say the ethical study of human skeletal remains should stop, since there are important aspects such as demography and pathology that can be learnt thereby. The main contribution to be made by Australian archaeology and physical anthropology to ideas about evolution is to demonstrate the range of variation, cultural and adaptive, and, yes, physical, of the single species Homo sapiens sapiens in colonising a new continent.

Acknowledgments Thanks are due to Jane Balme and Chilla Bulbeck for their helpful comments and suggestions. Notes 1 Kept in the Hunterian Museum of The Royal College of Surgeons, London, not to be confused with the Hunterian Museum of Glasgow, which housed the collection of his brother William Hunter, also an anatomist. 2 If anyone is wondering whence came the term “Caucasian” for white people from the Old World, it was coined by Blumenbach on the basis of his perception of inhabitants of the Caucasus region of (then) Russia as the most beautiful people in the world. 3 Topinard himself opened his study of Tasmanian Aborigines thus: “Les journaux nous ont appris que le dernier des Tasmaniens est mort il y a cinq ou six mois et que, de ces insulaires, au nombre de sept mille

31

Sandra Bowdler lors de la découverte de l’île de Van-Diémen, is le reste plus aujourd’hui qu’une femme; je crois meme qu’elle vient de succomber. Il m’a donc semblé que le moment était arrive d’étudier les quelques cranes de cette race rassemblés au Muséum de Paris, sans me préoccuper de ce qui a pu être écrit â ce sujet” (The newspapers have told us that the last of the (male) Tasmanians died about 5 or 6 months ago, and of the islanders who numbered 7,000 when Van Diemen’s land was discovered, only a woman remains today; I believe she may also have just died. It seems to me then that the time has come to study the few skulls of this race collected in the Museum of Paris, without concerning myself about what has been written on the subject) (Topinard 1868, 307). 4 Bearing in mind Australopithecines were not accepted as potential hominins until 1947 (Tobias 1992). 5 Talgai, Cohuna and a still undated skull from Java called Wadjak man.

32

Chapter 4 Smoke and Mirrors The Fossil Record for Homo sapiens between Arabia and Australia

Robin Dennell

Introduction A clear picture of when Homo sapiens first appeared across the southern rim of Asia will emerge only when there are reliably classified, well-dated and well-preserved fossil specimens from secure contexts more than 40 ka (the age of the cranium from Niah Cave, Borneo) from the countries between the Red Sea and Australia. Likewise, a clear picture of whether the appearance of our species across this region resulted from one or more replacement events or from local, in-situ evolution (or both) will become clear only when there are also well-dated, well-preserved and accurately described fossil specimens from the late Middle and early Upper Pleistocene before the appearance of our species. Sadly, neither condition can presently be met. Here, I summarise what is known about early H. sapiens and its immediate predecessors from each region of southern Asia prior to 30 ka, after which we can (provisionally) assume that our species was – with the exception of H. floresiensis on Flores – the only hominin taxon in southern Asia.The Australian skeletal record is also included, and because the fossil hominin record from mainland Southeast Asia is so poor, the survey is widened to include the Philippines and China, especially from south of the Qinling Mountains. Following this review, I then discuss how this evidence has been, or can be, interpreted.

The Fossil Hominin Record from Southern Asia, 250–30 ka Any discussion of the earliest evidence for Homo sapiens outside Africa has to include the Levant, with the proviso that because it lies on the northern edge of the region between Arabia and Australia, it does not necessarily indicate the processes that occurred further south and east.

The Levant The Israeli skeletal record for the Upper Pleistocene is unique in Asia in four respects. Firstly, many of the finds are from burials, and thus there are several examples of partial and complete skeletons, in contrast with the much more fragmented evidence from South China and Southeast 33

Robin Dennell Asia. Secondly, evidence of both Neanderthals and H. sapiens has been found in cave deposits with Middle Palaeolithic assemblages, although never at the same site (see Table 4.1). Thirdly, H. sapiens is found both before and after Neanderthals, unlike in Europe, the Zagros Mountains and Central Asia, where Neanderthals always precede H. sapiens. Fourthly, the Israeli evidence has been exceptionally well dated (see Table 4.2); indeed, because there are so many dates, it is often difficult to assign precise ages to finds, although the overall picture seems clear. Put briefly, there is a small amount of evidence that Neanderthals were present during Marine Isotope Stage (MIS) 6; and substantial evidence that H. sapiens was present during MIS 5 and MIS 4, from circa 125 to 75 ka, and then Neanderthals again from 75 ka to circa 45 ka, when H. sapiens re-appears and Neanderthals likely became extinct (Shea 2008). Only a small part of an extensive literature on the human remains from Israel is sampled here. Shea (2003) provides an excellent discussion of this evidence, and Schwartz and Tattersall (2003) provide detailed descriptions of key specimens. Evidence that Neanderthals were in the Levant before 125 ka comprises a mandible (Tabun II) from layer C in Tabun Cave.This layer has been dated to 165 ± 16 ka by thermoluminescence (TL), and slightly less by electron spin resonance (ESR) (early and late uptake models) and U-series dating (see Table 4.2). There is also an almost complete female skeleton (Tabun I) that was found in layer C but was probably intrusive from the overlying layer B (Bar-Yosef & Callander 1999), which has been dated by ESR (late uptake) to 122 ± 16 ka (see Table 4.2). This age estimate and others from layer B place the skeleton in MIS 5 or the end of MIS 6. Most experts regard both Tabun I and II as Neanderthal (e.g., Schwartz & Tattersall 2003, 384; Stefan & Trinkaus 1998), although Howell (1999, 217) placed the Tabun C mandible in the H. sapiens group of Skhuˉ l and Qafzeh. Homo sapiens is first evidenced in the Levant either at the end of MIS 6 or during MIS 5e (125–115 ka). The remains of seven adults (Skhuˉ l II–VI and IX) and three juveniles (Skhuˉ l I,VIII and X) were found in layer B of Skhuˉ l Cave (a few metres from Tabun Cave), and these have been dated to 119 ± 18 ka by TL. At the cave of Qafzeh, the remains of four adults and two juveniles were discovered in layer L in front of the cave in 1933–1935, and another two adults and five juveniles in units XV–XXII during excavations in 1965–1977 (Shea 2003, table 4.1). Depending on which suite of dates is preferred (Table 4.2), these remains may be the same age as, or slightly younger than, those from Skhuˉ l. Neanderthal remains younger than the evidence for H. sapiens from Skhuˉ l and Qafzeh are known from the caves of Kebara and Amud. At Kebara, a fragmented infant skeleton (K1) was found in layer F in 1964. In later excavations (1984–1991), this layer was sub-divided into Units VII–XII, and a partial skeleton of an adult was found in level XII, and numerous isolated teeth and bones in levels VII–XII. These have been dated to circa 55–65 ka. At Amud, the remains of two adults and four juveniles were found in layer B, for which several dates are available, mostly in the range of 50–70 ka (Table 4.2). The earliest post-Neanderthal example of H. sapiens in the Levant is the partial skeleton of “Egbert” (in fact, a young female, 7–9 years old) from layer XXVI at Ksar Akil, Lebanon, dated to 43.7 ± 1.5 by 14C and associated with an initial Upper Palaeolithic assemblage (Bergman & Stringer 1989; Mellars & Tixier 1989). This was found in 1938; a second skeleton, of a child, was lost in World War II. Several isolated teeth from the cave of Uçagizli in Southwest Turkey and dated by 14C as from 41.4 ±1.1 to 29.1–0.4 ka have been attributed to H. sapiens (Güleç et al. 2007) (although at least one is more Neanderthal-like [Kuhn et al. 2009, 108]). On this basis, H. sapiens re-appeared in the Levant by circa 45–42 ka (uncalibrated).

South Asia The only pre-sapiens fossil hominin from South Asia is part of a skull cap and left parietal from Hathnora in the Narmada Valley of Central India. Estimated cranial capacity is circa 1,200 cc 34

Table 4.1.  Dates for Levantine hominin skeletal remains Context

TL

35

H. sapiens 119 ± 18 Skhuˉ l B Skhuˉ l B Skhuˉ l B Qafzeh XVII–XXIII 92 ± 5 Qafzeh XV–XXI Qafzeh XIX Neanderthal Tabun B Tabun B Tabun C/Unit 1 165 ± 16 Tabun C Tabun C Kebara VII 51.9 ± 3.5 Kebara VII, Sq Q19 Kebara VIII 57.3 ± 4 Kebara IX 58.4 ± 4 Kebara X 61.6 ± 3.6 Kebara X Kebara XI 60 ± 3.5 Kebara XII 59.9 ± 3.5 Amud B1 57.6 ± 3.7 Amud B2 65.5 ± 3.5 Amud B4 68.5 ± 3.4 Amud B1/6–B1/7 Amud B2 Amud B4 Amud B1–B2 Post-Neanderthal H. sapiens KsarAkil XXVI KsarAkil XXVI KsarAkil XXXII KsarAkil XXXII

ESR EU

ESR LU

U-series

80.8 ± 12.6 59.7 ± 6.3

101 ± 17.9 76.7 ± 8.2 49.0

96 ± 13 104 ± 10.5

115 ± 15 120 ± 15

97.48

102 ± 17 76 ± 14

122 ± 16 85 ± 18

104 + 33, −18 50.69 + 0.23, −0.23

C

14

Comment

Source

Mean of 6 Mean of 7 Mean of 6 Mean of 20 Mean of 16 Mean of 2

Mercier et al. 1993, 172 Stringer et al. 1989, 757 McDermott et al. 1993, 254 Valladas et al. 1988, 615 Schwarcz et al. 1988, 735 McDermott et al. 1993, 254

Mean of 7

Grün & Stringer 2000, 602 McDermott et al. 1993, 254 Mercier & Valladas 2003 Grün & Stringer 2000, 602 McDermott et al. 1993, 254 Valladas et al. 1987, 159 Bar-Yosef et al. 1996, 301 Valladas et al. 1987, 159 Valladas et al. 1987, 159 Valladas et al. 1987, 159 Schwarcz et al. 1988, 657 Valladas et al. 1987, 159 Valladas et al. 1987, 159 Valladas et al. 1999, 265 Valladas et al. 1999, 265 Valladas et al. 1999, 265 Rink et al. 2001b, 713–714 Rink et al. 2001b, 713–714 Rink et al. 2001b, 713–714

Mean of 7 Mean of 8 Mean of 3

120 ± 16 140 ± 21 135 + 60/−30 117.6 ± 29.3 127 ± 34.3 > 44.8 (14C)

60.4 ± 8.5

64.3 ± 9.2

Mean of 11

53 ± 7 61 ± 9 70 ± 11

47 ± 9 51 ± 4 49 ± 5

Mean of 6 Mean of 8 Mean of 5

61 ± 8.6

43.7 ± 1.5

Mean of 4 Mean of 2

Plicht et al. 1989 Mellars & Tixier 1989 Plicht et al. 1989 Plicht et al. 1989

Table 4.2. The fossil hominin evidence from southern Asia, 250–30 ka 36

Country

Locality

Specimen

Context

Age (ka)

Identification

Israel

Tabun

Tabun I: almost complete skeleton Tabun II: mandible molar and femur fragments

122 ± 16 ka if layer B (Grün & Stringer 2000) 165 ± 16 (Mercier & Valladas 2003) Ca. 300 ka 119 ± 18 (TL) 92 ± 5 (TL) 96 ± 13 (ESR EU) 115 ±15 (ESR LU)

Neanderthal Indicates presence (H. sapiens: Howell of Neanderthal 1999) in MIS 6 in the Levant; see also Table 4.1. H. sapiens See also Table 4.1. H. sapiens See also Table 4.1. H. sapiens All Neanderthal

See also Table 4.1.

(B1) 57.6 ± 3.7 (B2) 65.5 ± 3.5 (B4) 68.5 ± 3.4 47 ± 9 43.7 ± 1.5 (14C)

Neanderthal

See also Table 4.1.

Lebanon

7 adults, 3 juveniles Skhuˉ l B Qafzeh L 4 adults, 2 juveniles Qafzeh XV–XXII 2 adults, 5 juveniles, several isolated teeth Kebara F 1 juvenile skeleton Kebara VII–XII 1 partial adult skeleton, numerous isolated bones and teeth Amud B 2 adults, 2 juveniles 2 juveniles, numerous fragmentary remains Ksar Akil XXVI? “Egbert”

Layer C but probably intrusive from layer B Layer C Layer E

India

Hathnora

Right half of skull cap and part of left parietal

Layer F Layers VII–XII Layer B Layer XXVI

Fluvial conglomerate

> 48 ka; if reworked, 131 ± 5– 236 ka (Patnaik et al. 2009)

H. sapiens

Comments

Earliest postNeanderthal indication; see also Table 4.1. H. erectus (Sonakia Homo sp. indet. 1985; Sonakia & seems the most Lumley 2006); reasonable archaic H. sapiidentification ens (Kennedy et al. 1991) H. heidelbergensis (Cameron et al. 2004) Homo sp. indet. (Athreya 2007)

Sri Lanka

Fa Hien Batadomba lena

Remains of several Layer 4, cave individuals Fragmented human remains Layer 7c, cave

Beli lena Kitugala Fragmented human remains Layer 10, cave Thailand

Doi Ta Ka Thum Wikan Nakin Moh Khiew

Four cranial fragments One right upper premolar

Cave locality Cave locality

30,060 ± 380 BP

H. sapiens

28,510 +2150/−1710 BP

H. sapiens

Under sample dated to 24,520 +1500/−1270 BP ?500 ka > 169 ± 11 ka

H. sapiens

25,000 ± 600 BP

H. sapiens

> 46, < 63.6 ± 6 ka

H. sapiens

475 ± 125 ka BP

H. erectus ?Pongo sp.

Laos

Tam Pa Ling

Vietnam

Tham Khuyen

Partial skeleton, adult, prob- Layer 2, cave ably female Cranial and maxillae Cave fragments 5 teeth Cave fissure

Tham Hai

One tooth but likely Pongo Cave fissure

Lang Trang

Five teeth

Cave fissure

Likely same age range as Tham Khuyen 146 ± 2 ka to 480 ± 40 ka, but on faunal grounds, ca. 80 ka

Hang Hum Ma U’Oi Ngandong

Five teeth 2 teeth, 1 skull fragment 12 crania, two tibiae

Cave locality Cave fissure River terrace

Age uncertain 193 ± 17 ka 28–54 ka or 143–546 ka

Punung I or II Wadjak Niah

One premolar Cranium and femur Cranium

Cave locality Open locality Cave

Possibly 115–128 ka Holocene 45–39 ka cal BP

Indonesia (Java)

Malaysia (Borneo)

?H. erectus Homo sp.

Possible secondary burials Possible secondary burials Possible secondary burials Dating uncertain

Only one tooth definitely hominin

37

Only one tooth H. erectus or H. definitely sapiens, dependhuman ing on age estimate. Homo sp. indet. Homo sp. indet. H. erectus Older dates more likely than 28–54 ka ?H. sapiens H. sapiens H. sapiens Earliest unambiguous specimen of H. sapiens in southern Asia (Continued)

38

Table 4.2. (Continued) Country

Locality

Specimen

Philippines

Callao Cave Tabon Cave

3rd metatarsal Tibia fragment Rt. Mandibular fragment

China

Tianyuandong

Partial skeleton

Context

Age (ka)

Identification

Cave, layer 3

67 ka (Mijares et al. 2010) 47 +11/-10 ka 31 +8/-7 ka (Détroit et al. 2004) 40,328 ± 816 ka cal BP H. sapiens

Gaitou Partial cranium Ganjian Isolated teeth Tongtianyan Cave, Cranium Liujiang

Cave Cave Cave

39–44 ka H. sapiens Possibly ca. 100 ka ?H. sapiens H. sapiens Possibly 111–139 ka, but relation of date to skull is unknown

Bailiangdong

Two isolated teeth

Cave

Huanglong Cave

Seven isolated teeth

Cave

Zhirendong

Two molars and a mandible Cave

Ca. 30 ka, or possibly > 160 ka Age estimates range from 38–44 ka to ca. 104 ka Possibly 100–113 ka but could be younger

?H. sapiens ?H. sapiens ?H. sapiens

Comments

Earliest unambiguous specimen of H. sapiens in China Provenance of cranium within sequence remains unknown Major dating uncertainties Major dating uncertainties Relation of human remains to dated flowstones is unclear

The Fossil Record for Homo sapiens between Arabia and Australia (Sonakia 1985, 336)  or 1,155 to 1,421 cc (Kennedy et  al. 1991). Because these were found in cemented gravels and conglomerates directly by the Narmada River, they have been difficult to date, but most researchers assign them to the late Middle Pleistocene. U-series dating of a bovid scapula found near the cranium indicates a minimum age of 236 ka (Cameron et al. 2004, 419). A more recent assessment is that the minimum age of the hominin is circa 48 ± 1 ka, but if it was not re-worked, its age lies between 131 ± 5 and 236 ka (Patnaik et al. 2009). The associated fauna can be safely attributed to the late Middle Pleistocene, although Patnaik (2000) suggested that that some faunal remains have been re-worked. Taxonomic assessment of the cranium has been problematic. Recently, Athreya (2007) suggests that it is best regarded as an indeterminate “Middle Pleistocene Homo”. Given uncertainties over its dating, and the lack of any facial and dental features, this is probably the safest assessment of this find.

Sri Lanka The earliest skeletal evidence for H. sapiens in South Asia comes from three caves in Sri Lanka. This evidence does not indicate the earliest occupation of the island, as artefacts, albeit in disturbed contexts, have been found in coastal gravel and dune deposits belonging to the Iranamadu Formation and dated (by TL) to 64 ka and 74 ka (Deraniyagala 1992, 686; Kennedy 2000, 180). The hominins responsible are unknown. The most informative accounts of the Sri Lankan evidence are Kennedy (2000, 180–186), the two volumes by Deraniyagala (1992) and Perera et al. (2011). At Fa Hien, numerous human remains were found in Area B, at the rear of the cave in layer 4. The remains of a 5.5- to 6.5-year-old child (YF-86B-N7–3) were found, co-mingled with the remains of two infants (one < 1 year old), a juvenile and a young adult female. Secondary burial was indicated. Charcoal associated with the child was dated by 14C to 30,060 ± 380 BP (this date, and all other 14C dates from these caves, are based on a half-life of 5,730 years, and uncalibrated). Other human remains were dated to 24,470 ± 290 (YF-86/ B-M7–3), 32,060 ± 630 (YF-86/B-M7–5) and 33,070 ± 410 (YF-86/B-M6–6), but the number of individuals from these contexts is uncertain (Kennedy 2000, 181; Deraniyagala 1992, 696). In the basal layer (7c) of Batadomba lena (Perera et al. 2011), 17 geometric microliths, shell and bone beads and some fragmented human remains were found, dated to 28,510 +2150/−1710 BP. One accelerator mass spectrometry (AMS) (30,603 ± 400 BP, uncalibrated) and one conventional (27,700 + 2090/−1660 BP, uncalibrated) radiocarbon assay on charcoal from unit 7c date the onset of human presence to between 36,280 and 31,450 cal BP (Perera et al. 2011, 257).The remains included a thoracic vertebra covered in yellow pigment, a left parietal with traces of burning, a smaller parietal bone, a worn RM3, and a mandible that was probably from an adult female and burnt. As at Fa Hien, secondary burial practices are likely. The microliths are important because they have been seen as evidence for an immigrant, colonising population of H. sapiens from Africa (Mellars 2006b), although the recent discovery of earlier geometric microliths at Jawalapuram IX, central India, dated to 36 ka, have been interpreted as an indigenous development (Clarkson et al. 2009; Petraglia et al. 2009). Human remains were also recovered from Belilena Kitugala. In the tenth level from the top (= layer III from the base), remains of 12 individuals dated to 13 ka and microliths more than 27 ka (Deraniyagala 1992, 699–700) were found. Details of the human remains are sparse, but Kennedy (2000, 185) notes that the bones were less well preserved than at Batadomba lena. TL dates from this site broadly confirmed the radiocarbon dates (Deraniyagala 1992, 690).

Mainland Southeast Asia (Malaysian Peninsula, Thailand, Cambodia, Laos, Vietnam) This region, which covers circa 700,000 sq mi/1.5 million sq km (about the same size as Western Europe) “has surprisingly little to show for itself in the way of fossil human skeletal remains” 39

Robin Dennell (Marwick 2009, 51). This may be partly because most archaeological attention has been on later prehistoric and early historic monuments, but even when cave deposits have been investigated, fossil skeletal remains have so far been usually limited to only a few teeth or cranial fragments. This is often attributed to the activities of porcupines. These collect dry (degreased) bones to chew in order to wear down their incisors (as these grow throughout their life) to keep them a useable length (Brain 1981, 109), and they may also chew bone to ingest calcium (Bacon et al. 2008), or phosphorus (Brain 1981, 139). In Southeast Asian cave assemblages, often only the tooth crowns are preserved. At Lang Trang,Vietnam, for example, more than 10,000 tooth crowns, most with the roots gnawed, were preserved in breccias but little other fossil material (Vu The Long et al. 1996). However, Brain (1981, 109–117) reported that porcupines often collect more bones than they gnaw, and the proportion of unidentifiable bone fragments can actually be lower in porcupine-accumulated assemblages than in those left by hunter-gatherers.The intensity of bone gnawing thus depends upon the availability of dry bones.The extreme attrition observed at Lang Trang and other Southeast Asian caves may therefore indicate either a scarcity of bone in rainforest (Storm et al. 2005, 540) or the effects of calcium dissolution on bone in a humid karstic environment, or both. Another factor affecting bone survival in many of these karstic caves may also be the winnowing effect of stream action (Bacon et al. 2008). To date, the little evidence that has been found comes from northern districts of Thailand, Laos and Vietnam. Thailand One potentially important cave site is locality 1, Doi Ta Ka, near the village of Had Pu Dai, Lampang Province, northern Thailand. Here, four cranial fragments from the right frontal were found and identified as Homo erectus (Subhavan 2009). These were associated with Pongo, Gigantopithecus (one tooth each), sabre-toothed cats, hyaena, Ailuripoda melanoleuca, cervids and suids. The age of the hominin remains is stated as circa 500 ka on faunal grounds, but this estimate is little more than a guess as these taxa have a considerable time-range in South China (see Wang et al. 2006). Additionally, it is not clear whether the faunal assemblage is mixed or relates to a single phase. Further investigations are needed at this locality. Excavations at Thum Wikan Nakin (Snake Cave, “grotte de serpent”, Chaiyaphum Province) by a Thai-French team produced a late Middle Pleistocene (170–250 ka) fauna, characterised by the teeth of 31 species of large mammals and 30 species of small ones (Tougard et al. 1996). These included Pongo pygmaeus, Ailuripoda melanoleuca baconi, Crocuta crocuta ultima, and Ursus thibetanus. One human tooth (TF3467, RP4) was also recovered (Tougard et al. 1998). U-series dating of calcites produced an estimated minimum age of 169 ± 11 ka: U-series dates of fossil teeth proved unreliable because of uranium leakage, and this emphasises the point that in these complex, humid and water-rich contexts, calcites (travertines, speleothems) are more reliable as these are closed systems (Esposito et al. 2002). The tooth (TF3467, RP4) was similar to specimen MU18 from Mu O’oi in showing both erectus and sapiens features: although the crown shape looks like that of H. erectus, the complexity of fused root branches and occurrence of only one apex look more like that of H. sapiens (Marwick 2009; Tougard et al. 1998). A partial skeleton, comprising the calvarium, a nearly complete mandible, and most of the upper part of the body, was recovered from Moh Khiew Cave, Krabi Province.The individual was mature and probably female. It was found in a burial at the base of level 2, in a layer containing several flake and pebble tools and also debitage. An AMS charcoal sample from the burial was dated to 25,000 ± 600 BP (Matsumura & Pookajorn 2005). Laos Demeter et al. (2012) report an important discovery at the cave of Tam Pa Ling, of cranial remains, including the frontal, partial occipital, right parietal and temporal bones, and the right and left 40

The Fossil Record for Homo sapiens between Arabia and Australia maxillae with a largely complete dentition (right I2–M2 and left I2–M1). Dental evidence indicates that the individual was a young adult. The remains are described as representing an early population of H. sapiens in Southeast Asia. Dating by AMS 14C, TL and optically stimulated luminescence (OSL) showed that the remains were buried before 46 ka, and direct dating by uraniumthorium (U/Th) of the frontal bone produced an age estimate of 63.6 ± 6 ka. These remains are therefore older than the cranium from Niah Cave and are currently the oldest from mainland Southeast Asia and east of the Levant. Additionally, the inland location of Tam Pa Ling “suggests that Pleistocene modern humans may have followed inland migration routes or used multiple migratory paths” (Demeter et al. 2012, 14379) rather than simply following coastlines. Vietnam Some Vietnamese caves contain breccias with a Middle Pleistocene Stegodon-Ailuripoda fauna, including Pongo. In some cases, teeth attributed to H. erectus were probably from the orang-utan, Pongo, as the two are easily confused, especially if worn. At Tham Khuyen, Lang Son Province, nine teeth were initially identified as Homo erectus (Nguyen Lang Cuong 1992), but this number was later considerably reduced (Schwartz et  al. 1994; Ciochon et al. 1996; Marwick 2009, 54). Dating by U-series and ESR on teeth and speleothems indicated a date of 475 ± 125 ka. According to Schwartz et al. (1995, 3), only one tooth – an upper left canine, TK 65/167 – was “unequivocally hominid”, although a deciduous upper first molar (TK 65/8) could be from either Homo or Pongo. The remaining teeth were upper and lower molars and likely from a species of orang-utan that was not P. pygmaeus and from a new type of hominoid named Langsonia liquidens that is not affiliated with either Homo or Pongo. The presence of these hominoids at this site implies that large-bodied primates have a complex history in mainland Southeast Asia. At the adjacent cave of Tham Hai, one tooth was identified as Homo sp. in a faunal assemblage believed to be similar age to that from Tham Khuyen, but is likely to belong to a species of Pongo (Schwartz et al. 1995, 3). Five teeth (two molars, one premolar, one canine, one incisor) from the cave of Lang Trang, Thanh Hoa Province, were also attributed to Homo erectus, largely on account of their age rather than morphology. ESR dates on fossil teeth in breccias containing fossil teeth were dated from 146 ± 2 ka to 480 ± 40 ka (Olsen & Ciochon 1990; Marwick 2009, 54). However, Long et al. (1996) state that the Lang Trang faunal list is similar to that from Lida Ayer, Sumatra, which is dated to circa 80 ka, so the tooth of Homo may derive from H. sapiens. According to Schwartz et al. (1995, 6), only one (an upper right canine) of the 200 hominoid teeth from Lang Trang is hominid (and “typical of Homo sapiens”), and all the rest are from Pongo pygmaeus. Five isolated teeth were identified as from Homo sp. indet. at the cave of Hang Hum (Yen Bai Province) in a faunal assemblage that includes Stegodon orientalis, Rhinoceros sinensis, Tapirus augustus and Pongo pygmaeus (Nguyen Lang Cuong 1992; Schwartz et al. 1995, 5). Four teeth from the cave of Tham Om (Nghé An Province) are probably attributable to Pongo than Homo, as initially thought (Schwartz et al. 1995, 9) and likely 250–140 ka in age. The caves of Ma U’Oi and Duoi Oi (Hoa Binh Province) are small, complex fissures that were investigated by a joint Vietnamese-French team (Bacon et al. 2008). One maxillary molar (MU57), one left M1 (MU18) and small fragment of skull occiput (MU88) were found in fossiliferous breccias (Bacon et al. 2006; Demeter et al. 2004, 2005), associated with dental remains of Elephas maximus, Rhinoceros cf. sondaicus, Rhinoceros cf. unicornis and other mammals. The fauna is considered characteristic of the Stegodon-Ailuripoda complex (although neither is present at Ma U’Oi), and similar to that from Tham Khuyen and Lang Trang in Vietnam and Thum Wiman Nakin in Thailand.The fossiliferous breccia was dated to 193 ± 17 ka by U-series, and speleothem over this breccia was dated to 49 ± 4 ka, also by U-series dating (Bacon et al. 2006). Tooth MU18 is described as heavily worn but displaying a mosaic of archaic and modern traits. Its size is within the range of H. erectus, and the crown is larger than that of Southeast Asian H. sapiens. However, the crown is square like those of H. sapiens, and there was no taurodontism.The 41

Robin Dennell maxillary molar MU57 is described as more sapiens-like in terms of size and absence of occlusal wrinkles, peripheral placement of cusps and taurodontism. The skull fragment was too indeterminate to allow a taxonomic identification.

Island Southeast Asia The most important issues raised by the skeletal evidence from Java are the dates for when H. erectus became extinct and H. sapiens first appeared on the island. Here, the key sites are Ngandong and Punung. Ngandong Excavations of the terrace deposits beside the Solo River at Ngandong by the Dutch Geological Survey in 1931–1933 (Oppenoorth 1932; Koenigswald, 1933) recovered the remains of 12 hominin crania (Solo I–XI; Solo III was later found to represent two individuals [Schwartz & Tattersall 2003, 450]) and two tibiae, along with more than 25,000 other mammalian fossils that define the Ngandong Fauna. Parietal and pelvic fragments were later excavated by an Indonesian team in 1976 and 1978. The hominin remains from Oppenoorth’s excavation were described in monographs by Weidenreich (1951) and Santa Luca (1980), who attributed them to a late population of Homo erectus. This conclusion is shared by most researchers (e.g., Antón 2003, 144), although Schwartz and Tattersall (2000, 20) suggested they probably belong to a sister taxon, and Hawks et al. (2000) suggested they should be re-classified as H. sapiens because of their similarities to WLH-50 from Australia. The dating of the hominin assemblage has proved highly controversial following the dating by Swisher et al. (1996), who dated bovid teeth from Ngandong to circa 28–54 ka and suggested that H. erectus had therefore been contemporaneous with the earliest populations of H. sapiens in Southeast Asia. This dating was questioned on technical grounds by Grün and Thorne (1997). However, a similar range of dates was obtained by Yokoyama et  al. (2008), who attempted to date two crania from Ngandong (Ng-1 and Ng-7) and Sambungmacan 1 by uranium-thorium (234U/230Th), uranium-protactinium (235U/ 231Pa) and thorium-thorium (227Th/230Th) series dating. These produced dates as young as 39.6 ± 9.5 kafor Ng-1 by thorium-thorium and a maximum age of 88.1 ± 10.8 ka by uranium-thorium (early uptake model). They thus suggested that their likely age was between 40 and 60–70 ka, in which case H. erectus and H. sapiens may have overlapped in time. Nevertheless, uranium leaching was observed in the dated specimens, and despite the ingenuity and thoroughness of the investigators, their dates are not problem-free. On faunal grounds, there are sound reasons for placing the Ngandong Fauna (and the associated hominins) in the late Middle Pleistocene. The Ngandong Fauna contains several extinct species and no indications of rainforest. In contrast, the succeeding Upper Pleistocene Punung Fauna consists entirely of extant species and provides the first indications in Java of dense, humid forest. Storm (2001a, 372) points out that the expansion of rainforest indicated by the Punung Fauna covered the whole of Java, and so open woodland would not have found a refuge in east or south Java from which it could re-expand in post-Punung times. On faunal grounds, therefore, H. erectus was probably extinct on Java several millennia before the arrival of H. sapiens (see Storm 2001a, 2001b; Storm et al. 2005).This assessment now seems confirmed by a recent and thorough re-dating of the Ngandong sequence that used 40Ar/39Ar dating on water-lain pumices and ESR and U-series dating on in situ fossil teeth (Indriati et al. 2011). This study concludes that “it now seems possible to bracket the age of the deposits at Ngandong . . . with a maxima of 546 ka based on the argon results and a minima of 143 ka based on the oldest of our fully modelled combined ESR/U-series ages. It is certainly possible that the age of the hominins more closely approaches one than the other of these extremes, which given the geochemical issues with U-series at this site and the apparent site formation processes, some of us suspect is more likely to be the argon 42

The Fossil Record for Homo sapiens between Arabia and Australia age” (Indriati et al. 2011, 9). In the light of this re-assessment as well as the faunal evidence, it now seems highly probable that H. erectus in Java was extinct before the arrival of H. sapiens. If so, population continuity between the Ngandong population and ones in Australia now seems unlikely. Punung Storm et al. (2005) reported the discovery of a left P3 (PU-198) in the Punung faunal collections at the Senckenburg Institute, Frankfurt. These were excavated by Koenigswald in the 1930s from two caves (Punung I and II) that appear to have had similar faunal assemblages, which were unfortunately thereafter mixed together. The size of the premolar falls within the range of modern H. sapiens from Australasia (N = 46), and outside that of H. erectus (N = 7). At a new and nearby cave site (Punung III), the Punung Fauna (including orang-utan, siamang and sun bear) is now dated by TL and OSL to between 128 ± 15 and 118 ± 3 ka (i.e., MIS 5e), which “would imply that H. sapiens arrived in Southeast Asia during the Last Interglacial” (Westaway et al. 2007a, 715). However, human remains have not yet been found at Punung III, and Barker et al. (2007) and Bacon et al. (2008) were sceptical of both the dating and identification of the Punung I/II premolar as that of H. sapiens. The timing of its arrival in Java thus remains unclear. Wadjak This site deserves mention for historical reasons. The cranium (Wadjak I) was discovered in 1888 and sent to Dubois, who examined the site in1890 and found a second cranium and associated post-cranial bones (Wadjak II).When discovered, they were thought to be of Pleistocene age and for many decades were regarded as “proto-Australian” (Dubois 1921) or a part of a continuous lineage from Ngandong to modern humans (Weidenreich 1945) (see Storm and Nelson 1992). Thanks to radiocarbon dating, the site is now known to be Holocene, and a human femur from Wadjak II has been dated to 7670–7210 cal BP (Shuttler et al. 2004).Wadjak I is likely of the same age. Contra Dubois (1921) and Weidenreich (1945), the Wadjak I cranium has no relevance now to documenting the ancestral population of Aboriginal Australians. Niah Cave The best-known and least ambiguous specimen of H. sapiens from Southeast Asia is a partial cranium that was found in the “Deep” or “Hell” Trench at Niah Cave in 1958 by a team led by Tom Harrison. It was later radiocarbon dated to circa 40 ka and was for many years the oldest example of our species world-wide (see Barker et al. 2007, 251; Hunt & Barker, this volume). An almost complete left femur and a right proximal tibia fragment were found near the cranium and are probably from the same individual, as well as a human talus. Following a thorough reinvestigation of the deposits and materials from Niah, the cranium has now been dated by AMS procedures and U-series dating to 41–34 14C ka BP (ca. 45–39 ka cal BP), some five millennia after the cave was first occupied. For those who rely solely upon cranial evidence (excluding isolated teeth), the Niah Cave cranium provides a vital anchor point of circa 40 ka for the appearance of our species in Southeast Asia. The Philippines Tabon Cave on the island of Palawan, immediately north of Borneo, has an occupation sequence dated from 9 to more than 30 ka BP. Its human remains comprise a frontal fragment (P-XIII-T-288) and a problematic left mandible attributable to either Homo sp. or Pongo that were dated by U-series to 16,500 ± 2000 BP (Dizon et al. 2002), a right mandibular fragment (PXIII-T436 Sg 19) and 11 other specimens (Détroit et al. 2004).These are a right temporal bone (IV-2000-T-188), an occipital bone (IV-2000-T-372) and nine fragmentary post-cranial remains. At least three individuals are represented. Direct U-series dating produced dates of 31 + 8/−7 ka for the right mandibular fragment, and 47 + 11/−10 ka for a tibia fragment (IV-2000 T-197) (Détroit et al. 2004, 710). Although a maximum age of 58 ka for the tibia might be considered as 43

Robin Dennell unacceptable, an age of 50 ka or less is consistent with recent evidence from New Guinea (see Summerhayes and Ford, this volume). Ongoing research should result in further discoveries at this important site. A recent potentially highly important discovery is that of a hominin third metatarsal dated to circa 67 ka from Callao Cave, Luzon, in the northern Philippines, that is described as similar to that of H. sapiens, H. habilis and H. floresiensis (Mijares et al. 2010; Pawlik et al., this volume). This find is intriguing as it is the first clear evidence of humans in the Central Philippines and indicates that humans (or other types of hominins) could make sea crossings before 65 ka. Further discoveries are keenly awaited from this site.

China Because there is so little human skeletal evidence from mainland Southeast Asia, Chinese material should also be considered when discussing how and when H. sapiens appeared in East Asia. Unfortunately, hardly any Chinese sites between 30 and 125 ka are free of doubts over the accuracy of dating, the precise stratigraphic provenience of finds within a layer or the identification of remains as H. sapiens. To date, the best and currently earliest unambiguous evidence for our species in China is from the cave of Tianyuan (Tianyuandong), near Zhoukoudian, North China. Tianyuandong Excavations in 2003 produced 34 remains of H. sapiens: most of the front and right side of a mandible, and parts of upper and lower limbs, all likely from one individual (Shang et al. 2010; Shang & Trinkaus 2010). These were found clustered in layer III, an unconsolidated breccia. Six bones were dated by AMS 14C to between 30,500 ± 370 and 39,430 ± 680 BP (uncalibrated). A sample from the human femur was dated to 34,430 ± 510 ka BP uncalibrated, or 40,328 ± 816 ka BP calibrated, close in age to the cranium from Niah Cave, Borneo. Analysis of wear on the toe bones indicated that Tianyuan may also provide indirect evidence for footwear (Trinkaus & Shang 2008). Gaitou Cave, Qilinshan Hill Here, excavations in 1956 produced a fragmentary human skull: a palate with part of the maxilla and several teeth, a right zygomatic and an occipital fragment (Wu & Poirier 1995, 193–194). These were found between two flowstones. Recently, the lower flowstone was dated to 112 ± 1.4 ka, and the upper to 38.5 ± 1.0 ka. Analyses of calcite veins below the human specimens indicated a likely age of 44 ± 0.8 ka. The human specimens therefore are probably 39–44 ka, but cultural remains within the same layer may have a maximum age of 112 ka (Shen & Michel, 2007; Shen et al. 2007). Ganjian Cave, Tubo Ganjian Cave (30 km from Liujiang) is another site where human remains have been dated by U-series techniques on flowstones above and below human remains. Here, human teeth (upper incisor II; left M1; 3 right M1; right (?) M2; left M3; right lower M, left dm2) were found, associated with a Stegodon-Ailuripoda fauna. Another four human teeth that probably came from Tubo were acquired from a nearby drugstore (Wu & Poirier 1995, 210–213). Thorium-uranium dating of stalagmites above and below the layer with the teeth were dated to 94 and 220 ka respectively. Two fossil non-human teeth were dated by 230Th-234U and 227Th-U230U to 85 ka and 139 ka. On this basis, the human teeth are therefore circa 100 ka old (Shen et al. 2001a). Tongtianyan Cave, Liujiang This discovery is one of the most contentious finds of H. sapiens in South China. In 1958, local farmers found when digging for fertiliser an almost complete human skull, several post-cranial 44

The Fossil Record for Homo sapiens between Arabia and Australia bones and a Stegodon-Ailuripoda faunal assemblage. Because of the circumstances of discovery of the skull, “its exact stratigraphic provenance can hardly be unequivocally fixed” (Shen et al. 2007). A recent re-examination of the cave (Shen et al. 2002) showed a complex stratigraphy, with three depositional units of silty clays with flowstones, and a younger intrusive breccia, in which the skull was probably (but not certainly) found. Dating of flowstones by U-series and TIMS (thermal ionization mass spectrometric techniques) indicated that the breccia is at least 68 ka but more likely circa 111–139 ka. If, however, the skull derived from the Middle Unit that was cut by the breccia, its age could be greater than 53 ka. The over-riding uncertainty over Liujiang is its stratigraphic provenance, as it may have derived from a younger, intrusive burial cut into the breccia after it had formed.The only likely means of resolving its age is by dating the skull itself. At Bailiangdong Cave, 5 km northeast of Liujiang, two teeth of H. sapiens were found in a layer with remains of 30 mammalian species and 500 artefacts. 14C dating of calcites and shells and AMS dating of shell and grains of carbon dated the entire sequence to between 7 and 37 ka, with the layer containing the human teeth to circa 30 ka. However, Th-U dating of the stalagmite floor of layer 6 (overlying the layer with the teeth) and a stalagmite contemporary with the same layer indicated an age of 160 ka, which would thus be the minimum age of the human teeth (Shen et al. 2001b). Huanglong Cave, Yunxi County At Huanglong Cave, seven human teeth (upper central and lateral incisors, I2, upper canine, M3, M2 and M3) were excavated in layer 3 (Liu et al. 2010a). Two discordant sets of dates were obtained. U-series dating of two rhinoceros teeth gave an age range of 94.7 ± 12.5 ka and 79.4 ± 6.3 ka. Speleothem dating by U-series indicated ages of 103.739 ± 16 ka and 103.119 ± 13 ka. However, ESR dating of a rhinoceros tooth produced an age of 44.18 ± 4.54 ka (LU model) and 34.78 ± 3.28 ka (EU model). To complicate matters further, it is unclear whether the speleothem sample from the section came from above or below where the hominin teeth were found. This point highlights the difficulty of correlating speleothems exposed in a cave section with material found in the cave floor. Zhirendong, Mulanshan (Mulan Mountain) This cave recently produced a potentially important piece of evidence for early H. sapiens in South China (Jin et al. 2009; Liu et al. 2010b). Here, two molars and a toothless mandible were found in the upper part of layer 2 of Unit B, associated with a late Middle or early Upper Pleistocene fauna. A series of flowstones in the overlying layer 1 was exposed in the cave section and dated by U-series techniques. The upper two flowstones produced dates of 28.4 ± 6.6 ka (Sample S1 ML-1A) and 51.8 ± 22.7 ka (Sample S2 ML-1B). The lower set of flowstones was dated to between 74.1 ± 21.6 ka (Sample Sa ML-6a) and 106.2 ± 6.7 ka (Sample Sb-ML-6b). The minimum age of the human remains is thus estimated as circa 100–113 ka. The mandible is described as exhibiting a mix of archaic and modern features, which are interpreted as implying “early modern human dispersal or gene flow across at least southern Asia sometime before the age of the Zhiren Cave human remains” (i.e., before 100 ka, the age of the oldest dating sample 6B), but also that “any ‘dispersal’ involved substantial admixture between dispersing early modern human populations or gene flow into regional populations” (Liu et al. 2010b, 19205). Two caveats are needed here. First, the flowstones overlying the human remains were tightly packed within a thin stratigraphic zone with a vertical thickness of only 10 cm between Sample 2 at circa 51 ka and Sample 6b at circa 106 ka (see fig. S6 in Liu et al. 2010b), which suggests a “number of small-scale sedimentation or repeated erosional and depositional events at this portion of the cave sediments” (Kaifu & Fujita 2012, 3). Without knowing how the flowstones extended across the cave floor from the section, it is not possible to know which relates to the human mandible. It could therefore be considerably younger than 110 ka. Secondly, although there are several 45

Robin Dennell Chinese mandible specimens from the early part of the Middle Pleistocene, there are none from the later Middle Pleistocene, and consequently we do not know how gracile the mandibles of H. erectus became in China between 300 and 150 ka (Dennell 2010). Longlindong and Maludongdong Curnoe et al. (2012) report recent discoveries of cranial and post-cranial remains from these two caves and dated by radiocarbon to the late Pleistocene, circa 14.3–11.5 ka cal BP. Although the remains are classified as H. sapiens, they exhibit several archaic features not seen in extant populations. They suggest that these might indicate an indigenous late-surviving archaic population, or the descendants of an early population of incoming H. sapiens that remained isolated and did not contribute genetically to recent East Asians. Either possibility implies that the demographic history of East Asians is more complex than currently envisaged.

Australia The most problematic aspects of the rich but complex Pleistocene human skeletal record of Australia are its dating and likely origin(s), both of which impact directly upon wider discussions about the appearance of our species outside Africa. Most Australian Pleistocene skeletal evidence dates from the Late Pleistocene (Brown 1992; see Table 4.3), although the stratigraphic context is not always clear (Bowdler 1992). The earliest evidence comes from the long-extinct Lake Mungo in the Willandra Lakes region of New South Wales and comprises a remarkably early cremation of two individuals (Mungo 1 and 2) (Brown et al. 1970) and an extended burial of an adult (Mungo 3), presumed male (Brown 2000; Thorne & Curnoe 2000). The dating of Mungo 3 has generated considerable controversy. Initial estimates suggested a radiocarbon age of between 28 and 32 ka. Later studies involving TL, OSL and 14C dating techniques indicate an age of 40 ± 2 ka for Mungo 1 and 3. The earliest indications of humans at Mungo are 11 flakes dated by OSL to 46–50 ka (Bowler et al. 2003). However, a considerably older estimate for Mungo 3 was obtained by Thorne et al. (1999), who suggested a burial age of circa 62 ±6 ka, derived from OSL dating of sediments and ESR analysis of tooth enamel. Their estimate implies that Australia was colonised in oxygen isotope stage 4 (57–71 ka). Critics of Thorne and his colleagues maintain that their results are incompatible with previous results and ignore crucial field stratigraphic details (Bowler & McGee 2000; Gillespie & Roberts 2000) – charges that are strenuously denied (Grün et al. 2000). Even so, more evidence is needed to demonstrate clearly that Australia was colonised before 60 ka. This dispute mirrors a similar one over the age of the oldest archaeological sites in Australia. If one accepts OSL dates from Malakunanja II (61 ± 10 ka) and Nauwalabila (55 ± 11 ka) in Arnhem Land, North Australia, Sahul could have been colonised before 60 ka (Roberts et al. 1994). However, there are no archaeological sites dated by 14C that are older than 40ka, even though there are numerous seemingly reliable 14C dates for geological samples extending back to 54 ka (Allen & Holdaway 1994; O’Connell & Allen 1998), and this evidence may simply imply that the OSL dates are too old.Yet it would seem rash to reject all OSL dates more than 40 ka: the mtDNA extracted from the Mungo 3 skeleton implies a deep branching event with no modern descendants (Adcock et al. 2001; Relethford 2001), and according to Schillaci (2008), Australian crania are more like those from Skhuˉ l and Qafzeh than later ones, which might imply a dispersal from Africa as early as MIS 5. The date when humans first arrived in Australia remains an open question. Debates over the origin of indigenous Australians have been as contentious as over when humans first landed in Sahul. At various times, researchers have proposed one, two or three waves of immigration. A two-wave model has been popular amongst multi-regionalists as a way of explaining the differences between the gracile and more robust specimens. Some researchers (e.g., Hawks et al. 2000; Wolpoff et al. 2001) interpreted this distinction as implying two episodes of colonisation: an earlier one, derived from a gracile Mongoloid population evidenced in China, 46

The Fossil Record for Homo sapiens between Arabia and Australia Table 4.3.  Suggested dates of Australian Pleistocene skeletal remains Locality

Specimen

Age

Technique

Mungo 1 and 2 Mungo 3

24,700 ± 1,270 BP 32,750 ± 1,250 BP

14

Mungo 3

Cremations Extended burial, male See above

42–45 ka

Mungo 1 and 3 Mungo 3

See above See above

40 ± 2 ka 62 ± 6 ka

WLH-50

Cranium

Bowler & Price 1998; Oysten 1996 14 C, TL, OSL Bowler et al. 2003 OSL, U-series Thorne et al. 1999 OSL, U-series Grün et al. 2011

Kow Swamp

Several (c. 40) skeletons Several skeletons

> 12.2 ± 1.8, < 32.8 ± 4.6 ka 13,900 ± 280–9,590 14C ± 130 BP 22–19 ka OSL

Kow Swamp Nacurrie Keilor Coobool Creek

Almost complete 11,440 ± 160 BP skeleton Cranium and femur 12,000 ± 100 BP fragments Femur associated 14,300 ± 1000 BP with cranium

C C

14

Reference Brown 1992 Bowler et al. 1970

C, TL, OSL

14

AMS 14C

Brown 1992 Stone & Cupper 2003 Brown 1992

C

Brown 1992

U/230Th

Brown 1992; Bowdler 1992

14

234

and a later, robust one derived from late Homo erectus in Java. One crucially important find in these discussions is the Late Pleistocene cranial specimen WLH50 from Willandra Lakes. The cranium is extremely robust, and this has been regarded as either pathological (Webb 1990) or indicative of descent from a robust ancestor. Hawks et al. (2000) applied pair-wise analysis (which measures differences between specimens) to WLH50 and other fossil crania and argued that because WLH50 resembled those from Ngandong more than the Qafzeh specimens, the WLH50 individual may have derived from the Ngandong hominins (or a population like them); in view of these similarities, they also proposed that the Ngandong H. erectus hominins should be re-classified as H. sapiens. However, Stringer (1998) used principal components analysis to show that WLH50 was closer to African specimens than Ngandong, and Collard and Franchino (2002) argued that pairwise analysis is inappropriate for indicating phylogenetic relationships. Additionally, there is no evidence that hominins like those at Ngandong (i.e., Homo erectus) inhabited Sunda in the Upper Pleistocene, and they were likely extinct long before Australia was colonised. A further problem is that some of the robust specimens were artificially deformed (Lahr 1996, 294; Brown 2010). The most parsimonious model that explains the variation in the Australian Pleistocene skeletal sample is one that assumes only one major wave of immigration and thereafter a complex history of expansion and contraction out of and into glacial refugia, depending upon prevailing levels of rainfall and temperature (Veth 1993). Under such conditions, genetic drift and bottlenecks would generate the morphological diversity amongst early Australians (Brown 1992; Lahr 1996, 295).

Discussion The study of human evolution and demography in the Upper Pleistocene in southern Asia between Arabia and Australia is still in its infancy. Three points can be made about what needs to be done. 47

Robin Dennell

The Need for More, and Better, Data There are still enormous gaps in our knowledge of human populations across southern Asia during the Upper Pleistocene. The largest by far is the million-square mile Arabian Peninsula, where there is not a scrap of human skeletal evidence before the Holocene. Continuing eastwards, there is no skeletal evidence more than 30 ka and unequivocally less than 125 ka from southern Iran, Pakistan, India, Bangladesh and Myanmar,1 although Sri Lanka has the best skeletal evidence from southern Asia between 25 and 30 ka.The main restriction on clarifying the appearance of H. sapiens in East Asia is the lack of well-dated, taxonomically unambiguous skeletal specimens between 100 and 30 ka (Gao et al. 2010, 1936; Wu 2004, 138). The earliest unambiguous examples of H. sapiens from Tianyuandong in North China and Niah Cave in Borneo, 3,000 miles to the south, date from circa 40 ka but are most unlikely to have been the earliest examples of H. sapiens in East Asia because genetic studies of modern populations indicate that the earliest populations of our species were present circa 50–60 ka (Liu et al. 2006; Su et al. 1999). Unfortunately, the significance of earlier finds such as Ganjian, Liujiang, Zhirendong and Huanglong remains unclear because of doubts over their identification as H. sapiens (especially when the only evidence is isolated teeth), dating (as when finds from a cave filling have to be correlated with closely spaced flowstones in the cave section, as at Zhirendong and Huanglong) and stratigraphic context (as at Liujiang). The handful of hominin teeth from caves in the highland regions of northern Thailand and Vietnam lend some support to Schepartz et al.’s (2000) suggestion that upland regions of southern China and Southeast Asia were first colonised in the late Middle Pleistocene. Outside the northern parts of Thailand and Vietnam, there is virtually no evidence that mainland Southeast Asia was even occupied during the later Middle Pleistocene, and we should perhaps consider the possibility that the alluvial plains of this region’s large rivers – the Chao Phraya, the Red and Pearl Rivers, the Salween and Mekong – were not inhabited until the Upper Pleistocene. If so, H. sapiens might have colonised a vacant region, rather than replaced its indigenous inhabitants.

Interpretative Frameworks: Where Paradigms Clash In discussions over the origins of our species in southern Asia, there have been two opposing paradigms, each deriving its main support from different parts of Asia and usually arguing over a small number of poorly dated specimens found many years ago. A (probable) majority of researchers prefers Replacement (or Out of Africa) Models, based on the inference that H. sapiens originated in Africa and then dispersed outwards across Asia and Europe, and ultimately colonised Australasia and the Americas (see, e.g., Stringer & Andrews 1988; Mellars 2006b). Under that umbrella term, there are numerous subordinate debates over the timing and number of dispersals, and the degree of competition, co-existence and hybridisation that might have occurred between incoming populations of Homo sapiens and the indigenous populations (see Stringer 2001; 2002). In contrast, a (probable) minority of researchers favours multi-regional models of recent human evolution, whereby hominin populations evolved in parallel towards H. sapiens and avoided speciation via gene flow between neighbouring populations (see, e.g., Wolpoff et al. 1984; Wolpoff et al. 1994; Hawks et al. 2000). Multi-regionalism is also an umbrella term, in that its adherents place varying degrees of emphasis upon assimilation with different populations. As example, some Chinese researchers now argue for “continuity with hybridization” (e.g., Wu 2004), and some adherents of replacement models now incorporate the likelihood of interbreeding between H. sapiens and indigenous communities, as indicated by recent genetic data showing the persistence of Neanderthal DNA in Eurasian (but not African) human populations (Green et al. 2010). As Gao et al. (2010, 1937) point out, “the ‘Continuity with Hybridization’ theory of East Asia and the ‘African Origin of modern Chinese’ theory are not mutually exclusive”. The situation is further complicated by the discovery of a previously unrecognised population of “Denisovans” from the 48

The Fossil Record for Homo sapiens between Arabia and Australia analysis of the ancient DNA of human remains from Denisova Cave, Siberia (Krause et al. 2010), as a probable sister population of Neanderthals. As Denisovan DNA is apparently present in modern human populations in Southeast Asia and Melanesia, H. sapiens may have interbred with both Neanderthals and Denisovans (see Oppenheimer, this volume). As the only skeletal evidence of this Denisovan population is the phalange and molar that were analysed for their aDNA, some cranial specimens from East Asia might be Denisovan rather than H. erectus or H. sapiens. If we turn beyond polemic about the strengths and weaknesses of each approach, neither type of explanation copes particularly well with the limited and problem-laden amount of skeletal data from southern Asia. For example, those favouring a replacement model cannot yet explain what type of populations were replaced in Arabia, southern Iran, Pakistan, India, Myanmar or mainland Southeast Asia. Nor can it yet indicate how many dispersal events might have occurred, under what climatic conditions, or with what degree of co-existence, competition and hybridisation with local populations. Perhaps most importantly of all, the date at which H. sapiens left Africa and dispersed eastwards towards Australia remains unclear, with estimates ranging from circa 60 ka (Mellars 2006b) to perhaps as early as 125 ka (Dennell & Petraglia 2012; Scally & Durbin 2012; Boivin et al.2013). Multi-regionalists have their own problems, particularly in East Asia, which is the region most commonly cited as demonstrating local evolution from Homo erectus to H. sapiens. One thorny problem is the likelihood that gene flow was extremely restricted between populations in East Asia and those further west. North China would probably have been isolated from Central and Western Asia by the deserts of the Gobi,Taklamakan, Kara Kum and Kyzl Kum, especially as these became more extensive in the later Middle Pleistocene (Dennell 2009, 257; 2013). To the south, the Ganges-Brahmaputra and the mountains of northeast India and northern Myanmar were (and are) long-standing barriers between South and Southeast Asia. It is thus likely that gene flow between populations in East Asia and those further west was limited and intermittent throughout much of the Pleistocene. A second problem for multi-regionalists is that there is so little reliable skeletal evidence from the early Upper Pleistocene (ca. 125–50 ka) of East and Southeast Asia that it is extremely hard to demonstrate population continuity from the late Middle Pleistocene to 40 ka. Additionally, many researchers doubt claims of morphological continuity in East Asia from H. erectus to H. sapiens (e.g., Wolpoff & Caspari 1997a; Wolpoff et al. 2000) on the grounds that the morphological traits used to demonstrate continuity are either plesiomorphic (Stringer 2002) or not regionally unique (Lahr 1996).

Towards the Future First and foremost, more taxonomically diagnostic specimens need to be found in secure, accurately dated contexts and properly described and published. Until then, assessments of when and how our species appeared across southern Asia have to rely heavily upon proxy genetic and archaeological evidence. Explanatory frameworks need also to develop beyond repeating the respective merits and drawbacks of Replacement versus Multi-regional Models. What we need instead are regional demographic histories that are tied into climatic and environmental sequences. Gao et al.’s (2010, 1929) comment that “ancient humans in East Asia may have differentiated into a number of regional groups  – local extinction, interregional migrations, and even population replacements could occur as well, and show an evolutionary pattern similar to a river network” is likely applicable to most of southern Asia. Population histories of Asia were probably complex – “a repeated theme of regional expansion and contraction, colonisation and abandonment, integration and isolation” (Dennell 2009, 475) – and both local evolution and population movements were probably involved (see also Norton and Jin 2009). A useful beginning is to think in terms of regional population histories and palaeo-demes (Howell 1999) instead of larger scale, continental-wide models of replacement or population continuity. Here, the type of 49

Robin Dennell source-sink model (Dennell et al. 2011) developed for Europe might be usefully applied to areas such as Southeast Asia (Louys & Turner 2012) and provide ways of incorporating archaeological and genetic information into regional models of Pleistocene demography. Note 1. Schepartz et al. (2000, 7) mention a claim by Ba Maw (1995) of a maxillary fragment that was attributed to H. erectus, with a claimed age of 200 ka, but point out that both claims are unlikely as associated material included domesticated dog and Neolithic artefacts. It seems prudent to discount this find as evidence of H. erectus in Myanmar.

50

Chapter 5 An Arabian Perspective on the Dispersal of Homo sapiens Out of Africa

Huw S. Groucutt and Michael D. Petraglia

Introduction The privileged position of Africa in human evolution has long been hypothesised and was largely confirmed, after many years of often heavily polarised debate, in the later twentieth century (e.g., Willoughby 2007). Recent developments have shown that the later stage of this process, the evolution of Homo sapiens in Africa and their subsequent dispersal into Eurasia, was a spatially and temporally complex process, involving, for instance, interbreeding with ‘archaic’ hominin species (e.g., Green et al. 2010; Reich et al. 2010).The rapidity with which such changes to our understanding of the dispersals of H. sapiens out of Africa have taken place suggests a need for interpretative caution, the development of a solidified interdisciplinary approach, and a move towards nuanced and explicit models of the dispersal process (see Blinkhorn & Petraglia, this volume). A large body of research from the interlinking disciplines of Palaeolithic archaeology, genetics and palaeontology  – contextualised in terms of palaeoenvironmental variation and biogeography – addresses the character of Upper Pleistocene dispersals out of Africa into Asia. In this chapter we describe four main models for this process (Table 5.1).We focus on the evidence from the Arabian Peninsula as a means of testing the four models, placing particular emphasis on the archaeological record. Given that Arabia may have been the first place dispersing African H. sapiens encountered in Eurasia, the peninsula has a special position, in that it was here that the genetic and behavioural foundations of populations that would go on to populate the rest of the world were assembled, tested and filtered by natural and cultural selection. Quaternary Arabia saw dramatic environmental oscillations, and these most likely played a pivotal role in determining the timing and character of H. sapiens dispersals. Although the prehistoric record of Arabia remains understudied compared to many other areas, recent years have seen significant advances (see, e.g., Petraglia & Rose 2009; Groucutt & Petraglia 2012).The time has come for the rich Arabian record to play a more central role in debates on hominin dispersals out of Africa.

51

52

Table 5.1.  Key models for the dispersal of Homo sapiens out of Africa Dispersal model

Summary/key points

Timing of dispersal

MIS 3,~50– Upper Similar to traditional ‘Upper 40 ka Palaeolithic Palaeolithic Revolution’ Model Archaeological focus, particularly Levantine Early/Initial Upper Palaeolithic Reasons for behavioural changes and dispersal range from a neural mutation creating capacity for modern culture to the demographic effects of the development of complex projectile technology MIS 4,~65– Microlithic Environmental changes in 60 ka Model southern Africa ~80–70 ka, which the model leading to significant demographic and behavioural changes, most clearly seen in the Howiesons Poort and Still Bay industries, and a single coastal dispersal out of Africa Chronology based on genetic evidence

Evidence and predictions Hominin fossils

Genetic structure

Archaeology

Route out of Africa

Key references

Bar-Yosef 2007; Northern, ‘Trait lists’ of Crania at Ksar Akil Conservative Klein 2009; perhaps ‘modern’ and Qafzeh. interpretation Shea 2011e associated behaviour (35–30ka) of mtDNA with (e.g., blade Teeth at Üçagizli coalescence Heinrich technology, (> 40ka) suggests last Event 5 symbolism, common longcausing local ancestor Neanderthal distance raw of modern extinction material humans dates to exchange) MIS 3

Macaulay et al. Southern, Primarily based Predicts Lake Mungo, 2005; Mellars coastal route on similar age discovery of Batadombalena 2006a, 2006b along Indian (~36ka, Sri estimates for M microlithic Ocean rim, assemblages Lanka) and N mtDNA crossing Bab in SW Asia, Predicts fully macroal Mandab, also ochre, modern fossils haplogroups interior perforated in SW Asia in very arid in shells, MIS 3/4 MIS 4 and and other avoided. indicators of ‘modern behaviour’

MIS 4+3,~70 Structure of modern human ka, ~45 ka populations reflects multiple Upper Pleistocene dispersals Out-of-Africa dispersals occurred in both MIS 4 and 3, following population subdivision within Africa Originally a fossil model, reinvigorated by genetic findings Middle Climatic amelioration MIS 5,130–75 Palaeolithic allowing Homo sapiens to ka Model enter Arabia in MIS 5e, and South Asia by MIS 5a Dispersals marked by presence of Middle Palaeolithic technology, similar to that of African MSA MIS 5 Homo sapiens in SW Asia as seen in the Levant may not be a ‘failed dispersal’, even if a localised extinction occurs in the Levant Two-stage Model

Lahr & Foley Lake Mungo, Australian ~70 ka dispersal Southern, coastal route 1994, 1998; ~70 ka dispersal – genomic associated around Rasmussen robust anatomy. sequence with mode Indian et al. 2011 ~45 ka dispersal supports model 3 (Middle more gracile of at least two Palaeolithic) Ocean rim ~70 ka dispersals, technology. similar evidence ~45 ka Northern, emerging associated through elsewhere with Mode Sinai ~45 ka 4 (Upper Palaeolithic) Skhu¯l /Qafzeh in Demographic Technological Broadly Petraglia SW Asia and fluctuation analysis southern, et al. 2010; Lake Mungo means genetic reveals split Armitage in Australia record complex similarities between et al. 2011; as upper Open questions of S and SW Sinai and Rose et al. and lower in relation Asian lithic Bab al 2011; Scally & constraints to genetic technology Mandab Durbin 2012 on age evidence, but to the routes Predicts MIS 5 finds support in African Homo sapiens new mutation MSA fossils in S Asia rate estimates

53

Huw S. Groucutt and Michael D. Petraglia

Models for the Dispersal of Homo sapiens Out of Africa Table 5.1 presents four key models (sensu lato) for the dispersal(s) of Homo sapiens out of Africa. Each model represents a broadly coherent interpretive framework, reflecting evidence from more than one discipline and, where not explicitly making predictions, at least implying certain features of the various records shown in Table 5.1. Proponents of the Upper Palaeolithic Model suggest that the dispersal of H. sapiens out of Africa occurred ~50–40 ka. In the most explicit version of this model, Klein (2009) argues that a neural mutation gave an African population such a selective advantage that it rapidly dispersed into Eurasia. Alternative viewpoints that can be encapsulated within this model suggest rather differing narratives but at least a similar chronology. For instance, scholars have recently hypothesised that the development of projectile technology in the later Middle Stone Age (MSA) gave a selective advantage that resulted in dispersal ~50 ka (Shea & Sisk 2010; 2011; Brooks et al. 2006). The Microlithic Model shares some similarities with the Upper Palaeolithic Model, but its advocates suggest that broadly similar processes occurred earlier in time. It emphasises genetic evidence, primarily patterns of modern mitochondrial DNA (mtDNA) variation, to determine the chronology of dispersal. To Mellars (2006b; 2006c), the archaeological evidence of southern Africa, particularly the derived features of the Howiesons Poort industry, are congruent with the genetic story. Proponents of this model suggest dispersal followed a coastal route (see, e.g., Bulbeck 2007), so we should expect to find distinctive archaeological features along the coastline of Africa and southern Asia. The Two-Stage Model combines aspects of the three other models. Advocates of this model suggest that the key phases for dispersals out of Africa were both MIS 3 and 4. It shares with the Middle Palaeolithic Model the suggestion that the earliest dispersals of H. sapiens out of Africa were associated with a Middle Palaeolithic technology but disagrees that this occurred in Marine Isotope Stage (MIS) 5.The Two-Stage Model was initially primarily a reflection of fossil and anatomical variability (Lahr & Foley 1994; 1998) and has been reinvigorated by recent genetic studies, including those in Australia (Rasmussen et al. 2011) and Southeast Asia (Reich et al. 2011). The Middle Palaeolithic Model primarily relates Middle Palaeolithic assemblages in Arabia and South Asia to dispersals from Africa and argues that there are problems with the genetic evidence, particularly in terms of how confident we can be in the accuracy of dating estimates (e.g., Petraglia et al. 2010). The archaeological evidence is placed within the climatic amelioration of MIS 5, the last interglacial, a time when we would expect population expansion and dispersals to occur. The key support for this model comes from rapidly emerging evidence for African-like lithic industries in Arabia dating to MIS 5 (Armitage et al. 2011; Rose et al. 2011).

Hominin Fossils Spatial and temporal variability in the morphology of hominin fossils was a central element in the formulation of the out-of-Africa hypothesis (e.g., Stringer 2002) and is in theory a key way to test different models. The African fossil record seems to demonstrate the gradual, mosaic and regionally variable evolution of Homo sapiens from populations classified as either Homo heidelbergensis or H. rhodesiensis. The Upper Pleistocene then saw H. sapiens disperse into Eurasia. As Arabia lacks a known pre-Holocene hominin fossil record, we must here try and situate Arabia within the wider hominin fossil record (e.g., Rightmire 2009). The Upper Palaeolithic Model predicts the absence of successful (i.e., continuous) populations of fully modern H. sapiens before circa 50–40 ka in Arabia and more widely in southwestern Asia. The disappearance of H. sapiens from the Levantine fossil record circa 75 ka is taken as evidence for the regional extinction of an earlier population (e.g., Shea 2008). The cited fossil evidence (see Table  5.1), however, is from an extremely limited geographical area (the Levant), and we 54

An Arabian Perspective on the Dispersal of Homo sapiens must be cautious in generalising from this to the rest of Asia. Specimens from further east, such as Niah Cave, Borneo (Barker et al. 2007), dating to circa 40 ka, might be associated with an MIS 3 dispersal of H. sapiens from Africa, but without demonstrating the prior absence of H. sapiens in South Asia, this evidence merely provides an absolute terminus ante quem for dispersal from Africa. The confirmation of dates before circa 45 ka in Southeast Asia and Australia will be problematic to the Upper Palaeolithic Model. Even the older dates associated with the ‘short chronology’ for the peopling of Australia involves, for this model, a spectacularly rapid dispersal through the whole of southern Asia.The Upper Palaeolithic Model likewise suggests that findings such as the circa 67 ka metatarsal from Callao Cave, Philippines, must represent either a species other than H. sapiens or a continental scale ‘failed dispersal’ (Mijares et al. 2010). Likewise, the publication (Demeter et al. 2012) of a fully modern human crania dating to ~50–63 ka from Laos contributes to a small but growing collection of Asian fossil material that contradicts the Upper Palaeolithic Model. The Tam Pa Ling fossils, however, are broadly congruent with the other three models discussed in this chapter. Proponents of the Microlithic Model also see the presence of H. sapiens in the Levant in MIS 5 as a failed dispersal. The Microlithic Model will be strengthened or weakened by the discovery of pertinent fossil material around the Indian Ocean rim. At present its absence means fossil evidence does not provide an important line of evidence for this model. The ~45ka Lake Mungo burial in Australia is cited as supporting evidence by Mellars (2006b), but it is unclear what this fossil, at the far southeastern end of the dispersal arc from Africa, really demonstrates in terms of when humans dispersed across southern Asia.The Lake Mungo and other Australian discoveries make the Upper Palaeolithic Model rather unlikely but do not help us to distinguish between the other models. The Two-Phase Model is built on the hypothesis that the southern Asian fossil record will demonstrate dispersals in both MIS 3 and 4 with rather different morphological characteristics (Lahr & Foley 1994; 1998). In this model, MIS 5 saw the expansion and diversification of H. sapiens in Africa (and the Levant). The end of the last interglacial led to population decline and fragmentation. Yet an initial successful dispersal at broadly 70 ka saw a specific regional population, probably from East Africa, follow a southern and possibly coastal route all the way around the Indian Ocean rim. Hence Lahr and Foley (1994, 55) suggest that modern Aboriginal Australians may “present a high degree of morphological continuity with early and more robust populations”. We should expect then that Arabia will produce robust H. sapiens fossils sharing some of the features associated with modern Aboriginal Australians dating to circa 70 ka. The distinctive feature of the Two-Phase Model is to combine this dispersal with another at circa 45 ka, involving populations with more gracile and derived morphologies. The latter dispersal, essentially reflecting the view of the Upper Palaeolithic Revolution Model, is seen as the most important for the peopling of Eurasia and as involving considerably more people than the earlier dispersal. Morphological variability of Upper Pleistocene and Holocene populations will then reflect their relationship to these two dispersal processes and subsequent adaptation to local conditions. Proponents of the Middle Palaeolithic Model suggest that the Levantine MIS 5 occupation was part of a broader process of population dispersal into southwest Asia.The absence of a known pre-Holocene hominin fossil record in Arabia, as well as the general paucity of the Asian hominin fossil record, makes elucidating the spatial and temporal dimensions of the MIS 5 dispersal on fossil evidence currently problematic. The Levant is a tiny area, and the Middle Palaeolithic Model predicts the discovery of early Upper Pleistocene H. sapiens fossils in Arabia. Claims for MIS 5 H. sapiens fossils in East Asia are congruent with the Middle Palaeolithic Model (e.g., Liu et al. 2010b), but these specimens are either of questionable morphological status or poorly dated (see Dennell, Chapter 4, this volume). The Tam Pa Ling fossils (Laos, ~50–63ka) are congruent with several models for the dispersal of H. sapiens. The fact that they are found far inland perhaps suggests they represent a population that had long adapted to the region, congruent with the Middle Palaeolithic Model, rather than rapidly adapting from coastal populations arriving in MIS 4/3 emphasised in other models. 55

Huw S. Groucutt and Michael D. Petraglia At present then, the poverty of the southern Asian Middle and Upper Pleistocene hominin fossil record makes palaeontology largely mute on the spatial and temporal character of dispersals into Eurasia. Given the small size of the Levant it is at best problematic to generalise regional and subcontinental scale processes from a few sites in this area. It is unlikely that the earliest H. sapiens populations dispersing from Africa moved solely towards the Levant (see, e.g., Groucutt & Petraglia 2012). Whether such dispersals were successful or failed remains an open question that requires testing in areas such as Arabia. The discovery of a relatively continuous Upper Pleistocene fossil record of H. sapiens in Arabia for instance, would strongly support the Middle Palaeolithic Model. Given the proximity of southwest Asia to Africa it is unlikely that the fossil record will be able to distinguish between repeated dispersals and longer periods of occupation until the record has a considerably higher resolution.

Genetics In recent years patterns of genetic variation in modern populations have become central to debates on the dispersals of early Homo sapiens (see Oppenheimer, this volume). The consistent presence of the most primitive (basal) genetic lineages in modern African populations is congruent with the origin of H. sapiens somewhere within this continent.The key issue then, as with the fossil evidence previously discussed, comes down to the chronology and the nature of dispersals out of Africa. Assigning a chronology to patterns of genetic variability reflects calibration against other sources of evidence (see, e.g., Soares et  al. 2009; Endicott et  al. 2009), such as the homininchimpanzee divergence, estimates for which vary by millions of years. Other possibilities include calibration against the peopling of areas such as Australia. However, what exactly this tells us about the temporal and spatial context of dispersals from Africa, thousands of kilometres away, is unclear, unless we factor in the assumption of constant dispersal rates through Arabia and South Asia. The assumption of constant mutation rates is another potentially problematic source of chronological error. Recent studies suggest slower mutation rates (e.g., Scally & Durbin 2012), casting considerable doubt on many narratives of dispersals that are largely premised on a particular interpretation of genetic structure and a faster mutation rate. In addition to the above problems, the issue of demographic fluctuation in the Upper Pleistocene is critical and is stressed in the Middle Palaeolithic Model (see, e.g., Petraglia et al. 2010). Given the frequently difficult environments of MIS 4–2, it is likely that many local extinctions of H. sapiens populations occurred, meaning that their genetic variation will not be represented in modern populations. Likewise, rapid growth and dispersals of certain populations in the Holocene will have swamped the legacy of earlier Pleistocene dispersals.We must not assume that modern populations have remained glued to the same areas as their Pleistocene ancestors. Additionally, a dispersal rate of just 3 km a year would mean that in 2,000 years, well within error ranges of genetic coalescence age estimates, a population could have moved 6,000 km. Recent studies on genetic diversity in Arabia have suggested that contemporary populations of the Arabian Peninsula are mostly derived from post Last Glacial Maximum dispersals from western Asia, with small contributions from sub-Saharan Africa, North Africa and South Asia (e.g., ˇ erný et al. 2011). Southern Arabia has Abu-Amero et al. 2007, 2008, 2009; Cadenas et al. 2008; C higher levels of ‘ancient’ lineages than areas further north. Patterns of variation are, however, by no means simple. For instance, populations in eastern Yemen are more ‘African’ than those in western ˇ erný et al. 2008).The presence Yemen, closer to the hypothesised Bab al Mandab crossing point (C of African lineages in Arabia may relate to ancient dispersals, but in many cases such lineages have been correlated with recent historical movements, particularly those associated with the slave trade ˇ erný et al. (e.g., Richards et al. 2003; Kivisild et al. 2004; Luis et al. 2004; Abu-Amero et al. 2007; C 2008). Possible dispersal models, then, are congruent with the Arabian genetic evidence. 56

An Arabian Perspective on the Dispersal of Homo sapiens While the Upper Palaeolithic Model is based primarily on archaeological evidence, it does in addition reflect a conservative interpretation of genetic evidence. Klein (2009), for instance, reproduces the mitochondrial DNA (mtDNA) phylogram of Ingman et al. (2000), which suggests the oldest branch including both Africans and non-Africans dates to ~52 ka, but this estimate comes with a prodigious error range of ± 27,500 years. Nevertheless, some recent and more conservative reinterpretations of genetic evidence are congruent with the Upper Palaeolithic Model (e.g., Endicott & Ho 2008). The chronology of the Microlithic Model is centrally based on genetic evidence. The key studies referred to by advocates of the Microlithic Model suggest a dispersal out of Africa around 60ka (e.g., Quintana-Murci et al. 1999; Metspalu et al. 2004; Macaulay et al. 2005; Thangaraj et al. 2005). Such studies, however, are based primarily on studies in South and Southeast Asia. It is debateable what evidence from, say, the Andaman Islands, more than 5,000 km from the Bab al Mandab as the crow flies, tells us on the precise nature of dispersals into Arabia, even if we ignore various issues with genetic evidence, such as the assumption of constant mutation rates. How supporters of this model will react to recent changes in mutation rate estimates (Scally & Durbin 2012) remains to be seen. The Two-Stage Model predicts a complex genetic structure in Eurasia, of a rather different character from that which would result from a single dispersal. Such a perspective seems to be increasingly supported by geneticists, as shown by recent studies in Australia (Rasmussen et al. 2011) and Southeast Asia (Reich et al. 2011; but see Oppenheimer, this volume). Such studies, and a number of others (e.g., Ghirotto et al. 2011; McEvoy et al. 2011), strongly suggest that the peopling of Eurasia took place in multiple waves. The key aspect that differentiates this from the Middle Palaeolithic Model is the start date of dispersals into Asia, and dates are precisely the weak aspect of genetic evidence. The Middle Palaeolithic Model primarily reflects archaeological evidence contextualised in palaeoenvironmental terms. Nevertheless, some recent genetic studies are congruent with this model. The occasional presence of haplogroup L4 in Arabia (e.g., Abu-Amero et al. 2008), for instance, may reflect the existence in MIS 5 of a pre-L3-bearing population in Arabia or, perhaps more likely, reflects more recent movements. However, recent studies suggest genetic differentiation of modern humans associated with dispersal(s) into Eurasia during MIS 5. For example, in the ‘delayed expansion’ hypothesis, Gao et al. (2010) suggest that populations dispersed into southwestern Asia circa 100 ka but did not expand further until about 40 ka. It is possible that coalescence ages suggesting MIS 4 dates are systematically biased by the sex-biased nature of these dispersals, which could have actually taken place sometime earlier than proponents of the Microlithic Model suggest (Keinan & Reich 2010). Recent changes to estimates of mutation rate remove an important obstacle to the Middle Palaeolithic Model (Scally & Durbin 2012).

Archaeology Although the Arabian Peninsula is still poorly known in comparison to many areas of the world, its archaeological record perhaps offers the best potential for testing different models for human dispersals out of Africa. The elucidation of these dispersals rests on demonstrating patterns of similarity between the material culture and chronology of human populations in neighbouring regions (Figure 5.1). Advocates of the Upper Palaeolithic Model suggest that laminar industries analogous to the Upper Palaeolithic/Late Stone Age will be found in Arabia dating to MIS 3. So far such evidence is limited in the Arabian record, as discussed in a recent review by Maher (2009). Most of the relatively small number of putatively Upper Palaeolithic–like sites are known from surface contexts, and often their attribution, on typological grounds, is problematic. Edens (2001) reports a seemingly unique assemblage from the al-Faw of southern Saudi Arabia. This assemblage is 57

Huw S. Groucutt and Michael D. Petraglia Year (ka BP) 0

50

100

150

200

250

300

Oman/Yemen (‰ VPDB)

Archaeological sites Al Hatab Shi-bat Dihya 1 Jebel Qattar 1 Aybut Auwal Jebel Faya Assemblage A Jebel Faya Assemblage B Jebel Faya Assemblage C

–12

–8

–4 640 600 4

560 520

Insolation 30°N (Wm–2)

0

480 5.5

SPECMAP

–2

7.1 7.3

5.1 5.3

–1

440 7.5

0 1

1

5

7

2 0

50

100

150

200

250

300

Year (ka BP)

Figure 5.1. The correlation of paleoenvironmental change and key archaeological sequences in Arabia (illustration by the authors; environmental information adapted from Fleitmann et al. 2011).

characterised by microblades and blades seemingly produced using a soft hammer. Three cores were analysed and are congruent with the debitage.The assemblage suggests possible connections with the late Upper Palaeolithic or Epipalaeolithic in the Levant, but it is undated and from a surface context. Loosely Upper Palaeolithic–like sites are generally found in the west and north of the peninsula, and given this distribution and their rarity, we take this pattern as suggesting occasional, short-lived dispersals from the Levant in MIS 3 as well as autochthonous developments in Arabian refugia, unrelated to dispersals from Africa. The recent publication of the excavations at Jebel Faya included information on Assemblage A, which is dated to 40–38 ka (Armitage et al. 2011).This establishes an MIS 3 occupation in Arabia, but the technology represented is incongruent with the Upper Palaeolithic Model. The lithic assemblage presents a form of technology that is unknown elsewhere but is more broadly Middle

58

An Arabian Perspective on the Dispersal of Homo sapiens Palaeolithic than Upper Palaeolithic.Various reduction strategies, most commonly the working of orthogonally shaped multiplatform cores, were employed in the production of flakes, generally of small size.There is almost no blade component, faceting of platforms, or retouch.The underlying Assemblage B dates to somewhere between MIS 3 and MIS 5a and, again, does not demonstrate a form of technology that would be predicted by the Upper Palaeolithic or Microlithic Models. Its Middle Palaeolithic–like technology is focussed on flake production. Recently discovered sites in the Wadi Surdud of western Yemen also date to MIS 3 (Delagnes et al. 2012). These sites, most significantly Shi’bat Dihya 1 (SD1), which is dated to 55 ka, represent a firmly Middle Palaeolithic technology. Excavations uncovered thousands of lithics and numerous faunal remains over an area of 21 m2.Whilst ‘Levallois’-like lithics are present in the assemblage, they are only a small component, and this and other technological characteristics do not allow easy association with either contemporary African or Levantine assemblages. Isotope and phytolith analyses at SD1 suggest an open and relatively arid landscape at the time of occupation.The slightly younger SD2 is characterized by a more simple form of technology. As at Jebel Faya, the evidence from Wadi Surdud provides no evidence for MIS 3 dispersals from Africa, but rather some form of autochthonous evolutionary trajectory.The lithic technology at SD1 can be seen as an early manifestation of a long-lived southern Arabian tradition of hard hammer reduction of flat debitage surfaces to produce blades and points, elsewhere described as the Nejd Leptolithic (e.g., Rose 2006; Hilbert et al. 2012). The Microlithic Model likewise faces an absence of evidence for the type of archaeological patterning it predicts. Mellars (2006b; 2006c) suggests we should find a trail of microlithic technology along the coast linking southern Africa and South Asia. Such technology has not been identified in Arabia. Of course, some surface assemblages that have been ascribed to the early/ middle Holocene may be somewhat earlier, but their general lack of patination and the absence of specific types such as segments argues against this. Even if Pleistocene microlithic technology were to be discovered in Arabia, what would this tell us? The fact that microlithic technology developed independently in numerous spatial and temporal contexts in the Upper Pleistocene and Holocene clearly suggests that it does not make a good indicator of dispersals. It is possible that sea level rise will have submerged microlithic evidence for an MIS 4 coastal dispersal, but along most of southern Arabia the coast drops away steeply, and there is very little coastal plain. Although we should expect populations to have left some kind of evidence along the southern coastal area, such evidence at present appears to be conspicuously absent. What of the Two-Stage Model? It seems increasingly clear that dispersals from Africa into Arabia occurred in MIS 5. However, in MIS 4 and 3 there seems to be a conspicuous absence of evidence for connections with Africa, at least as can be recognised in terms of lithic technology. The evidence for this period seems to suggest instead autochthonous trajectories in regional refugia, perhaps combined with periodic population pulses from the Levant. On present evidence, the archaeological record of Arabia does not support the Two-Stage Model. It might be contended that the absence of predicted archaeological material in Arabia was because the MIS 3 phase of dispersal occurred to the north of Arabia. Even if this were the case, we should expect to see some evidence for Levantine-like Upper Palaeolithic technology spilling into Arabia. The Middle Palaeolithic Model is finding increasing support as the Arabian archaeological record becomes better known. As excavations have begun to be conducted over the past 10 years, the record of numerous surface sites is beginning to be anchored to dated reference points. At a broad level it is possible to make certain observations on the Arabian Middle Palaeolithic. For instance, many sites have a bifacial component (e.g., Zarins et al. 1981; Whalen et al. 1981; Armitage et al. 2011). This stands in stark contrast to the Upper Pleistocene Levantine record, which appears to entirely lack a bifacial component and arguably orientates Arabia as a whole towards Africa. Certain suggestions have been made about the industrial affiliation of Arabian assemblages. McClure (1994), for instance, tentatively suggests Aterian connections for one site in southwest Saudi Arabia, where he collected about 300 lithics dominated by large tanged points

59

Huw S. Groucutt and Michael D. Petraglia and scrapers (see also Beyin 2006). A recent reanalysis, however, suggests that this assemblage probably dates to the Holocene and is related to Fasad technology, with apparent similarities to the Aterian resulting from convergent evolution (Scerri 2012). Only excavation of stratified assemblages will finally address the character of this assemblage, which shares some broad characteristics with the North African Aterian. Assemblage C at Jebel Faya has been dated by optically stimulated luminescence (OSL) to 95–127 ka (Armitage et al. 2011). The assemblage consists of around 500 lithics, demonstrating a variety of reduction strategies. Of particular interest is the production of bifaces. Such types, in combination with the presence of Levallois and some volumetric blade technology, lead the Jebel Faya team to argue that the assemblage demonstrates a dispersal of Homo sapiens from North East Africa around the transition from MIS 6 to 5. Retouched pieces in Assemblage C include a variety of ‘scrapers’ and denticulates. The presence of two assemblages beneath Assemblage C is briefly mentioned by the authors, but no information is given on their technological characteristics, and no ages are provided. Surface sites in the region proximal to Jebel Faya demonstrate a broadly similar technology to Assemblage C (e.g., McBrearty 1993, 1999; see also Wahida et al. 2009). Whilst we must be mindful of possible connections to areas other than Africa, such as to the east (Petraglia 2011), the closest parallels with Assemblage C are found in Africa. In Oman’s Dhofar province Rose et al. (2011) and Usik et al. (2013) report the discovery of 260 sites characterized by a core reduction process very like that of the MIS 5 Nubian industry of northeast Africa. At one Omani site, Aybut Auwal, sediments containing a Nubian Type 1–like core and part of a point, but seemingly also relating to additional artefacts on the surface, were dated by OSL to 106 ± 9 ka (Rose et  al. 2011). The seemingly homogeneous technology in Dhofar provides provocative evidence for population connections between southern Arabia and Africa in MIS 5 and, as the authors point out, provides the first example of a specific ‘African’ technocomplex in Arabia. Sites in Oman seem to demonstrate a rather different form of technology, characterised by the presence of hard hammer blades (Rose 2006, 2007; Jagher 2009; Hilbert et al. 2012). These assemblages are characterised by unidirectional reduction, and there is little platform preparation or retouch. Geomorphological context in at least some cases suggests these assemblages are younger than the ‘Nubian’ sites and so date to MIS 5a or younger (Rose et al. 2011). Given the evidence from findings from Jebel Faya and SD1, these can be hypothesised as representing a post-MIS 5 regional autochthonous trajectory. The pre-MIS 3 picture in Yemen is still emerging, and thus the contribution of its Middle Palaeolithic record to debates on Upper Pleistocene dispersals remains to be made clear. Nevertheless, the presence of technological features rather different from those just discussed is significant. Some sites in surface contexts share certain features, such as the absence of bifacial technology, the rarity of centripetal Levallois reduction in contrast to frequent unidirectional and sometimes bidirectional Levallois (typically point) reduction (e.g., Inizan and Ortlieb 1987; Crassard 2007, 2008, 2009; Crassard and Thiébaut 2011). These characteristics may associate these assemblages with dispersals from the Levant, but such hypotheses need to be tested, and some of these assemblages demonstrate Nubian reduction techniques. At present we can merely say that the Middle Palaeolithic assemblages of southwestern Arabia are another element in the increasing complexity of Arabian Upper Pleistocene technological evolution and demography. Recent archaeological research in Saudi Arabia has lagged behind that in the south of the peninsula. An international, interdisciplinary team including the authors conducted fieldwork at Jubbah, northern Saudi Arabia in 2010 and 2011. Our initial and brief reconnaissance visit in 2010 led to the discovery of multiple archaeological sites, with the Middle Palaeolithic the most prominent period. An excavation at the site of Jebel-Qattar 1 revealed an environmental sequence extending from terminal MIS 5 to at least MIS 7 and demonstrating oscillations between arid and wetter periods (Petraglia et al. 2011). A lithic assemblage was collected at Jebel Qattar 1 that we associated with the terminal MIS 5 deposits. The assemblage is characterised by discoidal and centripetal Levallois reduction, predominantly of local ferruginous quartzite. Lithics at Jubbah 60

An Arabian Perspective on the Dispersal of Homo sapiens include Levallois and discoidal cores, points and a bifacial component. Analysis of our discoveries is ongoing, and as considerable variability seems to exist between sites, it is difficult at present to associate Jubbah with variability elsewhere in Arabia.

Developing an Interdisciplinary Perspective In addition to the evidence of the fossil, genetic and archaeological records, we can elucidate dispersal processes by exploring other factors, critically including the palaeoenvironmental context associated with the different chronologies of dispersal (see Figure 5.1). Recent studies have done much to elucidate Arabian palaeoenvironmental variation, particularly for the Upper Pleistocene (e.g., Parker 2009; Fleitmann et al. 2011; Rosenberg et al. 2011, 2012; Drake et al. 2013).The central environmental characteristic consists of the repeated oscillation between humid and arid conditions. Repeated studies demonstrate that MIS 5 in particular was a time marked by dramatically increased precipitation, in line with the Middle Palaeolithic dispersal model (e.g., Fleitmann et al. 2011; Rosenberg et  al. 2011, 2012). This alternating pattern of humidity and aridity is broadly consistent with global glacial (dry) and interglacial (wet) patterns, but there are some differences, such as the increasing evidence for humidity in Arabia in MIS 6 (Parker 2009; contra Anton 1984) and evidence for short-term environmental fluctuations. Opinions on the route out of Africa are divided between the Sinai and Bab al Mandab routes.The latter involved a water crossing, but this would have been reduced from the present 30 km – although small islands mean a minimum single water crossing of 18 km – during times of low sea level (e.g., Fernandes et al. 2006). Lambeck et al. (2011) suggest the southern Red Sea may have narrowed to around 4 km at repeated points of the Upper Pleistocene. The environmental characteristics of MIS 3 are currently the subject of some debate, being dubbed the Debated Pluvial by Parker (2009). Radiocarbon dates suggested a wet period between circa 35 and 20 ka (e.g., McClure 1976; Garrard et al. 1981; Schultz & Whitney 1986). This, however, probably reflects inaccurately young radiocarbon dates (e.g., Parker 2009). This apparent wet stage is not repeated in other records such as speleothems and marine cores, although some evidence does suggest that there were at least short wet periods. Recent redating of lacustrine deposits has shown that previously claimed MIS 3 lakes actually date to MIS 5 (Rosenberg et al. 2011). In contrast to the traditional 35–20 ka wet stage, recent studies in both the UAE and Saudi Arabia demonstrate a short- lived wet stage at circa 60–55 ka (McLaren et al. 2009; Parton et al. 2010, in press). The role and extent of such short wet phases is at present unclear. In MIS 4 Arabia witnessed a profound aridification. Although few records are known from this time, studies have demonstrated extensive dune accumulation in MIS 4, although this also reflects the availability of sediments in the Persian Gulf area at a time of low sea level (e.g., Preusser et al. 2002). Records such as speleothems indicate that this was a dry period, and rainfall was insufficient for speleothems to form in southern Arabia, and northern Arabia would have been even more arid. As a way to hypothesise population movements through a hyperarid Arabia, the ‘coastal oasis’ hypothesis of Faure et al. (2002) is widely cited. There is, however, no evidence in Arabia, or for thousands of kilometres around it, for use of marine resources in MIS 4, and whilst coastal springs may have been important in some areas, it is debateable how closely spaced they would have been. Studies have demonstrated that MIS 5, particularly MIS 5e as well as MIS 5c and MIS 5a, was a time of greatly increased precipitation across Arabia. This is clearly shown in the speleothem record of southern Arabia (e.g., Fleitmann et al. 2011), which consistently shows that MIS 5e was the wettest period of the whole of the Upper Pleistocene and Holocene, and the presence of vast lakes at this time, such as the circa 2,000 km2 Mudawwara palaeolake on the border of Saudi Arabia and Jordan (Petit-Maire et al. 2010). Other palaeolakes in southern Arabia, at Mundafan and Khujeymah, have recently been dated to MIS 5 (Rosenberg et al. 2011, 2012). Other records 61

Huw S. Groucutt and Michael D. Petraglia include the development of calcretes and palaeosols in MIS 5 at Jubbah (Petraglia et al. 2011). Such evidence parallels the amelioration of North Africa at this time (Osborne et al. 2008; Drake et al. 2011). In environmental terms, MIS 5 Arabia was a rich environment into which Homo sapiens could have dispersed into Arabia by both northern and southern routes. Another category of evidence to consider consists of biogeographical patterning. Populations of Papio hamadryas baboons in southern Arabia constitute the only natural populations of baboons found outside sub-Saharan Africa. Recent genetic studies demonstrate that the Arabian baboons appear to have dispersed from Africa in MIS 5 and 7 (Winney et al. 2004; Fernandes 2009). The lack of evidence for more recent dispersals perhaps indicates the absence of suitable windows for hominin dispersal in the Upper Pleistocene following MIS 5. Paralleling this evidence, studies of the giant clam species Tridacna costata in the Red Sea demonstrate a massive reduction in population size and body size following MIS 5, which the authors suggest demonstrates pressure from hominins (Richter et al. 2008). Further research is, however, needed to firmly separate a human role from changes in the timing and character of plankton blooming in relation to Red Sea salinity (see, e.g., Fenton et al. 2000) and factors such interspecific competition. Aside from perhaps supporting the presence of human populations around the Red Sea, this study shows that evidence for the use of coastal resources in northeast Africa actually dates to MIS 5, whilst such evidence is lacking for younger periods, although coastal ‘adaptation’ in MIS 4 is often hypothesised.

Evaluating the Models The Upper Palaeolithic Model can, perhaps, explain some of the characteristics of the Middle to Upper Palaeolithic transition in the Levant (e.g., Shea 2011e). However, as evidence from around the Indian Ocean rim accumulates, this model seems less supportable as an explanation for the broader picture of dispersals from Africa into Eurasia. It is, like other models, not strongly supported by fossil evidence. Numerous genetic lineages in Asia predate MIS 3 – even from a conservative viewpoint in terms of assumptions on mutation rates. The emerging archaeological record of MIS 3 Arabia suggests autochthonous trajectories, and nothing suggests connections to Africa after MIS 5. The Microlithic Model faces the same problems as the Upper Palaeolithic Model in terms of fossil evidence. Likewise, in terms of archaeology there appears to be an absence of compatible material in Arabia. Such statements, however, come with the usual caveats on how poorly known the Arabian record is. Much genetic evidence can still be seen as congruent with the Microlithic Model, but this seems to rapidly be changing. The insistence by many advocates of the Microlithic Model that the dispersal of Homo sapiens into Eurasia took the form of a single dispersal by a single population seems to be rapidly crumbling as new genetic evidence emerges (e.g., Rasmussen et al. 2011). Considering the problems associated with interpreting genetic evidence, it seems unfortunate to insist that dispersals occurred only within MIS 4 and 3. Surely a more parsimonious perspective is to look for periods of ameliorated environmental conditions within the error range of genetic estimates. The Two-Stage Model combines both the strengths and weaknesses of the previous two models. Given the rarity of fossils, it finds its main strength in genetic evidence. The archaeological side of this model is less developed, and there in fact seems to be a noticeable difference between African and Arabian lithic industries after MIS 5. Evidence from Jebel Faya and Wadi Surdud, the only two excavated MIS 3 localities in Arabia, both suggest autochthonous developments and not new dispersals from Africa. It should also be stated that the Middle Palaeolithic Model is not necessarily incongruent with subsequent dispersals in, for example, MIS 3, as it merely suggests that dispersals began in MIS 5. The Middle Palaeolithic Model is congruent with the earliest fossil evidence for the presence of H. sapiens outside Africa, in the Levant. Emerging genetic evidence is compatible with the 62

An Arabian Perspective on the Dispersal of Homo sapiens Middle Palaeolithic Model, whilst a variety of lines of arguments, some already discussed, such as assumptions on Upper Pleistocene demography, can explain the incongruence between genetic and emerging archaeological evidence. Joining evidence from India (Petraglia et al. 2007), stratified MIS 5 sites in Arabia are demonstrating clear connections with contemporaneous African industries. This is seen at Jebel Faya (Armitage et al. 2011), and with the discovery of more than 260 Nubian Complex sites in Oman, there are now connections to a specific spatial and temporal context in Africa (Rose et al. 2011; Usik et al. 2013). The Middle Palaeolithic Model is supported by other lines of evidence, such as the dispersal of baboons, and is associated with the environmental amelioration of the Saharo-Arabian belt (see Figure 5.1).

Conclusion Arabia is a crucial testing ground for models of dispersals out of Africa. At present the specific characteristics of the peninsula, namely the absence of a known hominin fossil record and the generally shallow age depth and complex structure of genetic lineages, necessitates a focus on archaeological evidence. There is currently no evidence supporting population connections between Africa and Arabia in the Upper Pleistocene after MIS 5, neither the microlithic assemblages predicted by the Microlithic Model nor the laminar technology predicted by the Upper Palaeolithic Model. In contrast, increasing evidence for connections in MIS 5 suggests this was a time when populations dispersed from Africa into Arabia. Considerably more research is needed to illuminate the evolutionary, demographic and cultural processes of MIS 5. Taken at face value, the record at present suggests at least two dispersals from northeastern Africa at circa 130/125 ka (Armitage et al. 2011) and circa 110 ka (Rose et al. 2011). Given wider biological and ethnographic information on dispersal processes, it is likely that the peopling of Arabia from Africa was a complicated process and should not be oversimplified to a single ‘event’.The key questions now include continuing to cast light on MIS 5 Arabia, exploring the possibility of dispersals during the short wet periods of MIS 3 and understanding the role of population refugia versus dispersals in structuring the prehistory of Arabia. At present, emerging evidence highlights the similarity of MIS 5 lithic assemblages in Arabia and Africa, during a time of climatic amelioration, followed by seeming autochthonous developments in southern Arabia in MIS 4 and 3 as climate deteriorated (see Figure 5.1).This pattern does not articulate easily with the Upper Palaeolithic, Microlithic or Two-Stage Models. On present evidence we suggest that the Middle Palaeolithic Model is the most congruent with theoretical expectations and empirical observations. It is also parsimonious in that this is when populations seem to have been expanding in Africa, whereas in MIS 4 in particular, they seem to be shrinking. Progress necessitates building a strong interdisciplinary framework and the explicit modelling of dispersal processes, which can then be tested by the procedure of falsification. The discovery of stratified archaeological and fossil localities in Arabia and elsewhere in southern Asia remains a key research priority.Whatever future discoveries show, it seems certain that Arabia will be of central importance in our understanding of the dispersal of modern humans.

Acknowledgments We thank the Saudi Commission for Tourism and Antiquities for assistance with our research in Saudi Arabia. For discussions on the role of the Arabian Peninsula in the dispersal of modern humans, we thank Abdullah Alsharekh, Rémy Crassard, Adrian Parker, Nick Drake and James Blinkhorn. We appreciate the invitation to contribute to this volume and thank Robin Dennell and Martin Porr for their helpful editorial comments. We acknowledge the financial support of the Arts and Humanities Research Council, the Leakey Foundation and the National Geographic Society. 63

Chapter 6 Assessing Models for the Dispersal of Modern Humans to South Asia

James Blinkhorn and Michael D. Petraglia

Introduction Models for the dispersal of Homo sapiens from Africa touch lightly upon evidence from the Indian subcontinent but only because of its geographic location as a stepping-stone to East Asia and onward to Australasia (Lahr & Foley 1994; Cann 2001; Oppenheimer 2009). Few, however, engage critically with the unique constellation of geographic, anthropological, genetic and archaeological phenomena that exist in South Asia, and the conflicts that arise when trying to synthesize such different lines of evidence from this diverse region (see Appenzeller 2012). This may be due, in part, to the fact that research on modern human origins in South Asia is still in its infancy. Current archaeological data regarding modern human dispersals into the subcontinent lack the levels of resolution available in other regions that are often incorporated into global narratives of dispersals from Africa. As a result, archaeological approaches to modern human origins in South Asia generally rely on using detailed evidence from a small range of excavated and dated sites to draw together data from undated excavations and surface surveys in order to provide sufficient spatial and temporal coverage. In contrast to this, South Asian populations have been broadly sampled for genetics studies, which suggest that this region has played an important role in the diversification and geographic differentiation of modern human populations (Kivisild et al. 1999, 2003; Macaulay et al. 2005; Sun et al. 2006; Atkinson et al. 2008). Given the limitations of the available data, the development of models for human dispersals from which testable hypotheses can be derived presents a means to integrate the South Asian evidence into more globalised narratives.

Models for Dispersal of Homo sapiens into South Asia Much of our understanding of the Palaeolithic of South Asia derives from research undertaken in the mid-20th century, when there were few explicit attempts to investigate modern human origins. Instead, research focused upon developing a general scheme of typo-technological development. Archaeologists often identified cultural continuity in the Palaeolithic record of South Asia (e.g., Allchin et al. 1978) and suggested that the appearance of an invasive population was not 64

Models for the Dispersal of Modern Humans to South Asia clearly marked in the stone tool record. Toward the end of the 20th century, researchers began to recognise blade-based technologies and the production of bone tools and ornaments (Murty 1979; Misra 2001). However, the association of these forms of material culture with modern humans remained implicit, and the need to define the Upper Palaeolithic of South Asia appears to relate more to the adoption of European Palaeolithic terminology in the subcontinent than to the use of trait lists for modern human behaviour that have been developed in Western Europe and Africa (Mellars & Stringer 1989; Klein 1992). Over the past decade, as the notion of modern human behaviour has received considerable scrutiny (McBrearty & Brooks 2000; d’Errico 2003; Henshilwood & Marean 2003), new approaches to identifying behavioural modernity and the presence of H. sapiens in South Asia have emerged (e.g., James & Petraglia 2005; Mellars 2006b; James 2007). Currently there are two main models for the dispersal of modern humans into South Asia, both of which draw on a diverse range of environmental, genetic and archaeological evidence. The first of these models, which we shall refer to as the Marine Isotope Stage 4–3 Boundary Model (MIS 4–3 Model), explicitly links the arrival of H. sapiens in South Asia with the appearance of crescentic microliths, backed bladelets and symbolic material culture at the Marine Isotope Stage 4–3 boundary, between 65 and 50 ka (Mellars 2006b; 2006c). The MIS 4–3 Model suggests that environmental pressures within Africa 80–70 ka promoted a suite of new ecological adaptations to increase productivity, exemplified by backed and crescentic artefacts seen in Howiesons Poort industries, which permitted demographic expansion (Mellars 2006c). It is suggested that this expansion is related to the appearance of mitochondrial (mt) DNA haplogroups (Hg) L2 and L3 (80–60 ka) (Mellars 2006c). This demographic pressure, supported by a sophisticated suite of behavioural adaptations, is argued to have permitted the earliest successful dispersal out of Africa. Patterns of mtDNA mismatch distributions and lineage expansions are used to suggest that a single dispersal of our species occurred, moving into Arabia and southern Asia by 50 ka, but not earlier than 65 ka (Macaulay et al. 2005; Mellars 2006b). Lithic industries based upon backed and crescentic artefacts are argued to have been transported from Africa to South Asia by dispersing H. sapiens groups, illustrated by typological similarities between southern and eastern African sites and industries from the subcontinent such as Batadomba Lena, Jwalapuram and Patne (Mellars 2006b) (Figure 6.1). This is supported by evidence for some of the earliest explicitly symbolic material culture occurring with the earliest microlithic technologies in the subcontinent, bearing similarities to engraved pieces and ornaments known from Howiesons Poort assemblages (Mellars 2006b). The MIS 4–3 Model, therefore, suggests that the arrival of H. sapiens in South Asia corresponds with the coalescence of Hg M and N circa 60 ka and marks a sharp technological discontinuity in the archaeological record, identified by the presence of microlithic industries and symbolic behaviour. The second model, the MIS 5 Model, proposes that the earliest colonisation of South Asia by H. sapiens occurred before the eruption of Toba in Sumatra, 74 ka, by populations using Middle Palaeolithic tool kits (Petraglia et al. 2007). It is argued that mtDNA Hg L3 populations may have dispersed from Africa between 100 and 80 ka (Cabrera et al. 2009), evidence for which may not be preserved in modern populations owing to extinction events or patterns of sampling. As a result, we observe reduced diversity of non-African L3 lineages represented by Hg M and N in modern populations (Petraglia et al. 2010). This supports the assertion of modern human authorship of Middle Palaeolithic assemblages from Jwalapuram dated to 78 and 74 ka that exceed traditional coalescence ages of Hg M and N, yet show technological affinities with Middle Stone Age assemblages from Africa associated with H. sapiens (Petraglia et al. 2007, 2010; Clarkson et al. 2012). As no genetic evidence for the MIS 5 dispersal to the Levant, evidenced from Skhu¯ l and Qafzeh, is preserved in contemporary populations, a similar scenario is postulated for South Asia. Patterns of modern genetic diversity are argued to be more complicated than previous models have acknowledged, as illustrated by the anomalous young coalescence age of Hg M in South 65

James Blinkhorn and Michael D. Petraglia

Figure 6.1.  Map of sites: (1) Site 55; (2) 16R Dune; (3) Pushkar Valley; (4) Singhbum and Wadri Atri; (5) Bamburi and Patpara; (6)Visadi; (7) Hathnora; (8) Patne; (9) Inamgaon; (10) Jwalapuram; (11) Fa Hien; (12) Batadomba Lena (illustration by the authors). Asia (45 ka), compared to Southeast Asia (56 ka), East Asia (65 ka) and Australia (76 ka) (Sun et al. 2006), which contradicts a simple rapid dispersal model with progressive founder effects (Petraglia et al. 2010). Instead, patterns of population expansion (Petraglia et al. 2009), and potentially back migration (Barik et al. 2008; Oppenheimer 2009), are highlighted as factors that may reduce coalescence age estimates for the earliest dispersals of H. sapiens into South Asia. Similarly, a rapid, coastal dispersal is brought into question, and this model instead highlights the potential importance of continental routes and resources for dispersing populations (Petraglia et al. 2010). Finally, evidence for rapid population expansions and the deterioration of environmental conditions from 35 ka onwards appear to have created increased population pressure, and it is in this context that an autochthonous innovation of blade and bladelet technologies, including backed and crescentic artefacts, occurs (Petraglia et al. 2009). 66

Models for the Dispersal of Modern Humans to South Asia The MIS 5 Model therefore suggests that H. sapiens arrived in South Asia by 78 ka, although these earliest genetic lineages may not have survived into the modern population; population expansions identified in the genetic record may mask the time depth of coalescence of the earliest lineages that are preserved in the modern population; the earliest populations of H. sapiens were associated with Middle Palaeolithic technologies; and the development of blade and bladelet technologies is a local development from Middle Palaeolithic industries, related to the environmental and demographic context of South Asia. The two models make specific inferences regarding the timing of human dispersals, the route(s) they followed, the technologies employed by dispersing populations and their impact upon modern genetic diversity, summarised in Table  6.1. We shall present the first direct comparison of these models for modern human dispersals into South Asia and investigate the validity of these hypotheses against the currently available body of evidence. In doing so, we hope to highlight where contradictions occur in the models, allowing us revise and refine the them, and ultimately target future research to corroborate or falsify particular hypotheses.

The Fossil Hominin Record of South Asia Because of the paucity of Pleistocene hominin specimens from South Asia, fossil evidence is not a major feature of either model.The only pre-sapiens hominin fossil from the subcontinent, a calvarium from Hathnora in the Narmada Valley, can most safely be identified as mid-Pleistocene Homo (Athreya 2010), the dating of which is problematic but is likely between 83 ka and ≥ 236 ka (Kennedy 2000; Cameron et al. 2004; Patnaik et al. 2009). The earliest specimens of H. sapiens derive from the cave sites of Fa Hien (35 and 28 ka) and Badatomba Lena (28.5 ka) in Sri Lanka (Kennedy and Deraniyagala 1989; Deraniyagala 1992). Skeletal evidence of H. sapiens is not present on the South Asian mainland until after the Last Glacial Maximum at Jwalapuram 9 in southern India (16–20 ka) (Clarkson et al. 2009) and remains scarce until the Holocene. As a result, we cannot directly corroborate the timing for the arrival of modern humans indicated by either model. However, higher levels of skeletal diversity can be observed between early-mid Holocene hunter-gatherer and agro-pastoral groups than are observed today (Lukacs 2007). This suggests our understanding of diversity from modern genetic samples may not accurately depict that of the range of early Holocene populations, let alone earlier, Pleistocene populations. If this is the case, ancient DNA studies have the potential to expand considerably our understanding of the population diversity of the subcontinent.

Current Models and the Human Skeletal Record Both the MIS 4–3 and MIS 5 models refer to fossil skeletal evidence from elsewhere. The MIS 4–3 Model suggests that the earliest evidence for modern humans beyond Africa, seen at the sites of Qafzeh and Skhu¯ l between 130 and 70 ka, marks only a short-lived and localised expansion beyond Africa (Mellars 2006b) and hence a ‘failed dispersal’ (Oppenheimer 2012a). In particular, the reliance on ‘archaic’ technologies is considered inadequate to have withstood the southward migration of the Neanderthal populations that replaced H. sapiens in the Levant by circa 70 ka (Mellars 2006b). The MIS 4–3 Model also suggests that a coastal dispersal circa 60 ka led to the rapid colonisation of Sunda and Sahul. This may be indicated by the modern human cranium from Tam Pa Ling, Laos, circa 50 ka (Demeter et al. 2012), the skull from Niah Cave, Borneo, circa 40–42 ka BP (Barker et al. 2007) and the burials at Lake Mungo, Australia, that are dated to 50–40 ka (O’Connell & Allen 2004), which broadly corroborate the suggested timing for exiting Africa. However, the lack of ‘advanced’ (i.e., microlithic) lithic technologies with the earliest human skeletons at Niah and in Australia is considered the “greatest enigma in the current archaeological record” (Mellars 2006b, 798). 67

James Blinkhorn and Michael D. Petraglia Table 6.1.  Summary of model predictions MIS 4–3 Model Ecological

MIS 5 Model

A rapid coastal dispersal most parsimonious with archaeological and genetic evidence Earliest dispersal from Africa to South Asia indicated by coalescence of mtDNA Hg M and N between 65 and 50 ka

Continental routes of dispersal offer humid corridors and wide resource base during MIS 5 Genetic Earliest dispersal from Africa to South Asia may not be preserved in contemporary population genetics; later population expansions mask genetic signatures of earliest dispersals Middle Palaeolithic technologies Archaeological: Use of crescentic and backed are sufficient for early dispersal Ecological efficiency technologies (e.g., Howiesons to the Levant 130–70 ka Poort) offer a competitive edge (e.g., Skhu¯l and Qafzeh) and over ‘archaic’ technologies South Asia by 78 ka (e.g., (e.g., at Skhu¯l and Qafzeh), Jwalapuram and 16R Dune) permitting a successful dispersal from Africa Technological similarity Archaeological: Typological similarity between between South Asian Middle Cultural heritage Howiesons Poort and South Palaeolithic and sub-Saharan with Africa Asian blade and bladelet Middle Stone Age industries, industries, including symbolic including Howiesons Poort designs Technological discontinuity Archaeological: Technological discontinuity may between Middle Palaeolithic Technological occur between Late Acheulean continuity in South and industries containing and Middle Palaeolithic Asia cresecentic and backed industries; continuity between artefacts Middle Palaeolithic and blade and bladelet industries Archaeological: Crescentic and backed artefacts Crescentic and backed microliths Appearance of appear in South Asia as a result in South Asia appear as the crescentic and of translocation by dispersal of result of local innovation and backed artefacts H. sapiens from Africa development from Middle Palaeolithic circa 35 ka The MIS 5 Model suggests that the occupation of the Levant during the last interglacial by H. sapiens cannot simply be refuted as a failed dispersal, matching the length of time the MIS 4–3 Model suggests H. sapiens have successfully dispersed beyond Africa. Over a 60,000 year period, H. sapiens may have had repeated opportunity to expand beyond the Levant and colonise South Asia, using a Middle Palaeolithic lithic technology. In contrast to the MIS 4–3 Model, the lack of ‘advanced’ lithic technologies with the earliest H. sapiens beyond South Asia is unproblematic for the MIS 5 Model.

Routes of Dispersal Understanding the nature and availability of potential routes into South Asia is critical to investigating dispersals. Many models map broad routes of dispersal (Oppenheimer 2009), particularly southern, coastal routes, but few are explicit as to specific patterns of movement. The broad, 68

Models for the Dispersal of Modern Humans to South Asia mountainous chain of the Himalayas, Karakorum and Hindu Kush that separates South from Central Asia provides a formidable barrier to dispersal and makes an east-west axis the only plausible route for hominin dispersals from Africa. Variation in environmental conditions in South Asia throughout the Late Pleistocene is likely to have played an important role in structuring freshwater resources, which may have significant impacts upon the viability of different routes of dispersals. Patterns of humidity and aridity are dominated by waxing and waning of monsoonal intensity, relating to both the global glacialinterglacial cycles and the impact of orbital precession and eccentricity upon insolation for the Arabian Sea. Broadly, glacial periods, such as MIS 2, 4 and 6, are noted for heightened aridity and interglacial periods, including MIS 5 and 1, are marked by peaked humidity (e.g., Jain et al. 2005; Juyal et al. 2006; Dhir et al. 2010; Singhvi et al. 2010). MIS 3 is unusual as patterns of insolation appear to offset the impact of glacial conditions experienced at a global scale, suggesting conditions were cooler than interglacial phases but more humid than true glacial conditions. The Indus River and the Thar Desert have been identified as two potential barriers to eastward dispersal into South Asia (Field et al. 2007; Petraglia et al. 2012a). Repeated iterations of a geographic information system (GIS)-based model of potential coastal dispersal routes into South Asia during MIS 4 saw ‘migrant populations’ diverted up the Indus corridor, rather than crossing the broad Indus delta (Field & Lahr 2006; Field et al. 2007). In reality, crossing the Indus may not have significantly slowed population dispersals, but the rich and reliable resource base available in the Indus corridor may have provided a significant distraction from a rapid, coastal dispersal if driven partly by resource depletion (as suggested by Oppenheimer 2012a). The Thar Desert presents a different kind of barrier to dispersals, as the extreme aridity experienced in MIS 4 resulted in a landscape that lacked the freshwater resources required to support hominin populations. However, during MIS 5 there is evidence for river systems penetrating the arid core of the Thar (e.g., Jain et al. 2005), indicating clear potential for hominin occupation during periods of enhanced humidity. The start of MIS 3 is marked by high levels of sand dune mobility (Andrews et al. 1997), but after 50 ka fluvial channel reactivation, widespread soil development and interdunal lake horizons indicate the return of more humid conditions (Jain et al. 2005) that may once more have been able to support hominin populations. Analysis of coastal dispersal routes around the peninsula saw modelled populations repeatedly turn towards the interior at the mouths of rivers, including the Ganges-Brahmaputra Delta where this analysis saw ‘populations’ divert back toward South Asia rather than pass onward to Southeast Asia (Field et al. 2007). High levels of Holocene sedimentation prevent a clear understanding of the character of now-submerged South Asian Pleistocene coastlines and their resource structures. However, geological and geographic analysis of the subcontinent has led Korisettar (2007) to highlight seven inland ‘Purana’ basins, exhibiting a suite of characteristics that probably provided more suitable habitats for Pleistocene hominins than the coasts. The geology of these basins include aquifers that, during periods of increased aridity, deliver summer monsoon rainfall from the Western Ghats to the central and eastern basins in the forms of springs and streams, while also receiving direct rainfall from the winter monsoon (Korisettar 2007).These basins would have provided refugia during periods of climatic decline, as well as centres for population growth in periods with more favourable environmental conditions.

Current Models and the Timing and Routes of Dispersals The MIS 4–3 Model suggests a single, coastal dispersal via the Bab el Mandeb strait circa 60 ka, supported by the assertion that this would require only limited socio-economic adaptations from one coastal location to the next (Sauer 1962) and be driven onwards by resource depletion. The expanded Thar Desert appears to have formed an arid barrier to dispersal during MIS 4, and although fluctuating conditions at the start of MIS 3 may not prevent a dispersal, instability of 69

James Blinkhorn and Michael D. Petraglia freshwater resources for either coastal or inland populations may not encourage populations to expand into new areas. A rapid dispersal along the coastline of South Asia during even humid phases may have been significantly slowed by the attraction of resources available in the peninsular river systems. Conversely, the extreme humidity, sharp relief terrain and dense rainforests of northeast India may also have discouraged further rapid, eastward dispersals, as even coastal groups may rely on access to continental sources of freshwater and raw materials. The MIS 5 Model suggests the use of various continental corridors into South Asia during MIS 5 and emphasises the importance of river valleys for movement (Petraglia 2005). Although such continental dispersals and colonisation events may have required diverse adaptations, populations would have sought reliable access to fresh water, lithic resources and favourable habitats, such as woodland and grassland eco-zones with related fauna (Petraglia et al. 2010). The integrated riverine network that existed in the Thar Desert during MIS 5 may have provided a suitable inland route of dispersal for human populations, regardless of whether they were previously based in coastal or upland regions. However, upland routes into South Asia that skirted the Thar Desert by dispersing into the Siwalik foothills and onward toward the Gangetic Plains cannot be ruled out. The ‘Purana’ basins of South Asia appear to offer rich contexts for human occupation, despite extreme climatic events such as the Toba eruption; because they are also most likely to have been inhabited by archaic hominins, they presented further adaptive challenges for colonisation of landscapes that were already occupied.

Genetic Evidence Over the past 30 years, biomolecular evidence has provided numerous significant insights into human evolution, particularly the likely recent African origin of H. sapiens (Cann et  al. 1987; Relethford 2008). The majority of this research has focused on mitochondrial and Y chromosome DNA, offering insights into uniparental lineages, but recent advances in sequencing technology and analysis have resulted in a new suite of evidence from nuclear genome studies, from which complementary patterns in the modern gene pool can be identified. Genetic evidence forms an important basis for inferences in both models, but caution is required to integrate this evidence from modern populations with Late Pleistocene environmental and archaeological data. Particular caution must be exercised in the interpretation of genetic age estimates for the expansion of modern humans from Africa, as these have been based upon a mutation rate that is twice that directly observed in contemporary populations (Scally & Durbin 2012). In spite of this, genetic evidence can provide some important insights into how, when and potentially where modern patterns of genetic diversity became defined. Studies of the nuclear genome currently offer a broad time span for the divergence between African and the ancestral Eurasian population, with a recent estimate of 112–88 ka and a confidence interval spanning 150–63 ka (Gao et al. 2010). Several genomic studies have suggested that at least two dispersals from the proto-Eurasian population are required to explain the observed genetic variation in modern populations. McEvoy et  al. (2011) suggest patterns of Linkage Disequilibrium highlight continued ancient (African) gene flow with proto-Europeans subsequent to their divergence with proto–East Asians, indicating at least two dispersals from the proto-Eurasian population. An ancient DNA study based on an Australian Aboriginal hair sample collected in the early 20th century suggests that Australian populations were the product of the earliest dispersals from Africa circa 75–62 ka, with a genetically distinct second wave identifiable in East Asian populations circa 38–25 ka (Rasmussen et al. 2011). Similarly, a survey of Southeast Asian populations for evidence of Denisovan admixture also separates populations into at least two waves of migration (Reich et al. 2011). The coalescence of mtDNA Hg L3, currently estimated to have occurred circa 72 ka, with a confidence interval spanning circa 56–87 ka (Soares et al. 2009), sets a mitochondrial benchmark 70

Models for the Dispersal of Modern Humans to South Asia for modern human dispersals from sub-Saharan Africa. L3 is the youngest node containing both sub-Saharan African and non-African sequences; the dispersal from Africa must have occurred after the differentiation of these populations from Hg L populations (Maca-Meyer et al. 2001; Gonder et  al. 2007). As all non-Africans may preserve evidence for interbreeding with Neanderthals (Green et al. 2010), this is likely to have occurred within the proto-Eurasian population, prior to the differentiation of Hg M and N. Given what is known about the distribution of Neanderthals, population contacts are likely to have occurred in southwestern Asia. The coalescence of Hg M and N should, therefore, occur after the dispersal from Africa. The Indian subcontinent currently houses one-sixth of the world’s population, but it has been suggested that in the Late Pleistocene more than half of the total population of modern humans lived in South Asia (including Southeast Asia), peaking at circa 60% by 38 ka (Atkinson et al. 2008). Furthermore, South Asia shows the earliest and fastest rates of population expansion amongst non-African lineages, with a five-fold increase in population size occurring by circa 52 ka (Atkinson et al. 2008). Over the past decade, the number of basal or deeply rooted haplotypes specific to South Asia, particularly for Hg M, has increased dramatically (Kivisild et al. 2003; Sun et al. 2006; Chandrasekar et al. 2009; Morlote et al. 2011). Thus far, more than 30 South Asia specific lineages are known to emerge directly from the root of Hg M, suggesting their local origin (Chandrasekar et al. 2009). Comparisons between India and Iran show a dramatic genetic cline, with more than 70% of Indian lineage samples belonging to Hg M, whereas only circa 5% of Iranian populations belong to Hg M, with a much higher proportion of West Eurasian lineages present (Metspalu et al. 2004). Because of the ubiquitous presence of deeply rooted M lineages throughout India, it is suggested that M may have first arisen, or at least become distinct from the proto-Eurasian source population, in Indian populations (Chandrasekar et al. 2009). This may be supported by the relatively large number of Indian specific lineages with individual coalescence ages greater than 60 ka (Chandrasekar et al. 2009). In contrast to this, South Asia appears to have played a more limited role in the differentiation of Hg N. This group is represented in the subcontinent primarily by regionally specific lineages of Hg R, and its derivative Hg U, and is observed at the highest frequency in the north (Maji et al. 2008). In particular, groups R5–7 and U2i are thought to be autochthonous (Maji et al. 2008), and the coalescence of clusters of South Asian Hg N variants occurs circa 45 ka (Barnabas et al. 2006). The distribution of Hg R and U in West Asia and Europe has led to suggestions that the presence of these lineages in South Asia is the result of a northern dispersal from Africa, occurring as a separate population from the migrants with M lineages (Barnabas et al. 2006).

Current Models and Genetic Coalescence Data Contrary to the MIS 4–3 Model, mounting evidence suggests that more than one dispersal has occurred from a proto-Eurasian population, one subset of which had continued contact with African groups after the earlier dispersal of another population (McEvoy et al. 2011). As a result, the coalescence of Hg M and N, circa 60 ka, may not provide a sound indication of when populations dispersed, and the similarity in ages between numerous M and N lineages may reflect similar selective pressures leading to lineage extinctions during MIS 4 or population expansions in MIS 3. The growing number of age estimates of greater than 60 ka for the coalescence of Hg M from South Asian specific lineages, as well as those further east, provides increasing support for a successful dispersal during MIS 5 into the subcontinent and beyond, as proposed by the MIS 5 Model. In contrast, age estimates for Hg N lineages from South Asia remain at circa 45ka, and support a model of at least two dispersals from the ancestral Eurasian population. It is in the context of a potential second dispersal to South Asia that some studies have highlighted rapid rates of population growth within South Asia during MIS 3 (Sun et al. 2006; Atkinson et al. 2008; 71

James Blinkhorn and Michael D. Petraglia Petraglia et al. 2009). This may provide support for the suggestion that increasing demographic pressures in MIS 3 were an important factor for the regional development of blade-based and microlithic technologies.

Archaeological Evidence Archaeological evidence provides the richest source of data from the Late Pleistocene with which to investigate modern human origins in South Asia.Whereas fossil and genetic evidence provides very clear and specific links with biological entities, such as H. sapiens, archaeological studies focus on cultural behaviour and adaptation. Linking particular stone tool industries to specific hominins is inherently problematic, particularly as demographic and environmental circumstances appear to play a significant role in patterns of cultural diversity (Powell et al. 2009). Although the relationships with particular hominins remain unclear, patterns of continuity, or at least gradual change, within archaeological assemblages are more likely to be the result of cultural descent within a population, whereas marked technological departures between industries suggest a significant break in cultural descent, potentially relating to a change in population. Late Acheulean industries in South Asia are typified by the presence of Large Cutting Tools that are notably smaller than those of the Early Acheulean, such as diminutive handaxes, and an increased use of prepared core technologies for flake production (Chauhan 2009; James & Petraglia 2009). Archaeological investigations in the Son Valley have identified some of the youngest known Acheulean sites known globally, with Late Acheulean industries from the MIS 6–5 boundary at Patpara and Bamburi (Haslam et  al. 2011). Late Acheulean industries are broadly distributed within South Asia, although some areas lack evidence for occupation, such as the Ganges Valley, the Western Ghats and all but the eastern margin of the Thar Desert, although site visibility may be a significant issue. Extensive Late Acheulean occupation occurs within the central and southeastern river valleys of the subcontinent and the more mountainous landscapes of the Aravallis and Siwaliks. Middle Palaeolithic assemblages are differentiated from preceding Late Acheulean industries principally by the scarcity of Large Cutting Tools (diminutive handaxes and cleavers), the diversification and decrease in size of prepared core technologies, the production of blades and more focused use of higher quality raw material resources (James & Petraglia 2005). In particular, Levallois and discoidal core types become much more prominent in the Middle Palaeolithic, although not ubiquitous (James 2007). It is suggested that blade production techniques became more intensive through time, with higher numbers of blade removals from each core, as observed at Patne (James & Petraglia 2005). Scrapers dominate flake tool components in most assemblages, but the presence of blades, knives, points, tanged points, backed artefacts and borers is spatially and temporally variable (James 2007).The earliest reported Middle Palaeolithic technologies appear at 16R Dune, recently re-dated to circa 130–109 ka (Misra 1995b; Singhvi et al. 2010), but the lithic assemblages are poorly reported. The best-studied and most securely dated Middle Palaeolithic industries come from Jwalapuram, displaying considerable technological continuity from 78 to 38 ka, although there are gaps in this sequence (Petraglia et al. 2007; Petraglia et al. 2012b). Subsequent South Asian industries are dominated by the production of blades, but variation can be observed between areas that focus on production of macroblades (e.g., Watru Abri and Singhbum) or bladelets and geometric microliths (e.g., Patne and Batadomba Lena, Jwalapuram 9)  and those containing a mix of these elements (e.g., Visadi, Inamgaon & Site 55) (James & Petraglia 2005). These have generally been described as Upper Palaeolithic, although microblade industries are clearly present by 35ka and are referred to as Microlithic. Proportions of prismatic blade and microblade cores appear to gradually increase in Upper Palaeolithic assemblages, as at Patne (James 2011), as do levels of reduction intensity and preferential selection of high-quality raw materials throughout time, as at Jwalapuram 9 (Clarkson et al. 2009). In addition to lithic 72

Models for the Dispersal of Modern Humans to South Asia industries, a range of new archaeological materials appear in Upper Palaeolithic/Microlithic sites, such as the production of bone tools, the manufacture of beads and the decoration of ostrich eggshell, as seen at Patne (Sali 1985), although these occurrences are infrequent and show no clear spatial or temporal pattern.The earliest microblade industries reported come from Sri Lanka circa 38 ka (Perera et al. 2011), and Jwalapuram circa 35 ka (Clarkson et al. 2009; Petraglia et al. 2009). Further north, Upper Palaeolithic/Microlithic industries tend to date less than 30ka, such as at Patne circa 25 ka and Inamgoan circa 21 ka (James & Petraglia 2005).

Current Models and Cultural Transitions The MIS 4–3 Model predicts that we should be able to identify the rapid demise of ‘archaic’ material culture and the arrival of the ‘modern African’ package of behaviours from 65 to 50 ka onwards. However, no such revolution is visible in the archaeological record of South Asia. The development of both systematic blade production and symbolic behaviour appear spatially and temporally disjunct, with clear evidence for the local development of Microlithic industries from the Middle Palaeolithic at Jwalapuram by 35 ka (Clarkson et al. 2009) and at Patne by 25 ka (James 2011). In terms of lithic technologies, the MIS 5 Model predicts some technological discontinuity between Late Acheulean and Middle Palaeolithic industries. However, as both prepared core technologies and Large Cutting Tools occur in both types of industry, a more nuanced approach than is currently possible may be required to separate these industries and the cultural associations of South Asian hominins, further complicated by the potential for biological and social admixture. Further comparisons between Middle Palaeolithic assemblages from Jwalapuram and subSaharan African Middle Stone Age industries, including the Howiesons Poort, confirms earlier evidence that these industries share greater technological affinities with one another than either do with the microlithic industries from Jwalapuram (Clarkson et al. 2012). However, the observed continuity within Middle Palaeolithic industries up to 38 ka, and between Middle Palaeolithic blade and bladelet industries, certainly at Patne and Jwalapuram, suggests microlithic technologies were developed by South Asians from Middle Palaeolithic industries, which may indicate population continuity (James 2011; Clarkson et al. 2012).

Evaluation of Models for Modern Human Dispersals to South Asia Despite using similar bodies of evidence, and at times similar theoretical approaches, the Microlithic Model and Middle Palaeolithic Model present very different predictions regarding the timing, nature and evidence required to identify the earliest presence of modern humans in South Asia. A range of new evidence has come to light since these models were first put forward that permit us to evaluate them, summarised in Figure 6.2. The single, rapid, coastal dispersal of modern humans from Africa to South Asia circa 65–50 ka and evidenced by crescentic and backed microliths, as proposed by the MIS 4–3 Model, cannot be supported by the current evidence. Suggestions of a coastal dispersal cannot be based on evidence due to the lack of Late Pleistocene coastal sites, partly, though not wholly, the result of sea level change. Recent genetic evidence is not reconcilable with a single dispersal event of modern humans from the proto-Eurasian population into Asia, and evidence for at least two dispersals into South Asia may be apparent in the divergent patterns of geographic distribution and coalescence age estimates for region-specific lineages of Hg M and N (McEvoy et al. 2011; Rasmussen et al. 2011; Reich et al. 2011). Typological similarities between Howiesons Poort and South Asian microlithic industries do not clearly show a pattern of cultural descent from one to the other. However, the autochthonous development of microlithic technologies in Australia during the mid-Holocene, occurring in a context of high population levels and deteriorating 73

James Blinkhorn and Michael D. Petraglia MIS

–2

0

1

2

–1

δ18O (‰ SMOW)

Genetic

Archaeology

0

Key

1

SPECMAP oxygen isotope curve

20

2

Indian Monsoon Index

40

Thar Desert sand dune mobility

3

A

B

C Microlithic

60 80 Age (ka)

4

Middle Palaeolithic Late Acheulean

100

MIS 4-3 Model predicts earliest H. sapiens in South Asia

5

MIS 5 Model predicts earliest H. sapiens in South Asia

120

MIS 5 Model suggests earliest potential H. sapiens in South Asia

140

6 230

220

210

200

190

180

Insolation difference (w/m2)

Figure  6.2.  Synthesis of recent environmental, genetic and archaeological evidence relating to the dispersal of Homo sapiens into South Asia, and the timing of arrival of H. sapiens as predicted by the MIS 4–3 and MIS 5 models. Left: SPECMAP oxygen isotope curve and Indian Monsoon Index (adapted from Leuschner & Sirocko 2003), with phases of Thar Desert sand dune mobility and Marine Isotope Stages. Centre: (A) Genomic estimate of split of Eurasian from African populations 112–88 ka (C.I.150–63 ka) (Gao et al. 2010); (B) mtDNA estimate of appearance of Hg L3 at 72 ka (C.I. 87–56 ka) (Soares et al. 2009); (C) genomic estimate of split of East Eurasians from proto-Eurasian group (Rasmussen et al. 2011). Right: Archaeological phases associated with robust chronometric dates (illustration by the authors). environmental conditions (Lourandos 1997), suggests that technological convergence can explain such typological similarities. The occurrence of microlithic technologies in the Indian subcontinent 10,000 years after the appearance of modern humans at Niah Cave and Lake Mungo circa 45 ka suggests they were not employed by the earliest dispersal of modern humans. Indeed, the development of blade and bladelet technologies in South Asia does not appear to directly relate to major population dispersals in the Late Pleistocene. The suggestion that the earliest modern humans in South Asia appear in MIS 5 with Middle Palaeolithic technologies is not entirely unproblematic, but a range of predictions from the MIS 5 Model can be supported to some extent by the available evidence. Coalescence age estimates for the earliest dispersals from Africa are gradually increasing, with some exceeding the age of the early Middle Palaeolithic evidence from Jwalapuram (e.g., Chandrasekar et al.2009). Suggestions of an earlier dispersal resulting in the Middle Palaeolithic horizons at 16R Dune 130–109 ka remain unsupported by genetic evidence, although palaeoenvironmental data suggest that humid continental corridors may have facilitated dispersals at this time. Identifying such a dispersal, as well as differentiating Late Acheulean and early Middle Palaeolithic industries, is currently problematic owing to the lack of analysed assemblages prior to 78 ka. The relationship between early and late Middle Palaeolithic assemblages is well established at Jwalapuram, and the gradual 74

Models for the Dispersal of Modern Humans to South Asia intensification of blade production both here and at Patne provides robust evidence for the regional development of microlithic industries, yet most Middle Palaeolithic sites occur more than 60 ka, another significant gap in the archaeological record to support this model. What is clear from this assessment is that more detailed archaeological research focused on the Late Acheulean to Middle Palaeolithic and Middle Palaeolithic to the Upper Palaeolithic/ Microlithic transition is required to support either model.While genetic research may offer further evidence for numerous dispersals from the proto-Eurasian population, more-intensive sampling strategies could elucidate deeply rooted lineages in support of early dispersal models. Currently, the lack of widespread continental palaeoenvironmental records spanning the Late Pleistocene prevents more detailed analysis of how the variety of habitats that occur across the subcontinent responded to climatic change and how this may have affected their hominin occupants.

Conclusion Despite the differences between the MIS 4–3 Model and MIS 5 Model, some degree of agreement can be identified between them. Both models suggest that complex demographic and environmental interactions can provide suitable contexts for the innovation of technologies, including microliths and symbolic behaviour, an approach that now has broader theoretical support (e.g., Powell et al. 2009). Similarly, both models suggest, to some degree, that patterns of cultural transmission may be informative for investigating population dispersals. Indeed, the fact that neither model neatly correlates with the available genetic evidence highlights the fact that cultural change operates in a significantly different manner to genetic inheritance. Identifying this common ground between the models potentially enables a new theoretic focus on the small, huntergatherer populations that passed technological knowledge on to new generations or allies rather than the more obscure biological populations that can be identified by geneticists. The process of model building is undertaken as a means to develop theoretical frameworks that can integrate interdisciplinary evidence and produce hypotheses that can be tested against different types of evidence or to identify where new research is required to differentiate between different models. Because of the nature of the various types of available evidence employed by the models, this assessment remains at something of a grand scale, drawing on genetic variations from thousands of individuals and spanning the entire subcontinent over a period of circa 100 ka. The success of model building occurs only with the rejection of hypotheses and formation of new, more precise models. As such, the predictions that are possible from our current models are a far cry from the microevolutionary processes that were at work as the earliest modern human populations began their gradual expansion into the exotic habitats of South Asia.

Acknowledgments Our research in India has been supported by the Archaeological Survey of India and the American Institute of Indian Studies. We have been generously funded by grants from the British Academy, the Leverhulme Trust, the Leakey Foundation, the Royal Anthropological Institute and the Kathleen Hardy Scholarship (St. Hugh’s College, Oxford). Numerous individuals have assisted us in our research in South Asia, and we would especially like to thank Hema Achyuthan, Ravi Korisettar and J. N. Pal.

75

Chapter 7 East of Eden Founder Effects and the Archaeological Signature of Modern Human Dispersal

Christopher Clarkson

Introduction The archaeological signature for the first arrival of modern humans in Europe has been a focus of intensive research and for certain regions is now relatively well understood. The archaeological signature for modern human arrival in Arabia, India and much of Asia, on the other hand, is much debated despite the fact that genetic research provides a temporal window for colonisation of 60–75 thousand years ago (kya) (Oppenheimer 2009, 2012b, this volume; Sun et al. 2006; Rasmussen et al. 2011). Indeed, the issue of modern human colonisation east of Africa is emerging as one of the most important and hotly debated topics in Palaeolithic studies (Appenzeller 2012; Balter 2010). Unresolved questions include the exit route from Africa (e.g., Nile corridor or Horn of Africa), the date at which modern humans arrived in each region, the speed at which they dispersed, the alterations to subsistence and technology required at each step of the journey and the extent of cultural and biological interaction between hominin species encountered en route (Beyin 2006; Field & Lahr 2005; Foley & Lahr 1997; Petraglia et al. 2010; Peer 1998; Vermeersch 2001). Such questions are unanswered, and vast lacunae exist in the archaeological records of regions such as Arabia, India and Southeast Asia. The story of modern human spread and cultural change east of Africa therefore remains poorly sketched, and major challenges lie ahead in fleshing out this crucial period in human evolution. Recent analyses of modern and ancient human DNA have made important inroads into solving this puzzle, documenting relationships between populations, approximate ages for branching events, likely dispersal pathways, and inter-species admixture and even identifying hitherto unknown species (e.g., Green et al. 2010; Liu et al. 2006; Oppenheimer 2009, 2012b; Rasmussen et al. 2011; Reich et al. 2011). Unfortunately, genetics research cannot yet provide precise dates for branching events or locate the geographic nodes at which such events took place. The fossil record obviously provides the acid test of modern human presence in a region, but unfortunately this record is sparse and often ambiguous in regions east of Africa. A single burial at Lake Mungo provides a terminus ante quem for human arrival in Australia by 40 kya (Bowler 76

The Archaeological Signature of Modern Human Dispersal & Price 1998), while evidence from the north of Australia at sites such as Malakunanja and Nauwalabila points to initial occupation perhaps as early as 65 kya (Roberts et al. 1990). Sites in the Philippines and China also hint at modern human arrival prior to 45 kya (Mijares et al. 2010; Pawlik et al., this volume; Shen et al. 2002).The Australian and Asian archaeological records therefore provide a chronological benchmark from which to infer earlier settlement of regions between Africa and Sahul where modern human skeletal remains are absent before ~30 kya (Kennedy & Deriyanagala 1989; Deriyanagala 1992; Field & Lahr 2005). This begs the question then of how we can identify early modern human settlement in the absence of a rich human skeletal record for these regions. To answer this question, archaeologists have often turned to lithic technology – or, in other words, the material culture signature of these founding populations  – for clues. While this approach is generally successful in Europe where modern human arrival is accompanied by distinctive Upper Palaeolithic lithic and osseous technologies, it is unclear whether such a distinctive signature exists east of Africa. For instance, archaeologists are now debating whether modern humans dispersed eastwards from Africa with distinctive microlithic/microblade (Mellars 2006b) technology complete with African symbolic elements, such as cross-hatched motifs, perforated beads and osseous point technology, or with a less distinctive Middle Palaeolithic/Middle Stone Age (MSA) Levallois/discoidal core and scraper tradition (Armitage et al. 2011; Clarkson et al. 2009, 2012; Petraglia et al. 2009, 2010; Rose et al. 2011). Technological comparisons are also hampered by a lack of detailed lithic analysis in many cases and the use of regional typologies (Vermeersch 2001). This debate is difficult to resolve in the absence of numerous well-dated sites with intact skeletal remains in regions east of Africa. However, one way to begin to tackle this problem is to review the lithic assemblage variability from sites east of Africa and determine whether any such signature for modern human colonisation can be discerned. This chapter therefore reviews the nature of lithic technology between Africa and Sahul for a critical temporal window of 45–100 kya. Sites that are not well dated but are suggestive of an antiquity greater than 45 kya are included in this review. While climate change likely played an important part in the timing of human dispersals by opening and closing potential dispersal corridors and by acting as a push or pull mechanism, climate change is beyond the scope of this chapter and is not reviewed in detail here.

Africa Africa is the birthplace of modern humanity and provides increasingly clear evidence of early complexity in all realms of cultural expression. Crucial to understanding and tracing the dispersal of modern humans out of Africa is a detailed inventory of the technological variability that existed in Africa at the time of exit. MSA lithic technology is diverse and varies markedly through time and space, complicating any attempts to identify an obvious homeland for such a migration. Distinct technological phases include the Still Bay and Howiesons Poort of southern Africa, the Aterian of North Africa and a range of other chronologically and technologically more or less discrete industries throughout Africa falling within this 45–100 kya period. Common forms of African MSA core technology include recurrent preferential, centripetal and bidirectional Levallois technology, single and multiplatform core reduction and bipolar reduction. Distinctive core technologies include Nubian point production in northeastern Africa and the Horn of Africa. Large blade manufacture is a distinctive feature of MSA1 sites along the southern coast of South Africa at sites such as Klasies River Mouth and Nelson Bay Cave but is rare elsewhere in Africa during the period of interest. Microblade technology is arguably a component of some Howiesons Poort sites but is generally rare in southern Africa (Clarkson 2010; Soriano et al. 2007). 77

Christopher Clarkson Retouched technologies are also regionally distinctive. Backed artefact manufacture is most common in eastern and southern Africa but dramatically oscillates in frequency at many sites where this technology occurs, with peaks in backed artefact manufacture typically associated with periods of climatic instability (Hiscock et  al. 2011). A distinct peak in backed microlith manufacture takes place during the Howiesons Poort in southern Africa, dated between 64.8 and 59.5 kya (Jacobs et al. 2008). Only in some sites is this technology associated with microblade core technology, such as at Rose Cottage Cave (Clarkson 2010; Soriano et al. 2007). Most sites throughout southern Africa display continuities with earlier and later MSA bipolar and radial core technology. Backed microliths make their first appearance in East Africa at circa 60 kya, as seen at Mumba rock shelter Layer V, alongside predominantly bipolar core technology (Gilganic et al. 2012). The underlying Layer VI, dated to 75–63 kya, consists primarily of single platform, discoidal and Levallois core technology (Diez-Martin et al. 2009). Production of points also varies dramatically in time and space across Africa (Brooks et al. 2006). Unifacial points precede and post-date Howiesons Poort backed artefact production at some sites in southern Africa (Singer & Wymer 1982; Wurz 2002), while beautiful bifacial foliate points make a brief appearance in Still Bay layers in southern Africa dated 71.9–71 kya (Jacobs et al. 2008; Villa et al. 2009). Unifacial and bifacial points are widespread and of longer duration in eastern Africa, and they are common at sites like Mumba in Tanzania (Mehlman 1979; Brooks et al. 2006), in the Horn of Africa and in Ethiopia (e.g., Porc Epic, Enkapune Ya Muto, Aduma and Somali sites) (Ambrose 2002; Clark 1984; Mehlman 1979; Pleurdeau 2005; Yellen et al. 2005). Various scraper, notch and denticulate forms also appear in the MSA, but these are mostly informal in character. The site of Porc Epic, although poorly dated (obsidian hydration 65–71 kya) is perhaps the most important MSA African site for a southern dispersal via the Horn of Africa, being essentially adjacent, though inland, of the Bab al Mandeb. This site contains mostly Levallois cores and flakes; few blades, burins and denticulates; some Levallois and Nubian points, though most points are bifacial (Figure 7.1, A, B). The same is true of Gogora rock shelter in the Ethiopian highlands where bifacial points dominate the retouched assemblage (Clark 1988). Tanged points are found across northern Africa from the Maghreb to Libya (Jacobs et al. 2011; Cremashi et al. 1998) from 100 to circa 55 kya, with occasional reports in Egypt and the existence of tanged Levallois flakes at Sodmein Cave (Mercia et al. 1999).The Aterian is unlikely to be associated with modern human dispersal out of Africa due its confinement largely to the central and western areas of North Africa and the presence of Nubian technology in the Nile Valley. Sodmein Cave and Taramsa Hill are especially important sites from the point of view of potential northern exit from Africa via the Nile Corridor (Figure 7.1, C, D). Upper Palaeolithic blade technology does not appear in the Nile Corridor until circa 33 kya, as at Nazlet Khatar 4, and is thus too late to document the spread of Upper Palaeolithic blade or microlithic technologies out of Africa via the Nile Corridor (Leplongeon & Pleurdeau 2011;Van Peer et al. 2010).Taramsa, dated to 55–60 kya, is within the temporal window of interest for out of Africa and displays classic Levallois and Nubian point core technology in association with a modern human child burial (Vermeersch et al. 1998). Sodmein Cave preserves a rich record of Levallois and Nubian point production and therefore demonstrates the presence of a non-microlith MSA along the Red Sea coast between 118 ± 8 and circa 40 kya (Mercia et al. 1999). The site of Bir Tirfarwi similarly contains bifacial foliates and Nubian Levallois technology, including the Nubian 1 subtype dated 120–96 kya (Wendorf et al. 1993). Bifaces are not a common feature of MSA assemblages after circa 150 kya, although bifaces were found with the early Homo sapiens remains at Idaltu in Ethiopia (Clark et al. 2003) and at Skhu¯l in the Levant (Garrod & Bate 1937). Bifaces appear to have largely disappeared from the African record by the time modern humans left Africa sometime between 45 and 100 kya. Bifaces might therefore be viewed as an archaic technology that we would not expect to see associated with an exit from Africa after circa 100 kya. 78

The Archaeological Signature of Modern Human Dispersal

Figure  7.1.  Examples of key artefact types from East African MSA sites, 45–100 kya. A–B: Ethiopian sites dated circa 50–60 kya (from Clark et al. 1984) (Gademotta, Kukeletti and Porc Epic). C: Taramsa (from Vermeersch et al. 1998). D: Sodmein Cave (from Van Peer et al. 1996).

The Levant Dozens of sites across the Levant preserve Middle Palaeolithic sequences beginning around 250 kya and represent a mix of Levallois, scraper, point, large blade and bifacial reduction technologies (Hovers 2009). A significant skeletal record also exists at some sites, pointing to an early presence of Homo sapiens outside of Africa between 120 and 85 kya followed by reoccupation by Neanderthals between circa 70 and 48 kya (Grün et al. 2005; Mercier et al. 1993; Rink et al. 2001a; Schwarcz et al. 1988; Stringer et al. 1989;Valladas et al. 1987, 1988, 1999). Not only did both species exploit similar animal species (Rabinovich & Hovers 2004; Rabinovich & Tchernov 1995; 79

Christopher Clarkson Speth & Tchernov 2001; Stiner 2006), but their Levalloiso-Mousterian lithic technology is often considered virtually identical (Garrod & Bate 1937; Hovers 2009) and often serves as a warning to those who would see different species possessing distinctive material culture. The lithic industries of the early anatomically modern humans found at sites such as Qafzeh, Skhu¯l and perhaps Tabun Layer C2, span roughly 130 to 100 kya and consist of a mix of recurrent centripetal, uniand bipolar Levallois flake and Levallois point technology, infrequent and non-systematic blade production and mostly lightly retouched scrapers (Hovers 2009; Shea 2003). Like the situation in Egypt, microlithic and systematic blade production does not begin in the Levant until after 47 kya when ‘transitional’ blade and point assemblages make their first appearance at sites like Boker Tachtit and Kebara Cave, with the Levantine Aurignacian appearing around 37 kya (Bar-Yosef et al.1996).

Arabia Arabia is also of crucial importance in identifying the human exit from Africa owing to its close proximity to East Africa and its role as a potential conduit between Africa and regions to the east during pluvial periods. Several publications have contributed vital new information about the timing and nature of occupation in three different parts of Arabia, although no human skeletal remains have yet been found at any of these locations. At the Jebel Faya rock shelter site in the United Arab Emirates, close to the Persian Gulf, three distinctive Palaeolithic assemblages are found dating between 123 ± 10 and 38.6 ± 13 kya (Armitage et  al. 2011). Assemblage C, the oldest, contains Levallois, volumetric blade and simple single platform core reduction techniques and includes formal tools such as handaxes, bifaces, scrapers and denticulates (Figure 7.2, A). Assemblage B consists of core reduction involving flat cores with parallel, converging and crossed removals, with scrapers, burins and perforators. Assemblage A, dating 38.6–40.2 kya consists of multiplatform cores with parallel removals, burins, scrapers and denticulates. Levallois, handaxe and bifacial point reduction is absent from assemblages A and B, while backing and microliths are entirely absent from all assemblages. Assemblage C is considered consistent with, but not identical to, African and Levantine Middle Palaeolithic industries, while it is suggested that industries A and B could be an autochthonous development. Armitage et al. draw on the suggested similarities to African sites to argue that modern humans may have been present at Jebel Faya by MIS 5e and hence well before the Toba eruption, likely also spreading across the Persian Gulf into Iran at favourable times. At the inland site of Jebel Qattar, situated next to an ancient palaeolake in the Nefud Desert in northern Arabia, artefacts located atop a pedocrete and buried beneath a sand dune date to a maximum of 75 ± 5 kya.These consist of discoidal, preferential and centripetal recurrent Levallois and single and multiplatform cores, with scrapers, notches, denticulates, a pseudo-Levallois point and a unifacial point (Figure 7.2 B). While cautious about hominin attributions, the authors of this study point out that these finds indicate that hominin groups were capable of living far from the coast around the time modern humans most likely exited Africa. The third region where Middle Palaeolithic assemblages are recently dated is the Dhofar region of Oman on the southern coast of Arabia (Rose et al. 2011). Assemblages from the site of Aybut Al Auwal date to circa106 kya and, like Jebel Faya, also fall within MIS Stage 5 (Figure 7.2, C, D). The excavators argue for very strong affinities between the Nubian technocomplex of northeastern Africa and the Horn of Africa on technological and chronological grounds. The technological similarity centres on the distinctive technique of distal core preparation used to produce points through the creation of a triangular core with a steep triangular distal guiding ridge. This technique differs from that of classic Levallois point production, and is apparently not found in the Levant. Other locations in Oman contain bifaces in association with undated Middle Palaeolithic sites (Rose 1994). 80

The Archaeological Signature of Modern Human Dispersal

Figure 7.2.  Examples of Arabian Middle Palaeolithic assemblages. A: Assemblage C, Jebel Faya, United Arab Emirates (from Armitage et al. 2011). B: Jebel Qattar, Saudi Arabia (from Petraglia et al. 2011). C–D: Aybut Al Auwal, Oman (from Rose et al. 2011). Finally, tanged points of probable Pleistocene age and closely resembling those of the North African Aterian are reported in the Rub’ al Khali area of Saudi Arabia by McClure (1994). Unfortunately, no radiometric dates were obtained for the assemblage found in association with palaeolakes of Pleistocene age.

South Asia The record from South Asia (Pakistan, India and Sri Lanka) has been pivotal in discussions of the archaeological signature of early modern humans east of Africa because of the well-excavated and well-dated sites that have recently been reported in this region and because of the central 81

Christopher Clarkson role South Asia played in early population expansion and dispersals to the east. Genetic studies have revealed that India was the gateway to subsequent colonisation of Asia and Australia and saw the first major population expansion of modern human populations anywhere outside of Africa (Atkinson 2008; Endicott et al. 2007; Metspalu et al. 2004; Rasmussen et al. 2011; Soares et al. 2009). South Asia therefore provides a crucial stepping-stone in early modern migration to Southeast Asia and Oceania. Two hypotheses are now being debated, which are here termed the microlithic first (Mellars 2006b) and the Middle Palaeolithic first (Petraglia et al. 2007, 2009; see Blinkhorn & Petraglia, this volume; Clarkson et al. 2009, 2012) models. The microlithic first hypothesis argues that modern humans entered India from the northwest along the coast circa 60 kya with distinctive microlithic technology (Mellars 2006b; personal communication 2011). This technology consisted of small stone bladelets steeply blunted along one edge and struck from specially prepared microblade cores. The microlithic is also argued to be associated with the first systematic use of bone and antler tools in India as well as the manufacture of perforated beads and cross-hatched incised objects. Mellars draws a direct link with similar objects in sub-Saharan Africa more than 50 kya, which he sees as strong evidence that modern humans brought this technology from Africa to India, just as they took southern African blade technology and split based bone points to Europe. Importantly, to explain the absence of microlithic technology outside of India in neighbouring regions such as Southeast Asia and Australia at around this time, Mellars argues that colonists lost the know-how to produce such items en route to Southeast Asia and Australia as a consequence of frequent founder effects and lack of appropriate raw material for making backed blades. Clarkson et al. (2009; 2012) and Petraglia et al. (2009; 2010; 2012b) argue against modern human colonisation with microlithic technology on several grounds: 1. The microlithic technology of South Asia is significantly younger than modern human colonisation of Australia, by at least 15 ky. Colonisation of Australia is now thought to have taken place circa 55 ± 5 kya (Hiscock 2009), whereas the oldest microlithic in South Asia is dated no older than 36 kya and overlaps in age with late Middle Palaeolithic technologies at sites in the Kurnool District (Petraglia et al. 2009; 2012b) and at Patne (Sali 1989). 2. The oldest backed microliths in South Asia are in fact found in the southern interior of Sri Lanka, and not in the northwest or on the coast where they would be expected if part of the colonising technology (Deraniyagala 1992; Perera et al. 2011). 3. Microliths found in Pleistocene sand dunes in coastal sites on the southern Tamil Nadu coast and on the northeastern Sri Lankan coast date to less than 26 kya (Gardner & Martingell 1990; Singhvi et al.1986), suggesting an earlier microlithic presence along the coast is unlikely. 4. While Mellars suggests a full Upper Palaeolithic package appeared simultaneously in South Asia, in fact key sites such as Jwalapuram 9 in Andra Pradesh and Patne in Maharashtra indicate staged introductions. At Jwalapuram 9 bead manufacture does not commence until 15,000 years after backed microliths first appear. At Patne, the engraved cross-hatched ostrich shell from Layer IIE sits above a date of 25 ± 2 kya, well above the first introduction of microliths (Sali 1989). Thus the cross-hatched piece in India dates to more than 25,000 years after such pieces are found on ochre (not shell) in sites such as Blombos and Klein Kliphuis in South Africa (Henshilwood et al. 2009; Mackay & Welz 2008). Claims for a direct cultural connection between finds made some 10,000 km and more than 25 kya apart are difficult to sustain. 5. No comparable microlithic tradition has yet been found between Africa and India during the period of likely human colonisation at anything like the circa 60 kya antiquity suggested for the microlithic by Mellars. 82

The Archaeological Signature of Modern Human Dispersal The possibility that older microlithic technology was located on the drowned continental shelf is possible, but an appeal to negative evidence seems implausible given the rapid colonisation of inland regions of Australia and New Guinea where colonists penetrated high mountain chains, deserts and glacial landscapes between 35 and 45 kya (Summerhayes et al. 2010; Summerhayes and Ford, this volume; Allen 1996b; Smith 1989). By contrast, the inland regions of India and Sri Lanka would have offered few challenges to human occupation where large river valleys and diverse ecosystems offered favourable settings for habitat expansion. The fact that the first large human population expansion to have taken place outside of Africa took place in South Asia is consistent with the suggestion of highly favourable conditions for early modern human expansions into the interior. Korisettar (2007) has argued that inland basins may even have offered more favourable environments for early colonists than did Pleistocene coastlines, as testified by high site densities and long-term occupation of inland basins by earlier hominins. The alternative to the microlithic model is the Middle Palaeolithic first hypothesis, proposed by Clarkson et al. (2009; 2012), Haslam et al. (2011; 2012) and Petraglia et al. (2007; 2009; 2010). Instead of bringing microlithic technology to South Asia, Clarkson, Petraglia and their colleagues argue that modern humans brought classic components of an African Middle Stone Age tool kit. These included unprepared cores, prepared radial cores and scraper technology, rather than microlithic technology. While not as distinctive as microlithic technology, Clarkson has shown that the core technologies found in Indian sites within the time range of likely modern human colonisation (45–50 kya) are similar to those found in Africa, Southeast Asia and Australia, suggesting the possibility of continuous cultural transmission in the system of flake production among modern human colonists dispersing eastwards from Africa (Clarkson et al. 2012; Haslam et al. 2010, 2012). This combination of prepared and unprepared radial core technology and informal scraper assemblages is found from above and below the Toba ash in the Jurreru River Valley (Clarkson et al. 2012), and in Middle Palaeolithic contexts lacking handaxes in the Middle Son River Valley (Haslam et al. 2012). These technologies are also found at other locations throughout India and Pakistan. The Jurreru Valley sites span more than 74 to around 38 kya, while the Son Valley sites appear to be younger than 100 kya, as those sites with handaxes date to more than 100 kya (Figure 7.3, left). Some rare finds of retouched points are noted at several Middle Palaeolithic sites in India (Costa 2012; James 2007; Misra 1995a), and retouched points with possible small impact fractures are found at Locality 22 in the Jurreru Valley (Haslam et al. 2012; Clarkson et al. 2012). At Site 55 near Riwat in Pakistan (Dennell et al. 1992; Rendell & Dennell 1987), a similar suite of single, multiplatform and radial cores are found with informal retouched flakes and some elongated flakes, dated to circa 45 kya, but there is no evidence for systematic blade production (Figure  7.3, right). Sites in the Thar Desert and around Karachi near the mouth of the Indus River show even more pronounced Levallois characteristics, although these have not been dated (Biagi 2006). Many other sites are identified as having substantial Middle Palaeolithic and overlying microlithic assemblages, but in general these remain poorly dated and are not well described (Petraglia et al. 2009). Under the Middle Palaeolithic first model, the microlithic is seen as a local innovation within India that appeared a long time after first modern human colonisation. The microlithic arises in response to rapidly deteriorating climatic conditions in South Asia around circa 36 kya and leading up to the Last Glacial Maximum when microlithic technology became most abundant (Clarkson et al. 2009). This climatic shift was tied to the downturn from peak Northern Hemisphere summer insolation, resulting in a strong reduction in summer (southwest) monsoon rainfall across India and in a semi-glacial period mosaic environment with an expanded Thar Desert acting as a likely barrier to population influx. This technological change involved switching from technologies dominated by Middle Palaeolithic prepared cores and retouched scrapers to the use of more standardised, composite, hafted tool kits during a period of higher mobility 83

Christopher Clarkson

Figure 7.3.  Examples of artefacts from South Asia. Left: Sites above and below the Toba Ash in the Jurreru Valley, central India (from Clarkson et al. 2012). Right: Site 55, Riwat, Pakistan (from Rendell et al. 1989).

Figure 7.4.  Hypothetical correlation between phases of colonisation and technological change from MSA/Middle Palaeolithic to microlithic in South Asia (modified from Rasmussen et al. 2011). and reduced predictability in resource location and capture. This scenario is consistent with the appearance around the world of microlithic technology at times of climatic fluctuation (Hiscock et al. 2011). An alternative to an indigenous origin for the South Asian microlithic is that the microlithic was a later introduction from the north. A second wave of population interaction seems to be implied by the recent Rasmussen et  al. (2011) genetic analysis which suggests two colonising waves, an earlier M lineage circa 62–74 kya and a later N lineage circa 25–38 kya. Their analysis 84

The Archaeological Signature of Modern Human Dispersal confirms a common recent modern human origin in sub-Saharan Africa for all extra-African populations but suggests the possibility of two dispersal routes – a northern and southern route – with later admixture in South Asia (Figure 7.4). This could explain the appearance of an earlier Middle Palaeolithic technology shared between the first modern human colonists between Africa and Sahul, as well as the later introduction of microlithic technologies at the time they also appear in the Levant, Iran and northern Asia.

Southeast Asia There are very few sites in Southeast Asia dated older than 42 kya and fewer still with large or well-reported lithic assemblages. Of those very few sites, Jerimalai in East Timor, has perhaps the largest and best-described lithic assemblage for this early time period (Balme & O’Connor, this volume; O’Connor 2007b; O’Connor et al. 2011a).The site has only been radiocarbon dated and likely did not reach the base of cultural materials, leaving open the possibility that older cultural materials remain to be found at the site. The Jerimalai lithic assemblage from the 38–42 ka levels is almost entirely manufactured from small nodules of high-quality chert (Figure 7.5C). Faceted radial cores (Levallois-like) made on small nodules or larger flakes and side and end scrapers dominate the formal component of the assemblage. Bipolar and anvil rested core reduction is also present as well as informal single platform and multiplatform core reduction. All cores are highly reduced, although the size of available nodules also appears to have been limited to smallsized pieces. The flake assemblage reflects the frequent use of faceting and radial flake removals from small nodules and larger flakes, with Levallois-like flakes, pointed blades with faceted platforms and other faceted flakes struck from the ventral surfaces of larger flakes all present in the assemblage. The scraper assemblage shows no recurring formal types, but notching is common and rarer examples of pointed beak-like retouch and burination are also present. Another site of similar antiquity is Lang Rongrien from northwest Thailand (Anderson 1990). The assemblage from the more than 45 kya layer (Layer 10) is very small and quite ambiguous but contains single and multiplatform cores, radial and bifacially worked flakes and cores, scrapers, burins, an antler artefact and one possible microblade, but the assemblage is too small to be very informative (Figure 7.5, A). The only other site in Southeast Asia of such antiquity is Niah Cave on the island of Borneo, dated at circa 46 kya (Harrisson 1959c; Barker et al. 2007; Barker and Hunt, this volume). Very little is written about the very few artefacts from the base of the site, which Harrisson (1959c) termed Middle Palaeolithic in nature. The illustrated examples, however, are of flakes that could derive from virtually any form of core technology, and little more can be said about lithic technology on Borneo without better description of the lithics.

Australia and Melanesia A large number of sites date to circa 40 kya in Australia, but only six have dates of 45 kya or greater. These are Malakunanja II, with the lowest artefact layer bracketed between thermoluminescence (TL) dates of 52 ± 7.11 and 61 ± 9.13 (Roberts et al. 1990), Nauwalabila 1 with the lowest artefacts bracketed between optically stimulated luminescence (OSL) dates of 53 ± 5.4 and 60.3 ± 6.7 (Bird et al. 2002; Roberts et al. 1994), Lake Mungo with the oldest artefacts dated more than 47 ± 3.4 kya by TL (Roberts et al. 1998; Bowler et al. 2003; Bowler & Price 1998) and Devils’ Lair (48 kya Cal BP), Nawarla Gabarnmang (45 kya Cal BP) (David et al. 2011) and Carpenter’s Gap 1 (45 kya Cal BP), both dated by radiocarbon (O’Connor 2005; Turney et al. 2001a). In New Guinea, two sites are dated more than 45 kya. These are the site of Kosipe in the Highlands (Summerhayes et al. 2010) and the site of Bobangara on the Huon Terraces (Groube et al. 1986). 85

Christopher Clarkson

0 1 2 cm

A

Nauwalabila

Lake Mungo

Top

Left

Front

Back

Right

Bottom

Front

Left

B

Top

Back

Front

Bottom

Right Left

Top

Back

Bottom

Right

C

Figure  7.5.  Examples of stone artefacts from Southeast Asia and Australia. A: Flakes and cores from Lang Rongrien, northwest Thailand (from Anderson 1990). B: Three-dimensional scans of single platform cores (top) and radial cores (bottom) from Nauwalabila (left) and Lake Mungo (right). C: Stone cores (bottom), flakes and scrapers (top) from the Jerimalai assemblage (from O’Connor et al. 2011a). A familiar range of core technologies is present at these sites, including single platform, multiplatform and radial cores, with and without faceting. The same range of core types is found from Arnhem Land in the north to Lake Mungo in the south (Figure 7.5, B). Scrapers are likewise informal side and end scrapers and notched pieces, and there is so far no evidence of points or other formal artefact types. Levallois appears to be absent from Australia, although sample sizes from all these sites remain very small at present, and flakes with radial scar patterns and or faceted platforms are present in some sites in low numbers. Few reports describe early artefacts from New Guinea and nearby Melanesia, but these refer mainly to multiplatform cores and informal scrapers (e.g., Spriggs 1997). The main departure from this picture of simple core and scraper technology, and one that largely defines Sahul as a distinctly different region from that of Southeast Asia, is the presence of both flaked and ground stone hatchets at sites in New Guinea and northern Australia. Flaked and waisted axes are known from sites such as Bobongara on the Huon Peninsula in Papua 86

The Archaeological Signature of Modern Human Dispersal New Guinea (PNG) dated by TL to circa 40 kya (Groube et al. 1986), and from Kosipe in the Highlands of PNG dated to 44–49 kya (Summerhayes et  al. 2010; Summerhayes & Ford, this volume). In northern Australia, fragments of ground axes are known from 35.4 ± 0.4 cal BP in Arnhem Land (Geneste et al. 2010). Flaked and ground axes are also now reported from Honshu Island in Japan, dating between 32 and 36 kya (Takashi 2012), begging the question of whether the axe tradition in these locations shares a common origin in mainland or Southeast Asia, or whether these were independently invented.

Diminishing Diversity East of Africa This review of major lithic assemblage characteristics in sites from MSA Africa to Oceania identifies no single recurring dominant element in these technologies that could be used as a fingerprint for modern human arrival in each new location. Microlithic technology is not present in Arabia, occurs in Egypt (ca. 33 kya), the Levant (36–47 kya) and India (36–37 kya) far too late and appears in East African assemblages only after circa 60 kya (Gilganac et al. 2012). Thus, while East Africa provides a possible source of microlithic technology for a spread eastwards to India, no microlithic technology is found in the intervening regions until much later. The oldest transitional Upper Palaeolithic with microlithic technology outside of Africa appears in the Levant at circa 46/47 kya (Bar-Yosef et al. 1996). This means microlithic technology must have skirted both the Nile corridor and Arabia in order to arrive in the Levant more than 10 kya after its first appearance in East Africa. No process has yet been advanced to explain how such a leapfrogging of microlithic technology took place from the Horn of Africa to the Levant and thence to Iran and Sri Lanka at circa 36–40 kya. In fact, assemblages from sites examined here represent a mix of prepared and unprepared radial and non-radial core technologies along with informal scrapers that are found from Arabia to Oceania after 100 kya. The absence of microlithic technology could be explained by a spread out of Africa before 60 kya – that is, before microlithic technology arrived in East Africa – and indeed this is what the latest genetic analyses suggest (Rasmussen et al. 2011), pointing to an exit between 62 and 75 kya. Another important observation is that the further from Africa we look, the less technological variation we find (Figure 7.6), with novel variants appearing on the margins of modern human expansion where water crossings took place (e.g., axes in Australia, New Guinea and Japan). East of Africa, Levallois also appears to decline in importance and is either absent or substantially altered by the time moderns reached Southeast Asia and Australia. I would argue that the core technology at Jerimalai bears enough resemblance to Levallois in terms of the key concepts laid out by Boëda (1995) to justify at least the label of ‘Levallois-like’. Radial core technology persists all the way to Australia, but single platform and multiplatform technologies are dominant there. African assemblages can therefore clearly be identified as the most diverse technologically, with unifacial and bifacial points, microliths, tanged points, microblades, blades, and Levallois, discoidal, single platform, multiplatform and bipolar cores all present across Africa and with much of this variation present in East Africa alone. Arabia lacks microliths but retains the rest of the range (possibly including tanged points), but with the addition of bifaces in some parts. South Asia loses bifaces circa 100 kya but displays the remaining technological range. Southeast Asia lacks points, while Australia lacks points and Levallois, but gains flakes/ground axes. This pattern of a gradually declining technological diversity with distance from Africa can be interpreted as a consequence of frequent founder effects, resulting in loss of technological diversity among small daughter populations and resulting from population bottlenecks or frequent population collapse. This phenomenon is well documented for the process of modern human expansion out of Africa in genetics (Prugnolle et al. 2005) and cranial variation (Manica et al. 2005) and was also suggested as a cause of reductions in technological diversity in small dispersing 87

Christopher Clarkson 10 9 Points

Technological Diversity

8

Microliths

7

Handaxes Tanged Points

6

Microblades 5

Levallois

4

Discoidal

3

Single Platform Scrapers

2

Multiplatform

1 0

Flaked/Ground Axes Africa

Levant

Arabia

South Asia SE Asia

Oceania

Figure 7.6.  Diversity of technological elements found in assemblages from 57 sites associated with presumed earliest modern humans in each region. Assemblages arranged in order of increasing distance from Africa (illustration by the author).

Homo erectus populations (Lycett et al. 2006; White 2000). The variable nature of assemblages in each region might also be explained as the effects of guided variation in small dispersing populations (Boyd & Richerson 1985; Bettinger & Eerkens 1999), whereby inherited traits and concepts for toolkit construction underwent trial and error as colonising populations entered new landscapes with new resources.

Conclusion Resolving questions of modern human origins and dispersals using primarily lithic evidence is far from easy. Previous studies have examined the evidence for eastward modern human dispersal using a range of approaches including multivariate analysis of core morphometrics (Clarkson et al. 2012; Haslam et al. 2011) and cladistics (Clarkson et al. submitted), while this study examines assemblage composition analysis. The striking pattern that emerges in all of these studies is one of decreasing diversity with distance from Africa, the absence of microlithic technology outside of Africa until well after modern human arrival in Australia, and the presence of a mixed technology of non-microlithic sub-Saharan Africa Middle Stone Age character in Southeast Asia and Australia. While this technology is not distinctive in the same way as the microlithic, this does not mean it does not have a sub-Saharan African origin. In my view, consistent with genetics and skeletal record, the evidence points to Africa as the most likely source for this technology sometime prior to 60 kya. The hypothesis is therefore of a pre-60 kya modern human exit with an essentially MSA technology undergoing reduction in diversity with distance from Africa. Unfortunately it will be difficult to test this hypothesis without new and unambiguous fossil evidence of certain age and found in direct association with lithic assemblages at key points along the putative dispersal route east of Africa. Much work remains to be done, but this review has at least served to propose a coherent argument for a non-microlithic dispersal of anatomically 88

The Archaeological Signature of Modern Human Dispersal modern humans before 60 kya.While fossil finds provide our greatest hope of shedding new light on this question, improved dating and characterisation of lithic assemblages from Arabia, India, Southeast Asia and Australia in the critical time period of 50–60 kya remain critical to better understand the signature of modern human colonisation of these poorly studied areas.

Acknowledgments My thanks to the editors for inviting this contribution and for organising the session at the 2009 IPPA meetings, Hanoi. My thanks also to Ceri Shipton and Jacqueline Matthews for helpful comments on a draft of this paper.

89

Chapter 8 Missing Links, Cultural Modernity and the Dead Anatomically Modern Humans in the Great Cave of Niah (Sarawak, Borneo)

Chris Hunt and Graeme Barker

Introduction During the course of his expedition to Borneo in 1855 to make collections of “shells, insects, birds, and the orang-utan” (Wallace 1913, 27), the Victorian naturalist Alfred Russel Wallace was informed about enormous caverns at Niah and Mulu (Figure  8.1). Though he was unable to visit them, he was convinced that such caves were likely to be the best locations to search for the ‘missing link’ between humans and apes (Harrisson 1958; Sherratt 2002). In 1864 he wrote to Charles Darwin with the information that the new British consul going out to Sarawak had informed him that he intended to explore caves near the capital Kuching “and if anything of interest is obtained, a good large sum can no doubt be raised for a thorough exploration of the whole country” (Harrisson 1958, 551). In the event, it was A. Hart Everett, an amateur naturalist and collector who came out to work for the Sarawak government in 1869, who led the first scientific expedition to Sarawak caves in search of the ‘missing link’, backed by a committee formed of leading members of Victorian science in the United Kingdom: the geologist Charles Lyell, the archaeologist William Pengelly, and the anthropologist George Busk. In 1873 he visited the Bau Caves near Kuching and the Niah Caves, publishing an anonymous account of his visit to Niah’s main cavern, the Great Cave, in the Sarawak Gazette in the same year. In the article, he described how he arrived at the spectacular West Mouth of the Great Cave: after a difficult walk through the swamp forest, “we found ourselves standing at the mouth of a large arched cavern, several hundred feet broad, and over two hundred feet high, huge stalactites were pending from the ceiling, and a fringe of vegetation dropping from its outer edge” (Anon. 1873, 60; Figure 8.2). Everett collected some of the human remains lying on the surface of the West Mouth and other entrances to the Great Cave and took them back to London. Although George Busk (1879–80, 321) ascertained that they were likely to be recent in age and “the information they afford is very meagre”, they raised sufficient interest for the British Association for the Advancement of Science to pass a resolution at its Dublin meeting in 1878 “that Mr John Evans, Sir John Lubbock, MajorGeneral Lane Fox (General Pitt-Rivers), Mr George Busk, Professor Boyd Dawkins, Mr Pengelly, and Mr A. W. Franks be a Committee for the purpose of exploring Caves in Borneo; that Mr Evans be the Secretary; and that the sum of £50 be placed at their disposal for the purpose”. 90

110°

115°

120°

5000 3000 2000 1000

Mount Kinabalu

500 5°

200 Sea level

Niah SARAWAK



Schwaner Mountains K A L I MA N TA N



0

300 km

Figure 8.1.  Borneo, showing places mentioned in the text (illustration by Dora Kemp).

Figure 8.2. The West Mouth of Niah Great Cave, looking west; the archaeological zone is on the right (photograph by Graeme Barker). 91

Chris Hunt and Graeme Barker Further sums were raised, including from Charles Darwin, to meet the expedition’s estimated costs of £370, and Everett was dispatched to the Niah Caves again. He faced local opposition to his plans for excavation because of the disturbance to known burial grounds, and after undertaking a small amount of bone collection he concluded that the caves “were too recently raised above the waters of the sea to render it probable that future discoveries will be made” and “no further expense should be hazarded”. In a bizarre twist of fate, his Niah Caves collections, donated to what is now the Natural History Museum in London, almost certainly provided the orang-utan mandible and human skull that formed the Piltdown Man, the spectacular ‘missing link’ fossil reported in 1912 but demonstrated in the 1950s to be a crude forgery (Oakley & de Vries 1959).

The Discovery of the Deep Skull The famous swiftlet and bat populations of the caves, rather than their potential archaeological significance, first attracted Tom Harrisson (1911–1976), a passionate ornithologist, to visit the caves in 1947 (Harrisson 1958, 564–565). After being expelled from Cambridge University for disorderly conduct, he took part in a series of expeditions to remote places (including to Sarawak) and then in the Mass Observation project in Britain (Harrisson 1937, 1943, 1961; Heimann 1997). During World War II he trained as a commando and parachuted into the interior mountains of Sarawak in March 1945 to organise resistance against the Japanese occupying forces (Harrisson 1959a), earning a Distinguished Service Order (DSO) for the success of this mission. In 1947 he was appointed curator of Sarawak Museum and government ethnologist, a position he held until his retirement in 1967. Alongside an abiding interest in the anthropology of the Sarawak peoples, and animal conservation, he embarked on a major programme of archaeological excavation given the complete absence of such work since Everett’s expeditions. Having no training, he enlisted the help of Michael Tweedie, curator of Singapore Museum, and after initial work in the Bau Caves and elsewhere near Kuching, they started work in the West Mouth of the Niah Great Cave in 1954, with an initial two-week season. They excavated a large trench with considerable rapidity  – and with a lack of control or recording that has meant that unfortunately little of this work can now be reconstructed. They dug through a succession of burials which we can now recognise as likely to have been Metal Age and Neolithic, finding underneath them deep deposits rich in evidence of early human occupation: charcoal, ash, animal bone fragments, and occasional primitive stone tools. Though Harrisson described the results as “incredible – just been digging there, fantastic”, he realised that it was “at once evident that to tackle this cave properly, we were going to need personnel by the score, financial resources by the tens of thousands, and a long-term programme of continuing work both in the field and with excavated material back in the Museum” (Heimann 1997, 291). In 1957 he returned to Niah with his newly wed wife Barbara, beginning a ten-year campaign of major excavations of several months’ duration each season, most of which Barbara supervised for the duration of the work with Tom dividing his time between Niah and Sarawak Museum in Kuching. By the end they had conducted extensive excavations in most of the entrances to the Great Cave, and in many other small caves around the Niah limestone massif, but the most extensive excavations, and the most spectacular discoveries, were in the West Mouth. The 1957 season concentrated on the West Mouth. A large trench was excavated around the 1954 pit, dug in horizontal ‘spits’ measured in relation to the original ground surface (Figure 8.3). At the base of the trench the Harrissons excavated a deep sounding, which they termed Hell because of the difficult working conditions under the full afternoon sun. A quantity of charcoal collected at depth in this trench (106 inches, or 260 cm, below the 1954 ground surface) and sent to the University of Groningen in the Netherlands for the new method of radiocarbon dating yielded a date of around 40,000 years ago, the maximum age range of the method at that time. On 92

Anatomically Modern Humans in the Great Cave of Niah

Figure 8.3. Tom Harrisson overseeing excavations in the Hell Trench of the West Mouth, Niah Great Cave. Note the system of digging in horizontal ‘spits’ (photograph reproduced with permission of Sarawak Museum). 7 February 1958, early in the 1958 season, Barbara’s team started to uncover fragments of human skull at the same depth as the 1957 charcoal sample (Figure 8.4). Tom Harrisson was in Kuching to receive a visit from one of the world’s experts on human origins, Professor von Koenigswald. Summoned by telegram by Barbara, he and Koenigswald travelled to Niah (by helicopter, courtesy of Shell) to witness the full excavation of the find. The skull, referred to subsequently as the Deep Skull, was studied by Don Brothwell at the British Museum and identified as that of teenage girl or young adult female, anatomically modern, of Australoid type (Brothwell 1960; Figure 8.5). Although Barbara Harrisson remembers Koenigswald, who was hoping for a primitive fossil, dismissing the Deep Skull as “not interesting” as its physical features became apparent during excavation, its discovery brought the Great Cave to international attention because, if it was of the same antiquity as the 1957 charcoal sample, it was the earliest modern human fossil known that that time anywhere in the world. Human remains found near the Deep Skull, at the same depth, included an almost complete left femur and a right proximal tibia fragment (Krigbaum & Datan 1999; 2005) and a human talus (Hooijer 1963). 93

Chris Hunt and Graeme Barker

Figure 8.4. The discovery of the Deep Skull. The label to the left of the partially excavated skull refers to the location of the charcoal found the previous year and dated to circa 40,000 years ago (photograph reproduced with permission of Sarawak Museum).

94

Anatomically Modern Humans in the Great Cave of Niah

Figure  8.5. The Deep Skull after restoration (left) and a human skull of recent age (right) (photograph reproduced with permission of Sarawak Museum). The Harrissons also found prolific evidence for human occupation at similar depths to the Deep Skull extending several metres to the north of the Hell Trench into a small rock overhang or shelter formed at the northern cave wall. They obtained a series of radiocarbon dates indicating that this part of the West Mouth, which they termed the ‘habitation’ or ‘frequentation’ zone, had been regularly occupied through the Late Pleistocene and into the Holocene. Further into the cave entrance they found a dense collection of some 200 graves (representing about 400 bodies) dating to the Neolithic and Metal Age, circa 4,000–2,000 years ago. The Harrissons and their collaborators published numerous papers on their discoveries, especially in the Sarawak Museum Journal, but never a final report with detailed stratigraphic or contextual documentation, so doubts about the reliability of the Pleistocene finds, especially the status of the Deep Skull, were raised regularly by scholars attempting to incorporate the Niah finds into studies of the region’s prehistory (e.g., Bellwood 1997; Bulbeck 1982; Kennedy 1979; Solheim 1983; Storm 2001a;Wolpoff 1999).Were the radiocarbon dates, at an early stage in the development of the method, reliable? A key component of the Harrisson dating was extrapolation by depth, anchored by the radiocarbon dates, and the most common sediment excavated in the West Mouth, described by Harrisson as the ‘pink and white layer’, was interpreted by him as formed by a constant drizzle of pink ‘cave earth’ from the cave roof mixed with fragments of white limestone lumps. Age-depth extrapolation of this kind, however, ran counter to the experience of most cave excavators dealing with complex cave sediments. Indeed, had the excavators unknowingly mixed material of different ages because of their spit method of excavation? Was the Deep Skull in fact from a Neolithic or Metal Age burial? Small-scale excavations were undertaken in 1976 by the Malaysian archaeologist D. P. Zuraina Majid, for her Yale University PhD, to try to resolve such doubts, but although she secured new radiocarbon dates from the Hell Trench and other soundings in the West Mouth and demonstrated the validity of the broad sequence of Late Pleistocene and Holocene occupation and burial reported by the Harrissons (and added important new data to it), she was unable to resolve the underlying stratigraphic questions about the Harrisson discoveries (Zuraina Majid 1982). This was the 95

Chris Hunt and Graeme Barker context of the resumption of fieldwork at the site in 2000 by the Niah Caves Project (NCP), with the objectives of clarifying the nature and chronology of the stratigraphic sequences in the major cave entrances, and of associated human activity; locating these sequences in regional climatic and environmental frameworks; and using the new information to inform the re-study of the substantial archive of records and finds held in the Harrisson Excavation Archive at Sarawak Museum (Barker 2005, 2013; Barker et al. 2002, 2007).

The West Mouth Sedimentary Sequence and the Location of the Deep Skull The Niah Caves Project has established that the sedimentary deposits of the West Mouth are heterogeneous and have a variety of origins, including deposition by airfall, running water, mudflow and slope processes. Gilbertson et  al. (2005) identified four archaeologically significant lithofacies in the area adjacent to the findspot of the Deep Skull in the Hell Trench (Figure  8.6; Table 8.1). The age relationships of these sedimentary bodies have been resolved by accelerator mass ­spectrometry (AMS) dating of charcoal fragments contained within them, the reliability of these dates significantly improved by the use of the acid-base-wet oxidation (ABOX) pre-treatment technique (Higham et al. 2009). The sediments were laid down in a basin formed between the cave entrance lip, the north wall of the cave, and the toe of the guano mound filling the interior of the West Mouth. The basin lies partly under and in front of the rock overhang, extending to where the Hell Trench was located. Any water flowing down the guano mound drained into this basin and flowed northwards along it down a rock channel parallel to the cave lip, into the overhang, where it drained away through a sink-hole (inferred, not excavated). Between circa 50,000 and 38,000 cal BP, Lithofacies 2C and 2 accumulated in the basin, the former from the exterior (western) side of the cave entrance and the latter from the interior, intermingling in the channel. Lithofacies 2C consisted of a colluvium formed of collapsed speleothem and other debris from the cave lip. It supported the development of temporary surfaces that were sufficiently stable to be burrowed into by insects such as robber wasps (Sphex diabolicus) and vertebrates, and on which people deposited cultural debris, including the residues of fires, meals and butchery activities. In the process of slipping down from the cave lip, Lithofacies 2C became interbedded with Lithofacies 2, complex red-brown silts and sands up to 2.5 m thick formed by the episodic occurrence of streams, ponds, mass movement and soil formation, separated by periods of desiccation. At least 13 episodes of fluvial erosion have been recognised within Lithofacies 2 from geochemistry, granulometry and micromorphology, some of which reworked desiccated and cracked muds on the floor of the existing channel (Gilbertson et al. 2013). The duration and intensity of these alternating wetting and drying episodes are not known, but the major oscillations that can be detected in the sedimentology and palynology can be broadly correlated with the isotope climate signals in the North Greenland Ice Core Project (NGRIP) (Hunt et  al. 2012; Figure  8.7). Cultural debris also accumulated on the Lithofacies 2 surfaces. Lithofacies 2C has continued to accumulate on the exterior side of the basin to the present day, but between circa 38,000 and circa 35,000 cal BP a major hydro-collapse in the interior guano mound caused a massive mudflow of wet guano up to 3 m thick to flow downslope into the basin, where it struck, flowed into and mostly covered Lithofacies 2 and 2C (Figures 8.6 and 8.7). This more or less instantaneous mudflow, probably forming in hours or days, categorized as Lithofacies 3, is Harrisson’s ‘pink and white layer’ that he assumed had accumulated over many thousands of years as a drizzle of roof fall. Capping it in places was a related sediment derived from Lithofacies 3 by weathering, Lithofacies 3R. Lithofacies 4, which formed on top of Lithofacies 3 and 3R between circa 35,000 and circa 8000 cal BP, consisted of brown fine-grained silt-rich sediments with plentiful evidence of human activity, the ‘frequentation deposits’ described by Harrisson as being an extremely rich mixture, 96

Anatomically Modern Humans in the Great Cave of Niah West Mouth, Niah Great Cave West

East

E

F 0

2Ch

5m

2Cc

Overhang

Fissure

4

2C

2Cm

2

F

E C

ridge of limestone rock

V

N

C

1 B

0

WNW

10 m

ESE

A

C

A standing column of speleothem

a r ch

graves

aeo

lo gi

ca l

GG r es

er v e

p e r i m e t e r f e n ce

GW 5

V

C

V

C

2Ch

3R B

2Cc

5

Channelized sands and silts

4

Brown silts with anthropogenic deposits

3/3R

4 3

2C

‘Pink and white’ silts

2

Red-brown silts and sands

2C

Yellow colluvium

2Cc

2Ch

1 0

Collapsed speleothem

2Cm V C Midden of vertebrates, molluscs, ceramics

palaeosurfaces

10 m

Leaf peat

GG

Guano associated with graves and grave-infill deposits

GW

Guano with wood ash

G

Guano (undifferentiated)

1

Gritty sands with selenite

M

Boulders with clays and sands Cave wall or boulder of limestone Find spot of the ‘Deep Skull’

A

B Section lines

Figure 8.6.  Schematic representation of the Late Quaternary lithostratigraphy of the northern part of the West Mouth of Niah Great Cave. The relative height and position of the findspot of the Deep Skull are based upon descriptions in the Harrisson Archive records, but the stratigraphic position illustrated assumes that the lithostratigraphy at that location was the same as that evident in the surviving nearby excavation faces (from Gilbertson et al. 2013, fig. 3.6; illustration by Tim Absalom). including ash, charcoal, butchered animal bone and stone tools. Most of this lithofacies was removed by the Harrissons, but mapping the vestiges that remain as plinths of sediment under the rock overhang and as a few standing walls, and correlating these with photographs in the Harrisson Excavation Archive, indicated that it extended over some 150 m2 from the back of the rock overhang across the basin at the front of the West Mouth and was up to 4 m thick in the centre of its distribution. Human activities during the accumulation of Lithofacies 4 included dumping large quantities of ash and digging pits into the underlying sediments probably for storing, and in the process leaching out toxins from, the nuts and tubers that were being collected for food (Barker et al. 2007; and see Piper and Rabett, this volume). According to the radiocarbon dates of charcoal 97

Chris Hunt and Graeme Barker Table 8.1.  Lithofacies adjacent to the findspot of the Deep Skull in the Great Cave Lithofacies

Description

Origin

2

Trough cross-bedded silts, sands, diamicts and gravels, containing lenses rich in bone and charcoal; occupying a channel-like feature Silty diamicts containing bone, charcoal, ash lenses dipping into the cave from the entrance rampart and interdigitating with Lithofacies 2 Silty diamict with diffuse gypsum nodules Silty diamict with abundant ash, bone, charcoal; evidence of pit digging

Deposition in an ephemeral streamway; occasional human activity Colluvial deposition in the cave mouth; occasional human activity Mudflow deposits Colluvial deposition in the cave mouth; occasional human activity

2C 3 4

Note: Following Gilbertson et al. 2005 and Gilbertson et al. 2013.

ls

ka

BP –44

NGRIP δ18 O –42

–40

dia

sta

IP er GR Int

–38

5 6

35

4

Deep Skull

OxA-11034

7 8

40

2

11

45

OxA11302

3

9 10

Pit Fills

Niah 311

13

50

Hell Sequence

2C OxA-V2059-11

OxA-V-2076-16 OxA-V2076-15 OxA-V2057-31

?H6 107"

12

OxA11303

Niah 310

?

Key: Samples

Radiocarbon dates

U/Th date

2

Lithostratigraphic units

Figure 8.7. The principal Pleistocene sediments (Lithofacies 2, 2C, 3, 4) identified in the northwest corner of the West Mouth of Niah Great Cave, their inferred ages from radiocarbon dates obtained by the new investigations, the age of the Deep Skull indicated by direct U/Th dating and the correlations with the climatic (isotopic) sequence in the GRIP Greenland ice core indicated by sediment characteristics and palynology (compiled by Chris Hunt). in these pits, these activities dated back to 33,790 ± 330 bp or 37,341–39,550 cal BP (OxA-11302) and 29,070 ± 220 bp or 33,121–34,518 cal BP (OxA-11303). The archive photographs show that such pits extended right across the Lithofacies 4 zone. The geochemistry and palynology of the Late Pleistocene sediments of Lithofacies 4 indicate that they developed in a climatic regime that was slightly cooler and drier than today, interspersed with wetter episodes, a sequence that again can be broadly equated with the NGRIP isotope curve (Hunt et al. 2012; Figure 8.7). 98

Anatomically Modern Humans in the Great Cave of Niah From the depth measurements, descriptions and photographs in the Harrisson Excavation Archive, it is clear that the Deep Skull and the associated human bones were found near the top of the channel sediments, in the zone where Lithofacies 2C and 2 interbedded, and below the Lithofacies 3 mudflow (Figure 8.6).Two radiocarbon ABOX dates were obtained from Lithofacies 2 sediments immediately below the Lithofacies 3 mudflow, of 42,600 ± 670 bp or 44,695–47,005 cal BP (Niah-310) and 41,800 ± 620 bp or 44,344–46,137 cal BP (Niah-311). A charcoal sample found in the Harrisson Excavation Archive with a label in Tom Harrisson’s handwriting “charcoal adjacent to Deep Skull” was dated to 35,510 ± 350 or 39,676–41,503 cal BP (OxA-V-2076–16). The latter was from square H19 at 106 inches, and the skull was found at the same depth across squares H6 and H19, squares measuring only 12 by 12 inches each, so the charcoal was clearly only a few inches from the skull. All three dates are in accord with almost a dozen NCP dates obtained from charcoal taken either from surviving faces of the Harrisson trenches or from the archive than can be definitely associated with the Lithofacies 2/2C ­channel-fill sediments. Geometrically the skull lay within the highest local sediments of Lithofacies 2 and should therefore be of the same age as them, but in fact it is significantly younger than these altitudinally equivalent sediments on the evidence of direct uranium-­thorium (U/Th) dating of a part of its mandible by Alisdair Pike, which indicates an age of 35,200 ± 2600 years ago (Barker et al. 2007).

The Sediments within and around the Deep Skull Following its discovery in 1958, the Deep Skull was lifted within a block of sediment, encased in plaster and shipped to the British Museum for cleaning and investigation. The sediment cleaned from it by Don Brothwell in 1959 was stored thereafter in the British Museum and was rediscovered there by John Krigbaum in 2002. A sub-sample of this material was made available for analysis in 2005. For comparison, we also analysed some sediment from the Harrisson Excavation Archive labelled by Tom Harrisson “Soil from around Skull at H/6, 107”; Niah 15–2-58”. Even though the second sample derives from the same 12-by-12” grid square where much of the skull was found and at the same depth, there are significant differences in the sedimentology, geochemistry and palynology of the two samples. The samples were visually examined under a low-powered dissecting microscope, and the colour was established using Munsell charts. The material was then lightly disaggregated in a pestle and mortar and passed through a 2 mm stainless steel sieve to remove coarse particles before further analysis, following Gale and Hoare (1992). Geochemical analysis of the two samples was carried out by inductively coupled plasma mass spectrometry (ICPMS) at the University of Wales, Aberystwyth. Analysis of major elements was done on the sample from the Deep Skull by X-ray fluorescence using a Spectro-X Lab, loss on ignition was by the low-temperature method of Gale and Hoare (1992), and magnetic susceptibility was done using a Bartington MS2 meter. Although there was insufficient material in the H/6, 107” sample for X-ray fluorescence and magnetic susceptibility analyses, enough material was available in the sample from the Deep Skull for two replicate X-ray fluorescence (XRF) analyses.1 The material from the Deep Skull consisted of 36.9 g of angular fragments (~1–12 mm diameter, average about 3–4 mm) of a consolidated sandy diamict/muddy pebbly sand with ~5–10% porosity. The colour of the material was strong brown (7.5YR4/6). The matrix made up ~55% of the sediment and consisted of structureless, clast- to matrix-supported muddy sand, some areas with openwork texture, with occasional tubular voids probably reflecting ancient fungal hyphae or more likely fibres of vegetable matter.The sand fraction was sub-rounded to well rounded and mostly composed of dark materials in the very fine sand range, though some (usually openwork) pockets are in the 1 mm range and richer in quartz and mudstone. The material was generally fairly weakly aggregated. One fragment of sediment was clearly micritised by calcite induration 99

Chris Hunt and Graeme Barker Table  8.2.  ICPMS geochemistry of the material from the Deep Skull and from the H/6, 107" sediment sample (ng/g [ppb]) Element Lithium Scandium Titanium Vanadium Chromium Cobalt Nickel Copper Zinc Arsenic Rubidium Strontium Yttrium Cadmium Tin Antimony Caesium Barium Lanthanum Cerium Samarium Holmium Ytterbium Lead Thorium Uranium

H/6, 107"

Deep Skull

6560 7657 1127212 75719 52931 10549 8546 52388 35607 628 21402 48999 11255 296 403 16 910 432413 15761 26163 2155 374 969 2125 2146 611

5125 9664 1504785 93955 58425 11632 9578 781228 49012 829 18916 53589 13794 563 1375 178 842 449546 17721 27811 2599 500 938 4873 2973 670

and six fragments were indurated by a slightly irregular layer of dark amorphous mineral, probably manganese. About 45% of the material was composed of clasts larger than 2 mm and ~40% of the sediment was quartz in the form of sharp euhedral and irregular crystals, translucent, coloured strong brown to reddish yellow (7.5YR6/6 to 7.5YR5/6) and 5–10 mm in diameter. There were also fewer than 0.5% sub-rounded to well-rounded clear glassy quartz grains in the 2–4 mm range. Other clasts included ~3–5% sub-rounded to sub-angular mid-grey mudstone fragments up to 9 mm diameter. One mudstone fragment had a whitish quartz vein. Fragments of calcite were less than 2%, mostly irregular whitish, opaque fragments, plus three rhomboid whitish opaque crystals. Phosphate was less than 0.5%, in aggregates of dark-brown stumpy crystals. Bone was less than 1% and consisted of flakes of large bones, long-bone fragments probably of a swiftlet-sized bird or bat, plus irregular fragments, all well rounded, dark brown and heavily mineralised. Three small fragments of charcoal were present (< 0.1%). The H6, 107” sample was also strong brown in colour (7.5YR4/6) but much more even in texture, being a clayey sand in hand specimen. Under the low-power microscope no large clasts or charcoal fragments were visible, but it must be stressed that this was a very small sample, weighing ~10 g. Many elements, particularly copper, tin and antimony, but also scandium, titanium, vanadium, cobalt, nickel, zinc, arsenic, strontium, yttrium, cadmium, lanthanum, holmium, lead, thorium and uranium, are more concentrated in the sample from the Deep Skull; only lithium, rubidium, caesium and ytterbium are more concentrated in the sample from H/6, 107” (Table 8.2). 100

Anatomically Modern Humans in the Great Cave of Niah Table 8.3.  Major element geochemistry of the material from the Deep Skull by X-ray fluorescence (%)

Aluminium Silicon Phosphorus Sulphur Chlorine Potassium Calcium Iron Manganese

Replicate 1

Replicate 2

37.19 56.09 0.7505 0.3732 0.0757 0.6605 2.616 4.476 0.1221

35.39 55.49 0.853 1.448 0.1227 0.605 3.176 5.006 0.1637

Mean 36.29 55.79 0.80175 0.9106 0.0992 0.63275 2.896 4.741 0.1429

Table 8.4.  Magnetic susceptibility of material from the Deep Skull Sample Low-Frequency Mass Susceptibility 252.9

% Frequency Dependent 11.15

The XRF sub-samples from the Deep Skull show reasonable replication (Table 8.3); given the heterogeneity of the sediment, a closer replication would not be expected. Aluminium, potassium and some silicon will have been derived from clay minerals from eroding soil profiles or from windblown loess; the remaining silicon will have been from silica (silt and possibly sand), largely ultimately derived from aeolian sources (Stephens et al. 2005). Calcium may have been present as calcium carbonate, but the presence of sulphur suggests that some was in the form of gypsum. Further calcium may have been present as calcium phosphate, a major component of bone, but also present as crystals in the sample. Chlorine is often a component in wood ash. Iron oxides are pervasive at Niah and are likely, at least in part, to be derived from tropical soil profiles, along with the clay minerals, but in the context of the cave, some may be diagenetic. The manganese is likely to have been derived from diagenetic manganese oxides. The concentration of metals and metalloids in the Deep Skull sediment, compared with the H/6, 107” sample, is consistent with its high concentration of clay, though the concentrations of zinc, cadmium, tin, antimony and lead are inconsistent with a simple model of increased clay content: it suggests different sources – different stratigraphic levels – for the two samples despite their adjacency to each other. The XRF analyses of the Deep Skull sediments also indicate that they are significantly richer in aluminium than the altitudinally similar sediments from the upper part of the Hell Trench sediments (Tables 8.2, 8.3, 8.5), indicative of its higher clay content. It is significantly less rich in phosphorus and potassium and richer in iron than the surrounding sediments, and its magnetic susceptibility is two orders of magnitude higher than anywhere within the Hell Trench sequence (Tables 8.4, 8.5). Palynological sub-samples were decalcified in dilute hydrochloric acid, then deflocculated in sodium pyrophosphate and sieved on nominal 6 µm nylon mesh. Remaining silicates were removed using cold hydrofluoric acid. The residues were neutralised, stained with basic fuchsin and safranin and then mounted in Gurr Aquamount. All pollen in each sample was counted, together with all algal microfossils under transmitted light using X 1000 magnification. In each sample, an aliquot of about 200 organic particulates was also counted for palynofacies analysis 101

102 Table  8.5.  Geochemistry by X-ray fluorescence of the top half metre of the Hell Trench sequence in the West Mouth of Niah Great Cave (%) Depth (cm)

Aluminium Silicon Phosphorus Sulphur Chlorine Potassium Calcium Titanium Manganese Iron Copper Zinc Barium Strontium Low-frequency magnetic susceptibility

2.5

7.5

12.53

17.5

27.5

32.5

37.5

42.5

47.5

52.5

18.91 34.01 23.93 4.751 0.558 1.923 16.41 0.213 0.103 1.827 0.059 0.168 0.0068 0.059

20.97 40.66 25.2 1.5 0.475 11.01 1.784 0.227 0.135 1.784 0.055 0.129 0.0044 0.514

23.65 45.41 22.25 0.517 0.369 1.438 7.042 0.226 0.1 1.773 0.048 0.87 0.0038 0.368

24.08 49.86 19.81 0.529 0.313 1.49 4.366 0.265 0.066 2.017 0.047 0.61 0.0038 0.044

25.61 50.44 19.05 0.514 0.312 1.453 3.088 0.263 0.052 2.047 0.039 0.046 0.0033 0.032

24.82 51.82 18.81 0.532 0.293 1.397 2.487 0.263 0.052 1.986 0.039 0.045 0.003 0.036

24.49 52 18.93 0.667 0.28 1.272 3.041 0.256 0.045 1.848 0.039 0.043 0.003 0.035

23.07 43.75 23.61 0.849 0.286 1.422 7.714 0.233 0.123 1.689 0.044 0.104 0.002 0.046

24.54 47.06 21.74 0.578 0.231 1.514 4.939 0.248 0.077 1.861 0.044 0.075 0.0023 0.044

23.04 45.12 23.11 0.759 0.245 1.362 7.03 0.242 0.098 1.748 0.046 0.088 0.0024 0.048

1.11336

0.510204

0.665722

0.682731 0.678204

0.936874

1.808396

1.496388

Note: Data from Hunt et al. 2007.

0.666996 0.675926

Anatomically Modern Humans in the Great Cave of Niah Upland

D ryland

R egeneration

S

P

A

ru

nu ue s rc te u s r C cu l a s ia u E ar r ic i n ac a ea e M yr i A ca lb iz A z ia st e C rac al li ea La c ar e ct p a M u ca ac e ar an ga

Low land Fores t/C atholic

Q

S w am p F

cr o M st i e l ch ia u O ce m nc a au e re R o sp um hi z er S op m a on ho ne ra C am ra ti S pn a c an o as t s S ir ia p e r e o l ph m ar is a a E gn la e um E oc up ar h p Ly or u s c o b ia Ly po c e c o d iu a e M po m yr d i ce u i P s t ic m r n u ol se m y a la R po go ub di ia u m R ut c ea t y a c e pe S ap ea i e S nd el ac a e S gin ae m el a l U ll s a rt m ic o A ac ot ln e a h u e t r il C s et as e t C an ya o p t P h a s is od c e oc a e ar pu s

Mangrove

Skull C o n te n ts H6 107" 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10 10

10 10 10 10

20 10 10 10 10 10 10

ry o on c o ce cc u e r n tr s i d ic P in y s s i io la i d t e s c S te a e a ys ts p lg S to a l p i d i cy n ro s gy ium t s T yp r e a Z y g 119 n R em ec a yc le d P

ot

C

B

ar T r de av e T t . P ll e d ot a l o ll e n P ol le n

us

in

F

ae ex P

um

R

Algal

In

e et ol a

la

id

cu

ps

un

ro

an R

P

C

te

en

C

ce

m

on

ta la he ure c e n o a ae yp p o e d L a r a ia c bi c e ea a a L i ta e e lia e c P ea la n e P ta g oa o ce ae

yl ph yo ar

C

C

40

FT

M S ang w L o a mr o v U wl p e pl an Fo a D ndd F res ry or t / la es Sw R nd t/ C a eg m at p en ho er lic H at er io ba n ce ou s

Herbac eous

20

Skull Co n te n ts H6 107" 20

40

20

40

S oil D erived

lg ae In s P ec la t R nt e c A c y c e ll m l w F o r p e d a ll un h s & ga ou cu l s F un h y tic le F ga p h un l a V ga s p e A l o In M z o re e os M r tin po a t it re ur e e Im m at ur e

Im m ature

A

T

he

rm

al

ly

m at ur e T M T a M mo P p rp ol o h le ll e o n n u s

M ature

20 10 10 10

60

80

100

10

200

400 10 10 10 10 10 10 10 10 10

Inert

S o In i l- d e P r t e r iv al yn e d of ac ie s

10 10 10 10 10 10 10

Skull C o nte nts H6 107" 2 0 4 0 6 0 10 10 2 0 4 0 6 0

2 0 10

10 2 0 10 10

10 2 0 4 0 6 0 8 0 1 0 0

200 400 600

Figure 8.8.  Palynology of the sediments in the Deep Skull and in the H/6, 107" sample, analysed by C. O. Hunt (illustration by the authors). (Hunt & Coles 1988). Pollen was identified using the type collection made by Bernard Maloney at Queen’s University Belfast, plus reference to the published literature. Taxa were calculated as percentage of total pollen and spores, but excluding Justicia, which in other work at Niah is excluded from the percentage calculations throughout because it ‘drowned’ all other pollen signals (Hunt et al. 2007; 2012). Pollen is the most common component (67%) of the particulate organic matter in the sample from the Deep Skull but constitutes only some 13% of the organic material in the H/6, 107” sediment (Figure 8.8). In the Deep Skull sample algae are also important (14%), and there are lesser amounts of thermally mature material, plant cell walls and cuticle, recycled palynomorphs, inertinite (chemically inert amorphous carbon derived from bedrock), amorphous organic matter, fungal hyphae, spores, zoospores and vesicular arbuscular miccorhyzae. By contrast, the organic material in the H/6, 107” sample is dominated by thermally mature (charred) material, including charcoal fragments, thermally mature amorphous material and pollen; there are also some plant cell walls and cuticle, fungal hyphae and spores, amorphous organic matter and very rare insect fragments and algae. The pollen in the Deep Skull sample is dominated by herbaceous taxa, mostly Poaceae, but also Pteropsida, Cyperaceae, Labiatae, Caryophyllaceae, Chenopodiaceae, Ranunculaceae and Rumex. Dryland species include Ericaceae, Casuarina and Myrica, and upland species include Podocarpus, Prunus, Sterculia and Quercus. Also present is a wide variety of taxa from mangroves (Acrostichum aureum, Meliaceae, Oncosperma, Rhizophora, Sonneratia caseolaris), swamp forest (Campnosperma, Santiria) and lowland forest (Myristica, Myristicaceae, Rutaceae, Sapindaceae). Ecologically indeterminate taxa include Euphorbiaceae, Elaeocarpus and Lycopodium spp. By contrast, the sparse pollen assemblage from the H6, 107” sample is dominated by upland taxa, principally Podocarpus but also Alnus, with some dryland (Casurina, Ericaceae), regeneration (Albizzia) and herbaceous (Cyperaceae, Poaceae, Pteropsida monolete) taxa, and mangrove, lowland and swamp forest taxa are all absent. 103

Chris Hunt and Graeme Barker The pollen assemblage from the Deep Skull sample appears to be ecologically mixed. The component that includes abundant herbaceous taxa and altitudinally limited taxa such as Prunus and Alnus might suggest stadial conditions and temperatures below 23° C compared with the present day 27° C. The other component includes back-mangrove, lowland forest and swamp forest taxa suggestive of the kind of local lowland and coastal vegetation that developed at Niah during interstadials and during the Holocene (Hunt & Premathilake 2012; Hunt & Rushworth 2005a, 2005b; Hunt et al. 2007, 2012). In general terms, the flora in the Deep Skull sediment is similar to that of Pollen Zone H-7 of Hunt et al. (2012) about a metre below the Deep Skull findspot, rather than to the interstadial assemblages in the upper part of the Hell Trench sequence. The small size of the pollen assemblage from the H/6, 107” sample makes its interpretation uncertain, but it is likely to reflect dry, somewhat open, vegetation. The presence of Albizzia is climatically important, because today this is characteristic of hill and lower montane environments in Borneo between 600 and 1,600 m above sea level with annual temperatures in the range of 24 to 19º C (Hunt et al. 2012). It is probable, therefore, that the H6/107” sediments were laid down during a stadial. Unlike the Deep Skull assemblage, it appears ecologically ‘unmixed’ and is very similar to the Podocarpus-dominated basal assemblage Pollen Zone H-1 in the Hell Trench (Hunt et al. 2007; 2012), 2 m lower than the position of the H/6, 107” sample in the stratigraphy. It is rather different in composition from the pollen assemblages found at the top of the Hell Trench sequence, in monoliths taken from the same approximate altitude and at most 3 m removed horizontally. They indicate that this part of the Hell Trench sequence contains a major interstadial, characterised by abundant mangroves, lowland forest and an assemblage of brackishwater diatoms with signs of climatic deterioration at the very top (Hunt et al. 2007; 2012). The likelihood is that the H/6, 107” sample was located at a slightly higher layer stratigraphically than the Hell Trench monoliths, possibly because of cut-and-fill deposition in the stream channel that laid down Lithofacies 2, and relates to the following stadial.

The Niah Deep Skull and Cultural Modernity in Southeast Asia The age and stratigraphical position of the Deep Skull have been contentious for many years, but the work of the Niah Caves Project since 2000 has shown that the skull is undoubtedly Pleistocene in age rather than a Neolithic or Metal Age intrusion and is broadly similar in its antiquity to the age of circa 40,000 BP indicated by the original 1958 radiocarbon dating (Barker et al. 2007). At the same time, however, as we have described, ongoing sedimentary, geochemical and palynological studies of the surviving sediments in the West Mouth and of two archived sediment samples from the original excavations – one directly associated with the skull and another taken a few inches from its findspot – chime with the dating evidence to suggest that the skull is not coeval with the upper part of the Hell Trench sequence, where spatial considerations suggest that it was located. New ABOX radiocarbon dates on charcoal indicate that the channel-fill sediments (Lithofacies 2) here date to around 40,000–47,000 cal BP, whereas the direct U/Th date for the skull, altitudinally at the same level as these sediments, places it ~32,600–37,800 BP. The channel-fill sediments at this level appear to have been laid down during interstadial conditions, whereas the sediments attached to the Deep Skull contain a mixed assemblage that in part reflects stadial and in part interstadial conditions. The Harrissons termed the rich assemblage of butchered vertebrate bones mixed with abundant ash and charcoal that they found in the lower part of the Hell Trench and extending to the base of their excavations under the rock overhang the ‘bone and ash layer’. Plotting the grid squares and spit depths of the vertebrate bone fragments in the Harrisson Excavation Archive material has shown that the ‘bone under ash layer’ was primarily located along the channel that carried the Lithofacies 2 ephemeral streams across the front of the West Mouth into the overhang and its sinkhole (Piper & Rabett 2009b). Given this stratigraphic context it is unlikely that 104

Anatomically Modern Humans in the Great Cave of Niah much of this material was found in situ where it had been discarded. However, the discovery by the Harrissons of several sets of articulated animal bones and the only slight evidence for water abrasion or erosional damage noted by Piper and Rabett in their analyses of the Harrisson vertebrate fauna indicate that the latter had not travelled very far before it ended up in the channel – presumably just 2–3 m down from the cave lip as part of the Lithofacies 2C colluvium, or the same sort of distance from the interior side of the basin, or along the channel itself at the time of stream activity. It is possible, therefore, that the Deep Skull and the few associated limb bones arrived at their find location through similar slippage from a place of deposition a few metres away, but the evidence discussed earlier suggests that it is more likely that they were found in situ.The presence of the tibia and femur, both substantial limb bones, in more or less direct association with the skull also indicates that substantial fluvial transportation of the human skeletal remains is unlikely. The significant differences in the sedimentology, geochemistry and palynology of the upper part of the Lithofacies 2 channel-fill sediments compared with those of the Deep Skull sediments, along with the dating discrepancies, provide strong hints that the Deep Skull may have been located in some kind of pit dug down from the Lithofacies 4 sediments, sediments that started to be laid down from circa 38,000 cal BP, around the time of the Deep Skull. Pit digging was certainly widespread across the West Mouth archaeological zone during the formation of the Lithofacies 4 sediments, though those that have survived and been studied by the project appear to have been related primarily to food processing and storage (Barker et al. 2007). The interstadial-stadial mixed pollen assemblage would certainly make sense in terms of a pit fill mixing the Lithofacies 4 surface sediments from which a pit was dug and the upper Lithofacies 2 sediments into which its excavation would have penetrated. The material derived from cavities in the Deep Skull is visually almost indistinguishable from the sandier sediments from Lithofacies 2, and its sandy texture and particularly the patches of openwork sand are consistent with a broadly fluvial origin, as is the Lithofacies 2 sediment as a whole (Gilbertson et al. 2005; Hunt et al. 2007; Stephens et al. 2005; and see the preceding discussion). This is particularly indicated by the rounding of many components including the sand grains, mudstone clasts and bone fragments and the material with openwork texture. The calcite and manganese induration are diagenetic. Both are common in the cave and in Lithofacies 2 (Stephens et al. 2005).They probably reflect a water-table feature that ran through the skull in the past.The material from the skull is, however, significantly richer in large quartz than Lithofacies 2 sediments, which contain no quartz more than 0.2 mm diameter (Stephens et al. 2005; in press). No quartz more than 0.2 mm has in fact been identified anywhere in the cave deposits at Niah, although silt-sized quartz of aeolian origin is widespread. The quartz crystals in the Deep Skull sediment sample cannot have been introduced by running water because if so they would have been rounded by transport processes. The most parsimonious explanation, therefore, is that these bright, attractive, crystals were placed in or adjacent to the skull by human agency. The implication is that the skull and other bones, and the crystals, were part of a secondary burial not recognized by the original excavators. The limestone massif of the Gunung Subis, in which the Niah Caves are situated, and the alluvial lowlands around them, are extremely unlikely to have been the source for these crystals, which were almost certainly derived from a granitic igneous rock. The nearest suitable granitic rocks include the summit of Mount Kinabalu and lower altitude locations in Sabah, about 400 km away, isolated massifs in interior Borneo near the KalimantanSarawak border about 200 km away and the Schwaner Mountains in southwest Borneo about 510 km away (Hutchison 2005; Figure 8.1). Intriguingly, the putative secondary burial represented by the Deep Skull and associated limb bones, as well as the quartz crystals, may not be the only early example of processing the dead at Niah. A further 27 human cranial fragments were found by Piper and Rabett amongst the archived vertebrate fauna from the deepest levels of the Harrisson excavations, from spits some 10 m to the south of the Deep Skull findspot. A red wash, probably from a tree resin, has been found 105

Chris Hunt and Graeme Barker on the inner surface of some of these (Pyatt et al. 2010).Whether human skulls were used as convenient receptacles for the colouring medium or were deliberately coloured as part of funerary rituals cannot be ascertained, but it is noteworthy that the use of pigmentation has been cited as part of the package of behaviours associated with early Homo sapiens in other parts of the world (McBrearty & Brooks 2000). Recent years have seen a lively debate about the nature and degree of ‘cultural modernity’ amongst the prehistoric inhabitants of Southeast Asia and Australasia, part of the context of course of this volume. Authors such as Brumm and Moore (2005), O’Connell and Allen (2007) and Habgood and Franklin (2008) have suggested that large components of the range of behaviours that have come to be associated with cultural modernity in Europe and Africa (e.g., Bouzouggar et  al. 2007; d’Errico 2003; Henshilwood & Marean 2003; Hiscock & O‘Connor 2006; Klein 2000;Vanhaeren et al. 2006; Zilhão 2007) are rare or absent in the archaeological record of Island Southeast Asia and Australasia.The apparent ‘primitiveness’ of lithic material culture in the region has also been remarked upon (e.g., Pawlik 2010), though there is a rich ethnography for the sophisticated use of non-durable materials by people in the region (e.g., Satterthwaite 1986), and many of the ‘culturally modern’ behaviours defined by archaeologists are associated with clothing, which largely seems to have been unnecessary for most indigenous Australasian and Island Southeast Asian people, apart from the Tasmanians during the Last Glacial Maximum (Gilligan 2007c; this volume). There are, however, other signs of cultural modernity in the region, including art, for instance, hand stencils from Tasmania (Cosgrove & Jones 1989; Harris et  al. 1988) and East Kalimantan (Chazine 2005), and figurative and animal drawings in Thailand (Anderson 2005); bone technology, for instance, at Niah from ~45,000 BP (Barker et al. 2007; Barton et al. 2009; Rabett & Piper 2012; Rabett et al. 2006); complex foraging behaviours and plant processing technologies associated with landscape manipulation, reported from Niah (Barker et al. 2007; Hunt et al. 2007, 2012; and see Piper and Rabett, this volume), Papua New Guinea (Summerhayes et al. 2010), New Ireland (Leavesley 2005) and Australia (Mooney et al. 2011); the use of ochre, for instance, at Lake Mungo (Bowler et al. 2003) and Kimberley (O’Connor & Fankhauser 2001); and the use of beads, for instance, at Devil’s Lair (Dortch 1984) and Cape Range (Morse 1993a). In the contexts of these debates one of the most interesting lines of evidence for cultural modernity is the relationship between people and their dead, which includes the archaeologically visible processing and sometimes burial of human remains. This is seen early in Africa and the Middle East with the processed skulls at Herto, Ethiopia (White et al. 2003) and burials at Qafzeh and es-Skhu¯l (Vandermeersch 2006) and is also known with some later Neanderthals (Langley et al. 2008; Solecki 1975, 1977). In Island Southeast Asia and Australasia, burial seems to have been part of the human repertoire from very soon after the first signs of human presence in the region, most notably the burial at Lake Mungo associated with red ochre (Bowler et al. 2003; Habgood & Franklin 2008) and the later burials at Willandra Lakes (Grün et al. 2011) and Roonka (Robertson & Prescott 2006).The Niah Deep Skull burial(?) and the cranial fragments reported by Pyatt et al. (2010) may be part of a wider tradition of the cultural treatment of the dead, therefore. Such a tradition has certainly persisted in Borneo through the prehistoric and early historic periods (Barker et al. 2011; Chazine 2005; Cole 2012; Krigbaum 2005; Lloyd-Smith 2009; Szabó et al. 2008), with ethnohistoric and ethnographic reports affirming the long-lived cultural significance of the head (e.g., Freeman 1979; Evans 1990; King 2007; Schiller 2001; Sellato 1994; Schiller 2001; Szabó et al. 2008;Wadley 2004) together with mortuary rituals involving complex processing of the dead and secondary reburial (Metcalf 1991; Nicolaisen 2003; Sather 2003).

Acknowledgments This chapter draws on the work of a large number of researchers in the Niah Caves Project, but we would like to acknowledge in particular the contributions of Tim Reynolds to our 106

Anatomically Modern Humans in the Great Cave of Niah understanding of the Pleistocene archaeology of the West Mouth; of David Gilbertson, Alan Dykes, Helen Lewis, Sue McLaren, Garry Rushworth, James Rose and Mark Stephens in terms of the geoarchaeology/geomorphology; of Michael Bird, Tom Higham and Alisdair Pike for the dating programme; and of Philip Piper, Ryan Rabett and the Earl of Cranbrook for the zooarchaeology. We would also like to acknowledge the enthusiastic support of the staff of Sarawak Museum, especially its Director Ipoi Datan, and of Barbara Harrisson, without which the Niah Caves Project could never have developed successfully; and the essential financial support of the Arts and Humanities Research Board, Arts and Humanities Research Council, British Academy, British Academy Committee for Southeast Asian Studies and the Natural Environment Research Council (Radiocarbon Dating Facility, University of Oxford). The paper was written while COH was on sabbatical leave from Queen’s University Belfast at the McDonald Institute for Archaeological Research, University of Cambridge. He thanks St John’s College for dining rights and all in the McDonald for their hospitality. Note 1. The Deep Skull sample also included loose contaminant material that was removed with tweezers prior to the analysis: a paper fragment; several reddish human hairs; two fresh-looking termite fragments and vegetable fibres and monocotyledonous leaf-lamina fragments probably from palm leaves, presumably derived from the packing in which the skull was sent to the British Museum; and seven semi-fresh fungal hyphal strands probably relating to the slow drying of the skull after excavation.

107

Chapter 9 Faunal Biogeography in Island Southeast Asia Implications for Early Hominin and Modern Human Dispersals

M. J. Morwood

Introduction Over the past 10 years, research projects in Island Southeast Asia (ISEA) have yielded information on island faunal sequences and have demonstrated that some oceanic islands were occupied by early hominins – including an endemic species, Homo floresiensis (Brown et al. 2004; Brumm et al. 2010; Morwood et al. 2005, 2008; Bergh et al. 2009). Such projects have also documented the dispersal of modern humans across the region and subsequent economic, technological and demographic changes (Barker et  al. 2007; Bellwood 2005; O’Connnor 2007b). Much of the significance of these findings, however, lies in their context; including evidence for the history of faunal/hominin succession, dispersal, evolution and extinction in continental Southeast Asia (Sunda) and Sahul. The basic premise for this chapter is that the discipline of biogeography (i.e., the study of the distribution of species and the determining factors) and information on the geographical and temporal distribution of other faunal species provides an essential holistic context for explaining the history of human dispersals, impacts and connections. Such an approach is particularly apt in Island Southeast Asia, where biogeography, as a scientific discipline, began with the work of the 19th-century biologist Alfred Russel Wallace (1876). Also, despite the strategic importance of ISEA, on the ‘doorstep to Australia’ as the source area for the First Australians, and for its enormous potential to shed further light on the evolutionary and dispersal history of extinct and living animal species – including hominins – archaeological research in the region has been undertaken only in a few sites, in few areas and on a relatively few islands. A biogeographical perspective may value-add to the available database and possible Editors’ note: We, and many others in the palaeoanthropological community, were very saddened to hear of Mike Morwood’s death in July 2013. He was a friend and colleague of many of the contributors to this volume, and at the time of his death, he was probably the best-known archaeologist based in Australia. His discovery of H. floresiensis at Liang Bua, Flores, Indonesia, was one of the most remarkable and least expected finds in the history of palaeoanthropology and deservedly attracted global attention. As editors, it was both a pleasure and a privilege to have known him, and he will be greatly missed.

108

Faunal Biogeography in Island Southeast Asia interpretations. More specifically, differences between the regional distributions of hominins at different evolutionary and cultural stages (e.g., early hominins, the first modern humans, Neolithic farmers) and other terrestrial mammals provide proxy evidence for the development of cultural dispersal mechanisms and a measure of hominin technological prowess.

Faunal Dispersals Generally land animals reach oceanic islands by single- or multiple-stage swimming or rafting from adjacent mainlands, but relatively few land animals have the required capacity to cross substantial water barriers. As a result, oceanic islands have depauperate, unbalanced terrestrial faunas with no large carnivorous mammals and with the number of species and degree of endemism determined by island accessibility, age and size (Flannery 1995; MacArthur & Wilson 1967). Animals more adept at water crossings are also over-represented in island faunas – for example, modern humans, rodents, tortoises and proboscideans (mammoth, Stegodon, Elephas). However, islands vary greatly in colonisation difficulty depending on length of required water crossings, ocean currents and winds, as well as island size, age and geological history – all with consequences for the differential distribution of plants and animals.The distribution of animal species on islands compared to mainland sources, therefore, provides a quantitative measure of island accessibility and faunal dispersal abilities, as well as evidence for dispersal mechanisms – with the latter including ocean currents, long-term tectonic movements, land connections at times of low sea level and human transportation of commensal species. The distribution of terrestrial mammals on the 17,000 islands between Sunda and Sahul well illustrates these points, with the overall dispersal pattern indicating at least four levels of colonisation difficulty: 1. Continental islands that were part of the Asian mainland at times of glacially induced low sea levels (e.g., Sumatra, Java, Borneo). At such times, animals could walk to these now-island areas that potentially acquired the full range of Indo-Malay faunal species, including ~240 types of terrestrial mammal (Heaney 1986; Meijaard & Groves 2006). However, with past and present environmental constraints to continent-wide faunal dispersals and gene flow, these islands still have some endemics (e.g., the Javanese deer Axis javanicus). In the Indonesian region, the eastern edge of the Sunda continental shelf at times of low sea level is delineated by the Wallace Line – the major biogeographical boundary in Southeast Asia – while the western edge of the Sahul continental shelf is delineated by another biogeographical boundary, the Lydekker Line. The island zone in between is commonly referred to as Wallacea. 2. Wallacean islands that have never been connected to the Asian mainland are characterised by impoverished, unbalanced (lacking top carnivores) and highly endemic, terrestrial animal suites that are mainly derived from Sunda. Sulawesi, for instance, has ~120 types of land mammals, ~60% of which are only found there, including two marsupials derived from Sahul by tectonic rafting – the Little Sulawesi Cuscus and the Bear Cuscus. 3. Wallacean islands requiring more difficult sea crossings for colonising animals have very impoverished, unbalanced and highly endemic land mammal suites that are primarily derived from Asia. For instance, before the arrival of modern humans and associated impacts, Flores had just three types of land mammals (Stegodon, rats and early hominins), while Timor had just two (Stegodon and rats). 4. Sahul was an almost impossible colonisation prospect for Asian land animals, and only rats and modern humans made it unassisted. Before the latter arrived, the continent had a balanced, endemic fauna that included ~237 types of marsupials, monotremes and placental rodents. 109

M. J. Morwood

Figure 9.1. The distribution of land mammals from Sunda to Sahul.This provides a coarse measure of island accessibility. Note figures do not include modern humans and rats are counted as 1 (illustration by the author). Differences in faunal distribution across ISEA show rapid west-to-east and north-to-south decreases in number of land mammals (Figure 9.1). These differences reflect increasing distance from the main faunal source area – continental Asia, as well as sweepstake dispersal of animals by the predominantly north-to-south flowing ocean currents, the Indonesian Throughflow (Figure 9.2; Kuhnt et al. 2004; Hautala et al. 2001). Tucked away in the southeast corner of ISEA, Flores, Timor and Sumba were thus very difficult colonisation prospects with very depauperate land faunas.

Island Faunal Sequences and Evolution Because island faunas exhibit diagnostic evolutionary, phylogenetic and behavioural trends worldwide, faunal sequences, as evident in the extant, palaeontological and archaeological records, represent another measure of island accessibility and how this has changed over time. They also provide evidence for insular evolutionary pressures, climate change, natural disasters, long-term tectonic movements and human impacts. The last-named includes reduction in endemic biodiversity and the introductions of exotic species from which human movements and contacts across the region may be inferred – for instance, the Pacific rat in Neolithic sites, cuscus in Timor, the Sulawesi warty pig in Flores and dingo in Australia (Anderson & O’Connor 2008; Heinsohn 2001; Matisso-Smith 2007; Bergh et al. 2009). In combination, these lines of inquiry can be used to predicatively model colonization sequences and identify islands or sites of high faunal, paleontological and archaeological research potential. However, the size of an island relative to the size of the evolving species greatly influences the course evolution will take. On islands that are relatively large relative to the size of an incoming species, adaptive radiation may occur to fill empty habitat niches to a spectacular degree, as was 110

Faunal Biogeography in Island Southeast Asia 130°E

120°E

110°E Hainan

Luzon

Pacific Ocean

Palawan

10°N

10°N

Mindanao

Moluccas Borneo



0° Sulawesi Seram

SUNDA Java

West Irian Aru

Bali Flores

10°S

Sumba

Timor

SAHUL

10°S

Indian Ocean Australia

Pleistocene Coastline 110°E

120°E

130°E

Figure  9.2.  Predominant ocean currents in ISEA from the Pacific to the Indian Ocean (the Indonesian Throughflow) would have facilitated north-to-south dispersal of land animals but made west-to-east crossings difficult (after Kuhnt et al. 2004; Hautala et al. 2001; illustration by the author). the case with lemurs in Madagascar, finches in the Galapagos Islands, rodents in the West Indies and moas in New Zealand. And some descendant species may become larger: giant rats and giant tortoises evolved on many islands worldwide. In contrast, large mammals tend to downsize on relatively small islands – for example, pygmy elephants, mammoths, hippos and deer (Vos et al. 2007b; Foster 1964; Sondaar 1977). In the absence of predators and inter-specific competition, 111

M. J. Morwood smaller individuals will be advantaged by their reduced food requirements and shorter pregnancies. Associated changes may occur in locomotion, dentition, brain size and behaviour. Island isolation and the adaptive advantages of established species over any stray newcomers can also lead to retention of early faunal lineages elsewhere long extinct, for example, lemurs in Madagascar, Babirusa in Sulawesi and monotremes in Sahul. Such relict animals often indicate how little we actually know about the early history of animal dispersal and evolution in the region. On the other hand, animals that have evolved in largely predator-free, island isolation generally do not fare well when people arrive, and this vulnerability to humans is another characteristic of island faunas. But a complex of factors determines the timing and severity of human impacts. For instance, modern humans have devastated native faunas on the continental islands of Java, but evidence from a 2-million-year faunal sequence shows that fluctuations in climate and the arrival of new species from other parts of Asia were the main determinants of change and that human activities became the prime cause of faunal extinctions only after the onset of Neolithic farming (Morwood et al. 2008; Westaway et al. 2007a). Since that time, human population increases, with associated forest clearance, habitat destruction and hunting, have caused the staggered disappearance of many animal species, including elephant, tapir and tiger. On Borneo and Sumatra, there is a similar late Holocene and ongoing demise of large animals, including Javan rhino, tiger, orang-utan, sun bear and tapir (Cranbrook & Piper 2007a; Meiri et al. 2008). The shared extinction pattern reflects the fact that these are large, fertile islands with high rainfall that supported a greater range of habitats, including extensive tracts of rainforest; and that they were periodically connected to the Asian mainland with a greater range of faunal species (including large carnivores and early hominins).This pattern is very different from the ‘blitzkrieg’ scenarios indicated for the dry, resource-poor islands of East Indonesia, such as Flores  – and Australia  – where megafaunal extinctions occurred soon after the arrival of modern humans (Roberts et al. 2001; Bergh et al. 2009).

Early Hominin Dispersals For the last ~1.5 million years, hominins have been part of the faunal suite in ISEA, beginning with Homo erectus, a species that almost certainly migrated to what is now Java when it was part of the Southeast Asian mainland and did so at the same time as a major influx of other newcomer species (Vos 1985; Bergh 1999). In contrast, it was never possible to walk to Wallacean islands, such as Flores, where the well-documented faunal sequence has all the biogeographical and evolutionary hallmarks of profound insular isolation (Figure 9.3). Before the arrival of modern humans, tortoises, Komodo dragons, Stegodonts, rats and hominins were the only large terrestrial animals to colonise the island; some tortoises and rats became very large; Stegodonts and (probably) hominins decreased in size (Bergh et al. 2001; Kaifu et al. 2011; Musser 1981); Stegodonts and a late representative of an early hominin lineage continued long after their Asian continental brethren had become extinct (Morwood et al. 2008, 2009; Westaway et al. 2007b); and the arrival of modern humans was associated with extinction of the only two large-bodied mammals on the island – the endemics Stegodon florensis and Homo floresiensis. Until relatively recently, it was assumed that only modern humans had the capacity to make the sea crossings required to reach such oceanic islands (e.g., Birdsell 1977; Davidson & Noble 1992). However, stone artefacts at Wolo Sege in the Soa Basin of central Flores have been dated to 1.1 million years BP (Brumm et al. 2010), so, at least one early hominin species from continental Asia managed to successfully colonise Flores. But this seems at odds with the natural dispersal patterns of other Asian animals across the region, indicating that early hominins may have indulged in behaviours that increased the chances of being accidentally swept out to sea – use of simple bamboo rafts for exploitation of near-shore coastal resources, for instance. 112

Lithostratigraphy Age

Sites

Colossochelys sp. Stegodon sondaari Stegodon florensis subspec. Hooijeromys nusatenggara Papagomys armandvillei P. theodorverhoeveni Murine rodent spp. Varanus hooijeri Varanus komodoensis Archaic Homo Homo sapiens Sus celebensis Sus scrofa Macaca fascicularis Pa P arra adoxurus hermaphrodites Hy Hyysstr trriix javanica Dog Bovids Cervu vus timorensis

Faunal Biogeography in Island Southeast Asia

Recent

West Flores

4 Cave sediments

Liang Bua

7

10 Erosional hiatus White tuff Black tuff 17

Volcanic eruptions 17-11 V 17-11 kyr

Liang Bua

Cave sediments 100 Hiatus Conglomerate

Hiatus 190

Liang Bua

Ola Bula Formation

Hiatus Gero

Sandstone 700 Interval 880

Kobatuwa Boa Lesa Mata Menge

900

Tangi Talo

V Volcanic eruptions 900 kyr

Tuff Interval 1,050

Ola Kile

Soa Basin, Central Flores

Hiatus Limestone 650 Interval

Breccia

Wolo Sege

Hiatus

1,860

Figure 9.3. The Flores faunal sequence over the past 1.1 million years is characterised by very few species and long-term phylogenetic continuity. But there were also two faunal turnovers – one associated with a huge volcanic eruption 900,000 years ago, and the second associated with the arrival of modern humans (after Brumm et al. 2010; illustration by the author).

Also given its geographical position, Flores is unlikely have been the only Wallacean island settled by early hominins. One obvious candidate for early hominin occupation is Sulawesi immediately to the north, which had a much greater range of land animals, including elephants, pigs, bovids, squirrels, civets, shrews, tarsiers and monkeys, showing that this island was a much easier colonisation prospect for Asian animals. In fact, hominins could have occupied Sulawesi much earlier and it may be the immediate source population for the first Flores colonists: at times of low sea level, Greater Sulawesi, which included Selayar Island, extended to within 80 km of Flores, while the predominant north-to-south ocean currents, the Indonesian Throughflow, would have assisted accidental dispersals in that direction but proved a formidable obstacle for west-east movement between islands (Morwood & Oosterzee 2007). This may be why premodern hominins apparently did not reach the island of Timor to the east of Flores (personal observation) – or Australia. 113

M. J. Morwood

Modern Human Dispersals Recent genetic studies indicate that the first modern humans dispersed out of East Africa by crossing of the Hormuz Strait between 75,000 and 60,000 years ago and then migrated along the Arabian and South Asian coasts to Southeast Asia. The rake-like structure of deeply rooted mtDNA lineages for extant groups along this route suggests that the dispersal occurred relatively rapidly (Macaulay et al. 2005; Oppenheimer 2004, this volume) and that resulting modern human interactions with earlier hominin populations are likely to have been complex and variable: for instance, some interbred with long-established ‘Denisovan’ populations before their descendants occupied ISEA, Sahul and West Melanesia. There is also evidence that peopling of Southeast Asia occurred in multiple waves, with modern East Asians mainly descending from later migrants (Reich et al. 2011). The first modern human immigrants were, therefore, familiar with coasts and marine resources long before encountering ISEA and crossing to Sahul. This is supported on biogeographical grounds, with Sahul an almost impossible colonisation prospect for Asian land animals: over the past ~9 million years only rodents, presumably adrift on flotsam, managed to cross to the Papuan section of Sahul, where they underwent rapid adaptive radiation before dispersing into Australia ~6 million years ago (Aplin 2006; cf. Flannery 1988). Initial colonisation of ISEA, then Sahul and West Melanesia as far as the southern end of the Solomon Islands, by modern humans was definitely not achieved by people being washed out to sea on flood or tsunami debris. To the contrary, it was accomplished with powered, navigable craft and complex technologies for harvesting marine resources, as is apparent from the capture of tuna, a fast swimming, pelagic species, by the occupants of Jerimalai Cave in East Timor 42,000 years ago (O’Connor 2007b, 2012; O’Connor et al. 2011a) and by the minimum 200 km over-horizon crossing required to reach Manus Island to the northwest of the Bismarks (Fredericksen et al. 1993). If so, then more recent, better-documented colonisations of ISEA by other peoples with sophisticated maritime craft and technologies may provide useful, testable analogies for the first modern human immigrants. Two examples will suffice – the mid- to late Holocene expansion of Austronesian speakers into Southeast Asia and Oceania; and Japanese military campaigns in the region during World War II.

Austronesian Colonisation of ISEA Austronesian languages are geographically spread from Taiwan to ISEA and the adjacent mainland, east to Micronesia, Melanesia and Polynesia in Oceania, and west to Madagascar off the coast of East Africa. Lexical relationships between these languages indicate movement throughout the area of Neolithic farming peoples with a range of plant and animal domesticates (pig, dog, chicken), pottery and developed skills for maritime exploitation, including watercraft with increasing capacity for long-distance sea crossings and island colonisation. Dates for the earliest Austronesian sites range from ~5000 BP in Taiwan, 4000 BP in the North Philippines, 3500 BP in East Indonesia, Micronesia and West Melanesia, and 3000 BP in West Polynesia and then to between 2,000 and 800 years ago in the far-flung extremities of East Polynesia (Bellwood 2005, 2010; Diamond & Bellwood 2003; cf. Donohue & Denham 2010). The geographical and chronological patterning of this expansion, as based on the linguistic and archaeological records, indicates rapid movement of peoples across sections of this vast area – generally from north to south and from west to east. However, the rake-like structure of the Austronesian language family tree (Pawley 1999), plus near synchronous dates for the first Austronesian sites in ISEA, Micronesia and Melanesia, indicate that occupation occurred as a result of multi-directional, multi-stage dispersals, that could involve lengthy sea voyages, as in the case of the 1,000 km crossing required to colonise Fiji in East Melanesia. 114

Faunal Biogeography in Island Southeast Asia Extant language distributions indicate that Austronesians initially occupied only offshore islands and coastal enclaves, as along the north coast of Papua, where there were long-established food-producing communities; while archaeological evidence shows strategic pauses occurred during which social and technological innovations occurred. In the Bismarks of West Melanesia, for instance, a cultural complex associated with distinctive Lapita pottery emerged. Lapita communities then established settlements along the south coast of Papua New Guinea, and the Solomon Island chain south to New Caledonia, and voyaged east to occupy Fiji and West Polynesia, where there was another pause of ~1,000 years before the appearance of distinctive cultural traits associated with initial peopling of East Polynesian islands (Bellwood 1979).

Japanese Colonisation of ISEA Comparative evidence for Japanese occupation of ISEA and West Melanesia is much more detailed but shows remarkable parallels with inferred Austronesian dispersals (Figure 9.4; Marston 2005; Rottman 2005). For instance, deployment of troops and occupation of key sites in ISEA occurred very rapidly in a series of north-to-south, multi-pronged movements of troop-carrying fleets between December 1941 and May 1942. In the western sector, these were initiated from an established base on the Southeast Asian mainland to take Peninsula Malaya, Singapore, Sumatra and West Java. In the east, conquest involved task forces from Taiwan moving south into the Philippines, then East Borneo to West Java, North Sulawesi to South Sulawesi to East Java and the Moluccas to Timor; with later movements east to take the few towns along the north and south coasts of Papua. Only later were these gains consolidated with more widespread occupation of inland areas. At the same time Naval Task Forces based in Micronesia undertook lengthy sea voyages to occupy the Bismarks and establish a major military base at Rabaul in New Britain. From here troops were despatched to occupy Lae on the north coast of Papua New Guinea then initiate a campaign to take Port Moresby on the south coast. Other forces moved south down the Solomon Island chain, the aim being to take New Caledonia, then move east to occupy and establish bases in West Polynesia to keep American naval forces at bay. Japanese defeats at Guadalcanal, the Coral Sea and Kokoda disrupted these plans.

Conclusions As with later seaborne dispersals by Austronesian and the Japanese, the first modern human colonisation of ISEA was probably rapid and multi-directional, rather than via unilineal northern or southern routes (cf. Birdsell 1977; O’Connor 2007b). If it involved initial use of coastal enclaves with rich and familiar habitats, such as estuaries, then the punctuated distribution of such areas along sections of inhospitable coastlines between East Africa and ISEA required the capacity to sometimes undertake lengthy sea voyages, as well as the means to transport portable water (Bulbeck 2007). The first modern arrivals were, therefore, technologically, socially and psychologically preadapted for successful colonisation of ISEA. By ~50,000 BP, modern people had reached Australia (e.g., Roberts et al. 1994; Turney et al. 2001a). Subsequent movements around the coasts of Sahul are likely to have been similarly rapid, with the peopling of West Melanesia occurring as part of the initial colonisation impetus, rather than as a later afterthought – a variant of the Coastal Colonisation Model (Bowdler 1990). On the western margin of Sahul, sites documenting initial forays along the coastal margins of wide plains would have been submerged or destroyed by post-Pleistocene sea level rises. However, the steep (and in some cases uplifted) topography of the north Sahul coast and offshore islands offers much better potential for preservation of sites from this period. Rapid 115

M. J. Morwood Taiwan Caroline Islands Philippines

Bangkok Saigon

Palau

Malaya

PACIFIC OCEAN

Singpore

at m Su

Moluccas

Borneo

ra

Bismarks Makassar

Sulawesi

Java

Timor Air raids

INDIA OCEAN

Papua

Rabaul Lae

Guadacanal

Port Moresby Darwin

Kalumburu

Battle of the Coral Sea

Broome

Australia New Caledonia

Figure 9.4.  Japanese troop movements in Southeast Asia and Oceania between December 1941 and May 1942 (Marston 2005; Rottman 2005; illustration by the author).

population increases required for successful colonisation of virgin lands, would also have led to wide-ranging reconnaissance with early occupation of prime inland habitats, such as major river systems (see Birdsell 1977). As well as parallels between dispersals across the region by very different seafaring groups, there are also informative differences. For instance, the first modern humans, mariners with fisher-gatherer-hunter economies, found uninhabited Australia, Papua and other islands in West Melanesia to be attractive colonising prospects. Presumably, the paucity of terrestrial resources on islands beyond the Solomons constrained their further dispersal into Oceania. In contrast, there is no evidence for attempted Austronesian occupation of Australian coastal areas, despite their proximity to ISEA and the fact that Asian seafarers did visit these shores during the initial phase of Austronesian expansion into adjacent East Indonesia, as the appearance of the dingo in Australia ~3500 BP shows (Gollan 1984). Austronesian peoples with fisher-gatherer-hunterfarmer economies seem to have needed access to cultivatable land as well as marine resources to support settlements rather than just visits; and cultivatable land was not, and is still not, a feature of contact areas along the north Australian coast. Similarly, the Japanese did not attempt to occupy northern Australian coastal areas despite undertaking reconnaissance visits to the Kimberley and bombing Darwin, Broome and Kalumburu from newly acquired bases in East Indonesia. Northern Australia lacked resources crucial for sustaining the Japanese Empire in East Asia (e.g., oil, tin, rubber, agricultural surpluses) and was of limited military value for its defence. It is significant that in each of these case studies involving peoples with watercraft capable of transporting founder groups considerable distances, the limits of island colonisation extent do not seem to have been constrained by limitations of maritime technology, but by other factors – by local group subsistence requirements in the case of the first modern humans and Austronesians and by the imperial requirements and ambitions of a distant homeland in the case of the Japanese. 116

Faunal Biogeography in Island Southeast Asia

Acknowledgments Adam Brumm and Gert van den Bergh helped prepare illustrations for the paper and also commented constructively on the text. The Australian Research Council funded much of the associated research.

117

Chapter 10 Late Pleistocene Subsistence Strategies in Island Southeast Asia and Their Implications for Understanding the Development of Modern Human Behaviour

Philip J. Piper and Ryan J. Rabett

Introduction In 1958 the discovery of the Deep Skull in the West Mouth of the Niah Caves, Sarawak, represented, for that time, the earliest radiometrically dated evidence of anatomically modern humans anywhere in the world (Harrisson 1959c; Hunt & Barker, this volume). After almost half a century as the subject of considerable debate in terms of its authenticity, now lately resolved and confirmed (Barker et al. 2007), the Deep Skull and other subsequently discovered early modern human remains (e.g., Demeter et al., 2012; Détroit et al. 2004; Pawlik et al., this volume) attest to the presence of Homo sapiens in Southeast Asia for at least the last 45,000 years (all dates herein are calibrated), and possibly as long as 60–70 ka if the Callao human remains from northern Luzon, Philippines, are confirmed to be those of an anatomically modern human. Less well known is the fact that the hominin fossils at Niah were excavated (meticulously for the time) from within rich cultural deposits of similar antiquity. Examination of the artefacts and biological remains from Niah suggests that human inhabitants of the cave were already eminently capable of exploiting tropical environments. It is unlikely that this evidence came from people who were new arrivals in Island Southeast Asia (ISEA); nor were their cultural practices or subsistence strategies directly drawn or directly comparable to the Palaeolithic record as it appears in western Eurasia – the traditional point of comparison. The data emerging from Southeast Asia during the past decade echoes that which is appearing in Greater Australia (see, e.g., papers by Balme & O’Connor; Cosgrove, Pike-Tay & Roebroeks; Summerhayes & Ford; and Davidson, all in this volume), elsewhere in Asia (e.g., Brantingham et  al. 2004) and in Africa, starting with McBrearty and Brooks (2000), who questioned the evolutionary significance of changes seen in the European Palaeolithic and whether these can legitimately continue to be seen as a model of our species’ modern character. In this chapter we summarise the Late Pleistocene (defined as 126–11.7 ka) environmental history of Southeast Asia and review the current state of knowledge about early tropical forager subsistence here. We consider in what ways the development of hunter-gatherer adaptations were influenced by the structure and composition of the flora and fauna of this region, and the challenges of living in the diverse ecological settings that existed in Southeast 118

Late Pleistocene Subsistence Strategies in Island Southeast Asia Asia. We conclude with the contention that there is need for more regionalised explanations for the emergence of complex human behaviours.

Palaeoenvironment of Late Pleistocene Southeast Asia Rapid warming at the end of the penultimate glacial, during Marine Isotope Stage (MIS) 6, raised global oceans to a level comparable with the present circa 130 ka (Lambeck et al. 2002). This brought insularisation to much of Southeast Asia, a region covering 5 million km2 – an area larger than Western Europe and here defined as comprising those countries bordering or within the South China Sea and Wallacea. This period was also marked by changes to regional climate and vegetation through the onset of warmer, humid conditions and a strengthened monsoon (van der Kaars & Dam 1995; Sun et al. 2003) and, on the basis of evidence from Java, a major faunal turnover. Much of the region’s megafauna became extinct around this time, and an almost entirely extant mammal community suited to closed canopy rainforest, including large numbers of primates, appeared (Bergh et al. 2001; Storm 2001a). Known as the Punung Fauna, this has been dated recently by thermoluminesence to have been established no earlier than 128 ± 15 ka and no later than 118 ± 3 ka (Westaway et al. 2007a). Apparent correspondence between the onset of the inter-glacial and the expansion of rainforest species, though, presents something of a problem, as the spread of arboreal taxa is unlikely to have occurred at the same time as continental flooding was fragmenting the Middle Pleistocene ISEA landscape into island settings (Bergh et al. 2001, 391–392). If the mid-latitude records for the early Holocene are broadly representative, regional warming may have reached inter-glacial levels some thousands of years before the encroaching sea had totally severed land links (Lambeck et al. 2002, 201). Rainforest communities are more likely to have taken advantage of this early window of opportunity – still within the error margin of Westaway and her colleagues’ (2007a) earliest date – to colonize Java than to do so later (Rabett 2012). Alternatively, colonization could have occurred when land bridges were re-established during eustatic drops within the even sub-stages of the climatically variable last inter-glacial (Sémah & Sémah 2012). Accompanying the Punung fauna is the appearance of controversial evidence (in the form of an isolated P3 from the site of Punung III) for an extremely early presence of Homo sapiens in the region (Storm et al. 2005). There is growing support for an early modern human colonisation of Island and Mainland Southeast Asia with discoveries such as the Zirendong fossils from South China (Liu et al. 2010b) dated to 110 –113 ka, the human right third metatarsal found in Callao Cave, northern Luzon, Philippines (Mijares et al. 2010; Pawlik et al., this volume) and the cranium from Tam Pa Ling, Laos, dated to circa 46–51 ka (with a maximum age of 63 ka; Demeter et al. 2012). An increasingly securely dated modern human presence at this time in this region will probably require significant re-appraisal of H. sapiens’s pattern of dispersal from Africa, though all these discoveries are still probably younger than the Punung specimen by some tens of thousands of years, leaving the matter of the earliest colonisation still to be resolved. At the end of the last inter-glacial globally and regionally the climate cooled progressively through a series of pronounced stadial phases (MIS 5d and 5b, 116–110 and 100–87 ka) interceded by major inter-stadials (MIS 5c and 5a, 110–100 and 87–74 ka) (Lambeck et al. 2002). Many of the islands and peninsulas of the South China Sea re-formed into a single sub-continental landmass, Sunda Land (herein ‘Sundaland’) (Molengraaf & Weber 1921), that covered circa 1,850,000 km2 and extended the Asian mainland into the Southern Hemisphere. Kaars and Dam (1995) reported that conditions between circa 126 and 107 ka in West Java remained humid, warmer and somewhat wetter than today. Indeed, such conditions appear to have persisted until circa 81 ka when a change from freshwater swamp forest to swamp vegetation dominated by grasses and sedges at Bandung, West Java (630 m asl.), signalled a shift to a much drier climate, possibly as a result of a reduction in the effectiveness of the northwest monsoon. This does present a degree of mismatch with the regional isotope record that will require further 119

Philip J. Piper and Ryan J. Rabett investigation to resolve, but a similar persistence of tropical forest through late glacial stadial conditions has been reported for southwestern India (Campo 1986) consistent with a southward displacement of the summer monsoon and a strengthened winter monsoon providing some precedence for local variations within this cooling trend. The onset of the last glacial at the end of MIS 5 is recorded in the sedimentological and palynological sequence at Bandung, by the expansion of the lower montane oak forests (Kaars et al. 2000; Newsome & Flenley 1998). The cooling and expansion of montane ecosystems between MIS 4 and MIS 2 indicate that temperatures were as much as 4° C cooler than at present (Hope et al. 2004). As a result, the rainforest formations of the last inter-glacial appear to have contracted, and a savannah-like corridor opened through central Sundaland from the Thai/Malay Peninsula to eastern Java (Bird et al. 2005; Heaney 1991; Wurster et al. 2010). The faunal record from sites within this corridor, such as Punung (ca.128–118 ka; Bergh et al. 2001; Storm et al. 2005) and Song Terus (ca.120–80 ka; Ansyori 2010) that pre-date the circa 81 ka horizon, contains components typical of a high-canopy tropical rainforest. At Song Gupuh (Morwood et al. 2008), a site that post-dates this transition, the faunal community of the period consists primarily of bovines, cervids and suids and is overall more indicative of just such open environments. On the basis of marine core data obtained off the north coast of Borneo (Wang et al. 2009), MIS 3 (59–29 ka) revealed a considerable quantity of tree pollen, especially mangrove forest in the vicinity of one of the major river systems draining Sundaland. Although this might be a biased riparian signal, conditions here changed significantly during early MIS 2 (29–11.7 ka) when herbs and ferns came to dominate, suggesting the catchment area did not serve as a refugium for forest taxa. One of the few pollen sequences from archaeological sites in the region, the record obtained from the West Mouth of the Niah Caves, northwestern Borneo, suggests rapidly fluctuating climate and changes in environment between circa 52 ka and circa 45 ka (Barker et al. 2007; Hunt et al. 2007). At circa 52 ka the local landscape was dominated by Podocarpus conifers with Picea, Pinus and Dacrydium and some oak, indicative of cooler temperatures and a lowering of the submontane forest elements into the lowlands.This is succeeded by a predominance of open-ground taxa with plants such as goosefoots (Chenopodiaceae), Lactuceae and Artemesia linked to major drought stress and perhaps a retreat of forest communities. From circa 49 to 48 ka, reinvigorated forests combined taxa typical of lowland, montane and coastal swamp forests and a period of higher rainfall and probably warmer temperatures. At circa 47 ka there is a sharp increase in mangrove pollen suggesting that sea levels were rising (see Lambeck et al. 2002), and between circa 46 and 45 ka these brought coastal mangrove swamps close to the cave entrance. This latter period also coincided with the earliest comprehensive faunal records at Niah, which include orang-utan (Pongo pygmaeus), leaf monkeys (Presbytis sp(p).), tree squirrels (Sundasciurus) and some arboreal civets (Viverridae); all are indicative of high-canopy, dense forest as are terrestrial forest browsers such as the Sumatran rhinoceros (Dicerorhinus sumatrensis) (Cranbrook & Piper 2007a, 2007b; Cranbrook 2010). During the extreme conditions of the Last Glacial Maximum (LGM) circa 26–19 ka (Clark et al. 2009), rainforest refugia are thought to have existed in several parts of this region, particularly along the western margin of Sumatra, northeastern and southern Borneo (Bird et al. 2005; Hope et al. 2004; Wurster et al. 2010). The ‘savannah corridor’ appears to have covered an area from southern Thailand, through Peninsular Malaysia, western and southern Borneo, eastern and southern Sumatra and Java (Bird et al. 2005; Gathorne-Hardy et al. 2002; Hope et al. 2004). Many of the species that were representative of the Punung fauna, including the orangutan, had now disappeared from parts of Java (Bergh et  al. 2001), implying that rainforests were fragmenting or in decline. The northern part of the Sundaland coastal plain also appears to have been open Artemisia-dominated grassland (Sun & Li. 1999; Sun et al. 2000). Bird et al. (2007) argued that the cooler climate and open environments could have facilitated colonisation by migrating people across large parts of Sundaland and limited the amount of dense forest that would have needed 120

Late Pleistocene Subsistence Strategies in Island Southeast Asia to be traversed by hunter-gatherers. It is during the period starting mid-MIS 3 that the earliest robust evidence of confidently assigned H. sapiens remains begin to appear for the first time in Southeast Asia, most notably in Borneo at the Niah Caves and at Tabon Cave on the Philippine island of Palawan (Barker et al. 2007; Détroit et al. 2004). Regardless of whether the more open central landscapes of Sundaland did indeed help facilitate dispersal, these mobile hunter-gatherers were already well adapted to the full range of tropical habitats that prevailed here by the time the archaeological record picks up their presence.

Foraging Strategies in the Late Pleistocene of Southeast Asia Reconstruction of the local environment surrounding the Niah Caves during the earliest phase of human occupation here circa 45 ka (associated with and just preceding the deposition of the Deep Skull) suggests that it was surrounded by a dense tropical forest, though not necessarily completely analogous to a modern rainforest (Figure  10.1; Barker et  al. 2007; Stimpson 2011, 2012). Foraging within this ecosystem demanded the employment of a range of tactics to obtain the wide variety of the sparsely distributed resources available. Rainforest animals are generally smaller and harder to locate than those taxa that inhabit open habitats and occupy a wide range of ecological niches beyond the terrestrial, including a stratified range of arboreal habitats and diverse aquatic ones (Crowe 2000, 74–75). One of the major challenges facing many modern forest foragers and a point of considerable debate in the late 1980s and early 1990s (e.g., Bailey et al. 1989; Brosius 1991) is the lack of easily obtainable carbohydrate sources. Most stored energy in rainforests exists in the form of inedible woody tissues, and many of those plants that are edible contain toxins and require processing before they can be eaten. Evidence for just such processing has been identified archaeologically at Niah (Zuraina Majid 1982, 112–113; Barton 2005), including sago palm (Eugeissona utilis and/or Caryota mitis) as well as yams (Dioscorea sp.) and aroids that contain an irritant. Seeds of the Pangium edule (Malay: Kepayang), a mangrove-adapted tree, which are also inedible until detoxified (Johns & Kubo 1988), were also recovered from some of the earliest occupation units at Niah.The processing requirement of these resources suggests that foragers were well aware of plant properties and had already perfected the methods needed to make them edible. Such sophisticated knowledge is further indicated by evidence for the use of plant matter to create dyes and pigments, also during this period (Barker 2013; Pyatt et al. 2010). Foragers at Niah also appear to have employed more than one approach to hunting game animals.The forest-adapted bearded pig (Sus barbatus) (Cucchi et al. 2009) dominates the archaeological fauna, with an age profile including juvenile, sub-adult and adult pigs in the Late Pleistocene (Table 10.1). Barker (in press) argues that, rather than pursuit hunting in the thick understorey of the forest at this time, foragers were more likely taking advantage of the fact that social groups of bearded pigs use well-established tracks to move through undergrowth. Linkage between the bearded pig’s dominant position in the faunal assemblage at Niah and the well-known mass migrations of this animal has been posited (Rabett & Barker 2007). Traps set along the trails that these animals habitually maintain would have captured not only pigs but also, at lower frequencies, a wide range other terrestrial game. The low recorded incidence of other large mammals in the Niah fauna, including sambar deer (Rusa unicolor), barking deer (Muntiacus sp(p).), Sumatran rhinoceros (Dicerorhinus sumatrensis) and Malay tapir (Tapirus indicus) is consistent with just such a scenario (Piper & Rabett 2009a; Cranbrook & Piper 2007a). As well as hunting, probably with traps for terrestrial game, the range of arboreal and semi­arboreal prey in the Niah fauna also demonstrates an ability to catch not only primates – macaques, leaf monkeys and orangutan  – but also small and nocturnal carnivores, like the leopard cat (Prionailurus bengalensis), bear cat (Arctictis binturong), masked palm civet (Paguma larvata) and common palm civet (Paradoxurus hermaphroditus). Baited traps could have been utilised to catch these as well as monitor lizards (Varanus sp(p).). Other reptiles include the reticulated python (Python 121

Philip J. Piper and Ryan J. Rabett

Figure  10.1.  Map of Mainland and Island Southeast Asia with the sites discussed in the text. The dates or date ranges presented are those for the earliest-known occupation of each site (illustration by the authors). reticulatus) and hard- and soft-shell turtles (Geoemydidae and Trionychidae) from a range of different freshwater habitats.The latter are generally caught using a baited hook and line, although another modern technique is to wade through the river stabbing a barbed harpoon into the mud, which will penetrate the leathery carapace of the turtle (Pritchard et al. 2009). Raptors and hornbills also consistently appear within the Niah Caves Late Pleistocene assemblages. For example, the bathawk (Macheiramphus alcinus), the brown wood owl (Strix leptogrammica), crested goshawk (Accipiter trivigatus), Brahminy kite (Haliastar indus), changeable hawk eagle (Spizeatus sp.), rhinoceros hornbill (Buceros rhinoceros) and bushy-crested hornbill (Anorrhinus galeritus) have all been recorded in the 45 to 35 ka deposits (Stimpson 2011). Although it is possible that these large attractive birds could be part of the natural death assemblage (see Stimpson 2009), their recurrence throughout the cultural deposits suggests that they were at least sometimes being deliberately targeted. They may not have necessarily made up an important part of the diet but been coveted for their ornate feathers and skulls. Until fairly recently, the bright colourful head and bill and the black and white feathers of hornbills were particularly sought after by local inhabitants in Borneo (e.g., Bennett et al. 1997). The Pleistocene molluscan fauna introduced to Niah likewise came from an equally diverse range of habitats (Barker 2013). Some species occupy the same general niche as the Asian softshelled turtle, buried in the sediment at the bottom of rivers and streams. For example, the large corbiculid bivalve Polymesoda erosa is associated with brackish waters and similarly buries into the soft bottom-sediments, whereas Paludomus everetti and Clea nigricans can be found in clear running water. Alternatively, species within the genus Neritodryas can almost be considered arboreal, congregating on the leaves and stems of riverside vegetation and amongst leaf litter often 122

Table 10.1.  Number of identified specimens (NISP) of each reptile, bird and mammal taxon (excluding the chiroptera and other smaller mammals such as murid rodents) recovered from the earliest, Late Pleistocene deposits recorded at Niah Cave (ca. 45–30 ka) Class

Reptilia

Order

Testudines

Family

Taxon

English vernacular

NISP

%

Trionychidae

Amyda cartilaginea/Pelocheyls cantorii Contains Cyclemys/Heosemys Cuora ambionensis Cyclemys dentata Notochelys platynota Orlitia borneensis Geoemydidae spp. Manouria emys Amyda cartilaginea Dogania subplana Trionychidae sp(p). Varanus sp(p). Python sp(p). Arborophila sp. Lophura ignita Lophura erythropthalma Lophura sp. Macheiramphus alcinus Haliastar indus cf. Accipiter trivirgatus Spizeatus sp. Accipitridae sp(p). Strix leptogrammica Strigidae sp. Anthracoceros sp. Buceros cf. rhinoceros Anorrhinus galeritus Bucerotidae spp. Cissa chinensis

Asian soft-shell turtle/Giant soft-shell turtle Spiny hill/Asian leaf turtle Malayan box turtle Asian leaf turtle Malayan flat-shell turtle Malaysian giant turtle Hard-shell turtle Asian brown tortoise Asian soft-shell turtle Malayan soft-shell turtle Soft-shell turtle Monitor lizard Reticulated or Borneo blood python Partridge Crested fireback Crestless fireback Gallopheasants Bathawk Brahminy kite Crested goshawk Changeable hawk-eagle Hawk Brown wood owl Owl Pied/black hornbills Rhinoceros hornbill Bushy-crested hornbill Hornbills Green magpie

3

0.12

Geoemydidae

Testudinidae

Squamata Aves

123

Galliformes

Varanidae Pythonidae Phasianidae

Falconiformes

Accipitridae

Strigiformes

Strigidae

Coraciiformes

Bucerotidae

Passeriformes

Corvidae

5 1 5 13 1 383 3 4 12 60 269 10 2 1 1 1 4 1 3 1 9 1 1 1 1 1 4 1

0.20 0.04 0.20 0.52 0.04 15.28 0.12 0.16 0.48 2.39 10.73 0.40 0.08 0.04 0.04 0.04 0.16 0.04 0.12 0.04 0.36 0.04 0.04 0.04 0.04 0.04 0.16 0.04 (Continued)

Table 10.1. (Continued) Class

Order

Family

124 Mammalia

Primates

Estrildidae Cercopithecidae

Pholidota

Ponginae Manidae

Rodentia Carnivora

Sciuridae Hystricidae Ursidae Mustelidae Viverridae

Felidae Perissodactyla Artiodactyla

Rhinocerotidae Tapiridae Suidae Tragulidae Cervidae Bovidae

Taxon

English vernacular

NISP

%

Cissa sp(p). Corvidae sp. Lonchura c.f. fuscans Presbytis (presbytis)sp(p). Presbytis sp(p). Macaca fascicularis Macaca nemestrina Macaca spp. Cercopithecidae spp. Pongo pygmaeus Manis javanica Manis cf. palaeojavanica Sciuridae spp. Hystricidae sp(p). Helarctos malayanus Mustelidae spp. Viverra tangalunga Paradoxurus hermaphroditus Paguma larvata Arctictis binturong Viverridae spp. Prionailurus bengalensis Felis/Prionailurus sp(p). Dicerorhinus sumatrensis Tapirus indicus Sus cf. barbatus Tragulus napu Tragulus sp(p). Muntiacus sp(p). Cervus unicolor Bos cf. javanicus

Green magpies Crow Dusky munia Leaf monkey (not silvered langur) Leaf monkey Long-tailed macaque Pig-tailed macaque Macaque sp. Monkey Orang-utan Pangolin Giant pangolin Squirrel Porcupine Sun bear Mustelid Malay civet Common palm civet Masked palm civet Bear cat/Binturong Civet cat Leopard cat Cat Sumatran rhinoceros Malay tapir Bearded pig Greater mouse deer Mouse deer Barking deer Sambar deer Banteng Total

2 1 3 3 32 11 2 6 188 112 38 9 4 39 2 2 1 3 9 11 10 2 8 1 2 1,144 6 9 19 14 11 2,506

0.08 0.04 0.12 0.12 1.28 0.44 0.08 0.24 7.50 4.47 1.52 0.36 0.16 1.56 0.08 0.08 0.04 0.12 0.36 0.44 0.40 0.08 0.32 0.04 0.08 45.65 0.24 0.36 0.76 0.56 0.44 100

Note: The table includes all identified vertebrate remains from the deposits in Hell, EA/UR 1 & 2, and the basal deposits of Area A; see Barker 2013. The identified bird fauna is reproduced with the kind permission of Christopher Stimpson.

Late Pleistocene Subsistence Strategies in Island Southeast Asia Table  10.2.  Number of identified specimens (NISP) of each reptile and mammal taxon (excluding the chiroptera and other smaller mammals such as murid rodents) recovered from the earliest deposits recorded at Ille Cave (ca. 14–13 ka) Class

Order

Family

Taxon

English vernacular

Terminal Pleistocene NISP %

Reptilia

Testudines

Squamata

Mammalia Primates Rodentia

Carnivora

Geoemydidae

Cuora amboinensis Cyclemys dentata

Varanidae

Geoemydidae sp(p.) Varanus sp(p.)

Serpentes

Serpentes

Cercopithecidae Macaca fascicularis (subsp.) Sciuridae Sundasciurus sp(p.) Hystricidae Hystrix pumila Canidae Mustelidae

Felidae Artiodactyla Suidae Cervidae

Malayan box turtle Asian leaf turtle Hard-shell turtle Monitor lizard Unidentified snake Long-tailed macaque Tree squirrel

Palawan porcupine Unident. Canid* Dog Herpestes Short-tailed brachyurus mongoose Panthera tigris Tiger Sus ahoenobarbus Palawan bearded pig Axis calamianensis Calamian deer Cf. Rusa sp(p.) Unidentified deer Cervidae Deer Total

1

0.25

7

1.72

58

14.29

7

1.72

6

1.48

13

3.20

5

1.23

1

0.25

2 1

0.49 0.25

2 15

0.49 3.69

1

0.25

3

0.74

284 406

69.95 100

Source: After Ochoa 2009.

some distance from water. The broad diversity of ecological niches from which mollusca were sought suggests the human foragers had an intimate knowledge of where each species could be found within its respective aquatic and terrestrial environment and how to extract it (Szabó & Amesbury 2012). At Niah, the evidence suggests foragers were exploiting a wide variety of forest resources and possessed the technology and artifice to do this by at least 45 ka and probably earlier. Many other sites in the region have been shown to contain a similar ‘broad spectrum’ (after Gorman 1971) of resource exploitation. What the work at Niah and other recent excavations are showing, though, is that within the wide species representation is a quite targeted emphasis on particular animals. Sus barbatus was consistently the most dominant single species being taken throughout the Pleistocene at Niah and probably formed the bulk of the protein intake. Monkeys (Cercopithecidae) and other arboreal fauna formed a long-standing and over time increasingly important component 125

Philip J. Piper and Ryan J. Rabett of that diet; their rising significance commensurate with changes in availability and innovations in technology (Piper & Rabett 2009a; Rabett & Piper 2012). At other sites we see a similar pattern, though the target resources vary. The earliest modern human remains from Tabon Cave, near the southwest coast of Palawan island, also date to circa 45–50 ka, demonstrating that humans had also reached the Philippine archipelago by this time (Détroit et al. 2004; Fox 1970), and perhaps much earlier if the fossil human remains from Callao Cave can eventually be demonstrated to be unequivocally those of anatomically modern humans (Mijares et al. 2010; Pawlik et al., this volume). Very little animal bone was recovered from Tabon with the exception of a few pieces of deer and pig (Fox 1970), leaving currently the best zooarchaeological record on the island from the end of the Pleistocene at Ille Cave in the north of the island (Ochoa 2009; Piper et al. 2008, 2011). The foragers of Ille were also using a range of techniques that probably included trapping to capture a diversity of mammal taxa, including the long-tailed macaque (Macaca fascicularis philippinensis), porcupine (Hystrix pumila), short-tailed mongoose (Herpestes brachyurus) and some tree squirrels (Sundasciurus spp.), but the hunted faunal community was dominated by two locally extinct species of deer and the Palawan bearded pig (Sus ahoenobarbus) (Table 10.2). Although detailed zooarchaeological studies are still to be applied to the many other sites with Late Pleistocene components in ISEA, further hints of resource targeting can be drawn from the literature. These are predominantly of Early to Mid-Holocene age, and the patterns portrayed will have been affected by changes in the availability of local resources in response to climate and environmental change, though there is reason to suppose that this strategy, which extends even to the present, is of considerable antiquity and related to the fundamental structure of tropical environments. Among the more anecdotal reports of forager prey-preference, Erdbrink (1954, 298) recalls how the fauna from Gua Lawa, one of the earliest excavated cave sites in east central Java (van Es 1930) and now dated 9–3 ka (Simanjunak & Asikin 2004), contained “an astonishing number of bones of the common monitor lizard (Varanus salvator)”. He further noted that comparable cultural deposits at another un-named cave in the nearby Gunung Tjantelan were dominated by bones of the crab-eating (or long-tailed) macaque (Macaca fascicularis [= irus]). More recent and detailed work in the Gunung Sewa region of southeast Java has found that faunal assemblages there show major shifts in dominant taxa. This is especially evident across the Pleistocene-Holocene transition, where emphasis on suids, cervids and bovines (though not to the exclusion of other taxa) is replaced by one focussed on macaques (Macaca sp.) and, in some cases, the first systematic exploitation of molluscs related to the encroachment of the seas to within foraging range of the caves (see, e.g., Morwood et al. 2008; Sémah & Sémah 2012; Simanjunak & Asikin 2004 with relation to the caves of Song Terus, Song Keplek, Braholo and Song Gupuh). Evidence from sites in Mainland Southeast Asia, where detailed zooarchaeological analysis is also moving beyond taxonomic identification to assess and quantify abundance, echoes these findings. For example, at Lang Rongrien, in southern Thailand, the Late Pleistocene component of the occupation (42–32 ka) shows that, while various cervids and bovines were identified, a specific emphasis was placed on turtles and tortoises; Mudar & Anderson (2007) argue that chelonids might have provided a small but regular protein resource to supplement the erratic success of pursuit hunting. Geoemydidae turtles in particular are also common in the zooarchaeological record of Niah and Ille Cave, suggesting that this could have been a common strategy. The relatively impoverished faunal community recorded at Lang Rongrien contrasts markedly with the species representation recorded at Moh Khiew, a cave site in the same limestone karstic formation just 13 km from Lang Rongrien. Occupied from circa 26 ka, just slightly later than Lang Rongrien, the human inhabitants of Moh Khiew hunted a variety of terrestrial and arboreal taxa, including leaf monkeys (Presbytis sp(p.)) and Prevost’s squirrel (Callosciurus prevostii) (Pookajorn 1996). Whether variations in the composition of the faunal assemblages reflect differences in the prevailing environment between the two sites, or whether the caves were utilized for different purposes, it does 126

Late Pleistocene Subsistence Strategies in Island Southeast Asia

Figure  10.2. The cave mouth at Hang Boi, Vietnam, showing the archaeological compound (photograph by R. Rabett). demonstrate that human subsistence strategies varied even within the same geographic location (Anderson 2005). In northwestern Thailand, the recently excavated Tham Lod rock shelter in Mae Hong Son Province has yielded thermoluminesence and 14C dates indicating occupation spanned 35 to 13 ka (Shoocongdej 2006). Throughout the archaeological sequence, the inhabitants of Tham Lod appear to have focussed their hunting on large ungulates and, in particular, deer, bovines, pigs and mountain goats, with lesser numbers of smaller mammals and reptiles. Situated 200 m from the modern Lang River, Tham Lod would have provided an ideal spot to observe animal movements and perhaps participate in ambush hunting. In the same district as Tham Lod is Spirit Cave (Gorman 1970; 1971), whose earliest occupation at circa 13 ka overlaps with the final stages of habitation at the aforementioned site. No values are available to estimate the structure of the vertebrate assemblages recovered from Spirit Cave, but the composition is very similar to that recovered from Tham Lod (Gorman 1970, 95), supplemented by the collection of a range of terrestrial and freshwater mollusca. This may suggest a local continuation of similar hunting strategies to those at Tham Lod, but a thorough analysis would be needed to demonstrate this. In Vietnam, ongoing work at two cave sites in a sub-coastal limestone massif and location of Tràng An park, near the Song Hong River delta, has revealed occupations spanning the period from the LGM to the Early Holocene (Rabett 2012; Rabett et al. 2009, 2011).The archaeology of the older of these sites, Hang Trống, is in its early stages of post-excavation analysis. That of the younger site, Hang Boi (Figure 10.2), spans the Pleistocene-Holocene transition (ca. 13.6–10.6 ka); like other sites from this period in Vietnam (e.g., Hoáng Xuân Chinh, 1991; Yi et al. 2008), it has revealed a long-term emphasis on the collection and consumption of mollusca, in this case land snails (with most midden remains being those of Cyclophorus theodori and C. unicus) and freshwater crab (including Villopotamon sp.) (Figure 10.3; Table 10.3). Although a range of 127

Philip J. Piper and Ryan J. Rabett

Figure  10.3.  A section of the west-facing profile of the Hang Boi excavation, illustrating the stratified land-snail midden (photograph by R. Rabett). arboreal and terrestrial vertebrate fauna is also recorded from the site, their frequency suggests that they have been more opportunistically taken, probably to supplement more reliable staples (Table 10.4). A notable feature of the Hang Boi record is that despite the encroachment of coastal habitats towards the massif during the post-glacial period, there is no indication they were being exploited at this site; though evidence of coastal contact has been recovered in the form of a 128

Late Pleistocene Subsistence Strategies in Island Southeast Asia Table  10.3.  Minimum number of individuals (MNI) and number of identified specimens (NISP) of invertebrates recorded within a sample of Early Holocene deposits from Hang Boi, northern Vietnam Order/class

Family

Taxon

MNI

% MNI

NISP

Gastropoda

Thiaridae

Brotia sp. Melanoides tuberculata Cyclophorus theodori Cyclophorus unicus Cyclophorus spp. Scabrina cf. denudata Viviparus costatus Pupina sp. Cypraea sp. Odontartemon costulatus Camaena sp. Corbicula sp. Cristaria herculea Unionidae spp. Villopotamon sp. Potamidae sp(p). Total

5 1

0.22 0.04

-

1527

68.20

-

512

22.87

-

160 1

7.15 0.04

-

5 1 1 1

0.22 0.04 0.04 0.04

-

16 4 1 4 2,239

0.71 0.18 0.04 0.18 100

-

Cyclophoridae

Vivaparidae Pupinidae Cypraeidae Streptaxidae Mollusca: Bivalvia

Camaenidae Corbiculidae Unionidae

Decapoda

Potamidae

401 696 1,097

Source: After Rabett et al. 2011.

cowrie and a few neritid shells. Several scenarios can be entertained to account for this situation (see Rabett et al. 2011), but the salutary point is that proximity to the coast does not necessarily seem to lead to the immediate, preferential adoption of marine resources; indeed at this site, preexisting inland strategies were maintained and intensified. A similar situation has been observed at Gua Balambagan, an island site off the northeast coast of Borneo. Despite lying close to the coast throughout the period of its occupation, starting at 19.6 ka, foragers came to incorporate coastal resources into their economies only during the Early Holocene, prior to which large terrestrial fauna appear to have predominated (Zuraina Majid et al. 1998). The relationship that forager communities had with the exploitation of maritime habitats in this region was clearly a complex one: while there is apparent delay in adoption in some areas, in others there is strong indication that Late Pleistocene communities were well-adapted to exploit both near-shore and offshore marine fauna. Many of the Wallacean Islands have undergone substantial tectonic uplift, preserving ancient coasts. Alternatively, some of these islands have deep offshore marine trenches, and consequently current shorelines have great antiquity and potential to preserve Pleistocene occupations. In fact, Erlandson (2001) has argued that the one common feature of Pleistocene sites demonstrating early evidence of coastal resource exploitation is a steep offshore bathymetry that has preserved the ancient coastlines from post-Pleistocene marine inundation. For example, the north coast of Timor drops steeply to the continental shelf, and the Late Pleistocene cave sites of Jerimalai (ca. 42 ka) and Lene Hara (35–30 ka) have never been more than 1 or 2 km from the coast (O’Connor 2007b; O’Connor et al. 2002). Both have produced evidence for the collection of marine molluscs such as Strombus, Trochus and Lambis from rocky platform environments. The deposits at Jerimalai also produced evidence for the procurement of 129

130

Table 10.4.  Number of Identified Specimens (NISP) recorded for the Late Pleistocene, Pleistocene–Holocene boundary and Early Holocene from Hang Boi, northern Vietnam, during the 2008 field season Hang Boi Class

Pisces Chondrichthyes Reptilia

Order

– Rajiformes Testudines

Family

– Dasyatidae Geoemydidae

Taxon

– Dasyatidae spp. Cuora (= Pyxidea) mouhotii Cuora cf. trifasciata Cyclemys oldhami

Aves Mammalia

Squamata Galliformes – Scandentia

– – Phasianidae – Tupaiidae

Soricomorpha Primates

Soricidae Cercopithecidae

Chelonia spp. – Phasianidae sp(p). – Tupaia glis (= belangeri) Soricidae sp. Macaca arctoides Macaca cf. assamensis Macaca nemestrina Macaca cf. mulatta Macaca sp. Cercopithecidae sp(p).

Common Name

Late Pleistocene

Pleistocene– Holocene

Early Holocene

NISP

%

NISP %

NISP

%

Fish Sting ray Keeled box turtle

10 0 0

32.26 0 0

5 0 0

17.24 0 0

60 * 6

22.39 0 2.24

Chinese three–striped box turtle Dark–throated leaf turtle Turtle species Snake Wild fowl/Pheasant? Bird Common tree shrew

0

0

0

0

*

0

0

0

1

3.45

1

0.37

2 0 0 0 0

6.45 0 0 0 0

1 6 0 1 0

3.45 20.69 0 3.45 0

71 61 1 9 1

26.49 22.76 0.37 3.36 0.37

Shrew Stump–tail macaque Assamese macaque

1 0 0

3.23 0 0

0 0 0

0 0 0

0 * *

0 0 0

Pig–tailed macaque Rhesus macaque Macaque sp. Monkey

0 0 1 11

0 0 3.23 35.48

0 0 0 5

0 0 0 17.24

* 2 1 18

0 0.75 0.37 6.72

Rodentia

Sciuridae

Carnivora

Muridae Hystricidae Mustelidae Viverridae

Felidae Artiodactyla

Suidae Cervidae Bovidae

Ratufa (= Sciurus) bicolor Sciuridae sp(p). Rattus sp. Hystricidae sp(p). Arctonyx collaris Mustelidae sp(p). Paradoxurus hermaphroditus Herpestes sp. Viverridae sp(p). Panthera cf. pardus Felis sp(p). Sus sp. Cervidae spp. Bos sp.

Black giant squirrel

0

0

1

1.12

*

0

Squirrel Rat Porcupine Hog badger Mustelid sp. Common palm civet

0 0 0 0 1 0

0 0 0 0 3.23 0

1 0 0 0 0 0

3.45 0 0 0 0 0

6 2 1 * 1 *

2.24 0.75 0.37 0 0.37 0

Mongoose Civet Leopard Felid sp. Pig Deer species Cattle Total

0 1 2 1 1 0 0 31

0 3.23 6.45 3.23 3.23 0 0 100

1 2 0 0 2 3 0 29

3.45 6.90 0 0 6.90 10.34 0 98

0 2 0 1 5 19 * 268

0 0.75 0 0.37 1.87 7.09 0 100

* Taxa recorded in 2007 but not re-identified in 2008 and hence not quantified in this table.

131

Philip J. Piper and Ryan J. Rabett marine turtles and fishing for a variety of inshore reef fishes, and potentially some species found further offshore (O’Connor et al. 2011a). The extraction of a diversity of coastal resources, possibly including tuna, would have required a certain amount of forward planning and investment in the production of watercraft and fishing equipment (O’Connor 2007b). In the northeast of the region, on Talaud Island between the southern Philippines and northern Sulawesi, the earliest human colonization comes from Leang Sarru (ca. 35–32 ka). Here a similar suite of marine mollusca has been found to that on Timor, again coming from inter-tidal and sub-tidal environments (Ono et al. 2009). Apart from being an important record of human subsistence strategies on what are often fauna-impoverished islands, such occupations can provide insights into what has probably been lost in terms of human exploitation of coastlines in ISEA as a result of flooding of the low-lying coastal plains by the rising seas.

Discussion The palaeoenvironmental record for Late Pleistocene Southeast Asia indicates that, rather than comprising a single tropical vegetation type, it was composed of a mosaic of different ecosystems, from humid tropical rainforest to open woodland and savannah. Many of the coastlines would also have been periodically fringed by mangrove swamps (see, e.g., Woodroffe 1990). The major and sometimes rapid fluctuations in climate during the Late Pleistocene would have meant local and regional environments were extremely dynamic, with expansions and contractions of tropical rainforest, sub-montane and montane forests, savannah and coastal ecotones. The first arrival of Homo sapiens in Island Southeast Asia and their possible relationship with H. erectus in the region remains enigmatic (see Dennell, chapter 4, this volume). One of the crucial debates revolves around the age of the Ngandong H. erectus fossils. Swisher et al. (1996) used the results of chronometric dating from Ngandong, and more recently Yokoyama et al. (2008) applied direct gamma ray spectrometry dating of hominin fossils to argue for the survival of H. erectus into the Late Pleistocene. In contrast, Vos et al. (2007) and Bergh et al. (2001) suggested that, on the basis of their associations with a suite of extinct Middle Pleistocene fauna, the Ngandong H. erectus fossils could not be any younger than the MIS 5 environmental transition. This interpretation has been supported by the re-dating of the Punung stratigraphy, which has produced varying dates between 128 ± 15 and 118 ± 3 ka (Westaway et al. 2007a). Furthermore, on the basis of a study of the H. erectus fossils and the associated vertebrate remains, Westaway (2002) and Westaway and Groves (2009) have argued that their deposition at Ngandong has followed a very different taphonomic pathway that has included some degree of water transport. They suggest that the H. erectus fossils are likely to be older than the associated vertebrate remains and are of Middle Pleistocene age. There is growing evidence for the presence of hominins in Island Southeast Asia from MIS 5 onwards, and if the Punung III premolar is accepted as anatomically modern as claimed, then it possibly represents a replacement of the already extinct H. erectus by our own species at the end of the Middle Pleistocene. However, the site of Song Terus also has evidence of a human presence from 120 ka onwards. Sémah and Sémah (2012) and Sémah et al. (2004) argue that there is a marked change in the mode of cave occupation after 70 ka, but they cautiously do not interpret whether this represents a change in the species of hominin occupying the cave. Both the sites of Song Gupuh in Java (Morwood et al. 2008) and Kota Tampan (Zuraina Majid 2003) also have archaeological records dating to MIS 4, but neither of these sites has produced early hominin remains, and it is difficult to determine which hominin was responsible for the accumulation of cultural materials. If the Callao specimen is confirmed as modern human, then this will certainly support the presence of our species in Island Southeast Asia by MIS 4. Whether groups of H. sapiens entered an unpopulated landscape at that time, or later, is hard to say. The fossil record remains equivocal. The emerging picture of the genetic ancestry of people living in this region today, however, may be pointing to the early existence of another, presently elusive hominin population 132

Late Pleistocene Subsistence Strategies in Island Southeast Asia besides that of H. floresiensis.These are currently referred to as the ‘Denisovans’ after the cave site of Denisova, Siberia, where their distinctive genetic signature was first identified (Krause et al. 2010), but are now believed to have included Southeast Asia within their territory (Reich et al. 2011). At present, though, the trail for anatomically modern humans is picked up strongly in Southeast Asia only during the first half of MIS 3, from about 50 ka onwards, first at Niah, Tabon and Lang Rongrien and then more widely thereafter. In all likelihood these sites do not represent the first settlement of this region, as the hunting and gathering communities utilising these sites clearly already had a developed knowledge of the environments in Mainland and ISEA and how to live successfully within them. From almost the earliest period of occupation at Niah, botanical evidence indicates that a range of plants was being collected, including several that required labour-intensive detoxification to make them edible.The discovery of Late Pleistocene pit-like features at the site have been tentatively implicated in the techniques for achieving this result (Rabett & Barker 2007); descriptions in the historical literature talk of similar procedures (Skeat 1902, 128). Throughout human habitation at Niah, the forest-adapted bearded pig was the most commonly identified species, accounting for more than 90% of the ungulate remains recovered (Barker 2013). Even taking into consideration, and re-calculating for potential variability in the number of identifiable skeletal elements in the skeleton of different taxa (following Lyman 2008, 30), pig bones outnumber those of the next most common taxa in the archaeological record, the monkeys, by more than 2:1 (Piper et al., in prep.). It is unlikely that encounter hunting would have been particularly effective in the thick understorey of the forest, and it is considered more likely that the foragers set traps along the trails these pigs maintain through the forest undergrowth (Piper & Rabett 2009a). Such a system would have been effective at inadvertently catching a wide range of other ungulates that are known to take advantage of pig trails. Nocturnal ­mammals and reptiles could also have been effectively captured using a variety of traps, some of them perhaps baited. The presence of monkeys and orangutan also from the start of the Niah sequence, albeit in comparatively low numbers, is an indication that foragers were also technologically able to nullify the arboreal advantage that such primates would hold (Rabett & Piper 2012). In addition, the foragers at Niah were very aware of the ecological niches occupied by a range of reptiles, birds and invertebrates and had methods to effectively capture them. Overall the archaeological record from here suggests a human population with the knowledge and skills required to exist successfully within tropical environments. Furthermore, the processing of toxic plants and the preparation of traps is time-consuming, and these activities perhaps hint that at some sites, such as Niah, periodic occupations could have lasted more than just a few days. In other parts of Mainland and ISEA, such as Java, Peninsula Malaysia, northern Thailand and Palawan Island, where open woodland and grassland appear to have predominated during much of the Late Pleistocene, hunting was focussed more on large and intermediate-sized ungulate populations.This focus would have required different strategies to the dense tropical forests.These activities were supplemented by the capture of a range of smaller mammals and reptiles possibly with the aid of traps and snares that would have offset shortage resulting from unsuccessful large game hunts. Not all sites contain the same suite of fauna nor do they have equivalent diversities in hunted vertebrates; rather there appears to be variability between sites that suggests various strategies were being utilised dependent on local resource availability. This in turn was probably dependent on numerous ecological factors and the timing and duration of site occupation. The islands of Wallacea presented their own particular challenges to the forager populations that successfully reached and inhabited them. Some, such as the Philippine archipelago and Sulawesi, contained populations of endemic cervids and suids that were hunted along with smaller mammals and reptiles (Clason 1976; Piper & Mijares 2007; Pawlik et al., this volume). Other islands, such as Timor and the Talaud Islands, contained impoverished vertebrate faunas and no large game, and here there is evidence for the extensive use of coastal resources in the form of the collection of marine molluscs and fishing (O’Connor 2007b; O’Connor et al. 2002; Ono et al. 2009). 133

Philip J. Piper and Ryan J. Rabett The stable coastlines of Wallacea and evidence of coastal foraging also hint at the other resource procurement strategies that might have existed along the fringes of Sundaland during periods of stable sea level that have long been lost as a result of marine inundation (Anderson 2005; Ono et al. 2009). The form that forager subsistence strategies took in Southeast Asia was tailored to the structure, affordances and limitations of its many complex environments. Within these, human populations needed to adopt a range of different foraging strategies to most efficiently utilise the diversity of resources available. In more northerly latitudes, where Late Pleistocene subsistence choices were focussed on successfully procuring a narrow range of plants and often unpredictable but abundant animal resources – as the Chipewyan saying goes, “No one knows the ways of the wind and the caribou” (Urquhart 1989, 73), risk mitigation (and with it cultural practice) was probably geared to achieve this (see, e.g., Finlayson & Carrión 2009). In tropical Southeast Asia, where the environment accentuates variety over quantity, the needs were (and are) different. Economies by necessity had to be flexible, foragers needed to situate themselves in the right part of the landscape at the right time (seasonally or super-annually) to take best advantage of localised peaks in resource availability. People were not reliant on the exploitation of a single ubiquitous resource but on many different ones, and we can hypothesise that site and mobility strategies were at least in part geared to accommodate this. An ability to switch from one strategy to another according to the needs of the time continues to exemplify small-scale Southeast Asian communities (e.g., Endicott & Bellwood 1991; Oota et al. 2005) and provides its own guard against risk. There is a sense that during the Late Pleistocene a similar system existed: reliable dietary staples were often supplemented by more opportunistic hunting and foraging activities, and shifts in strategy accommodated changes in circumstance (Rabett 2012; Rabett & Barker 2010). Existence within the parameters of Southeast Asia’s tropical and changeable settings required innovation in technologies and the development of knowledge suites best suited to them. A hunter from the open landscapes of Europe would not have lasted long here (or vice versa) for the simple reason that the acquired knowledge such hunters possessed was geared, perhaps even specialised, to their own situation. Traditional approaches to the evolution of modern human behaviour have tended to see its emergence in terms of an early attainment of traits in one region followed by dissemination globally, subject only to cultural drift. We would argue that the evidence from Southeast Asia, like that appearing in Africa and elsewhere in Asia, calls for a different perspective to be taken: one that sees instead an evolutionary relationship between human dispersal into new environments and the selective pressures encountered as populations settled within them, with this in turn leading to regional trajectories of behavioural development. Rather than a faculty attained once in a particular place or at a particular time, we would contend that the behavioural evolution in H. sapiens is better portrayed as a process: one that is ongoing, responsive, varied and locally contingent.

Acknowledgments The authors wish to take this opportunity to thank Robin Dennell and Martin Porr for their offer to contribute to this important volume. We are also very grateful to the many colleagues we have worked with at Niah, Ille, Callao and Tràng An; the results of these and other local and international projects in Southeast Asia are considerably advancing our understanding of early human settlement of this unique and complex region. Philip J. Piper was partly supported by a Chancellor’s Grant from the University of the Philippines (Diliman), administered by the Office of the Vice Chancellor for Research and Development, and during the final stages of manuscript production by the ARC Future Fellowship Grant FT100100527. Ryan Rabett wishes to thank the McDonald Institute, the Evans Fund and the ASEASUK Research Committee on South East Asian Studies, the Templeton Foundation and Nguyen Van Truong for generous funding of project work. 134

Chapter 11 Modern Humans in the Philippines Colonization, Subsistence and New Insights into Behavioural Complexity

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares

Introduction The Philippines consist of 7,107 islands located at the northern limits of Wallacea and the northeastern fringes of the islands of Southeast Asia at latitude 13oN and longitude 122oE. It is separated from Borneo to the southwest by the Sulu Sea, from Mainland Southeast Asia to the northwest by the South China Sea, from Taiwan to the north by the Luzon Strait and from Sulawesi to the south by the Celebes Sea, and it is bounded to the east by the Philippine Sea. The Philippine archipelago straddles two distinct biogeographic zones, with Palawan located on the northeastern edge of the Sunda Shelf, and hence shares a fauna and flora with many of its closest relatives within Island Southeast Asia (Madulid 1998). A posited land bridge either in the Upper Pleistocene (Fox 1970; Cranbrook 2000) or more likely in the Middle Pleistocene (Heaney 1985; Pawlik & Ronquillo 2003) possibly facilitated the colonization of the island by Sundaic Island species, including perhaps hominins. The main archipelago islands of Luzon, the Visayas and Mindanao, situated in Wallacea, on the other hand, have never been physically linked to the Sundaic region, and a sea crossing has always been needed to reach them (Heaney 1993; Oliver & Heaney 1996; Esseltyn et al. 2010). Most of the oldest palaeontological sites in the Philippines have been identified on the island of Luzon (Figure  11.1). It possesses an impoverished island faunal community dominated by good successful open-sea migrants that once reached Luzon and diversified to produce the high endemism characteristic of remote archipelagos in Wallacea (Heaney 1986, 1993, 2002; Jansa et al. 2006; Morwood & van Oosterzee 2007; Oliver & Heaney 1996). Surveys of northern Luzon have produced an archaic vertebrate fauna containing giant tortoise (Geochelone), proboscideans (Stegodon and Elephas), bovines, cervids, suids and a rhinoceros (Rhinoceros luzonensis), and isolated fossil finds that could date to more than 500,000 years ago have been found in the Mindanao and Visayan areas of the Philippines (Koenigswald 1958; Fox 1978; Bautista & Vos 2001; Morwood & van Oosterzee 2007). The Philippines are also home to some of the earliest records of anatomically modern humans in the Island Southeast Asian (ISEA) and Australasian regions. Among them are the remains of 135

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares

Cabalwan Callao Cave

Pacific Ocean

Arubo

Manila

West Philippine Sea

PHILIPPINES

Pa l

Tabon Cave

aw an

Ille Cave

Sulu Sea

Sulawesi Sea

BORNEO

N

0

100

200 Km

Figure  11.1.  Map of the Philippine Islands with archaeological and palaeontological sites mentioned in the text (illustration by the authors). several individuals, all assigned to Homo sapiens, recovered at Tabon Cave, Palawan Island, during excavations in the 1960s by Robert Fox (1970; 1978), and more recently in a re-investigation of the cave by the National Museum of the Philippines and the Muséum national d’histoire naturelle, Institut de Paléontologie Humaine. These remains have been variously dated between 16 and 47 ka. Recent excavations at Callao Cave in the Peñablanca karst limestone region of northern Luzon have produced the third metatarsal of an enigmatic hominin, which has been provisionally ascribed to an anatomically modern human and directly dated using U-series ablation to 67 ± 1 ka (Mijares et al. 2010). If this specimen is eventually ascribed to Homo sapiens, it will provide further support for the posited early migration from Africa and colonization of Island Southeast Asia by modern humans suggested by sites like Punung in Java (Storm et al. 2005) and Lenggong in Peninsula Malaysia (Zuraina Majid 1994). If the Callao hominin should turn out to be that of a different hominin species, then it will indicate that non-sapient hominins were more widely distributed throughout Wallacea than just on the island of Flores. What has been made absolutely 136

Modern Humans in the Philippines clear through the evidence of cut marks identified on the animal bones recovered in association with the Callao specimen is that this was a tool-using hominin (Piper & Mijares 2007). In ISEA, lithic assemblages are characterized by simple, mostly unretouched flakes and cores without preparation, which has led to the interpretation that these implements were primarily made and used as part of an expedient technology (Mijares 2002; 2008) that showed no characteristics commonly designated elsewhere (e.g., Europe and Africa) as a part of a “package of modern human behavioural traits” (Mellars 1989a; Bar-Yosef 2002; Haidle & Pawlik 2010; Pawlik 2010). Recent studies of use wear on a lithic assemblage from Ille Cave in northern Palawan, however, have demonstrated that not all such seemingly simple lithic artefacts were used expediently and that evidence of residues on stone tool surfaces could indicate that some may have even been hafted (Pawlik 2010). In this chapter, we discuss how new discoveries of a hominin in northern Luzon, palaeoenvironmental studies and lithic use-wear analysis associated with the Philippine Palaeolithic are contributing to debates on the timing of the migration into, and colonization of, Island Southeast Asia; revise our understanding of technological functions of stone artefacts; and provide evidence for more advanced human behaviour in the preparation of complex composite tools in the Upper Pleistocene.

Early Human Colonization of the Philippines The presence of hominins in the Middle Pleistocene of the Philippines has been postulated for the Cagayan Valley in northern Luzon since the 1950s. Surface collections of artefacts and faunal remains have included extinct mammals supposedly associated with stone tools from an industry known as the Cabalwanian (Koenigswald 1958). A National Museum survey of the Cagayan Valley headed by Robert Fox claimed to have found at least 64 sites that contained mostly unretouched flakes, choppers and other unifacially retouched pebble tools and fossils of extinct mammals (Fox 1971, 1978, 1979; Fox & Peralta 1974). Although all the lithic artefacts were found on or close to the surface of eroded terraces and not necessarily in direct association with the animal remains, on the basis of the inferred chronological associations between the stone tools with their ‘archaic’ morphology and the extinct vertebrate faunas, Fox (1978) suggested that the lithic assemblages were produced by a precursor to modern humans who had reached Luzon during the Middle Pleistocene some 250,000 to 300,000 years ago. However, as yet no clear association between artefacts, tool-producing hominins and the extinct fauna has been demonstrated, and no geochronological sequence for the isolated fossil finds exists (Bautista & Vos 2001). However, Pawlik (2004) recently claimed the existence of a possible Early Palaeolithic lithic assemblage at the site of Arubo, in the province of Nueva Ecija in Central Luzon. The artefacts included several cores and flakes, a straight-edged chopper and two bifacial artefacts, a large cleaver fragment and a ‘proto-handaxe’ (Pawlik & Ronquillo 2003), which was then in following publications labelled as a (developed) handaxe (Pawlik 2009a; Dizon & Pawlik 2010) after its ‘Acheulian’ character was confirmed by the curator of the type locality’s assemblage of Saint Acheul, Claire Gaillard (personal communication 2006). Unfortunately, the Arubo site was ­heavily disturbed before excavation, caused by the dredging of a fishpond. Although that activity led to the discovery of the site, most of the artefacts were recovered from on or close to the ground surface, making it impossible to get a reliable chronometric date. Nevertheless, the morphological affinity with East Asian Early Palaeolithic material is striking. Unlike the Cabalwanian sites, Arubo shows a bifacial component. The evidence for curation and re-use was noted in microscopic use-wear analyses (Pawlik 2002; Teodosio 2006), as well as variation in core preparation and core reduction. Among the several core forms recovered at Arubo is a so-called horsehoof core, similar to those found in Java seemingly associated with Early Palaeolithic assemblages (Koenigswald 1936; Heekeren 1972; Bartstra 1984; Soejono 1984; Sémah et al. 1992; Simanjuntak 137

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares et al. 2001). Another similarity was noted with several excavated Early Palaeolithic sites in South China (Pawlik 2004, 10), where handaxes and other bifacial and unifacial forms similar to Arubo are frequently recovered (e.g., Huang 1989; Xiang Anqiang 1990; Xie Guangmao 1990; Schick & Zhuan 1993; Leng Jian & Shannon 2000; Peng Shulin n.d.) and where bifacial technology may have begun circa 800,000 years ago (Hou et al. 2000). The discoveries at Arubo may suggest that there is yet still an Early Palaeolithic colonization of Luzon Island by a hominin species that has so far evaded identification and characterization. Until recently, the earliest known anatomically modern human fossils recovered in the Philippines were those from Tabon Cave, located on Lipuun Point, Quezon Province, Palawan Island, and estimated to be as much as 40,000 years old (Fox 1970). Further study revealed that the human remains, including a frontal bone, two mandibular fragments and several teeth, actually represented several individuals. Radiocarbon-dated charcoal from the corresponding layer however pointed to a much younger age of approximately 22,000–24,000 uncal BP (Fox 1970, 40–44). The earliest date recorded by Fox (1970, 24) on charcoal, apparently associated with what he called Flake Assemblage IV, was 30,500 uncal BP (UCLA-958), but in a rough estimate of sedimentation and ‘age-depth’ relationships, Fox concluded that the oldest deposits and Flake Assemblage V would probably be circa 40,000 BP. Thirty years later, a frontal bone was directly dated by uranium gamma ray counting at the Institut de Paléontologie Humaine of the Muséum national d’histoire naturelle in Paris, and its date was corrected to 16,500 ± 2000 cal BP (Dizon et al. 2002). A human tibia found 21 cm below modern ground surface in a disturbed layer excavated during a re-investigation of Tabon Cave by the National Museum of the Philippines and the Institut de Paléontologie Humaine, Paris, delivered another uranium series date of 47,000 +11,000/−10,000 BP (Détroit et al. 2004). Although this is consistent with the estimate of Fox for the lowest cultural layer in Tabon Cave to ~40,000 BP, the very high standard error of the U-series dates demands a cautionary consideration of the absolute dates of the Tabon human fossils. The upstanding karstic limestone formations of Lipuun Point currently extend out from the coast and are surrounded on three sides by the sea, but during the early occupation of Tabon Cave between 37 and 58 ka it is estimated that sea levels would have been 60–80 m below modern levels and the coast would have been 32–36 km away. Bathymetric analyses and coastline ­reconstructions indicate that in order to reach Palawan from the nearest island of Borneo during this period it would have required a sea crossing of approximately 13 km (ibid.) and a water craft capable of crossing open sea would have been necessary. Palaeoenvironmental reconstruction indicates that the vegetation on Palawan would probably have been dominated by open woodland and savannah rather than the tropical rainforest that is dominant on the island today (Bird et al. 2007; Wurster et al. 2010). Unlike other sites close to Tabon such as Guri Cave, Sa’gung and Duyong, which all date to the mid-late Holocene, there is no evidence of coastal shell middens, which would support the sites distance from the coast, and it is likely that subsistence strategies by the inhabitants of Tabon would have been based primarily on inland resources (see Piper & Rabett, this volume). Unfortunately, very few animal bones were recovered from the Palaeolithic layers at Tabon, and most of these were those of cavernocolous bats and birds (Fox 1970, 38–39). The only large mammals identified in the assemblage were those of the endemic Palawan bearded pig (Sus ahoenobarbus) and extinct deer. A more complete record of Late Pleistocene and Holocene foraging strategies on Palawan Island has recently been recorded at Ille Cave, near El Nido in northern Palawan. The earliest recorded human occupation of the site is from approximately 14,000 cal BP and includes chert artefacts and a substantial animal bone assemblage (Ochoa 2009; Piper et al. 2008, 2011; Pawlik 2010;), which was radiocarbon-dated to an age between 13,890 and 14,048 cal BP (OxA-16666; Lewis et al. 2008). Hunting appears to have focussed on two species of deer that are now extinct on Palawan, including the Calamian deer (Axis calamianensis, which is still extant on the islands of Culion and Busuanga to the north of Palawan), as well as smaller proportions of the Palawan bearded pig, macaques and a range of smaller mammals and reptiles (Ochoa 2009; Ochoa & Piper in press).The tiger (Panthera 138

Modern Humans in the Philippines tigris) and possibly the wild dog (Cuon alpinus) are also both represented, and in combination with the other mammal fauna support other palaeoenvironmental records in suggesting that the dominant vegetation was open woodland or savannah during the Late Pleistocene. The oldest direct 14 C date of 9220 ± 45 uncal bp (OxA-21179) or OxCal: 8558 – 8303 cal BP on a Canarium nut shell indicates that this plant was being utilised by the foragers of Palawan at the beginning of the Holocene and probably earlier (Carlos 2010). Overall the record from Ille suggests that the early human colonists of Palawan used a diverse range of technologies and techniques to extract numerous different resources from various ecotones in the local and regional environment. Recent excavations at Callao Cave in the karstic limestone region of Peñablanca in northern Luzon have produced evidence that modern humans may have reached the Philippines some considerable time before the occupation recorded in Tabon Cave (Mijares et al. 2010). Callao is the largest and longest of the caves in the Peñablanca area, being more than 350 m from the main entrance to the innermost chamber. The passages are between 14 and 30 m wide with a floor to ceiling height varying between 10 and 45 m and an elevation of 85 m above sea level. Callao was first excavated by Maharlika Cuevas, and a team from the National Museum of the Philippines in 1979 and 1980 (Cuevas 1980) and again since 2003 under the direction of Armand Mijares (Mijares 2004, 2007, 2008; Mijares et al. 2010). In 2003, 10 different stratigraphic horizons were identified during excavation. Layer 8 (105  cm to 110  cm) produced chert flakes, fragmentary animal bones and evidence of the remnants of a hearth at the south end of Sq. 1. An AMS 14C determination on charcoal from this layer returned a date of 25,968 ± 373 uncal bp (Wk-14881), confirming that human populations had crossed Huxley’s modification of Wallace’s Line from Palawan or via the Sulu Sea and north into the main Philippine Archipelago during the Late Pleistocene. Excavations re-started in 2007 and continued in 2009, first with the extension of the trenches excavated in 2003 to greater depths and then with new excavations to the southwest of the original trenches. Between 160 and 250 cm below the modern ground surface, archaeological remains were scarce with just a single flake, a chert core and a few deer bones recovered from Layer 11. In Layer 13 (255 cm to 265 cm depth) the bones and teeth of animals became more abundant, and at a depth of 270 cm a cemented calcium carbonate deposit was encountered (carbonised breccias) containing a relatively dense concentration of animal bones. It was within this layer and associated with the faunal remains that the human right third metatarsal was identified. Initially, two cervid teeth (Callao 1 and 2) were chosen for U-series and electron spin resonance dating from the top and bottom of the breccia layer. They produced U-series ages of 52 ± 1.4 ka and 54.3 ± 1.9 ka respectively (Mijares et al. 2010). However, as uranium accumulation in bones may be delayed after burial, these dates were considered by the analysts to be absolute minima. Electron spin resonance (ESR) dating yielded a combined ESR/U-series result (see Grün et al. 1988) for Callao 1 of 66 + 11/−9 ka, while Callao 2 did not yield a result (the closed system ESR age estimate was younger than the U-series result). The discordant combined ESR/U-series results suggest some re-working of the faunal remains, as was confirmed by the taphonomic analysis of the bones. The hominin metatarsal was dated using U-series ablation to a minimum age of 66.7 ± 1 ka. The brecciated deposit containing the faunal remains and human bone is therefore considered to have a minimum age between 60 and 70 ka (Mijares et al. 2010). The enigmatic metatarsal has many morphometric characteristics similar to anatomically modern humans as well as some of archaic hominins, such as the size and shape of the proximal articular end that barely fall within the expected range of human variation, but the specimen has been provisionally assigned as that of a Homo sapiens (Mijares et al. 2010). The presence of an anatomically modern human in northern Luzon 60–70 ka raises important issues for the initial timing of the migration into and colonization of Island Southeast Asia. Conventional theory places the timing for the initial migration of modern humans into Southeast Asia at no more than 50 ka (e.g., Mellars 2005) on route to Australasia. Although an initial arrival of modern humans in the Sahul region as early as 60 ka has been proposed (Roberts et al. 1990; Chappell et al. 1996), 139

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares a review by Allen and O’Connell (2003) of the five sites in the Sahul region that have delivered dates beyond 45 ka and up to 62 ka concluded that these early dates were questionable and secure dating evidence for human occupation of Sahul can be considered to have occurred only around 45 ka (O’Connell & Allen 2004; Habgood & Franklin 2008). Hiscock (2009), however, has pointed out that this date likely represents an absolute minimum for initial colonization and the first peopling of Australia along the now submerged west coast of the continent was probably somewhat earlier than 45 ka. For modern human populations to have reached northern Luzon the most likely route would have been across the exposed Sunda Shelf linking Peninsula Malaysia to Sumatra, Java and Borneo. It would then have required at least two or possibly three sea crossings between Borneo and Palawan, Palawan to Mindoro and then from Mindoro or, on a longer southern route with one or two sea crossings, from Borneo to the Sulu archipelago, then Mindanao and the Visayas to reach Luzon (Voris 2000; Sathiamurth & Voris 2006).Thus, it is possible that Homo sapiens were in ISEA substantially earlier than has been previously considered likely. Storm et al. (2005) have argued for a modern human presence at Punung III in eastern Java as early as MIS 5, and Zuraina Majid (1994) has claimed that the lithic material from Lenggong in Peninsula Malaysia dating to circa 74 ka was produced by anatomically modern humans, though no human fossils have been recovered at the latter location. It is also possible that stone tools and animal bone assemblages recorded in Java during MIS 3 and 4 at sites like Song Gupuh (Morwood et al. 2009) and Song Terus (Sémah et al. 2004) where human remains are absent in the early sequences were also conceivably produced by modern humans already present in ISEA. Further afield, Liu et al. (2010b) have argued that two molars and an anterior fragment of mandible from Zhirendong dating between 100 and 113 ka have morphological traits comparable with those of anatomically modern humans. Dennell (2010), on the other hand, considers also the possibility of a gracile late H. erectus for the Zhirendong fossils. If the Callao MT3 specimen indeed belonged to an anatomically modern human, then the initial migration of modern humans entering the Philippine archipelago would have happened more than 20 ka earlier than the current earliest southern migration to Sahul and would support these other regional studies in suggesting that occupation of mainland Southeast Asia and greater Sundaland by anatomically modern humans probably occurred prior to 70 ka. The alternative scenario is that the Callao bone is not that of a modern human at all but that of a different species of hominin altogether. Morwood and van Oosterzee (2007) hypothesized that Homo floresiensis and some of the archaic vertebrate faunas with which this diminutive ­hominin was associated had not crossed from Java but were more likely to have arrived via a route from the Philippines and Sulawesi. They observed that the archaic Middle Pleistocene vertebrate faunas have a higher diversity in these northern islands than those to the south and contain giant tortoise (Geochelone), proboscideans (Stegodon and Elaphas), bovines, cervids and rhinoceros (Rhinoceros luzonensis) (Bautista & Vos 2001; Morwood & van Oosterzee 2007). The only species capable of reaching Flores, transported on the Pacific’s Black Current flowing past the Philippines and between Sulawesi and Borneo southwards, were those with the greatest dispersal abilities, such as proboscideans, Geochelone, monitor lizards, rats and Homo floresiensis (Bergh et al. 2009). This line of reasoning would suggest that the Philippines should have an older record of hominin occupation than Flores, which currently stands at a million years (Brumm et al. 2010). There is, however, as yet no solid evidence of a Middle Pleistocene hominin colonization of the Philippines, and further research will be required to determine how long humans have actually been in the archipelago.

The Paleoenvironmental Setting of the Callao Hominin So far no stone tools have been recovered associated with the Callao specimen, but it was recovered with a small but important animal bone assemblage (Piper & Mijares 2007). Most of the 140

Modern Humans in the Philippines

Figure 11.2.  A cut mark on an animal bone from the 67 ka layer in Callao Cave (after Manalo 2011; photograph by the authors).

bone assemblage was heavily fragmented and demonstrated the taphonomic effects of sub-aerial weathering and differential transportation and erosion by water. Nevertheless, seven unidentified long bone shaft fragments and a deer right tibial shaft fragment produced evidence of cut marks indicating that the Callao hominin was in fact a tool user and that the bone assemblage had principally been introduced to the cave by humans (Figure 11.2). The only taxon with enough skeletal elements represented to allow any comments on skele­ tal representation is the deer Rusa marianna (NISP = 157; based on counts from the 2004–2009 excavations). The articular ends of long bones such as humeri, femora, metapodials, radii and tibiae and complete extremities are all represented in the assemblage in small numbers. Loose teeth of both the maxilla and mandible indicate that both crania and lower jaws were originally introduced to the site. Taking into consideration the differential destruction of these less dense blade-like structures of the scapula, mandible and pelvis, and spongy cancellous bone of the vertebrae, it is likely that most body parts were being introduced to the site. At least six individuals were identified in the assemblage on the basis of maxillary and mandibular M3s. All the M3s are moderately worn indicating that they were from mature individuals. A single right lower dp4, a left distal tibia and distal metapodial shaft fragment with unfused epiphyses suggest that at least one sub-adult is represented. In addition a total of 35 fragments were identified as an endemic pig, which on Luzon is almost certainly the Philippine warty pig Sus philippensis. Two fragments of mandibular molar column recovered from Squares 1 and 2 at a depth of 280  cm are from an unidentified, extinct bovine. On Luzon the only records of a Bubalus sp. have been associated with the Middle Pleistocene fossil record (Croft et al. 2006). Two other Bubalus species are known from the Philippine archipelago: the extant B. mindorensis on the island of Mindoro and the extinct diminutive B. cebuensis on Cebu Island. Croft et al. (2006) suggest that the small size of both known Bubalus taxa might result from size reduction on isolated islands. The new fragmentary teeth from Callao imply that a now extinct bovine also once inhabited Luzon in the Late Pleistocene. 141

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares Even though just three mandibles identifiable to different species of rodent murid were recovered from Callao, they still provide some useful information of biodiversity and Pleistocene environments (Heaney et al. 2011), for example, the recovery of a species of Batomys in the lowland forests of Peñablanca. The only extant species known on Luzon inhabit the Central Cordillera on the other side of the Cagayan River Valley and live only in montane forest from circa 1350 to 2480 m above sea level dominated by trees such as oaks. Environmental records suggest a climatic minimum at around 70 ka (Oxygen Isotope Stage 4; Johnson et al. 2006; Wang et al. 2008), and pollen studies have produced evidence of elevational lowering of vegetation bands and different plant communities during the Late Pleistocene and early Holocene (Stevenson et al. 2010), but they do not record the presence of the montane oak-myrtle-conifer plant community at this low elevation. Perhaps it is more likely that the species of Batomys recorded at Callao was adapted to inhabiting different ecological communities from those of its montane relatives, but it does imply that there were much cooler lowland temperatures and, as a result, that vertebrate communities inhabiting these environments have no modern analogies.

Microwear Analysis and Modern Human Behaviour Despite the presence of anatomically modern humans in the Philippines’ Upper Pleistocene, possibly up to circa 70,000  years ago, visible evidence for traits of modern behaviour in the archaeological record is extremely scarce. It is generally considered that the lithic assemblages in Southeast Asia do not show clear signs of technological advancement, and a formal component as compared to the European and African record is mostly lacking (Haidle & Pawlik 2009; Pawlik 2009a). Most Late Pleistocene and Early Holocene flaked artefacts were manufactured with a simple and opportunistic reduction process, and the flakes are mostly unretouched. Advancement in morphology and technology is hardly observed, and stone tools appear widely unchanged from the Later Palaeolithic to the Early Neolithic, like the Hoabinhian in Mainland Southeast Asia or the Tabonian in the Philippines (Colani 1927; Pookajorn 1988; Anderson 1990; Moser 2001; Kamminga 2007; Fox 1970, 1978; Patole-Edoumba 2002; Pawlik & Ronquillo 2003; Mijares 2004; Xhauflair 2009). Also, a true blade technology does not appear in the entire Southeast Asian lithic record, and the few reported ‘blades’ or ‘microblades’ are often dismissed as accidental products, having no further characteristics than an elongated shape (Tulang 2000; Moser 2001, 33; Pawlik 2004, 2009a; see Patole-Edoumba 2006 for a dissenting view).The wide absence of ‘modern’ tool types and formal tools in Southeast Asia’s Palaeolithic industries in general and especially in comparison to the European lithic record has been considered as due to the lack of ability in early humans in eastern Asia to make sophisticated tools (Colani 1927; Movius 1944; Mijares 2002) and is nowadays explained by the existence of a wooden or bamboo tool industry replacing formal stone tools and/or the poor availability and difficult acquisition of lithic raw material (e.g., Narr 1966; Solheim 1970; Pope 1989; Schick & Dong Zhuan 1993; Reynolds 1993; Mijares 2002; Dennell 2009). Owing to taphonomic reasons, these ‘vegetal industries’ are hypothetical. Tools made of bamboo and wood have not so far been found in Pleistocene and early Holocene archaeological assemblages. On a practical point, it would be necessary to have at least some stone tools to produce those made of organic materials.The wood or bamboo tool hypothesis does not consider factors of tool mechanics or tool uses (Haidle & Pawlik 2010; Pawlik 2010), nor does it deal with the fact that large lithic assemblages have long been known in the Southeast Asian archaeological record (e.g., Saurin 1966; Gorman 1971; Heekeren 1972; Harrisson 1972; Hutterer 1977; Fox 1978; Anderson 1990). Certainly, it can be assumed that a wide variety of tools and utilitarian objects were made of vegetal materials, including bamboo and wood, but they were more likely complementing the lithic tool kit rather than acting as replacements, like the few bone tools found in Southeast Asia (Barton et al. 2009; Pawlik 2009a). Furthermore, the causality that the production of vegetal tools led to a simplification of lithic industries has not been 142

Modern Humans in the Philippines convincingly explained. Finally, artefacts made from rocks highly suitable for knapping (i.e., chert or even obsidian) are not uncommon in Southeast Asian archaeological sites (e.g., Beyer 1947; Gorman 1970; Charoenwongsa 1988; Pookajorn 1988; Moser 2001; Pawlik 2002, 2004; Mijares 2002, 2004; Neri 2002, 2005). Two technocomplexes belonging to the Upper Palaeolithic of the Philippines have been morphologically and technologically analysed and published so far, the so-called Tabonian Industry of Palawan (Fox 1970; Patole-Edoumba 2002) and the Peñablanca expedient technology in Northern Luzon (Mijares 2002). The human remains at Tabon Cave were associated with a lithic assemblage consisting primarily of flakes produced mostly on radiolarian chert (Schmidt 2008). Intentional modification of these flakes is rarely observed, and edge retouches and alterations are usually caused by use (Fox 1970; Ronquillo 1981; Patole-Edoumba 2002; Mijares 2004). A comparison of the Palaeolithic assemblage of Tabon Cave with the lithic materials from several Holocene sites in Palawan, for example, Duyong Cave, Guri Cave and Pilanduk Rockshelter (Fox 1970, 45–65; 1978; Kress 1979; Tulang 2000; Patole-Edoumba 2002; Pawlik & Ronquillo 2003), has demonstrated a continuation of the so-called Tabonian from the Upper Pleistocene into the Holocene until the Early Neolithic. The excavation at Callao Cave in the northern part of Luzon Island produced a small assemblage of flaked artefacts. Similar to Tabon, the Later Palaeolithic industry of Callao and several other sites of the same limestone formation at Peñablanca such as Laurente Cave, Minori Cave and Rabel Cave continue without significant morphological changes into the mid-Holocene (Ronquillo 1981; Mijares 2002, 2007, 2008; Pawlik & Ronquillo 2003). In general, the Peñablanca lithic technology, at more than 1,000 km distance from Tabon Cave, appears related to the Tabonian and consists of simple flake assemblages without formal elements, made predominantly of andesite and chert. On the basis of a technological study combined with microscopic use-wear analysis, these assemblages have been characterised as products of an ‘expedient technology’ where flakes were produced from locally available raw material by direct percussion and without further modification, used for single tasks and discarded afterwards (Mijares 2002). This interpretation corresponds with recent microwear studies on artefacts from Tabon Cave where the limited appearance of microwear traces suggest a similar strategy for the Tabonian industries on Palawan (Mijares 2004; Xhauflair 2009). Also the newly excavated chert assemblage from the Pleistocene layer of Callao Cave fits into an expedient technological tradition in terms of technology and use wear (Mijares 2008). If this kind of strategy for lithic tool production and use predominates in the Philippine Palaeolithic, then it is not surprising that the stone tools demonstrating the more complex manufacturing and reduction sequences observed in Europe and elsewhere are missing. Expedient technology lacks curation, core preparation, indirect percussion and therefore specialized blade production and geometric tool forms. Formal tools in general are extremely rare. The simple and indifferent technology that produced an overall amorphous small flake industry is dominant until the developed Neolithic and the beginning of the Austronesian expansion (Bellwood 1997). Non-lithic traits like tools made of bone, antler and shell, projectile points, figurative art, musical instruments and personal ornaments are also absent until the mid-Holocene. Only a few shell artefacts appear in the earlier stages of the Philippine Neolithic but not before 7,000 BP (Ronquillo et al. 1993; Szabó 2005). Although the Philippine Later Palaeolithic assemblages are almost certainly associated with modern Homo sapiens since at least about 50 ka ago, they do not comprise a distinctive package of modern traits and behaviour. This leaves us with two possibilities: Either the cognitive, cultural and technological behaviour of their makers was completely different from modern hominins in Europe and Africa and not ‘modern’ at all, or the hypothesis developed for and on cultural assemblages from Europe and Africa cannot be directly applied to materials from the ISEA and Pacific region and needs to be revised (Habgood & Franklin 2008, 2011; Haidle & Pawlik 2010). Furthermore, signs of behavioural and cognitive abilities of Southeast Asian hominins might be found elsewhere and with the aid of other methods. 143

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares One potential source of clues and traces of modern traits could be microwear analysis on lithic tool surfaces. Microwear analysis allows the determination of stone tool uses and functions and reconstructs prehistoric technology and behaviour (Semenov 1964; Keeley 1974; Tringham et al. 1974). This method applies basic physical principles of interacting surfaces in relative motion and studies the wear and tear created during such interaction between a working tool and the worked object (Yamada 1993).The effects are the same for modern as well as for prehistoric tools made from stone, usually chert and flint. Experiments have demonstrated that almost any kind of contact, even with much softer materials, will result in wear traces on the stone tool (e.g., Kamminga 1979; Keeley 1980; Odell 1981;Vaughan 1985; Unrath et al. 1986; Beyries 1988; Pawlik 1992; Anderson et al. 1993). As an initial case study, a lithic assemblage from the lowest layer of Ille Cave in the Dewil Valley in El Nido, northern Palawan Island, underwent a microwear analysis. Radiocarbon dates delivered an age for this layer of 13,890 to 14,048 cal BP (OxA-16666; Lewis et al. 2008). The morphology of the artefacts appears similar to Tabon and Peñablanca, with simple and irregular flakes manufactured by direct percussion and an absence of formal tools (Pawlik 2009b). Perhaps even more characteristic for modern behaviour are traces and residues that resulted from the working of shell and the use of pigment as indicated by residues of red ochre on some artefacts. On one endscraper-like flake, traces of red pigment appear in combination with hide working (Figure 11.3, A–C). Although it cannot be determined with absolute certainty whether the pigment stains are directly associated with hide processing or resulted from a different activity, the use of red ochre as a colouring or tanning agent for skins and leather in the Palaeolithic has frequently been observed (e.g., Vaughan 1985; Büller 1988; Juel Jensen 1988; van Gijn 1989; Pawlik 1995; Barham 2002). The surfaces of several artefacts from Ille Cave also carry so-called bright spots. They are commonly regarded as the result of non-intentional, repetitive rubbing contacts between siliceous artefacts, for example, when carried together in a pouch for some time (Unrath et al. 1986; Levi-Sala 1996).The appearance of such traces can, therefore, be considered as signs of curation, the process reflecting a tool’s actual use relative to its maximum potential use (Andrefsky 2008).This can also be interpreted as an advanced behavioural concept, contrary to the use-onceand-discard expedient technology model (e.g., Mijares 2002). Impact scars as well as the presence of polish spots on the tip and longitudinal striations on elevated parts of the microtopography of both faces on a triangular flake, (Figure 11.3, D–G), suggest that it was used as projectile implement (e.g., Fischer et al. 1984; Lombard 2005a, 2005b; Lombard & Pargeter 2008). The interior surface of the base displays polish that is not use-related but does conform to what is expected from minor movements of a tool within its haft (Cahen et al. 1979, 681). Blackish residues that appear along with such polishes are probably the remains of organic resin used as hafting mastic (Figure 11.3, H). The combination of impact wear, hafting traces and residues is quite remarkable and identifies these artefacts as hafted armatures that were attached to shafts. Two more working tools exhibit characteristic hafting polishes associated with a blackish-reddish residue film, one laterally retouched flake displaying traces of working harder organic material as well as bright spots (Figure 11.3, I–M). This kind of adhesive appears to be very similar to resin residues found on projectile points made of bone and stingray spine from the West Mouth of Niah Cave in Borneo, dated to 11,700–10,690 cal BP (OxA-12,391 and OxA-11865; Barton et al. 2009). Resins recovered from Niah Cave have been identified as deriving from either Shorea spp. or Canarium spp. (Lampert & Thompson 2002). These trees and their resins are also common in the Philippines. Residues with an appearance very similar to the resins from Niah Cave have been found as remains of appliqués on shell disk beads in the Neolithic layers of Ille Cave. Scanning electron microscope (SEM) imaging of the residues showed embedded remnants of plant tissues and structures, while energy dispersal X-ray analysis (EDX) confirmed the presence of metal elements in the residue matrix indicating that metal pigments were also a part of this multi-component coating (Basilia 2011). Shorea resin appears to be especially suitable for hafting purposes because it becomes soft again when heated up above 75° C (Tschirch 144

Modern Humans in the Philippines

Figure 11.3.  Stone tools from the Late Pleistocene layer of Ille Cave, Palawan with corresponding microwear traces and residues (after Pawlik 2010). A: Endscraper-like flake no. 41713 showing edge wear with traces of red pigment (B) and in combination with hide-working micropolish (C). D: Triangular flake no. 40406 used as projectile implement exhibiting impact scars (E, F) and micropolish with longitudinal striations (G) at the tip, and hafting residues associated with hafting polish at the basal part (H). I: Laterally retouched flake no. 41809 with ‘bright spots’ (J), resinous residues (K), and edge scarring caused by working harder organic material (L, M) (photographs by the authors).

145

Alfred F. Pawlik, Philip J. Piper and Armand Salvador B. Mijares & Glimmann 1896), which would make it an ideal binding material with regards to re-tooling processes and the replacement of worn-out implements. While the specialised bone points from Niah provide further proof for the availability of a Late Pleistocene hafting technology in Island Southeast Asia, the use of unretouched lithic flakes as hafted implements for multi-component tools at Ille Cave is unique and points to a technological concept that is beyond traditional morphological and typological models, but it is nevertheless a reflection of the constructive memory of its makers and their ability to perform complex sequences of action (Ambrose 2010). Use of pigments, resin, shell and hide working is a further indicator for the presence of modern behavioural traits in the Philippine Palaeolithic record.

Discussion Palaeolithic research in the Philippines is contributing to our understanding of the timing of the initial colonization of ISEA, the landscapes human populations encountered on the different landmasses across the archipelago and how they adapted to the diverse range of environments they encountered. The tentative identification of anatomically modern humans at Callao Cave suggests that our species reached these islands circa 60–70 ka, more than 10 ka earlier than previously recorded at Tabon Cave. Landscape reconstructions have shown that to reach Luzon would have required at least two and possibly three open-sea crossings, and suggests that a form of watercraft technology perhaps already existed in MIS 3. Similar arguments for advanced boat technology have been posited for the initial colonization of Australia (Mulvaney and Kamminga 1999), the capturing of inshore and potentially offshore fishes more than 40 ka on East Timor (O’Connor 2007b) and the intermittent habitation of the Talaud Islands between Mindanao and Sulawesi more than 30,000 years ago (Ono et al. 2009). The interpretation of human subsistence strategies on Palawan and Luzon have been instrumental in developing the larger regional picture of human adaptive responses to the mosaic of environments they encountered across ISEA and Wallacea. Late Pleistocene hunting strategies on Palawan and in northern Luzon were geared to take advantage of the open woodland and savannah environments and focused primarily on large game. On Luzon, in particular, the diversity of vertebrate fauna available was restricted by the impoverished nature of the animal community inhabiting the island. Existence in these environments would have required a very different range of skills from those appropriate to the tropical rainforests surrounding Niah Caves in Borneo at a similar time (Piper and Rabett, this volume), where a wide diversity of resources were acquired from a diverse variety of ecological zones within the forests. Microwear analysis of the stone tools from the 25 ka occupation horizons at Callao suggest these were used in a range of tasks including the processing of hard contact plant remains such as wood, bamboo, rattan and/or palm (Mijares 2008, 29). Although Mijares (2002; 2007; 2008) considers the majority of the flakes to represent parts of an expedient tool technology, he notes that two blade-like flakes could potentially have been used as spear or arrow points (Mijares 2008, 28). If verified this would represent the earliest known evidence for a projectile technology recorded in ISEA, and it would be several thousand years earlier than the first recorded bone projectile points at the Niah Caves in Borneo and other sites across the region (Rabett & Piper 2012). The hafting of artefacts in the Pleistocene is further supported by research on the lithic materials from Ille Cave where we have the first solid evidence for composite-tool production and complex tool design in the Philippine Palaeolithic. Hafted composite tools and the making of hafting mastic for fixing lithic armatures in wooden shafts have been observed in European Micoquien and Aurignacian assemblages (Pawlik & Thissen 2011; Dinnis et al. 2009). They are considered to be components of the European and African package (Keeley 1982; Wurz 1999; Deacon 2000; Ambrose 2010) and have been regarded as a significant trait of behavioural modernity for Southeast Asia and the western Pacific region (Barton et al. 2009; Haidle & Pawlik 2010; 146

Modern Humans in the Philippines Pawlik 2010). However, hafting traces are easily overlooked or neglected in microwear analyses (Cahen et  al. 1979; Keeley 1982). This analysis of relatively simple flakes from the Philippine Upper Palaeolithic shows that some were actually hafted armatures and parts of more complex composite tools. The dominantly small size of the flakes in Philippine lithic assemblages could even indicate the intention of the toolmakers to use them as hafted implements. This interpretation of lithic assemblages in the Philippines presents a different angle from the previously mentioned discussion of wood and bamboo industries to explain the absence of formal tools and lithic typologies in Southeast Asia. Considering bamboo and wood as prime material for the manufacturing of the shafts for composite tools and lithic armatures rather than replacements for stone tools opens up new avenues of research and provides a new view on the discussion of the ‘dilemma of missing types’ in Southeast Asia’s Palaeolithic and Epipalaeolithic (Haidle & Pawlik 2009). The emerging evidence from the Philippine Palaeolithic record suggests that during the Late Pleistocene human foraging populations were more technologically sophisticated than previously envisaged and that they demonstrated adaptive flexibility and ingenuity and were adept at successfully colonizing and inhabiting new and varied landscapes. Far from being peripheral to the expansion of H. sapiens into Sunda and Sahul, the Philippines are an integral part of that history and the Palaeolithic record of the country has an important role to play in further developing our understandings of hominin migration and colonization, adaptation, evolution and behaviour.

Acknowledgments This research was supported by an Emerging Interdisciplinary Research Grant of the University of the Philippines, OVPAA, code no. 2-002-1111212. Philip Piper was funded by the ARC Future Fellowship Grant FT100100527.

147

Chapter 12 Views from Across the Ocean A Demographic, Social and Symbolic Framework for the Appearance of Modern Human Behaviour

Phillip J. Habgood and Natalie R. Franklin

Introduction The continent of Sahul (Figure 12.1) was colonised some 50,000 years ago by anatomically and behaviourally modern humans. Reviews of the Pleistocene archaeological record of Sahul provide alternative perspectives on the debate surrounding the appearance of a “package” of archaeologically visible traits that have been argued to reflect modern human behaviour. We found (Franklin & Habgood 2007; Habgood & Franklin 2008, 2010) that the appearance of the individual traits revealed both chronological and geographical patterning within Sahul: • Chronologically, the package was gradually assembled over 40,000 years following initial occupation of the continent by behaviourally modern colonists, with four broad phases being recognised (Figure 12.2). • Geographically, seven “Zones of Innovation” were identified (Figure  12.1), and the Northern Zone could be further subdivided into three. From this pattern it was concluded that modern human behaviour cannot be automatically inferred from inventories of archaeologically recovered material, as the earliest Indigenous inhabitants of Sahul were behaviourally modern, yet the full package of traits was not present at the earliest sites (see also Brumm & Moore 2005; O’Connell & Allen 2007; Balme et al. 2009). Therefore, alternative explanations for the appearance of these archaeological features were sought. The potential causal influence of taphonomy on the patterning identified has previously been considered, but it was concluded that the pattern was not the direct result of taphonomic processes (Franklin & Habgood 2007; Habgood & Franklin 2008, 2010; contra Langley et al. 2011; Langley, this volume).Taphonomy does impact on site assemblages; however, we contend that the patterning for key components of the package was the result of material culture differences and cultural preferences of the Indigenous inhabitants. In this chapter we explore the inter-relationship of the broad chronological phases and the Zones of Innovation with the possible impact of ecological and demographic changes during the 148

Framework for the Appearance of Modern Human Behaviour 0

1000km

Papua New Guinea Northern Australia

Central western Australia

Central Australia

Murray-Darling

South western Australia

Tasmania

Figure 12.1.  Zones of innovation in Pleistocene Sahul. Shaded area indicates the continental shelf exposed at the Last Glacial Maximum (illustration by the authors). late Pleistocene on symbolic behaviour within Sahul. We also outline a demographic and social framework for explaining the appearance and disappearance of symbolic behaviour in Sahul evident in the form of art, personal ornaments and cemeteries. Finally, we briefly discuss the implications of this framework for Europe and Africa.

Demographic and Ecological Changes Population growth and demographic expansion have been viewed as possible reasons for the appearance of the package of modern human behaviour (e.g., Lindly & Clark 1990; McBrearty & Brooks 2000; Kuhn et al. 2001; Shennan 2001; Henshilwood & Marean 2003; James & Petraglia 2005; Powell et al. 2009). The general assumption is that the appearance of the package may be a consequence of population growth and the resulting increase in contact between groups or greater competition for resources. Furthermore, the transmission or uptake of cultural innovations may be more successful in larger populations than in smaller ones (Shennan 2001). Powell et al. (2009) argued that the geographic variation in the timing for the first appearance of modern human behaviour between Africa and Europe can be explained by demographic factors – skill accumulation and maintenance increasing with increasing population density. Henshilwood and Marean (2003, 632) proposed that a “prime mover” for the appearance of behavioural traits that 149

Phillip J. Habgood and Natalie R. Franklin relate to resource intensification (e.g., subsistence and technological changes) was population pressure. Lindly and Clark (1990, 252) suggested the appearance of symbolic behaviour related to “information-processing requirements created by increased social complexity, population density, and/or subsistence uncertainty”. Similar observations have been made to explain changes in the archaeological record within Australia (Lourandos 1997; Haberle & David 2004; Brumm & Moore 2005; O’Connell & Allen 2007). Brumm and Moore (2005, 168) suggested that the appearance of components of the package in the Holocene may reflect the reaching of an organizational threshold, that regional populations had reached a level at which new channels of information transmission became necessary to alleviate conflict and establish boundaries. O’Connell and Allen (2007, 404) argued for population growth being the “driving force” behind the increasing appearance of components of the package following the Last Glacial Maximum (LGM). It is difficult to establish population size and demographic change for Sahul. Relatively low population densities throughout the Late Pleistocene are suggested by the number of sites and intensity of site usage. Models based on Late Pleistocene, Mid-Holocene or bi-directional population growth on either side of the LGM have been proposed (Birdsell 1977; Beaton 1983; O’Connor et al. 1993; Smith & Sharp 1993). The demographic history of Sahul would have been very complex, involving population growth, expansion, replacement, contraction and differing amounts of population pressure that varied at a local or regional level. The pattern for the appearance of the package during the Late Pleistocene that we identified is not consistent with the suggestion that it was the direct result of continent-wide population increase, because individual components of the package appear at sites such as Riwi Cave and Mandu Mandu Creek (northwestern Australia) that reflect low-level or intermittent occupation at this time (Morse 1993a; 1993b, fig.  2; Balme 2000, fig.  6). Also, population increase cannot explain the sporadic and uneven distribution of individual components of the package through time. Therefore, whilst a continent-wide population increase cannot account for the identified chronological and geographical pattern in Sahul, the driving force may have been regionally or locally based population pressure (population density vs. resource availability) that reached critical thresholds at different times in different localities across the continent. Australia is an arid country, with water being a crucial determinant of Aboriginal subsistence and settlement, and so population numbers would have expanded or contracted at a regional level depending on resource availability (see CLIMANZ 1983; Gould 1980; Habgood 1991; Haberle & David 2004).The four phases we identified (Figure 12.2) can be broadly related to periods of climatic and ecological change and probable demographic transformations (Habgood & Franklin 2011): • Phase 1 (ca. 50–35 ka BP) equates with the colonisation of Sahul and the movement of small groups into a range of environments and landscapes across the continent, under a cooler and drier climate than at present and with lower sea levels. • Phase 2 (ca. 35–25 ka BP) incorporates the commencement of increasing aridity from late oxygen isotope stage 3 and the resulting climatic deterioration in the lead-up to the LGM. • Phase 3 (ca. 25–12 ka BP) encompasses the LGM with its increased aridity and reduced temperatures in much of Sahul. • Phase 4 (after ca. 12 ka BP) documents climatic amelioration with the waning of the LGM and sea level rise. O’Connor et al. (1993, 95) argued that late Pleistocene Sahul “experienced significant changes in demography and that these changes are fundamental to our understanding of postglacial 150

Framework for the Appearance of Modern Human Behaviour Ochre Long distance exchange Art

Phase 1 Colonisation off Australia

Burials Groundstone and waisted hatchets Freshwater Shellfish Beads

Phase 2

Grindstones Marine Shellfish Thumbnail Scrapers Mining & Quarrying

Phase 3

Notational pieces Bone tools

Phase 4

Microliths, Points,Tulas

ka

10

20

30

40

50

60

Figure 12.2.  Phases for behavioural innovations in Pleistocene Sahul (illustration by the authors). economies and territorial organisation”. We would contend that these changes in demography commenced prior to the LGM and that they impacted on economies, territorial organisation and symbolic behaviour. Mounting aridity from approximately 32,000  years BP would have impacted significantly on resource availability and therefore Aboriginal settlement patterns and population numbers. For example, the study of phytoliths from Carpenter’s Gap, northwestern Australia, documents increasing Spinifex grass in the area, indicating increasing aridity from 33,000 years BP (Wallis 2001). Variations in resource availability resulting from these climatic and ecological changes would affect population density and population pressure, which may be the trigger for the crossing of population pressure thresholds at different times in different regions and the appearance of symbolic behaviour. Veth (1993) used a biogeographical approach to identify refuge localities (providing networks of permanent water that could withstand climatic extremes), barriers (major deserts) and corridors between the refuges and barriers. The refuge areas encompass our Zones of Innovation. The major riverine systems of southeastern Australia and some coastal areas were also probably refuges (Morse 1993a, 1993b;Veth 1993). During the LGM, populations retreated to these refuge areas as they provided access to permanent water and other resources (Smith 1989;Veth 1993). With increasing aridity leading into and during the LGM, there were resultant changes in population distributions and density with some sites or areas being abandoned, while others experienced increased occupation intensity when people moved to refuge areas with networks of permanent water and reliable resources (Smith 1989; O’Connor et al. 1993; Veth 1993, 1995, 2006; Porch & Allen 1995; Brown 1987b). Veth (1993, 7) proposed that the “timing and magnitude of climate changes are unlikely to have been in parallel throughout the late Pleistocene and the Holocene in both northern and southern Australia”. O’Connor et al. (1993) detailed how, 151

Phillip J. Habgood and Natalie R. Franklin during the LGM, northwestern Australia was drier and colder, whereas southwestern Australia was wetter and colder. They also demonstrated that discard rates, occupation intensity and population varied between the two areas at this time – decreasing in the northwest while increasing in the southwest. Variable patterns of site usage would reflect different socio-demographic responses during increasing aridity and deviations in resource availability leading into and during the LGM  – either the concentration or packing of people within refuge areas with reliable resources (more intensive occupation) or the abandonment of some areas. There is also evidence for the congregation of peoples focused on the exploitation of particular resources during the LGM – Bennett’s wallaby (Macropus rufogriseus) in the southwestern highlands of Tasmania and freshwater shellfish and fish along and around the palaeo-river and palaeo-lake systems of southeast Australia. These demographic developments (changes in population distribution and density) would have resulted in a process of “population compression” (Witter 2007), “culture contact” and packing of more groups into smaller territories and, eventually, increased population pressure within refuge areas with confined geographic and cultural space. Such demographic packing would result in increased territoriality (more bounded territories with more “closed” social networks) and increasingly complex social systems, which, as observed by Lourandos (1997, 27), functioned in part to regulate or manage the relations between people and environment and between people and people in demographically and socially complex areas. Adaptational stresses involving subsistence and demographic factors can be responsible for increasing a population’s participation in symbolic activity (Conkey 1978). Wadley (2001) argued that modern human behaviour was evident in the archaeological record when items of external symbolic storage (art, ornaments, lithics) were used in the definition of individual or group identity, reflecting what we refer to here as “bonding” or “bounding” of groups. In the following sections, we explore these concepts in more detail, with a focus on the appearance of art, personal ornaments and cemeteries.

A Social and Symbolic Framework It has been proposed that in the ethnographic past Aboriginal people adopted sociological solutions to manage irregular aridity (risk minimisation strategies), such as more open social networks and the establishment of “Dreaming tracks” (e.g., Gould 1980; Peterson 1986;Witter 2007). Gould (1980, 85) observed that “what one sees . . . throughout the Western Desert are widely ramified and interconnected kin-sharing networks that serve to mitigate the uncertainties of rainfall and food”. Dreaming tracks or story or song lines reflect the activities of the “Dreamtime” ancestors during the creative era as they emerged from the earth and travelled across the country, creating various geographical features still present in the landscape today before re-entering the earth (see Franklin 2004 for further discussion and references). Dreaming tracks, the creation journeys of Ancestral Beings, frequently extended across group boundaries and provided “Chains of Connection” (Mulvaney 1976) and created webs of inter-connectedness between people and places (Peterson 1986), thus linking Aboriginal groups across vast distances. Many Dreaming tracks were more than 1,000 km long, so that sites along the tracks were associated with other sites well outside the particular “tribal” territory, creating non-genealogical relationships between different individuals, clans and sites many hundreds of kilometres apart. The landscapes of different groups were therefore linked into an even broader cultural landscape encompassing much of arid and semi-arid Australia (Gunn 1995; 1997; 2000). Art motifs were indissolubly linked with Dreaming tracks and custodianship of particular parcels of land. In the context of the recent bark paintings of northeastern Arnhem Land, Morphy (1981, 60) stated: 152

Framework for the Appearance of Modern Human Behaviour Aboriginal art is about meaning. Individual paintings and carvings are products of systems of communication which create meaning by encoding relationships between things, by relating people and place to the world of the land-transforming Ancestral Beings. Clan designs from the ethnographic past established connections between Aboriginal groups and Ancestral Beings. For example, a linked diamond design represented the Wild Honey Ancestor who travelled widely through northeastern Arnhem Land crossing the “Country” of many Aboriginal groups, but each group associated with the Wild Honey Dreaming had its own diamond design as a manifestation of that Ancestral complex (Morphy 1981) – variations on a shared design theme. Dreaming tracks also frequently coincided with the distribution of prehistoric rock art: • Morwood (1979; 2002) noted the association between the distribution of a particular painted rock art motif (a distinctive paired tortoise motif) in the central Queensland highlands and the Dreaming track taken by two Dreaming beings, a snake and a goanna. • Taçon (2008) analysed the rock art and associated oral history of Waanyi “Country”, northwest Queensland, documenting a large number of rainbow-like designs in an area that was a key junction for significant Dreaming tracks that covered vast areas of northern and central Australia. • The Dreaming path taken by Namarrkon, the Lighting Man, in Arnhem Land is restricted to the northern part of the plateau, and coincides with the distribution of distinctive beeswax figures (Gunn & Whear 2008). Rivers were created by Ancestral Beings during the “Dreamtime” and often became pathways for ceremonial exchange and trade that linked Aboriginal people across enormous distances, facilitating the construction of extensive social networks (Kerwin 2010; Ross 1997). Kerwin (2010) outlined the social importance of trade and detailed overlap in Dreaming tracks and trade routes, highlighting the cementing of alliances through shared Dreaming tracks and ritual. Elkin (1949) also noted that at least part of the extensive trade network that extended between northwest central Queensland and Lake Eyre was the track of the Mura Mura Dreaming beings. The frequent correlation between Dreaming tracks and trade routes suggests that both formed a means for the interaction of people across the landscape and for the diffusion of ceremonies, motifs and other material culture items. Although the discussion so far has focused on the ethnographic past, we would suggest that the adoption of sociological solutions such as more open social networks and the establishment of Dreaming tracks and trade routes to manage irregular aridity occurred during the late Pleistocene. This was a period of population expansion and the acquisition of knowledge about “Country”, and such sociological solutions facilitated formal linkages with “related communities” – the “bonding” of groups (inclusion). Further, we would contend that sociological processes also functioned to manage increasing population pressure caused by mounting aridity, population compression and intensive resource utilisation and control – the “bounding” of groups (exclusion). All of these sociological solutions resulted in changes in symbolic and other behaviours as manifested in the appearance of art, personal ornaments and cemeteries.

Bonding and Open Social Networks Bonding behaviour refers to long-distance links between groups of people with more “open” social networks and is reflected in the alliances that during times of resource stress and shortage enabled access to the resources of other groups (Gamble 1982; Lourandos 1997). For example, studies of recent Aboriginal territorial organisation have found that in areas with unpredictable 153

Phillip J. Habgood and Natalie R. Franklin and precarious resources, such as the arid zone, Aboriginal groups tended to be small, and social and territorial networks extended over large areas facilitating easy movement across fluid boundaries to gain access to the resources of other groups (Peterson 1986). Hayden (1987) has proposed that ritualised procedures shared between groups strengthened alliances between those groups. Rossano (2009) furthered this argument by stating that extended, open social networks (including inter-group trade and other alliances) that cut across traditional within-group boundaries required a repertoire of increasingly elaborate and demanding rituals. The Dreaming tracks just discussed would constitute examples of such elaborate ritual practices. Archaeologically, bonding behaviour would be reflected in the long-distance movement of materials and in similarities between items of material culture across large areas of the continent as communication between bonded groups was often facilitated by more homogeneous cultural traits (Lourandos 1997). In recent times, ochre, Pituri (native tobacco), pearl and Baler shell and many other items were exchanged up to 3,200 km across Australia (Mulvaney 1976, 1981; Kerwin 2010). There is also ample evidence during the Late Pleistocene for long-distance movement of materials, such as ochre, shell and stone throughout Sahul (Habgood & Franklin 2008). Examples include: • Fragments of Dentalium shell beads at Riwi Cave and Carpenter’s Gap Rockshelter 1 – the sea was some 300 km away at the time • Stencils of Baler shell pendants (che-ka-ra), most probably from the Gulf of Carpentaria, in the rock art at Carnarvon Gorge (Figure 12.3), central Queensland, and at Walkunder Arch Cave in the Chillagoe region of north Queensland, over distances greater than 1,000 km and 300 km, respectively • The Lake Mungo 3 grave covered with powdered red ochre – the nearest source some 200 km away • Ochre at Mandu Mandu and Puritjarra – the nearest sources some 300 km and 150 km distant • Cuddie Springs – stone source for grindstones 100 km away • Southwest Tasmanian sites – Darwin Glass transported up to 100 km • Bismarck Archipelago – late Pleistocene transportation of New Britain obsidian across sea distances in excess of 350 km Bonding behaviour is also evident in rock art, illustrated by the similarities between the widely distributed rock engraving sites that have been referred to as the “Panaramitee tradition” (Flood 1997; Franklin 2004). The Panaramitee has been represented as relatively homogeneous at a continental level in terms of technique, form, motif types and motif proportions (Maynard 1979). A particularly indicative example of the widespread similarities between Panaramitee tradition engravings is the Kybra site in Western Australia (Franklin 2007a; Figure 12.4), whose engravings are more different from other sites in that state than from engravings in the east of the continent. There are also broad similarities over thousands of kilometres between the engraved faces that are part of the tradition, suggesting widespread contact over vast areas (Franklin 2004; McDonald 2005; Ross & Smith 2009). However, the Panaramitee tradition is not identical throughout Australia in that there is interregional variation in terms of differing emphases on motifs or combinations of motifs within an overall restricted range across the continent (Franklin 2004; 2007a; 2007b). The Discontinuous Dreaming Network Model (Franklin 2004; 2007a; 2007b) explained the similarities between Panaramitee tradition engravings in terms of bonding behaviour in harsh, uncertain environments over large areas of the continent, with Dreaming tracks constituting the linking mechanism. The inter-regional variation found within the Panaramitee tradition potentially reflected the depiction in rock engravings of the motifs relevant to those particular parts of the Dreaming tracks. At N’Dhala Gorge in central Australia, for example, there are distinctive fern motifs and 154

Framework for the Appearance of Modern Human Behaviour

Figure 12.3.  Stencils of Baler shell pendants (che-ka-ra), Carnarvon Gorge, central Queensland (photograph by the authors). human figures with radiating headdresses (Figure  12.5)  – “variations on a theme” (discussed already for the bark paintings of northeastern Arnhem Land). We argue that trade networks and Dreaming tracks extended well back into the Late Pleistocene, constituting a linking mechanism or bonding behaviour and representing open social networks across the continent. This is supported by Late Pleistocene dates for long-distance movement of materials and for Panaramitee tradition rock engravings. The latter range from a minimum of 14,000 years BP in Cape York Peninsula at Early Man (Rosenfeld 1981) and Sandy Creek 1 (Morwood et al. 1995) to in excess of 21,000 years BP in the Kimberley (Watchman et al. 2001; Watchman et al. 2005). A bonding function for material culture traits was required at times of resource uncertainty, apparent in particular during the colonisation phase of the continent as it would have facilitated access to resources and assisted in the management of irregular aridity. It should be stressed however, that bonding behaviour has continued through to the ethnographic past, reflected, for example, in evidence that Panaramitee tradition rock engravings continued through time, showing a remarkable continuity in practice over thousands of years. For example, a “residual” Panaramitee tradition is present in the Sydney region to within the last few thousand years (Taçon et al. 2006, 2010; McDonald 2008).

Bounding and Closed Social Networks Bounding and emblemic behaviour is represented by items of material culture that transmit a message of group identity. It functions to mark the territorial boundaries of a group (Conkey 1982; Taçon 1994; Wiessner 1983). For example, in areas such as coastal Arnhem Land and southeastern Australia, the environment can support a significantly larger population density than in arid regions, so that in recent times Aboriginal groups tended to be larger, territories smaller and more clearly defined, social networks more closed, and boundary maintenance required 155

Phillip J. Habgood and Natalie R. Franklin

Figure 12.4.  Panaramitee tradition rock engravings at the Kybra site,Western Australia (illustration by the authors). 156

Framework for the Appearance of Modern Human Behaviour

Figure 12.5.  Panaramitee tradition rock engravings at N’Dhala Gorge, central Australia, showing the “classic” motif range as well as figures distinctive to this site (photograph by the authors). to a greater degree (Peterson 1986; Lourandos 1997). Emblemic behaviour is prevalent in such increasingly “crowded” social environments and, in this sense, is a “monitor of demographic stress” (Barton et al. 1994). It is also apparent in what Taçon (1994) has referred to as “socialising the landscape”, whereby people mark the land (i.e., rock art, stone arrangements), reflecting the group identity of the makers, whether that be clan, linguistic or some other form of identity (see also Franklin 1996). Bounding and emblemic behaviour would be evident archaeologically in the use of personal ornaments (beads and pendants), rock art and cemeteries. It would achieve higher levels of visibility in the archaeological record when there was accelerated demographic change, especially if there was an imbalance between local population and available resources (Barton et al. 1994), as would occur with population compression in refuge areas during the LGM (i.e., increased population pressure). Personal ornamentation can provide visual information about cultural identity  – a culturespecific code (Wadley 2001). It has been suggested that different bead types in different regions occurring at multiple sites over long periods of time represent integrated markers of individual 157

Phillip J. Habgood and Natalie R. Franklin and group identity (Vanhaeren 2005; d’Errico & Vanhaeren 2007; Kuhn & Stiner 2007). Personal ornaments (beads and pendants) from Late Pleistocene contexts within Sahul are all different and were used over a considerable period of time, which would be expected if they were being used as expressions of group identity. Examples include: • Dentalium sp. shell beads at Riwi and Carpenter’s Gap, in our Northern Zone, which date between 30,000 to possibly 42,000 years BP (Balme & Morse 2006; O’Connor 1995; Habgood & Franklin 2008, table 3) • Conus sp. beads at Mandu Mandu, in our Central Western Zone, which date between 21,000 to 32,000 years BP (Morse 1993a, 1993b; Habgood & Franklin 2008, table 3) • Beads made from macropod long bones at Devil’s Lair, in our Southwestern Zone, which date between 12,000 to 20,000 years BP (Dortch 1979, 1984; Habgood & Franklin 2008, table 3) • A perforated tiger shark tooth from Buang Merabak, New Ireland, in our Papua New Guinea Zone, which dates between 39,500 and 28,000 years BP (Leavesley 2007; Habgood & Franklin 2008, table 3) Cane (2001) argued that a fragment of abalone (Haliotis laegivata) from levels dating to 13,000– 14,000 years BP at Allens Cave on the Nullarbor Plain (in our Central Zone) may have been a pendant or other decorative item. Interestingly, this date coincides with the rapid flooding of the coastal plain, moving the coast from 130 km to 65 km from the site (Cane 2001). Wright (1971, 15) observed that “the constricting effects of this eustatic change must have had economic and social consequences”. These consequences appear to have included increased population pressure due to the loss of territory and resources, more closed social networks and territorial boundaries, and emblemic behaviour in the form of group-identifying personal ornamentation (abalone pendants). Other examples of personal ornaments have been recovered from late or terminal Pleistocene contexts within Sahul (note that the identification of some of them remains problematic): • A limestone pendant from a 14,000 BP horizon at Devil’s Lair (Dortch, 1984; Bednarik, 1998) • Fragments of Nautilus or pearl oyster shell (Pinctada sp.) and scaphopod shell (Dentaliidae sp.) from Late Pleistocene deposits at Mandu Mandu, which ethnographically are highly prized and traded commodities in northwestern Australia and are known to have been used as personal ornaments (Morse 1993a, 1993b; Balme & Morse 2006; O’Connor 2007a) • A headband of kangaroo incisors in a grave at the Kow Swamp cemetery (in our MurrayDarling Zone, Figure  12.1; Flood 1995), which dates to 9,000–14,000 if not 19,000– 22,000  years BP (see Habgood & Franklin 2008), possibly signifying ethnic, social or personal identity in a riverine refuge area • Three Money Cowry (Cypraeo moneta) shells with possible deliberate perforations from terminal Pleistocene deposits at Kafiavana, Papua New Guinea, which White (1972, 96) postulated may have been strung together as a necklace or sewn onto fabric The regional differences in bead types and use over an extended period argue for their role in group-identifying behaviour. The standardisation in form imposed by selection of particular natural media at sites (Dentalium shells, Conus shells, macropod bones, animal teeth) would reflect “established systems of expression” (d’Errico & Vanhaeren 2007). The different types of personal ornaments are found in contexts pre-dating and during the LGM (our Phases 2 and 3), when people would have been congregating in refuges at times of environmental stress. The impact of increasing aridity in the lead-up to the LGM is likely to have been felt much earlier in desert and semi-desert areas, including areas within our Northern, Central and Central Western Zones, where early examples of personal ornaments are found. 158

Framework for the Appearance of Modern Human Behaviour

Figure 12.6.  Simple figurative engraving of a “shark”, Sydney region (photograph by the authors).

For rock art, bounding behaviour manifests itself in the occurrence of distinct styles with spatially restricted distributions. On a continent-wide basis, Maynard (1979) labelled groups of sites with distinct, regionally varied styles of engravings and paintings the Simple Figurative (Figure  12.6) and Complex Figurative, and she claimed they were more variable and spatially restricted than the Panaramitee. A multivariate investigation confirmed that the Simple Figurative consisted of different styles with restricted distributions and a greater degree of variation than within the Panaramitee tradition (Franklin 1984; 1989, fig.  20.2). The Simple Figurative and Complex Figurative may reflect bounding behaviour as groups congregated in areas with high population pressure in the lead-up to and during the LGM, resulting in population packing and compression and culture contact. Bounding behaviour in rock art would be expressed differently in different regions and at different times according to regional or local conditions. Examples include Gwion Gwion paintings (Roberts et  al. 1997; O’Connor & Fankhauser 2001) and the ­possibly even earlier though undated Large Naturalistic Animal tradition in the Kimberley region (Flood 1997); Dynamic Figures in the Arnhem Land Plateau (Taçon & Chippindale 1994); distinctive Quinkan style anthropomorphic paintings (Morwood & Jung 1995; Cole & Watchman 2005) and indeterminate motifs identified in oxalate crusts in the Laura region (Cole et al. 1995); and geometric motifs (Campbell & Mardaga-Campbell 1993) and painting episodes identified in oxalate laminations in the Chillagoe region (Campbell et al. 1996; Watchman & Hatte 1996; Campbell 2000). Hand stencils at sites such as Wargata Mina, Ballawinne and Keyhole Cavern in southwest Tasmania may also be an example of emblemic behaviour and the marking of “Country” ( Jones et al. 1988; Cosgrove & Jones 1989; McGowan et al. 1993; Gunn 2007). This pattern of distinct rock art styles occupying more spatially restricted areas, plus the available chronological evidence (e.g., David 1991; Morwood & Hobbs 1995; Porch & Allen 1995), is consistent with the congregation of populations in the refuge areas during the period of 159

Phillip J. Habgood and Natalie R. Franklin environmental deterioration leading into and during the LGM, necessitating the need for boundary maintenance and emblemic behaviour. Importantly, stencils of pendants and figures wearing personal ornaments occur in rock art across northern Australia (Beaton & Walsh 1977;Walsh 1988; Chaloupka 1993; Welch 1996). Cemeteries can also be viewed as an indication of bounding behaviour. Pardoe (1988; 1995) argued that the appearance of cemeteries throughout the Murray River corridor following the LGM may reflect the need to legitimise control and ownership over restricted riverine resources in an area of high population pressure – more closed social networks and increasing territoriality in an area with escalating demand for resources. Furthermore, cranial modification as possibly practiced at Coobool Creek and Kow Swamp during the terminal Pleistocene may be another example of a system to enforce group identity, distinguishing one group from another in this region (Pardoe 1993). Evidence of emblemic behaviour therefore occurs in areas experiencing demographic stress, population packing and territorial contraction (high population pressure), and with increasingly crowded social environments. Conversely, in areas that did not experience these processes to the same extent (lower population pressure), emblemic behaviour would not be as evident.

Beyond Sahul: Views from Across the Ocean We contend that our demographic, social and symbolic framework would have a wider application than just Sahul and could be applied to Europe and Africa during the Middle and Upper Palaeolithic.

Europe Our framework has similar components to models proposed by Gamble (1982), Jochim (1983) and Barton et al. (1994) for the European Upper Palaeolithic. Gamble (1982) proposed that the distribution (across Europe) and dating (25,000–23,000 years BP) of Venus figurines are indicative of their function in more open social networks (bonding behaviour) at a time of climatic deterioration. By contrast, Jochim (1983) saw the development of Upper Palaeolithic cave art and its restricted distribution in southwestern Europe as an indication of relatively closed social networks (bounding behaviour) in a resource-rich refuge area during the LGM. Barton et al. (1994) integrated aspects of both of these models to examine the changing temporal and spatial distributions of mobiliary and parietal art across Europe from 30,000 to 7,000 years BP. They proposed (Barton et al. 1994, 191) that art, as a medium of inter- and intra-group information exchange, was under selective pressure from changes in demography in response to the major climatic fluctuations between 30 and 7 kyr BP. Like our model, Barton et al. (1994) see bonding behaviour and bounding behaviour (emblemic style) as monitors of demographic stress reflecting changes between more open and more closed social networks. We contend that our demographic, social and symbolic framework would have a wide application across the European Middle and Upper Palaeolithic and could be applied holistically to different classes of archaeological evidence, such as beads and cemeteries, and possibly stone artefacts, as well as mobile and parietal art. The model can also be useful in elucidating insights into the transition from the Middle to the Upper Palaeolithic. The Châtelperronian has been regarded as a heavily acculturated derivation of the Mousterian (see Harrold 1989; Mellars 1989a but see d’Errico et  al. 1998). Interestingly, some sites with Châtelperronian levels include examples of personal ornaments (e.g., pierced tooth pendants 160

Framework for the Appearance of Modern Human Behaviour and ivory rings), which may have provided visual information about the cultural identity of the Neanderthal groups at these sites and been used to mark cultural, social and territorial boundaries between Neanderthal and modern human populations.There are also potential ochre crayons from Châtelperronian levels at Grotte du Renne, Arcy-sur-Cure, France (d’Errico 2008), which may indicate that decoration was applied to the body or clothing – “clan designs” perhaps. We would explain this as an example of bounding and emblemic behaviour directly caused by culture contact between two very different groups, something to which Neanderthals had not been subjected previously. Lack of such contact may explain the paucity of personal ornaments within the Mousterian.Therefore, prior to competition for country with modern humans, the Neanderthals had a limited need to “mark territory” (cf. Taçon 1994) and use emblemic behaviour, such as personal decoration, as population pressure had not reached a threshold that stimulated such symbolic behaviour. Bead use is ubiquitous during the Upper Palaeolithic in Europe, with regional patterns of bead types inferring their use as “integrated markers of ethnic, social and personal identity” (Vanhaeren 2005; d’Errico & Vanhaeren 2007, but see White 2007). We suggest that the pervasive presence of personal ornaments and other symbolic behaviour during the Upper Palaeolithic reflects emblemic behaviour and the need for bounding of groups, more closed social networks and territorial boundary maintenance prompted by culture contact and increasing population pressure (migration, population increase, climatic and resource fluctuations).

Africa Integrated sociological, demographic and ecological explanations for changes in material culture have been proposed for sub-Saharan Africa during the Holocene (e.g., Brandt & Carder 1987; Barham 2002; Mazel 1989, 2009). However, can changing patterns of population pressure and bonding and bounding behaviour help explain apparently stochastic patterns for the appearance of symbolic behaviour in sub-Saharan Africa during the Middle Stone Age (MSA)? Barham and Mitchell (2008) detailed the significant environmental, demographic and archaeological patterns or changes in Africa during what they refer to as “The Big Dry” (Marine Isotope Stages 4–3), when large areas were abandoned while others became refuges. Personal ornaments are evident at sites in southern, eastern and northern Africa during this “Big Dry” (McBrearty & Brooks 2000; Barham & Mitchell 2008). Notational pieces and rock art also appear, and there is evidence of long-distance movement of stone (McBrearty & Brooks 2000; Barham & Mitchell 2008). McBrearty and Brooks (2000, 497) suggested: The diversification of MSA toolkits and the varying proportions of different artefact classes at different sites no doubt reflect regional traditions as well as different extractive activities. One of these regional traditions, the Lupemban, has a restricted geographical distribution in central Africa. Barham (2007) proposed that the distinct Lupemban bifacial lanceolates could be stylistic markers of social identity and signal the formation of coalitions. Barham (2002, 171) argued: The makers of the Lupemban with their distinctive points and pigment use may be a prime example of a regionally distinct response to climatically induced biogeographical changes. Using our model, the Lupemban lanceolates would be bonding groups together within the drainage pattern of the Congo River (Barham 2002, fig.  14.3), and bounding the Lupemban group(s) from groups utilising different regional stone artefact traditions. The “precocious” Still Bay and Howiesons Poort industries, which have restricted geographical and chronological distributions in southern Africa, have been a focus of discussion. The Still 161

Phillip J. Habgood and Natalie R. Franklin Bay industry is characterised by bifacially worked lanceolates, often made on fine-grained raw material, ochre use, beads and bone tools (Henshilwood 2008). Dated to ~71,000 years BP, it coincides with deteriorating environmental conditions (Henshilwood 2008; Jacobs & Roberts 2008). The Still Bay lanceolates and bone points may have been “exchange items used to promote social relations” (Henshilwood 2008, 38). The Howiesons Poort is characterised by the manufacture of standardised backed artefacts with an increased use of fine-grained raw material, the use of ochre and the appearance of incised ostrich egg shell (Henshilwood 2008). It has been proposed that the specific character of the Howiesons Poort stone artefacts was due to their manufacture by a new incoming ethnic group, “the Howiesons Poort people” (Singer & Wymer 1982); they were style-laden artefacts used as gifts in formal reciprocal exchange networks (Deacon 1989; Deacon & Deacon 1999); or they represent a technological risk minimisation strategy in response to changing environmental conditions and unpredictable resource availability (Deacon 1989; Ambrose & Lorenz 1990). Deacon (1989, 560)  observed, “The marking of boundaries and the intensification of social networks would explain the novelty of the Howiesons Poort industry”, and he suggested its demise may reflect reduced cultural selection for symbolic behaviour due to a “release of stress through amelioration of conditions”. Our social and demographic model would infer that the appearance and subsequent disappearance of the Still Bay and Howiesons Poort industries would relate to sociological responses to changes in population pressure caused by demographic and environmental variations. It would be tempting to suggest that the Howiesons Poort, with its style-laden artefacts and wide distribution in coastal southern Africa reflected bonding behaviour represented by reciprocal exchange of stone tools during a period of marked climatic variation. By contrast, the Still Bay, with its more limited distribution, lanceolates, engraved bone and ochre, use of personal ornaments and possibly “clan designs” drawn on the body or clothing as suggested by ochre crayons, indicated bounding and emblemic behaviour during an arid phase. Barham (2002, 170–171) hypothesised that variations in population pressure may explain the “intermittent flashes of innovation, and symbolic expression” during the Middle Pleistocene, especially in “arid phase refuges” such as the continental margins, equatorial interior and riverine and lacustrine areas – changing patterns of population pressure and associated bonding and bounding behaviour. Whilst much of the southern Cape was probably a refuge area during arid phases, Blombos, with its Still Bay industry, engraved bone and ochre and shell beads (d’Errico et al. 2001; Henshilwood 2007), appears to document evidence of strong bounding and emblemic behaviour at a time of severe climatic deterioration. It is also interesting that ostrich eggshell beads become more common in sub-Saharan Africa from ~40,000 years BP when the climate started to deteriorate leading to the LGM. Vanhaeren (2005) argued that ostrich eggshell beads were exchange items in gift-giving systems used to strengthen social networks – bonding behaviour during a period of climatic deterioration.

Conclusion There is both chronological and geographical patterning to the occurrence of symbolic behaviour in late Pleistocene Sahul that can be explained within a demographic, social and symbolic framework. Within this framework, socio-cultural and behavioural processes, demographic changes and climatic and environmental factors are inter-related. Sahul experienced significant climatic, ecological and demographic change throughout the late Pleistocene resulting in increasing population pressure from reduced resources and population compression and concentration in refuge areas. These processes necessitated socio-cultural responses involving the realignment of social networks and territorial boundaries reflecting what we refer to as bonding and bounding (emblemic) behaviour. These socio-cultural responses 162

Framework for the Appearance of Modern Human Behaviour resulted in changes in symbolic and other behaviours as manifested in the appearance of art, personal ornaments and cemeteries. This bonding and bounding behaviour became an integral component of the socio-cultural solutions adopted by Aboriginal people to manage irregular aridity, population compression and packing and resource control within Sahul. The proposed demographic, social and symbolic framework is not a monolithic continent-wide explanation but rather is based on local or regional demographic and social circumstances, and so the timing and pattern can be (and is) different in different parts of the continent. We also propose that a demographic, social and symbolic framework involving bonding and bounding behaviour can be useful in exploring or explaining the appearance of symbolic behaviour in sub-Saharan Africa during the MSA and the Middle to Upper Palaeolithic transition in Europe.

163

Chapter 13 Early Modern Humans in Island Southeast Asia and Sahul Adaptive and Creative Societies with Simple Lithic Industries

Jane Balme and Sue O’Connor

Introduction There is some debate about the timing of the first occupation of Sahul (Australia, New Guinea and the Aru Islands); however most would agree that the continent was colonised before 45 ka (Bowler et al. 2003) and perhaps as early as 60 ka (Roberts et al. 1994; Gillespie 2002;Veth et al. 2009). People were occupying inland high altitude sites in New Guinea by 45 ka and had reached the far southwest of Sahul (see Summerhayes et al. 2010 for mainland Papua New Guinea (PNG) and Turney et al. 2001a for southwest Australia) before 40,000 years ago. By 35 ka humans had successfully spread into the southern extremes of Tasmania (Cosgrove 1999) and even the very small islands of Southeast Asia (such as the Talauds), and also colonised the arid inland regions of continental Sahul (Figure13.1). However, the tool kits associated with this evidence for rapid adaptation to very different environments continue to be remarked upon as simple and unchanging in comparison to tool kits of similar antiquity in the Old World. For example, Klein (2009, 716–717) describes Australian stone tools as a loosely defined Core-Tool-and-Scraper Tradition that persisted basically unchanged until roughly 4 ka. Similar artifacts occur widely in southeast Asia in late Pleistocene and early Holocene deposits, and where they are found alone, the behavioural modernity of the makers can be questioned. However, at several Australian sites the flaked stones are accompanied by such advanced behavioural markers as formal bone artifacts. The implication here is that before the mid-Holocene, Australian stone artefacts were so simple that, were it not for the presence of other kinds of archaeological materials, they could be interpreted as not being the work of modern humans. This view of an unchanging ‘primitive’ undifferentiated stone technology (see also White 1977) lies deep in the history of archaeological research in Australia. Here we review the history of ideas relating to the tools used by the early colonists in this southern region and discuss the ways in which these ideas are now being overturned by an increasing recognition of the variety of stone artefacts, the role of organic tools and ‘intangible technology’. 164

Early Modern Humans in Island Southeast Asia and Sahul

Figure 13.1.  Map of Sahul and Island Southeast Asia showing sites and locations mentioned in the text (prepared by Kay Dancey).

Historical Views of Tool Use by the First Colonisers of Sahul and Island Southeast Asia The idea that change might have occurred over time in pre-European Australia at all is relatively recent. In the 19th century, change was generally thought of as a single trajectory using the European record as a benchmark. The historical context of the 19th century suggests that 165

Jane Balme and Sue O’Connor Australian Aboriginal society was in some kind of evolutionary stasis, as discussed in some detail by Bowdler elsewhere in this volume. Generally this was thought to be the Mousterian, with Sollas (1911, 161–162) referring to them as the “Mousterians of the Antipodes”, whereas the Tasmanians, whom he identified as physically different and with fewer items of material culture, were referred to as a more “ancient Mousterian race”. This European view of Aboriginal stasis and stagnation is summarized by the well-known comment of Pulleine’s (1929, 310) about an “unchanging people in an unchanging environment”. Soon after Pulleine’s remark, Hale and Tindale (1930) began excavations at Devon Downs in South Australia, the first systematic excavation in Australia in which the importance of stratigraphy was recognized. The resulting analysis of the artefacts from this and other sites, including surface sites on Kangaroo Island off South Australia, allowed Tindale to produce the first artefact sequence for Australia. He defined five cultures based on changes in artefact types from which he extrapolated an Australia-wide scheme. As more work was done, regional variations were identified with this scheme, and radiocarbon dates became available to provide an absolute chronology (see Bowdler 1993 and in this volume for a summary). In the 1960s it was resolved that the only pan-Australian change in the archaeological record occurred in the mid-Holocene. Initially the Holocene stone tool change was suggested to be the result of the introduction of hafting to make complex tools  – that is, a pre-hafting phase and a hafting phase (Mulvaney & Joyce 1965). However, the finds in 1967 of edge-ground axes, presumably hafted and dating to the Pleistocene (Schrire 1982), led to some discussion about the interpretation of the change. After some debate (Glover & Lampert 1969; Gould 1969), the term ‘Australian Small Tool Tradition’ (ASTT) was proposed by Gould (1969) for the post-mid-Holocene industry, and this term continues to be used. The components of this industry vary regionally but include backed blades, points and adzes.The ASTT does not occur in Tasmania.The stone artefact industry associated with pre-mid-Holocene occupation was named the ‘Core Tool and Scraper Tradition’ (CTST) (Bowler et al. 1970). It had rather loose defining characteristics but was said to contain large horse-hoof-shaped cores, amorphous scrapers and pebble choppers, with the addition of edge-ground axes in the north. In general it was believed that the earliest tools of the CTST were made on large heavy cores and large thick flakes retouched for use as scrapers, and that over time there was a progressive reduction in the size of tools with smaller cores, fewer core tools and scrapers increasingly made on thinner flakes (Lorblanchet & Jones 1979). However, more excavations throughout the different regions of Australia and additional radiocarbon dating revealed flaws in this evolutionary progression model of stone technology. In creating these Australian schemes, researchers no longer used Europe for chronological comparison; instead, they turned to Asia for antecedents to the Core Tool and Scraper Tradition, in particular the Hoabinhian of Malaysia and Vietnam – although less so once the comparatively recent age of the Hoabinhian sites were determined. Again a detailed summary of these comparisons can be found in Bowdler (1992, 133; also this volume). Apart from the influence of social evolution, assumptions that stone tools have become more complicated over time have influenced interpretations of the ‘primitiveness’ of Australian stone technology. Clark (1969) divided world stone tool technology into five technological modes. Mode 1 comprises flake industries and chopping tools; Mode 2, bifaces and bifacially worked handaxes; Mode 3, the prepared core technology; Mode 4, blade production and prismatic core reduction associated with the first anatomically modern humans in Europe; and Mode 5, microliths found in Late Stone Age industries of Africa and late Pleistocene or early Holocene of Europe and Asia. Because each succeeding mode is more complicated than the previous, the progression suggests increasing cognitive capacity. While Clark (1969) did not comment on Australian technology, he had previously remarked upon the Australian stone tool industry as being ‘crude’ (Clark 1968, 21). In Clark’s schema, Australian Pleistocene tool technology represents Mode 3 (Foley & Lahr 2003), a technology usually associated with Homo neanderthalensis in Europe and Homo heidelbergensis and H. sapiens in Africa. Later modes appear in the mainland Australian record only at the 166

Early Modern Humans in Island Southeast Asia and Sahul Pleistocene-Holocene boundary (Hiscock & Attenbrow 1998; Hiscock 2002; Slack et al. 2004) and are not widespread until the mid Holocene (Hiscock 2008, 156–158). Tasmania, lacking these stone tools, remained in Mode 3 up to the ethnographic present. On many of the Southeast Asian islands, simple flake-based industries lacking blades or regular patterned retouched forms occur throughout prehistory up to, and in some case into, the Metal Age (O’Connor 2007b; O’Connor & Bulbeck in press) and can be characterized as Mode 3 (O’Connor et  al. 2011a, online supplement). There has been criticism (see, e.g., Gamble 2007) of the use of Clark’s modes as a measure of cognitive ability, in particular because stone artefacts are only a small part of overall technology. Less visible evidence of technology, including items made from organic remains such as bags, nets, cordage, hooks, spears and digging sticks almost certainly had just as important a role. Foley and Lahr (2003, 117) recognize this problem but argue that, although modes do not tell the whole story, as each location has its own private history, they do tell an important one. The lack of Modes 4 and 5 in distant places such as Australasia could reflect a lack of contact with groups rather than the lack of cognitive capacity (see, e.g., Foley & Lahr 1997), and it may also reflect a lack of need. The implicit suggestion that Old World technology is more complex than that of Island Southeast Asia (ISEA) in much of this work is more directly stated by Mellars (2006b) as part of his explanation for the lack of blade technology in Australia – namely, that the repeated, successive and cumulative small-scale founding population groups associated with dispersal from Africa would lead to a progressive loss in the complexity and diversity of technology (Mellars 2006b, 799). Even if not explicitly stated, all of these interpretations seem to deny the ability to invent stone tool types beyond the Old World, to use inventiveness to adjust to a new environment or to recognize the complexity within Australian technology. In Australia, evidence for ground stone tools from some of the earliest sites may well be just as inventive a response to environmental circumstance as the blades of Mode 4 technologies in the Old World. That they are entirely absent in the islands to the north of Australia at this date would seem to be proof of their in-situ invention in northern Australia. In ISEA, edge grinding on stone tools first occurs in Neolithic and Metal Age assemblages when small fully polished adzes appear. Even then such tools are not found in many of the eastern Indonesian islands (e.g., O’Connor et al. 2005). New approaches to lithic analysis in Australia rely less on typology and discussions of modes and instead examine the reduction continuum through tool manufacturing, re-sharpening and discard to show that artefact shapes and sizes are largely a product of the “intensity with which tools were transported, used, maintained and recycled” (Hiscock 2008, 108). Thus, in areas where raw material is abundant, it may be used to make expedient tools that are discarded after a short use life. Conversely, in regions where raw material is unavailable or scarce, implements are more likely to be conserved by re-sharpening and re-shaping, and will have more retouch (Hiscock & Allen 2000; Hiscock 2006; Clarkson & O’Connor 2006). These approaches avoid the issues of cultural capacity, variety, aesthetics and inventiveness raised by the Old World approaches. What then can be said about these issues from the Sahul and ISEA evidence and what does the ‘private history’ of technology in this region suggest about environmental adaptation?

Sahul In Sahul, the longest continuous Pleistocene sequence is at Devil’s Lair in southwest Australia. Occupation of this limestone cave site began about 45.5 ka BP (~48 ka cal BP) (Turney et al. 2001a, 11). According to nearby pollen records and fauna and charcoal evidence from the cave, it was situated in open forest woodland for most of the Late Pleistocene, changing to tall open forest at the end of the Pleistocene (see summary of evidence in Dortch & Wright 2010, 1055). Little plant evidence apart from charcoal is present in the cave, but thousands of bone fragments representing nearly 6,000 individuals (Balme et al. 1978, table 3) are present and well preserved. The human contribution to the faunal remains is difficult to disentangle from those of other 167

Jane Balme and Sue O’Connor predators. However, evidence of charring (Balme 1980) and analyses of associations with other archaeological evidence (Dortch et  al. 2012) provide some conclusions about subsistence. The three most common species hunted are the southern bandicoot (Isoodon obesulus), a bettong (Bettongia penicillata) and a kangaroo (Macropus fuliginosus); the first two are small (less than 1.5 kg), slow, and relatively easy to capture as they are ground dwelling and hide in nests, whereas the third, a kangaroo, is large (adults weigh more than 30 kg) and fast, which implies planned rather than opportunistic foraging (Dortch et al. 2012). Only 1,182 stone artefacts were recovered from the main excavation trench at Devil’s Lair. These can best be described as a small flake-based assemblage (Dortch 2004, 95), of which about 2% are retouched (Dortch 2004, 94). Scrapers and notched flakes are the only artefact types identified. Bone artefacts, especially ground points made on macropod fibulae, are also present, although no analysis of their possible function has been published. Pleistocene occupation of nearby Tunnel Cave from about 22 ka BP (~27 ka cal BP) yielded similar results in terms of both fauna and artefact representation (Dortch 2004; Dortch & Wright 2010; Dortch et al. 2012). Although the small fauna represented in these sites may have been gathered from nests by hand, the technology associated with capture of the large kangaroos is not evident. Ethnographically, these kangaroos were speared and, in the absence of stone or bone tips, Pleistocene spears must have been wood. Pits and traps made of fibre, such as those recorded in Anell (1960) and Satterthwaite (1986), are other possibilities. Some of the most detailed analyses of fauna from Pleistocene Sahul derive from a series of limestone cave sites in southwest Tasmania containing evidence for sub-arctic occupation dating from about 35,000 years ago (Cosgrove 1999; Pike-Tay et al. 2008). This region is mountainous and cold, but occupation of the caves appears to have ceased about 12 ka as the climate ameliorated. In all of these Pleistocene sites the macropod Bennett’s wallaby (Macropus rufogriesus) by far dominates the fauna represented (Pike-Tay & Cosgrove 2002). These wallabies are medium sized – ranging from 11 to 15.5 kg for females and from 15 to 26.8 kg for males (Calaby 1983, 239)  – and are grazers. On the basis of age estimates of the animals represented in the site, Cosgrove and Pike-Tay (2004) suggest that they were exploited both seasonally between autumn and spring and sometimes through encounter hunting outside this season. The stone artefact assemblages from these Tasmanian sites are rich and have been described as ‘amorphous’ and variable (Cosgrove et al. 1990, 70), as might be expected from differences in raw material availability between the sites. One distinctive type, the ‘thumbnail scraper’, is present from about 29,000 years ago (~33 ka cal BP), but despite their small size, there is no evidence of hafting of these artefacts (Cosgrove 1999, 374). Other than charcoal, plant remains are not well preserved in the sites, but bone points primarily made on wallaby fibulae are common. The function of these is uncertain, although, because of the presence of tip damage and use wear, Webb and Allen (1990) suggest that some may have served as spear points. Cosgrove (1999, 382), however, points out the lack of evidence of hafting and suggests that this and the lack of correlation with prey species throws doubt on this interpretation. Thus, as in southwest Australia, the wallabies may have been caught with wooden implements or other perishable technology such as nets (Balme 2013). Recently, Gilligan (2010a, 45) has argued that “the focus on manufacturing scraper tools suitable for preparing hides, the advent of bone points or awls for piercing the hides and the targeted hunting of the major local furbearing species” all suggest that simple clothing was being manufactured out of wallaby skins at these upland Tasmanian sites.This interpretation is consistent with earlier studies of wallaby bone elements present in the sites that indicate deliberate removal of the wallaby skins (Cosgrove & Allen 2001) and dental growth patterns of the wallaby mandibles from Warreen Cave that indicate occupation of the cave during autumn and early spring – the coldest part of the year (Pike-Tay & Cosgrove 2002, 138), when fur clothing would be most needed. This would also explain the lack of evidence for hafting on the bone points. 168

Early Modern Humans in Island Southeast Asia and Sahul Other regional evidence for the early period of colonization in Sahul comes from semi-arid western New South Wales.The antiquity of sites around the Willandra Lakes in this area has been controversial with claims of up to 60 ka BP for the WHL3 burial (Thorne et al. 1999) at Lake Mungo. A more conservative and widely accepted optical date for this burial is about 40 ka BP, with a date for human occupation of the area between 50 and 40 ka (Bowler et al. 2003). A surface collection on a lunette sand dune at Lake Mungo, thought to be the same age as the first burial, revealed a cremation (WHL 1). At the time, the age was considered to be between 32 and 25 ka (Bowler at al. 1970, 57) (~37.6 ka to ~28 ka cal BP), although now its age is considered to be about the same as WHL 3 (Bowler et al. 2003). Of the 200 artefacts collected, 27 were said to be in situ and used to define the Australian Core Tool and Scraper Tradition (ACTST) (Bowler et al. 1970). Excavations carried out by Shawcross in sediments below WHL3 recovered hundreds of artefacts, and although no detailed analysis of these is available, Shawcross (1998, 196) states that the excavated industry is not typical of the ACTST. Instead it has more in common with worked material from quarries than with woodworking, which we take to mean that it is mainly a flakebased assemblage. Allen (1998) and Hiscock and Allen (2000) have analysed artefacts collected by Allen in this region between 1969 and 1972 and conclude that there is considerable variation in Pleistocene assemblages across the landscape that can be attributed to different proximity to stone raw material sources and other variation in the landscape. Not surprisingly, the fauna associated with Pleistocene sites surrounding these lakes consists of lake fauna including shellfish, fish and freshwater crayfish together with small numbers of mainly small to medium mammals (Allen 1998, 211–214). The high incidence of scrapers in the retouched artefact assemblage suggests that wood may play an important part in the technology of early sites in the region (Allen 1998). Fibre may also have been an important part of the technology associated with early exploitation of these lakes. The use of nets has been suggested to explain the abundance of fish of uniform size within single-use middens dated to 27 ka BP (~ 31.5 ka cal BP) sites in the nearby Darling River lakes (Balme 1995), and Kefous (1977) has suggested that the restricted size of fish found in Willandra Lakes middens may also result from net capture. The earliest evidence for occupation of the central arid region of Australia is at Allen’s Cave on the Nullarbor Plain for which there is an optically stimulated luminescence (OSL) date of about 40 ka from sediments above the lowest excavated artefacts (Roberts et al. 1996). However, Puritjarra rock shelter has a more or less continuous sequence dated with both radiocarbon and optical techniques from before 35 ka (Smith et al. 1997). This site is in a small rocky range in arid sand hill spinifex country (Smith 2006), and during the arid last glacial maximum (LGM) there would have been very little grass cover (Smith 2006, 372). The small numbers (303) of artefacts recovered from the deposits dated between about 18 and 35 ka suggests only small intermittent group visitation during this time (Smith 2006). Artefacts from these deposits are mostly not of local raw material and consist of three heavily reduced cores and flakes. The flakes are described as small (except for two large flakes, the mean length is ca. 22 ±10 mm) and narrow with narrow platforms (Smith 2006, 383–384). Retouched tools consist of scrapers – both steep-edged and convex – amorphous retouched tools, and notched pieces that Smith (2006, 390) interprets as hand-held woodworking tools. Unfortunately, the layers with early occupation at Puritjarra are acidic, and no organic remains have been preserved (Smith 1989, 97). However the fact that all the identified artefacts with use wear or retouch appear to be associated with woodworking suggests the importance of wood in arid zone technology from the earliest times. In the Cape Range area on the western margin of Western Australia’s arid zone, several sites, including Jansz, Mandu Mandu, C99 and Pilgonaman, date from about 35 ka BP (~40.6 ka cal BP) (Morse 1999; Przywolnik 2005). These sites provide rare evidence of coastal life during the Pleistocene, as here the coastal shelf falls rapidly away so that even at times of lowest sea level the sites were within 12 km of the shore (Morse 1988, 81). Stone artefacts are not abundant in the 169

Jane Balme and Sue O’Connor earliest levels of these sites and represent a flake assemblage with little retouch evident. In Mandu Mandu, for example Morse (1988, 43) describes artefacts dated to between 35 and 20 ka BP (see Morse 1993b for earlier dates than those published in 1988) (~39.6-~ 25 ka cal BP) as “large flakes and flaked pieces”, of which merely five are retouched. Faunal remains from the lowest levels of these sites are poorly preserved and fragmented marine molluscs, crab, sea urchin, wallabies and small arid zone marsupials (Morse1988, 1999; Przywolnik 2005, 190). Sites across monsoonal tropical northern Australia date from more than 40 ka and include Carpenters Gap 1 (O’Connor 1995) and Riwi in the south central Kimberley (Balme 2000), Nawarla Gabarnmang in southwestern Arnhem Land (Geneste et  al. 2010), Malakunanja II and Nauwalabila in northern Arnhem Land (Roberts et al. 1990; 1994) and GRE8 in northern Queensland (Slack et al. 2004). Occupation of the earliest levels in these sites is infrequent and the stone artefacts are predominantly small flakes with very few showing signs of retouch. Faunal remains in the deepest parts of the deposits are sparse and little has been published on their identification. However, Slack et al. (2004, 134) refer to a dense lens of freshwater mussels at GRE8 between dates of about 40 and 32 ka cal BP, and David (1993, 53) suggests that at Ngarrabullgan the faunal remains represent small to medium-sized animals. Many of the Pleistocene sites in northern Australia contain evidence for ground stone artefacts. The oldest of these comes from Nawarla Gabarnmang and dates to 35,400 ± 410 cal BP. This is a fragment (about 2.5 x 4.0 cm, Geneste et al. 2010, 67), which, because of the convexity of the ground surface, is interpreted to have been flaked from a ground-edge axe to thin the side of the axe. Few complete axes have been found dating to such an early stage of antiquity, but a quartzite waisted and grooved ground-edge axe was recovered from deposits dated to 32,000 BP from Sandy Creek (Morwood & Trezise 1989). Complete axes have also been recovered from Malangangerr in Arnhem Land in deposits dating from about 25,000 years ago (~29 ka cal BP) (Schrire 1982). These axes are bifacially flaked and then ground to create a thin edge (Schrire 1982, 106). Although they are weathered, some axes retain manufactured grooves that have been interpreted as created to facilitate hafting (Schrire 1982, 107, fig. 27d, 241). Ground flakes have also been recovered from Widgingarri 1 rock shelter from deposits dating 28 ka (~33 ka cal BP) (O’Connor 1999, 75). Some dolerite flakes and objects found in levels dated between 30 and 25 ka at Nauwalabila I are very weathered (Jones and Johnson 1985, 217) but are suspected to be ground-edge axe flakes (Jones 1985, 297). Unlike the early Australian axes, those from PNG are flaked rather than ground or polished. They often have a central indent or ‘waist’ indicating their probable use in hafts. Commonly referred to as ‘waisted axes’, they are found at mainland highland and coastal locations. Several open sites at ~2000 m near Kosipe have recently produced examples of these tools dated to between 42 and 35 ka BP (~45–39 ka cal BP) (Summerhayes et al. 2010). The Kosipe sites overlook a large swamp and are remarkable in that they contain well-preserved evidence for the exploitation of plant foods such as charred Pandanus drupes as well as starch from a yam, which is compatible with Dioscorea (Summerhayes et al. 2010, 79). Like the much larger flaked and waisted axes at the Bobongara site on the Huon Peninsula, the Kosipe axes are thought to have been used to cut down trees and, in combination with firing, to create open patches in the forest to promote the growth of useful plants (Groube 1989; Groube et al. 1986; Summerhayes et al. 2010, 78).While the waisted axes from the Huon Peninsula are massive, the examples from Kosipe occur in a range of sizes and raw materials. The PNG axes have been argued to have been predominantly used for forest clearance, but this probably involved a diverse range of tasks such as thinning smaller saplings, slashing undergrowth and ring-barking large trees.The smaller examples found near Kosipe may have been used for cutting timber to make wooden tools – such as the axe hafts. There have also been claims that seed-grinding stones may have been part of the stone tool technology associated with Pleistocene incursions into semi-arid grassland regions of Australia. The main evidence for this is the grinding stones excavated from Cuddie Springs, recovered from a level older than 35 ka and containing residues of plant use (Fullagar & Field 1997). However 170

Early Modern Humans in Island Southeast Asia and Sahul Grün et al.’s (2010) recent electron spin resonance (ESR) and uranium (U) series dating of fauna from this site suggests that there has been considerable mixing at the site and that the claim for Pleistocene grinding stones will need to be verified by further finds in well dated contexts.

The Southeast Asian Islands Although the early colonists must logically have passed through the islands of Southeast Asia to reach Sahul, only one site, Callao Cave in the Philippines (see Pawlik et al., this volume), is presently known which is as old or older than the archaeology of Sahul. Other early sites in Island Southeast Asia date to between ~ 42 and 30 ka BP, although this may be due to fact that most have been dated by only the radiocarbon technique (O’Connor 2007b). The earliest level at Callao Cave in northern Luzon, has a U-series age estimate of 67 ka on a single hominin foot bone, thought to be from a modern human. Interestingly, there are no stone artefacts whatsoever found in this level (Mijares et al. 2010), and Pawlik et al. (this volume) have raised the possibility that a technology based entirely on bamboo, bone and antler was used in place of stone in the immediate post-colonisation phase in the Philippines. Cut marks identified on animal bones from the earliest levels are consistent with those produced in experimental studies using knives made of bamboo (Mijares personal communication 2011). The oldest stone artefacts at Callao Cave are dated to ~ 30 ka cal BP. The artefacts from this level are predominantly made of chert and produced by simple direct percussion techniques (Mijares 2008, 103). The assemblage is said to contain a few “blade-like flakes” but the small size of the assemblage (25 artefacts) renders this assessment premature. This simple stone-working technology appears to characterise most Pleistocene ISEA assemblages and to continue through to at least the midto late Holocene (Mijares 2008, 103; Patole-Edoumba 2009; O’Connor & Bulbeck in press). At Jerimalai shelter and Lene Hara Cave in East Timor, the earliest occupation levels date between 43 and 30 ka cal BP. Both sites have stone artefacts that are predominantly made on chert. Jerimalai and Lene Hara also have excellently preserved faunal remains and artefacts made of shell and bone. The stone artefacts from the earliest layer at Jerimalai are mostly cores, flakes and retouched flakes produced by direct percussion, although bipolar and anvil rested core reduction is also evident. Cores include discoidal and rotated forms, as well as cores resulting from informal single and multiplatform reduction. The cores are highly reduced, although the size of available nodules appears to have been limited to small-sized pieces. The Pleistocene assemblage includes some retouched flakes and what may be scrapers (O’Connor et al. 2011a, figs. S5, S6, table S1). The scraper assemblage shows no recurring formal types. While the lithic assemblage from Jerimalai and Lene Hara appears simple, the organic technology shows that the people living at the sites were skilled in the acquisition of marine resources. They caught fast-moving pelagic fish such as tuna, as well as sharks, rays and reef fish. In addition to fish, they were eating a broad range of marine foods including turtle, shellfish, crabs and urchins. No direct evidence of fishing technology has been found in the earliest occupation levels of Jerimalai or other Pleistocene East Timor sites, so it is uncertain how the occupants captured the pelagic and other fish species. The northeast coast of Timor has a steep offshore profile and drops off rapidly into deep water. The small size of the bones of pelagic species (individuals of ca. 60–50 cm in length) indicates that mostly immature individuals were taken. As immature tuna sometimes come in close to shore, it is possible that they were caught by hook or spear from boats or even from the shore or reef edge. Simple fish aggregating devices such as tethered log floats or rafts can be used to attract tuna, which can then be caught using purse seine or leader nets (O’Connor et al. 2011a). Whatever method was used the high proportion of pelagic species in the 42,000- to 38,000-year level at Jerimalai indicates that their capture was planned rather than fortuitous. The lower levels of Jerimalai and nearby Lene Hara Cave also contain a range of reef fish such as parrotfish, unicornfish, trevallies, triggerfish, snapper, emperors and groupers (O’Connor et al. 171

Jane Balme and Sue O’Connor 2011a). Zoo-archaeological and ethno-archaeological studies in Southeast Asia and Oceania suggest the parrotfish and unicornfish would have likely been caught by netting or spearing, while trevallies, triggerfish, snappers, emperors and groupers are most commonly captured by angling using a baited hook, but netting and trapping can be successful depending on species and body size (O’Connor et al. 2011a). The earliest direct evidence of fishing technology in the East Timor sites is the tip of a fish hook made of shell. It was found in Jerimalai in levels dated between ~23 and 16 ka (O’Connor et al. 2011a). As it is incomplete, determining the type of species it was designed to target is not possible, but the presence of this hook indicates that strong fibre twine was also made at this time. The early levels of Jerimalai and Lene Hara contain a significant quantity of marine turtle bone. The turtle bones are mostly from the flippers indicating that butchery was multiple-staged, with initial butchery probably taking place at the shore. During early occupation between 42 and 35 ka cal BP, the sites would have been on a steep rocky limestone shoreline with a progressively falling sea level so there would have been a high cost involved in transporting such large animals whole. At the initial processing site the turtle would be butchered and fletches of meat removed. Only the flippers, presumably with meat attached, were transported to the shelters. By 35 ka cal BP, some of the smallest and most remote islands of ISEA, such as the TalaudSangihe group and the smaller islands in eastern Indonesia such as Gebe Island, Halmahera, had been settled. The Talaud-Sangihe Archipelago is a group of 77 islands lying between Mindanao and North Sulawesi. Of these, 56 are occupied today by peoples with an agricultural subsistence base. In the absence of agriculture, permanent settlement of these small islands would seem to be unsustainable, and yet the rock shelter Leang Sarru on Talaud was first occupied between 35 and 30 ka cal BP (Ono et al. 2009). The fauna in the Pleistocene-aged levels at Leang Sarru consists solely of marine shellfish, predominantly Neritidae, Turbinidae and Trochidae, and a few urchin remains (Ono et al. 2009). While it is extremely improbable that permanent human settlement could have been sustained on shellfish alone, no higher-ranked marine resources such as fish and turtle, terrestrial species or birds, or macrobotanic remains have been recovered. In view of the limited fauna, the stone artefact assemblage from Leang Sarru is large, with more than 14,000 artefacts recovered from the combined excavations (2,689 per m3) of Ono et al. (2009, 325) and Tanudirjo (2001). No publications are yet available that provide a quantitative breakdown of the stone artefacts by excavation unit or occupation phase or detailed descriptions of the assemblage, so it is uncertain what proportion of these belong to the 35 ka to 30 ka levels. The available brief description of the assemblage indicates that as at Jerimalai and Lene Hara, the artefacts are made on small chert cobbles that were locally procured from watercourses and beaches. The cores and flakes are small, with the average mean size of flakes being 2.5 mm x 10 mm. Retouched flakes constitute a low proportion of the assemblage and appear to have been expediently produced (Ono et al. 2009, 325). The reason for the knapping of stone flakes at Leang Sarru is a mystery in view of the fact that the shelter contained no food remains other than shell fish. There is a gap in occupation at Leang Sarru between 30 ka cal BP and circa 22 ka cal BP. The site witnesses its most intensive period of occupation during the LGM between 22 and 18 ka cal BP (Ono et al. 2009). At this time of lowered sea levels, the small islands of the Talaud-Sangihe group were joined to form one mega-island. After ~18 ka cal BP Leang Sarru has no occupation record until the Holocene. Whether this episodic pattern in the dates equates with discrete pulses of settlement, followed by population abandonment or extinction, cannot be ascertained on the basis of the current available data. It is even possible that the disrupted settlement history is apparent rather than real, resulting from limited sampling and dating (O’Connor et al. 2010). Regardless of whether occupation of these tiny islands was episodic or continuous it must have required an extraordinary level of skill in both seafaring and marine resource exploitation. In contrast with Leang Sarru, the Pleistocene artefactual assemblages from Leang Burung 2, South Sulawesi and Golo Cave, Gebe Island Halmahera, are remarkably sparse. While the pre30 ka levels of Leang Burung 2 contained some small to medium-sized mammal fauna such as 172

Early Modern Humans in Island Southeast Asia and Sahul the pig deer Anoa sp., only 49 stone artefacts were found, suggesting very occasional cave use (O’Connor & Bulbeck in press) or, alternatively, as at Callao Cave, a technological suite composed of organic materials that have not been preserved. Golo Cave, first occupied about ~35 ka, contains only 51 stone tools dating between ~35 and 13 ka BP (Bellwood et al. 1998; Szabó et al. 2007, 703, 708). These are made on a variety of metavolcanic rocks produced with the “primary goal of knapping . . . simple, sharp-edged flakes, most likely for use as unretouched cutting/scraping tools” (Szabó et al. 2007, 710). No retouched stone artefacts were found in the Pleistocene layers (Szabó et al. 2007, 709). As well as the sparse stone artefact assemblage, the oldest Pleistocene deposits at Golo Cave contained operculums from large Turbo marmoratus shells that were unifacially knapped to produce a steeply angled edge with a form reminiscent of a steep-edged scraper. Some of the removed flakes may have also been utilised (Szabó et al. 2007, 707). Similar worked operculums have been identified at Jerimalai shelter in the Pleistocene levels. It seems likely that such shell tools are widespread in Wallacea but have often gone undetected during excavation, as shell remains are often only sampled following excavation and the bulk discarded in the field. As at Leang Sarru the earliest occupation levels at Golo Cave preserve no food remains other than shell, so it is difficult to infer the purposes for which the flaked stone and shell tools were made or the type of activities that sustained settlement in these small islands.

Discussion In Sahul’s colonisation, the variable environments that were occupied quickly indicate the flexibility of the technology that allowed adaptation from tropical rainforest in PNG and sub-arctic conditions in Tasmania to very arid deserts inland. By and large the stone tool assemblage could be described as Mode 3, but the trajectory from large thick flakes retouched for use as scrapers to a reduction in the size of tools with smaller cores, fewer core tools and scrapers increasingly made on thinner flakes proposed by Lorblanchet and Jones (1979) has not held up. Some sites with the most massive tools are mid- to late Holocene in age, while many of the oldest sites such as Carpenter’s Gap 1 and Widgingarri 1 in the Kimberley, Devil’s Lair in southwest Australia and Upper Swan near Perth have lithic assemblages comprised almost exclusively of small unretouched and irregularly retouched flakes (Dortch 1984; Dortch & Dortch 1996; O’Connor 1995; O’Connor 1999; Pearce & Barbetti 1981). This is also true for some sites in New Guinea such as Lachitu on the north coast of PNG (O’Connor et al. 2011a) and Liang Lemdubu in the Aru Islands (O’Connor et al. 2005; Hiscock 2005) that formed part of Sahul in the Pleistocene. Very little specialised stone tool adaptation is evident except for the axes found in the tropical northern part of the continent. These appear to have been a distinctly Australian innovation as they are not found anywhere else in the world at this early date. Although few complete axes have been recovered, the examples we do have support the view that they were manufactured in different forms and sizes for different purposes. This is equally true of the robust Pleistocene flaked and waisted axes found in mainland PNG. This variation in Sahul stone tool assemblages has hitherto gone unacknowledged because of the focus on flaked stone assemblages. To these technologies must be added the bone tools found in southern sites, where they have been well preserved in caves, and organic technologies that have not been preserved but for which other evidence points to their existence. The importance of organic remains that have not been preserved cannot be underestimated in these sites. Where retouch is identifiable on flakes in Sahul sites, it is associated with woodwork, indicating the importance of this raw material from the earliest times.Wood may have been used to make spears to capture macropods found in some of the sites.As the function of spears with projectile points that are a marker of Upper Palaeolithic industries have little technological advantage over wooden spears (Waguespack et al. 2009) in this region, wood may very well be the technological equivalent. The evidence for fibre is indicated by the presence 173

Jane Balme and Sue O’Connor of hafted axes (although animal sinew may have been used as twine instead), by its association with watercraft (Balme 2013) and its suggested use for nets in western New South Wales. The stone artefact assemblages from all the stratified Pleistocene sites in ISEA can be characterised as Mode 3 (O’Connor et al. 2011a). In many respects the island assemblages are very similar to the earliest assemblages from Sahul, although lacking the large flaked and waisted axes found in PNG, as well as the smaller edge-polished axes and hatchets of northern Australia (Geneste et al. 2010). The island assemblages also bear comparison with flaked lithic assemblages in South Asia dated to approximately the same time (O’Connor et al. 2011a, online supplement). In terms of subsistence, the faunal suites from the islands exhibit some distinctive characteristics. The Philippines have a range of large mammal fauna; however, once the early colonists crossed the Wallace Line they would have been presented with significant adaptive challenges in terms of available game. This is especially true for the smaller islands of Wallacea that have unbalanced and depauperate faunas. Flores had a range of now extinct large fauna such as stegodons, a large varanid and a land turtle, as well as giant endemic rats, and at least some of these large fauna may have survived until as late as ~18 ka. Timor once had similar range of mega-faunas, but it would appear that they were extinct by 42 ka. The only non-marine fauna in the earliest levels of Jerimalai and Lene Hara are small reptiles, bats and large rats. The faunal resources available to the arriving colonists in the small islands such Talaud-Sangihe Archipelago were apparently even more limited. It is perhaps within this context that the remarkable maritime skills of the early people in East Timor should be viewed. In this chapter we have talked about the physical manifestations of technology. These extraordinary responses to the environment of ISEA and the adaptable use of stone, wood and fibre to different environmental conditions in this region is underscored by the intangible aspect of technology, including methods for maintaining information flow and planning associated with modern human cognition (Balme et al. 2009;Veth et al. 2011)

Conclusion Rather than seeing the evidence from Sahul and ISEA as indicating a progressive loss of tool types or as showing cognitive continuity in flake tool production over time and space following the dispersal of modern humans out of Africa, into India and thence to Sahul, it might be more profitable to see the lack of stone tool specialisation as an adaptation to new, unfamiliar and challenging environments. The basic reduction repertoire would not have required that the early colonists located fine-grained sources of stone or material of a required size for blade production. Organic materials could have been substituted for or combined with stone, when needed or when the opportunity arose. As we can see, in the earliest levels at Golo Cave the occupants were making their sharp flakes out of large Turbo shells; in East Timor and Talaud they were using small nodules of chert, whereas in the earliest sites in the Philippines it appears that people were relying wholly on expedient bamboo or other organic tools. The role of fibre in ISEA, as indicated by the presence of fish hooks, its critical role in the colonisation of Australia while strapped firmly to the watercraft and its probable subsequent use in Sahul, is further evidence of the remarkable adaptability of the people of this region. Instead of uniformity, the earliest assemblages in Sahul are now seen to be as diverse as the landscape itself.

Acknowledgments We thank Robin Dennell and Martin Porr for their invitation to write this chapter and for their useful editorial comments. Kay Dancey prepared the figure.

174

Chapter 14 Tasmanian Archaeology and Reflections on Modern Human Behaviour

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks

Introduction The constitution of modern human behaviour continues to be a hotly debated topic (Gamble 1992; Holdaway & Cosgrove 1997; Marean et al. 2007; McBrearty 2007; Shea 2011b). Opinion for its appearance has swung from enigmatic cognitive changes claimed at about 45 ka (Klein 2000) to social and economic intensification during the European Upper Palaeolithic, enshrined in notions of developing ‘cultural complexity’. Alternative arguments have been made for the Australasian region (O’Connell & Allen 2007), Europe (Powell et al. 2009) and Africa (Henshilwood & Marean 2003), suggesting that the earlier emergence of complex technologies, subsistence strategies and art can be explained by increasing population driven by resource competition or intensification. Ironically the smaller postulated population in Sahul (Brumm & Moore 2005) did not restrict the development of hafted axes at 44,000 BP (Groube et al. 1986; O’Connell et al. 2010), grinding technology at 35,000 BP (Fullagar & Field 1997), sea faring at 50,000 BP (O’Connell et al. 2010), backed artefacts at least 15,000 BP (Slack et al. 2004), symbolic expression at least 35,000 BP (Morse 1993a; Balme & Morse 2006; Balme et al. 2009), deliberate burial at 45,000 BP (Webb 1998), early deep-sea fishing (O’Connor et  al. 2011a) and animal translocation at 20,000 BP (Allen et al. 1988; Flannery & White 1991; Leavesley 2006), raw material movement over 300 km (Summerhayes & Allen 1993) and montane plant exploration by 45,000 BP (Summerhayes et al. 2010) (Table 14.1). The transition from the Middle to Upper Palaeolithic has also been characterised as an evolutionary process from simple to complex. This has been a strong organising principle used to categorise human societies in developmental stages, linking them in terms of human capacities and their position as evolving organisms. It stems from the European view that the Lower and Middle Palaeolithic is characterised as having a more transitory economic pattern, more conservative stone technologies, and lower populations with simpler social, economic and organizational networks than the Upper Palaeolithic, which is seen as more dynamic (e.g., Bar-Yosef 2007, 209). The notion of increasing socio-economic complexity from Lower and Middle Palaeolithic foragers to Upper Palaeolithic logistically organised systems is clear (Binford 1980). The Upper Palaeolithic is seen as a reflection of increased “efficiency” with greater cultural complexity based 175

176

Table 14.1.  Behavioural attributes of Sahul (Australia–New Guinea–Tasmania) compared with those of the European Middle and Upper Palaeolithic Attributes

Australia–New Guinea

Tasmania

Neanderthal

European Upper Palaeolithic

Boating technology Marine and lacustrine exploitation Stone tool technology

Yes Yes Mode 1,2,3,5, hafted axes, grinding technology Yes Regional

None detected None detected Mode 1,2,3

None detected Yes Mode 1,2,3,4

None detected Yes Mode 1,2,3,4,5

Techno complexes Stone movement Seasonality, land use patterns Prey selection Animal translocation Butchery and processing Small and elusive mammal and bird exploitation Plant exploitation Bone and shell technology Features, e.g., stone line hearths, spatial use. Storage Art, parietal and mobile

Yes Yes Yes Yes Yes

None detected Yes Mainly local, some regional Mainly local, some regional Yes Yes Yes Yes None detected None detected Yes Yes Yes Yes

Yes Regional

Yes Yes None detected

None detected Yes None detected

Yes Yes None detected

Yes Yes Present

None detected Polychrome figurative, hand stencil, geometric motifs, body ornamentation

None detected Monochrome negative red-ochre hand stencil motifs, notched bone

None detected None detected, but see Higham et al. 2010; Zilhão et al. 2010

Deliberate burial Skeletal

Yes Gracile and robust

None detected Gracile

Yes Robust

Yes Polychrome figurative, hand stencil, geometric motifs, engraving, sculpture, body ornamentation Yes Gracile

Yes Yes None detected Yes Yes

Sources: Gamble 1986; Holdaway& Cosgrove 1997; Cosgrove 1999, 2006; Balme et al. 2009; O’Connell et al. 2010; Conard & Richter 2011.

Tasmanian Archaeology on the presence or absence of stone types such as scrapers, handaxes, blades, projectile points and microliths in an archaeological assemblage. This in turn has been seen as a reflection of the behavioural differences between their makers. However, one of the problems of European Palaeolithic research has been its solipsism, where all else is measured against the array of stone tool typology to produce an evolutionary cultural continuum. Tasmanian stone tool assemblages have, for example, been described as Middle Palaeolithic, frequently drawing inappropriate analogies with Neanderthals, despite the fact that the Tasmanians were anatomically and behaviourally modern (Holdaway & Cosgrove 1997; Cosgrove & Pike-Tay 2004). This is paradoxical, because the Middle to Upper Palaeolithic transition is seen in Europe as a watershed that separated earlier archaic hominids from modern humans. Critics of these analogies have been calling for a greater scrutiny of what defines modern human behaviour (Holdaway & Cosgrove 1997; McBrearty 2007; O’Connell & Allen 2007; Brumm & Moore 2005; Shea 2011b). Given that many European regional analyses stem from this perspective, rarely is reference given to the areas such as Sahul, which do not have the classic cultural progression from handaxes, to scrapers, to blades. In an earlier paper Holdaway and Cosgrove (1997) pointed out the confounding nature of the Tasmanian late Pleistocene and the Neanderthal archaeological records. Here were two groups, so closely aligned by 19th- and early 20th-century scholars; one ‘modern’, the other ‘archaic’, making similar stone tools associated with a range of other behavioural attributes commonly identified as Middle Palaeolithic behaviour. However, when the Tasmanian, Australian, European Middle and Upper Palaeolithic records are compared (Table 14.1), the range of cultural attributes is highly variable, to some degree comparable to that seen now in the African record (McBrearty 2007; Shea 2011b, 2011c). The people of Sahul did not follow the expected classic evolutionary continuum. It is also clear that the absence of modern human behavioural ‘traits’ does not suggest that modern capabilities were lacking (O’Connell & Allen 2007). This is where the Australian and in particular, the Tasmanian evidence is important in debates about what material evidence represents human modernity and whether its appearance was gradual or punctuated. We suggest that to unravel these contradictions, a comparative approach is needed that focuses on identifying archaeological variability, comparing archaeological records at commensurate times and places. For example, in Tasmania there were Late Pleistocene human populations using a Middle Palaeolithic lithic technology; with non-hafted organic technology, they practised art and hunted systematically and seasonally with a structured approach to land use.The Neanderthal populations had a hafted lithic technology, systematic hunting and prime-age prey selection, seasonal acquisition of resources and burial of the dead. Either the Tasmanians were ‘archaic’, a description that does not fit with the emerging archaeological evidence, or the definitions of ‘modern’ and ‘archaic’ behaviour need reassessment. Given that the past 14 years have seen major advances in theory and material evidence, we felt it time to revisit the question of behavioural variability and modernity as it relates to these archaeological records.

Tasmanian Palaeoenvironment The importance of understanding the palaeoecological structure and its influence on the distribution of hunter-gatherer resources has been an important theme in Tasmanian Late Pleistocene archaeological research (Cosgrove et al. 1990; Holdaway & Porch 1995). It is undeniable that the environment provided a range of subsistence options from which choices were made regarding how and when to schedule their exploitation.This evidence is important in providing the framework within which to examine human responses to changing climates and environments. During the Late Pleistocene, Tasmania was connected by a land bridge to the mainland, when sea levels began to lower after 40,000 BP (Figure 14.1). People took this opportunity to cross and occupy caves, open situations and rock shelters in Tasmania. Although these areas have similar 177

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks

Figure  14.1.  Distribution of western and central Tasmanian Late Pleistocene archaeological sites. No published sites older than 10,000 BP have yet been identified in eastern mainland Tasmania. latitudes to those of southwest France and northern Spain, Tasmania experienced limited glaciations with a more maritime climate. Only small ice sheets covered the highest elevations and the climate during Marine Isotope Stage (MIS) 3 and MIS 2 was cool temperate, transitional between glacial and interglacial environments (Mackintosh et al. 2006). Average annual temperature depressions of at least 6° C occurred across the region at the height of the last glacial. Open grassland in central and eastern Tasmania and infertile soils supporting low shrubs in western Tasmania constituted the major vegetation structure (Colhoun 2000). In Europe at the same time, 178

Tasmanian Archaeology temperature depressions of at least 12° C and perhaps 15° C have been suggested (Gilligan 2007a). The remaining forests retreated to lower coastal situations. Pollen data from western and central Tasmania support a general but variable vegetation pattern across the region. Discontinuous areas of fertile soils underlain by dolerite and limestone supported herbs and Poa sp. grasses, while acid peats and infertile quartzite substrates supported low shrubs and button grass sedges elsewhere (Kirkpatrick & Fowler 1998). It is believed that this palaeoecological structure produced a patchy distribution of grasses attractive to grazing wallabies in western Tasmania (Cosgrove et al. 1990; Cosgrove 1999). The limestone geology that contained caves and rock shelters bordered grassy patches on which the wallabies fed, and humans occupied rock shelters from which they hunted the wallaby. The coldest periods were around 20,000 cal BP, and although rainfall was reduced by 50% of present values, western Tasmania would have still received more than 1500 mm annually, probably in the form of snow. Areas to the east would have suffered from droughts, as precipitation would have been reduced to between 200 and 150 mm annually. At this time, dune and lunette building was common because of reduced vegetation cover in the Midlands while in the west and east, pockets of Eucalyptus sp. and rainforest survived in protected locations with a shrub understory (Kirkpatrick & Fowler 1998). The prevailing westerly winds at this time likely brought rain and snow to the west coast as Tasmania would have had a moist maritime climate rather than a continental one, reflected more at inland locations like Lake Mungo on what is now the mainland (Bowler et al. 2003). Human impact on the environment appears to be coupled with deterioration in environmental conditions and landscape instability owing to reduced precipitation and vegetation cover (MacIntosh et al. 2009). Although a range of megafauna was present in Tasmania, the reasons for its extinction are contested. Evidence from Tasmania suggests that although there appears to have been a human overlap with two species of large kangaroos, Protemnodon anak and Macropus giganteus titan, at about 40,000 cal BP, no bones of these animals have been recovered from any of the archaeological sites in western or central Tasmania (Cosgrove et al. 2010). All of the human prey was smaller extant wallaby and wombat species. The only large carnivores recorded in Tasmania were Thylacaleo carnifex (marsupial lion), Sarcophilus laniarius (Pleistocene version of the Tasmanian devil) and Thylacinus cyanocephalus (Tasmanian tiger).Their bones have not been found in archaeological sites and, unlike the European hyenas, cave bears and wolves, they probably posed less competition to humans for prey resources and potential cave occupation. Both Sarcophilus laniarius and Thylacinus cyanocephalus were not much larger than a medium-sized dog and hunted or scavenged individually. Although Thylacaleo carnifex was bigger, between 87 to 130 kg (Wroe et al. 2003), it was possibly extinct by the time humans reached Tasmania, the current latest date being 53,000 ± 4000 BP for its occurrence (Cosgrove et al. 2010).

Human Skeletal Evidence The limited evidence of the skeletal morphology of Late Pleistocene Tasmanians indicates they were gracile and of modern appearance. A parietal bone from Nanwoon Cave in the Florentine River valley has a minimum age of more than 16,000 BP (Webb 1988). This bone was extremely thin walled, between 1 and 1.5  mm in thickness, and Webb observed few muscle markings. Unfortunately it was not in situ when recovered from the base of a small talus slope inside Nanwoon Cave. Its age was estimated from charcoal collected from a subsurface context near the top of the deposit, while the bone itself had fallen out of an exposed interior section of the cave deposit (Cosgrove 1999). No excavation was undertaken, but the remains are clearly of Late Pleistocene age. The gracile human burial found on King Island dates to circa 14,500 BP (Thorne & Sim 1994). It is unlike the more robust populations found at Kow Swamp on the mainland dating to 179

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks the same time period. These individuals have heavily buttressed crania and represent deliberate burials. The King Island skeleton is said to be modern because of its more gracile cranial form. It is interpreted as a reburial because there were only loose associations between the bony elements. The two tibiae and one femur were found together, but other elements were scattered in the deposit. There has been some discussion about its sex based on the robustness of the femur (Brown 1994; Thorne & Sim 1994). Thorne and Sim (1994) argue that its relatively short overall length and large femoral head reflect both its male attribution and an adaptation to high-latitude environments, similar to those found in glacial Europe, especially among Neanderthals. As an individual, it is not clear whether it is representative of the general Tasmanian Late Pleistocene population. Given the degree of morphological variability in the skeletal form found across south-eastern Australia at the time, larger samples are required to better characterise population morphology (Pardoe 1986; 1991a).

Chronology At present, Warreen Cave and Parmerpar Meethaner rock shelter are the two oldest sites (Figure 14.1) with basal ages of 34,790 ± 510 BP (39,906 ± 879 cal BP) (Allen 1996a, 154) and 33,850 ± 450 BP (39,310 ± 1151 cal BP) respectively (Cosgrove et al. 2010). The lowest occupation levels at Bone Cave date to 29,000 ± 520 BP (31,435 ± 554 cal BP) and 28,330 ± 630 BP (30,939 ± 638 cal BP) respectively (Allen 1996a, 113). Pallawa Trounta Cave is dated to 29,800 ± 720 (32,021 ± 637 cal BP), while Nunamira Cave and ORS 7 rock shelter have basal ages of 30,840 ± 480 BP (33,064 ± 475 cal BP) and 30,750 ± 1340 BP (33,410 ± 1518 cal BP) respectively (Allen 1996a). Sites in the Bass Strait islands are younger. For example, Cave Bay Cave dates to about 23,000 BP (Bowdler 1984), while Mannalargenna Cave on Prime Seal Island and Beaton Shelter on Badger Island date to 23,015 ± 210 BP (25,630 ± 457 cal BP) and 23,180 ± 1280 BP (25,747 ± 1583 cal BP) in their basal layers respectively (Sim 1998, 258). Although these are circa 12,000 radiocarbon years younger than Warreen Cave at 35,000 radiocarbon years, equivalent ages are likely, as the Bass Plain was the initial entry point for people into peninsula Tasmania. Spikes of 14C dates have been argued to indicate possible fluctuating population (Holdaway & Porch 1995), and discard rates of cultural material suggest a relationship between changing climatic and occupation intensities (Cosgrove 1995b). Increasing discard rates of bones and lithics occur between about 35,000–23,000 and18,000–13,000, with a hiatus around 20,000–18,000. It suggests lower rates of occupation at the height of the glacial although some sites like Kutikina at lower elevations have evidence of use at this time (Kiernan et al. 1983; Garvey 2006). Higherelevation sites were occupied over longer periods of the last glacial maximum (LGM), possibly because they continued to be more open from the time of initial human settlement. Lowerelevation sites were more prone to the influence of the expansion and contraction of unproductive rainforests and the restrictions placed on the availability of grassland patches. By 13,000 BP all the southwest cave sites were abandoned probably because of the encroaching rainforest and the disappearance of the prey animals (Cosgrove 1999).

Technology Although it has been observed that much of Tasmanian Aboriginal stone technology is similar to the European Middle Palaeolithic, it lacks points and Levallois technology, being mainly composed of multiplatform and single platform cores, primary flakes, retouched flakes and various sized scrapers (Jones 1977). Although there are additional items such as small denticulate flakes (Figure 14.2), delicate ‘thumbnail’ and end scrapers indicating a greater range of stone tool types, little if any technological change occurred for at least 25,000 years (Cosgrove 1999; Holdaway 2004). Organic technology probably comprised wooden spears and clubs that were a familiar 180

Tasmanian Archaeology

Figure 14.2.  A small pressure flaked, denticulate quartzite flake from Bone Cave recovered from below the level dated to 23,130 ± 460 BP (Allen 1989). It has a serrated edge about 12 mm in length (photograph by the authors). and flexible technology with some equivocal evidence of bone armatures (Webb & Allen 1990). There is no direct evidence of hafting or the use of resins or mastics for gluing. There were no large and dangerous prey animals or large carnivores to compete with. Thin-skinned marsupials without subcutaneous fat obviated the need for heavy-duty stone points and less need for scrapers for hide preparation. The Tasmanian ethnographic evidence suggests the use of static traps, falls and use of fire to capture prey animals. The possible range of hunting techniques is discussed by Cosgrove and Allen (2001) and deduced from the early ethnography. There is no evidence from any of the Late Pleistocene sites for the use of lined hearths, and most were simple fire pits for cooking. Evidence for an organized spatial use of cave sites is limited, although it appears greater use of space occurred at Nunamira, Bone and Stone Caves around 16,000 BP, suggesting 181

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks some organisation of activities. However, no large-scale spatial excavation has been undertaken to investigate this further, although the Mackintosh site has been investigated with open-area excavation, but no strong dispersal of activities has been detected (Stern & Marshall 1993).The limited evidence for organised space within these caves probably reflects the lack of research in this area, since only small amounts of deposit have been excavated (Allen 1996a, 31–39). The sub-Antarctic conditions would perhaps have encouraged fitted clothing, and bone points with use-wear polish have been recovered from archaeological deposits (Webb & Allen 1990). Early ethnographic descriptions of Aboriginal people suggest that their only form of clothing was a wallaby or kangaroo ‘cloak’ loosely worn across the shoulders (Plomely 1966, 531). However, research on ‘clothing’ and human energetic requirements in the Late Pleistocene suggest (Gilligan 2007a; 2007b; this volume) that fitted or ‘complex’ clothing did not develop because wind chill, a significant factor in human tolerance to the cold, did not reach a critical level. Gilligan (2007a) points out that the climatic conditions in Tasmania were more like those of the European MIS 5b and 4, rather than the Northern Hemisphere MIS 3 or 2, and this is supported to some extent by Mackintosh et al. (2006). The point made here is that the technological items that are needed to make fitted clothing such as blades and fine needles are absent from both European Middle Palaeolithic and Tasmanian assemblages (Gilligan 2007a). The suggestion is that these technological developments are more responsive to specific environmental and social needs that tend to appear when human tolerance to cold moves beyond the critical thermal threshold (Gilligan 2007a). Evidence for raw material movement across southwest Tasmania comes from three sources. Firstly, Darwin glass was obtained from Darwin Crater on the west coast of Tasmania and moved inland in straight-line distances of at least 100 km. This material has been found only in the limestone caves of southern Tasmania. It is quite brittle, but small thumbnail scrapers have been made out of this material. Secondly, blue-grey chert, a material common in the easterly sites of the southwest Tasmanian region and made into thumbnail scrapers, has its origin in the Ragged Range (Figure  14.1). This material was imported into Bone, Nunamira, Warreen and Pallawa Trounta Caves, 10, 15, 37 and 45 km distant respectively (Sheppard 1997). Recently, brecciated chert has also been recorded from Kutikina Cave (Burch 2007, 26). This material has its origins in northern Tasmania at the Parwee quarry, 130 km away (Cosgrove 2000). It is also present in the site of Parmerpar Meethaner 60 km southeast of the quarry. Although the amounts of these raw materials were relatively small, it suggests that Late Pleistocene artefact raw material movement was relatively extensive, particularly as distances would have been far greater when using river valleys and other indirect routes to move the material around the landscape. It also reflects a degree of social connectedness between the western regions during the last Ice Age.

Subsistence and Land Use Over the recent years, our understanding of Late Pleistocene Tasmanian Aboriginal economic behaviour has been enhanced by investigations into the seasonal cave use and studies of the economic utility of prey species (Pike-Tay et al. 2008; Garvey 2010). By studying the banded annuli in the wallaby teeth, a convincing picture has now emerged about the seasonal exploitation and butchery patterns of these prey animals. It is clear that caves more than 400 m above sea level were used in the summer, whereas those caves at lower elevations like Warreen and Kutikina were occupied at least during the winter. Most wallaby kills were recorded between autumn and spring at these locations, whereas at Bone and Nunamira Caves in the upper mountains, summer hunting was practised. The need to secure higher quality fat and marrow as well as thick fur pelts in the cooler periods is thought to have driven this scheduled hunting (Cosgrove & Allen 2001). However, seasonal and spatial studies of wallaby distribution via marrow, muscle and brain quality showed that they supplied a reliable, stable protein and fat supply irrespective of seasons 182

Tasmanian Archaeology (Garvey 2011).There appeared no great differences between male and female animals or between different altitudes or seasons. These results are very different from the highly variable meat and marrow quality of, for instance, Northern Hemisphere caribou herds that suffer from fat depletion through winter to early spring (Garvey 2011). It is also clear that selective butchery of the wallabies was carried out despite their small size. In Northern Hemisphere caribou studies, these large animals were selectively butchered and elements transported back to rock shelters and caves (Castel et al. 2002). In Tasmania the same pattern occurred, with a focus on the lower limb bones containing the highest amounts of meat as well as offering bones with the highest marrow content (Cosgrove & Allen 2001; Garvey 2006). This marrow provided a substitute for carbohydrates that played an important role in the metabolism of meat (Cosgrove et al. 1990; Cosgrove 1995a). The access to marrow was crucial for thermoregulation and digestion of lean meat and is evidenced by the high numbers of systematically smashed wallaby long bones. The wallaby and wombat are by far the commonest animals, composing 75% to 90% of faunal remains in all caves found at bone densities of over 250,000 in less than a cubic metre of deposit. No evidence for plant food exploitation has survived, although it presumably formed a component of the diet, despite cooler growing seasons and a lack of suitable geophytes. Bird bones have also been found within the archaeological deposits from several sites and include the medium-sized native hen (Gallinula mortierii) and the large Tasmanian emu (Dromaius diemenensis) at Kutikina and Nunamira. Emu eggshell was recovered from Nunamira Cave, and Cape Barren goose (Cereopsis novaehollandiae) eggshell was found at Mannalargenna Cave (Brown 1993; Cosgrove 1995b; Garvey 2006). The remains of the elusive platypus (Ornithorhynchus anatinus) were found at ORS 7,Warreen, Bone and Nunamira Caves and suggest people had a detailed understanding of their behavioural ecology. They live in burrows, are mainly nocturnal although they are known to be diurnal, lay eggs and forage for food on river bottoms and come to the surface to breathe (Bethge et al. 2009). Hunting strategies that allowed the killing of these animals would need to be largely planned rather than opportunistic. Marshall (1992) has argued that the platypus was an attractive source of fat and fine pelts. The other monotreme, the echidna (Tachyglossus aculeatus), has been found only at Warreen Cave, represented by just three bones. It is possible that these represent food remains, although their part in the human diet seems negligible. Similarly the eastern grey kangaroo (Macropus giganteus), barred bandicoot (Perameles gunnii) and Tasmanian bettong (Bettongia gaimardi) are rare or totally absent from the bone assemblages. These animals prefer, like the echidna, drier, open grassy woodland habitats. No megafauna was found in the food refuse (Cosgrove et al. 2010), and archaeological data suggest that medium sized (12–25 kg) animals were hunted during Late Pleistocene occupation. The reasons for this economic pattern have been discussed in terms of the patchy nature of the vegetation in southwest Tasmania during the Late Pleistocene (Cosgrove et al. 1990; Cosgrove & Allen 2001). This has been argued to be due to the limited distribution of fertile soils that could support grasses, the major food of the Bennett’s wallaby. It has been noted that this is the commonest human prey animal in the archaeological assemblages. Wallabies are solitary and relatively sedentary animals that come together in groups on restricted grassy patches that coincidentally were surrounded by limestone geology containing caves. This ecological ‘tethering’ of wallabies to patches allowed the hunters a degree of predictability and reliability. The seasonal evidence shows that people moved from patch to patch to obtain this highly ranked prey; winter at low altitudes, and summer at higher altitudes.The significance of the limited range of species detected in the assemblages is that people largely ignored the next lower-ranked prey animal and moved instead to the next patch on their hunting schedule. Studies based on prey choice and selection (Cosgrove & Allen 2001) show very few medium-sized or other types of animals were captured, supporting suggestions that there was a consistent focus on Bennett’s wallaby throughout the Late Pleistocene period. This is unexpected because this region experienced climatic perturbations that would have restructured the environmental and human resource conditions over this extended period. It illustrates the resilience of these animals as well as the overwhelming interest 183

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks the hunters had with their prey. Despite the fact that there were many more animals from which to choose such as, for example, the two possum species, wombat, potoroo, water rat, the larger grey kangaroo and possibly two species of megafauna, they appear to have deliberately taken wallaby at the expense of other animals (Cosgrove & Allen 2001). An obvious conclusion is that the pattern is inconsistent with an encounter hunting strategy and more likely reflects an optimal foraging one. Overall the evidence shows a structured economic approach and landscape use by Tasmanian Ice Age peoples that involved seasonal choices of when and where to visit the patches, and when to leave and travel to the next one.This would have involved timely scheduling between patch use, probably using these areas on a periodic basis rather than on a continuous cyclic one (Cosgrove & Allen 2001). It appears that during these times there was high residential mobility involved with hunting, with apparently little investment in spatial structures inside the caves. In addition, the data support the notion of deliberately scheduled short-term, seasonal use of the moist grassy patches to hunt ecologically ‘tethered’ wallabies. As already discussed, seasonal movements may have had less to do with procuring animals since wallabies were apparently less physiologically affected than their reindeer counterparts seasonally, nor is there evidence for the regular movement of different raw materials in any quantity.There is clear evidence, however, that these people used their landscape in a planned, seasonal way together with embedded behaviour that is not explained by economic concerns alone. Thus seasonal movement may have also been related to social connections, and art may have played a role in this, a topic to which we now turn.

Stencil Art It was not until early 1986 that confirmation of Ice Age antiquity of rock art was made.Tasmanian Aboriginal people had placed red ochre hand and arm stencils deep inside the limestone caves of southwest Tasmania (Figure 14.3). The stencilled hand is a regular motif of cave art, found across the prehistoric world. The Upper Palaeolithic French caves such as Peche Merle, Gargas and Cosquer (Clottes 2007, 86–87, 92–93, 98), for example, contain such stencils. In Tasmania there are at least five known sites: Ballawinne Cave (Harris et al. 1988, 94–95), Judds Cavern (Cosgrove & Jones 1989), Keyhole Cavern (Allen 1996a, 35–36), and an unpublished site, Mount Riveaux. The first discovery was at Ballawinne Cave, in a limestone doline, today deep within thick temperate rainforest. The cavern is about 20 m under ground reached by a narrow tunnel to which a large cavern is connected. Along the walls are about 18 human hands and splashes of red ochre. All, as far has been determined, are adult hands in negative stencils, executed by blowing ochre mixed with liquid and adhesives over the hand to make a stencil (Cosgrove & Jones 1989). Some are incomplete stencil hands where only certain fingers were held against the wall to produce various hand stencil shapes. In the large stream cave of Judds Cavern, approximately 30 m from the entrance on the western side is a level shelf area, about 8 m above the water level with evidence of cave art. On a sloping lintel about 2.4 m from the floor is an area with five hand stencils composed of four adult right hands and one left hand. Their height above the floor suggests that their execution required the person to raise his or her hand up against the rock face onto which red ochre paint was sprayed from the mouth (Cosgrove & Jones 1989). A further six hand stencils are located on the south wall. They are placed within two concave channels formed by differential erosion, separated by a raised harder formation. The channels are about 16 cm wide, and three hand red-ochre stencils have been placed within each of the two channels in a linear fashion. Two of them are right adult hands, one is a perfectly formed child’s hand and one of the stencils consists of a single finger. The stencilled application of the child’s hand suggests the activity of a family group. A further 1 km into the cave are seven complete arms outlined in red ochre but details remain unpublished (Cosgrove 1999). 184

Tasmanian Archaeology

Figure 14.3.  Five red-ochre hand stencils located on a limestone lintel in Judds Cavern, southcentral Tasmania (Cosgrove & Jones 1989; photograph by the authors).

Summary The archaeology of Late Pleistocene Tasmania has many of the qualities of the Middle Palaeolithic, particularly in terms of stone technology but with such a degree of variability as to simultaneously challenge the meaning of ‘archaic’ and ‘modern’ human behaviour as usually defined by European archaeologists. The Tasmanians show a remarkable variety in their cultural repertoire that allowed them to prosper during the glacial period on Sahul’s southern peninsula. What is interesting is that many of the attributes that are common to the Ice Age Tasmanians can be found among the behavioural attributes of Neanderthals (Table 14.1). That the Tasmanians were clearly anatomically and behaviourally ‘modern’ raises the question of using simple one-to-one correlations between technology and biology to define evolutionary developments between human groups.What separates the Tasmanians and Neanderthals are the clear differences in environmental structures and about half a million years of evolutionary divergence. However, the similarities in their technology, hunting behaviour, butchery and prey selection give us cause to re-evaluate our perceptions about what it means to be part of the human lineage. If the Tasmanians are modern humans technologically behaving like Middle Palaeolithic people, which seems to be the case, what does this say about Neanderthals and their classification and, what are the wider implications for understanding the archaeological attributes of modern human behaviour?

The Neanderthal Case The past decade has been a very productive one for our knowledge of our closest fossil relative, Homo neanderthalensis.A wide variety of studies has focused on various aspects of the Neanderthals’ skeletal record (reviewed in Hublin 2009; Weaver 2009), on the dietary signal coming from 185

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks chemical studies of their bones (e.g., Richards & Trinkaus 2009), on their geographical distribution and their archaeological record (Roebroeks 2008) and, very importantly, on their genetic characteristics. Genetic studies have increased our understanding of the Neanderthals’ evolution and their relationship with Homo sapiens. These indicate that modern humans and Neanderthals shared a common ancestor only 400,000–700,000 years ago, and a somewhat comparable picture is emerging from studies of their skeletal remains (Hublin 2009). Building on the same Bauplan, two different hominin lineages emerged: the ancestors of modern humans in Africa, and Neanderthals in western Eurasia, culminating in the classic Neanderthals of the last glacial. Until very recently these were thought to have vanished completely around 35,000 radiocarbon years ago. Comparison of the draft Neanderthal genome (Green et al. 2010) with the genomes of living people now suggests between 1% and 4% Neanderthal ancestry for present day humans outside Africa. Importantly, according to these studies Neanderthals were amongst the ancestors of some modern humans. Moreover, some people living outside Africa can also trace part of their ancestry to a thus-far-unknown Asian hominin group, the Denisovans, whose genome is almost as different from Neanderthals as the Neanderthal draft genome is from that of extant humans (Reich et al. 2010). Table 14.2 gives a very schematic overview of the biological, behavioural and cultural differences between Neanderthals and modern humans of the European Upper Palaeolithic. In the context of this chapter, such an overview has to be too short to do justice to the real complexity of the record, while we also have to acknowledge that Neanderthals and modern humans were very similar in many aspects of their biology and behavior. Nevertheless, the differences are also striking (see Table  14.2). They are usually interpreted in cognitive terms, often as the result of Neanderthals’ lacking “fully modern language” (see Roebroeks & Verpoorte 2009 for discussion). In such interpretations, the archaeological record is thought to show the presence of more complex language patterns and sophisticated cognitive abilities by the time of the Upper Palaeolithic of Europe – and even tens of thousands of years earlier during the Middle Stone Age in southern Africa. The problem with these cognitive and language-based explanations is that they can lead to tunnel vision, in which “modern human” accomplishments in any domain are treated as far more complex and superior to anything the ‘archaics’ ever accomplished (Corbey & Roebroeks 1999). For example, Brown et al. (2009) present evidence from the site of Pinnacle Point in Southern Africa, where humans regularly employed heat treatment to increase the quality and efficiency of their stone tool manufacturing process, 164,000 years ago.They infer that this technology required a novel association between fire, its heat and a structural change in stone with consequent flaking benefits that demanded “an elevated cognitive ability”. They also suggest that as early modern humans moved into Eurasia, their ability to alter and improve available raw material and increase the quality and efficiency of stone tool manufacture could have provided a behavioural advantage in their encounters with the Neanderthals. However, there exists solid and well-published data that European Neanderthals from at least 200,000 years ago onwards routinely used fire to synthesize from birch bark glue for hafting stone tools to their handles (Mazza et al. 2006). Hence, on the basis of current understanding of the archaeological record of fire use, and this evidence for pitch processing in particular, the hypothesis of Brown and colleagues is not tenable (Roebroeks & Villa 2011). Another problem with these cognitive explanations is their failure to address the fact that “fully modern” humans created very diverse archaeological signatures, sometimes strongly resembling what Neanderthals left behind in western Eurasian landscapes (Roebroeks & Verpoorte 2009). The record of Pleistocene Tasmanian Aboriginals as discussed in this chapter has many of the hallmarks of the Neanderthal record (Holdaway & Cosgrove 1997). Interestingly, the colonisation of wider Australia has been interpreted as the “earliest evidence of modern human behaviour” (Noble & Davidson 1996, 217). In the view of Noble and Davidson, when “the archaeological record shows that actions were taken upon materials that show evidence of forward planning to achieve a goal” (1996, 217), we are probably dealing with behaviour that can be identified as ‘linguistic’. In this vision, the arrival of humans in the Australian region is based on such behaviour, 186

Tasmanian Archaeology Table  14.2.  Biological, behavioural and cultural comparisons between the late Middle Palaeolithic and the Upper Palaeolithic in Europe Industry and time period

European Late Middle Palaeolithic (ca. 125–40 ka)

European Upper Palaeolithic (ca. 40–10 ka)

Species Body form and energy requirements Hunting efficiency and diet breadth

Neanderthals Robust, costly

Modern humans Gracile, less costly

Stable isotopes

Northern limits to distribution Lithic technology

Efficient, relatively narrow focus on Efficient, somewhat broader larger mammals, with evidence prey choice, including for the consumption of (cooked) smaller game, fish and plant material plants Top carnivores with heavy Comparable to Neanderthal emphasis on larger mammals signal, with some individuals consuming significant amounts of fish South of 55 degrees Range expansion Laminar reduction, discoidal and Levallois

Hunting weapon technology

Thrusting spears, little investment in projectiles

Investment in on-site structures

Limited, simple fireplaces

Burials Art, personal ornament and use of pigments

Without identifiable grave goods Use of ochre, possibly pigments

Variety of strategies, including bladelet production More investment in projectiles in bone, antler, ivory, and stone Remains of dwellings present, as well as some structured hearths Elaborate Figurative portable and parietal art, personal ornaments

Source: Modified, after Roebroeks 2008.

as this could not have happened in the absence of seagoing vessels constructed to plan. Thousands of years later and a few thousand kilometres to the south, the descendants of the planners of these seagoing vessels created the archaeological record discussed earlier, with strong resemblances to the European Middle Palaeolithic one (cf. Holdaway & Cosgrove 1997). A focus on the ecology of Neanderthal and modern human hunter-gatherers has yielded more productive and straightforward alternatives to the cognitive explanations of the differences in the record mentioned previously. These explanations focus on the costs and benefits of various behavioural strategies and, in contrast to the cognition-based explanations, do account for the diversity within the record of modern human hunter-gatherers (e.g.,Verpoorte 2006). By focusing on the different trade-offs hunter-gatherers had to deal with, such explanations reduce the cherished “proxies” for language to the outcome of cost-benefit analyses. We now have a rich picture of other aspects of Neanderthal life. This comes as a result of the new genetic studies, other cutting-edge methods such as isotope studies and detailed archaeological research, often stimulated by dichotomous views. Neanderthals were thin on the ground and subject to local extinction (Hublin & Roebroeks 2009; Roebroeks et al. 2011).They lived in a wide range of environments, from full interglacial to cold steppic ones. Unlike earlier hominins, the faunal evidence indicates that Neanderthals hunted and butchered large mammals in a manner that 187

Richard Cosgrove, Anne Pike-Tay and Wil Roebroeks can be compared to Upper Palaeolithic humans (Burke 2004;Voormolen 2008).Various large mammals were hunted, the dominant species including large herbivores that live in herds such as bovids, equids and cervids, and solitary animals such as rhinoceros (reviewed in Gaudzinski-Windheuser & Niven 2009). Nevertheless, while their hunting weapons included simple wooden spears, from the beginning of the Neanderthal lineage onward (Thieme 1997), they invested little effort in producing projectiles. The isotopic signal suggests that a large proportion of their dietary protein was obtained from meat, reflecting a relatively narrow diet (Richards et al. 2000a). A narrow diet has also been inferred from the scarcity of relatively fast moving game in Middle Palaeolithic sites in the Mediterranean (Stiner et al. 2000). However, smaller game stood at least occasionally on their menu too, including tortoises, rabbits and birds, as documented from some sites on the southern edge of their range, in detail at for instance Bolomor Cave in Spain (Blasco 2008; Blasco & Fernández Peris 2009). In the late Middle Palaeolithic some Neanderthals seem to have targeted birds for their feathers (Peresani et al. 2011).There is also some evidence from northern sites for exploitation of birds, for example, from Salzgitter-Lebenstedt in Germany (Gaudzinski-Windheuser & Niven 2009). Neanderthals also gathered plant foods, some of which were cooked, using fire in ways not unlike the Upper Palaeolithic hunter-gatherers (cf. Henry et al. 2011).

Conclusion There appears to be sufficient variability in the Pleistocene archaeological records of Neanderthals and Tasmanians as to require a reconsideration of the unhelpful labels ‘archaic’ and ‘modern’ as defined for the archaeological record. There is clear variability in Neanderthal behaviour across the very large time and space evidenced in the data from Israel and northern populations, in skeletal form and diet, perhaps reflecting different environmental adaptations (Conard & Richter 2011). There are also differences in the distribution of the Neanderthal populations across their geographic range. Variations in the archaeological records of the Tasmanian and Australian Late Pleistocene are also clearly detectable. These allowed the colonisation of a vast continent using a straightforward flaked stone technology with a component of ground-edge technology considered to be the oldest in the world (Geneste et  al. 2010). Other material items and cultural adaptations and developments such as the appearance of Late Pleistocene rock art further confirm the development of symbolism and language. The evidence for prey selection in both Middle Palaeolithic Europe and Late Pleistocene Tasmania would suggest planning depth and specific land-use focus. In addition, the lithic distribution, movement and technology of Neanderthal and Tasmanian Late Pleistocene populations suggest a degree of similarity that cannot simply be explained away as Tylor (1895, 336, 340) attempted to in claiming that,“so far as stone-implementmaking furnishes a test of culture, the Tasmanians were undoubtedly at a low paleolithic stage, inferior to that of the Drift and Cave men of Europe”. Indeed at the turn of the last century, Sollas (1911, 70) also characterized Tasmanian Aborigines as living relics of Paleolithic peoples, writing of “this isolated people, the most unprogressive in the world, which in the middle of the nineteenth century was still living in the dawn of the Palaeolithic epoch”. It has been repeatedly noted that the archaeological record of Late Pleistocene Tasmanians, the most southerly peoples at the time, resonated with the European Mousterian (Jones 1977, 191), with its prevalence of scrapers and the absence of blades, microliths, hafted bone tools and carved bone ornaments. Nonetheless, the first Tasmanians were anatomically modern people, like all the first occupants of Sahul. What this shows is that the models used to describe human groups as either ‘archaic’ or ‘modern’ are faulty and are clearly unhelpful in explaining issues such as the Middle to Upper Palaeolithic transition in Europe. To really comprehend the process of changing humanity, we need to uncouple the ‘simple’ to ‘complex’ social evolutionists’ paradigm and understand the archaeological record in terms of its cultural variability that reflects the wide range of responses to solving similar social and environmental problems that our species is capable of. 188

Chapter 15 Clothing and Modern Human Behaviour The Challenge from Tasmania

Ian Gilligan

Introduction The archaeological record of Australia presents fundamental problems for the concept of modern human behaviour, with Tasmania in particular offering a special challenge. Evidence from Tasmania exposes these problems and suggests the development of clothing is relevant to key aspects of behavioural modernity. This chapter summarizes the physiology of human cold tolerance and clothing and draws a distinction between “simple” and “complex” clothing. I consider strategies for addressing the archaeological “invisibility” of Palaeolithic clothing and outline the proposed relationships between clothing and markers of modernity (e.g., archaeologically visible adornment). I then argue that the routine lack of clothing in Aboriginal Australia is reflected in a relative paucity of signs of modernity, while a suite of developments in Late Pleistocene Tasmania can be linked to greater thermal requirements for clothing. Finally, I compare this adaptive pattern of behavioural modernity in Tasmania with similar trends in Africa and Europe and discuss the differing implications of simple and complex clothing.

The Importance of Australia An Afrocentric perspective has replaced earlier Eurocentric views of the emergence of behavioural modernity (McBrearty & Brooks 2000), but an Australocentric perspective questions the whole concept of behavioural modernity as a “package” of traits that accompanied the spread of modern humans out of Africa (Brumm & Moore 2005; O’Connell & Allen 2007; Habgood & Franklin 2008). Anatomically modern humans have been present in Australia for at least 45,000  years (O’Connell & Allen 2004) yet, prior to the mid-Holocene, archaeological evidence of behavioural modernity is distinctly patchy. Furthermore, with the notable exception of Late Pleistocene Tasmania, the few identifiable elements manifest no trend to accumulate into a “package”; instead, they occur sporadically and at widely separated times and places around the continent.

189

Ian Gilligan

The Importance of Tasmania The archaeological and ethnographic records of the Tasmanian Aborigines are particularly challenging, for a number of reasons. First, their material culture at the time of European contact was comparatively minimal, even by Australian standards – indeed, it was considered by some to be rather too minimal (e.g., Jones 1977, 202–203). This perception applied also in relation to their clothing (Gilligan 2007c, 8–9). Early European visitors, such as the surgeon on Cook’s expedition in 1777, often remarked that their use of clothing was seemingly inadequate given “the rigour of their climate” (Anderson, in Cook 1784, 112). Furthermore, unlike mainland Australia, Tasmania remained isolated from external cultural influences after the terminal Pleistocene.This prolonged isolation of a modern human population is almost unique (along with highland New Guinea) and makes Tasmania an ideal case for testing assumptions and propositions regarding the emergence of modern human behaviour. For example, the Holocene developments that constitute most of the evidence for behavioural modernity in Australia are conspicuously absent in Tasmania. However, during the late Pleistocene, this most southerly region of Sahul (the exposed land mass that included New Guinea, mainland Australia and Tasmania) presents strong evidence for the emergence of a “package” of traits, at a time when thermal conditions demanded increased use of clothing for human survival (Gilligan 2007b).

Clothing and Modernity In this chapter, it is argued that some key archaeological markers of behavioural modernity are related to the development of clothing for thermal reasons and hence a relative paucity of evidence in Australia reflects reduced thermal requirements for clothing (Gilligan 2010b). The limited and fluctuating pattern of archaeological signatures relates to the local environmental conditions that have prevailed since modern humans first arrived on the continent, and it is suggested that similar patterning can be discerned in the archaeological records of Africa and Eurasia during the Middle and Upper Pleistocene. Before exploring the extent to which these proposed relationships between clothing and behavioural modernity are borne out by evidence for the emergence of behavioural modernity in Tasmania and elsewhere, some basic concepts and propositions need to be summarized.

Thermal Physiology and Clothing The physiological principles and environmental limits of human responses to cold exposure are reviewed elsewhere (Gilligan 2010a, 21–22). Briefly, the optimal ambient temperature for lightly clothed modern humans is 25° C; shivering begins at around 13° C, and the safe limit – beyond which the risk of hypothermia can become acute – occurs at a still-air temperature of approximately −1° C. The added chilling effect of wind is evident in the wind-chill index (e.g., Steadman 1995). Cold tolerance is improved through acclimatisation, and routinely unclothed populations such as the Australian Aborigines show superior cold responses (e.g., Scholander et al. 1958) – allowing them to endure temperatures as low −5° C without clothing, if little wind chill is assumed. Physiological acclimatisation explains how Tasmanian Aborigines survived with minimal protection, for example, windbreaks, use of fire, and scanty wallaby-skin capes thrown over their shoulders (Gilligan 2007c, 8–10). Similarly, while the extent of cold tolerance among the indigenous inhabitants of Tierra del Fuego astonished Darwin (1839, 234–235), their behavioural adaptations included shelters made from guanaco pelts and seal skins, and clothing in the form of sealskin capes and robes of woolly guanaco skins (e.g., Lothrup 1928, 121–123). Clothing functions as thermal insulation by trapping small pockets and layers of air close to the skin surface, reducing the thermal gradient (and hence the rate of heat exchange) between 190

Clothing and Modern Human Behaviour Table 15.1.  Features distinguishing simple and complex clothing

Property and structure

Fitted Number of layers

Thermal physiology

Still-air protection (generally) Wind-chill protection Technology (palaeolithic) Scraping implements Cutting implements (generally) Piercing implements (generally) Repercussions Impairs cold tolerance Acquires decorative role Acquires social functions Promotes modesty/shame Becomes habitual

Simple

Complex

No 1

Yes 1+

1–2 clo Poor

2–5 clo Excellent

Yes No No

Yes Yes Yes

No No No No No

Yes Yes Yes Yes Yes

the body and its surroundings (see Gilligan 2010a, 22–23 for a detailed review). The most common measure of the thermal performance of clothing is the “clo” unit (Gagge et al. 1941, 429). Generally, each layer of clothing provides approximately 1 clo, with typical Arctic clothing (4 layers) providing about 4 clo of thermal protection.

Simple and Complex Clothing The primary basis of the distinction between “simple” and “complex” clothing (Gilligan 2007b, 103–104; 2010a, 24–26) is whether garments are hung or draped over the body (“simple” clothing) or instead are shaped and fitted – that is, tailored – to enclose the torso and limbs (“complex” clothing). Physiologically, simple clothing can provide limited protection, generally around 1–2 clo, but such open-style garments are prone to wind penetration. In contrast, complex garment assemblages can readily provide 4–5 clo and offer superior protection from wind chill, sufficient to allow prolonged human survival in polar and sub-polar environments. Although it is based on physiological and structural aspects, this distinction between simple and complex clothing has archaeological implications (Table 15.1). In Palaeolithic contexts, where raw materials comprise animal skins rather than textile fibres, the manufacture of simple garments entails mainly the cleaning and scraping of hides, utilizing scraper tools of various kinds. Complex garments require additionally that the skins be cut into specific shapes (e.g., to make the separate cylinders that enclose the limbs), and these shaped pieces must be joined together (by sewing). In multi-layered complex clothing assemblages, the inner layers require more precise cutting and sewing to achieve a close fit. For these reasons, the archaeological signatures associated with complex clothes will tend to include not only scrapers but also more specialized cutting tools (e.g., blade-based forms) and tools for piercing animal skins (e.g., bone awls and eyed needles; see Gilligan 2010a, 20). Another difference between simple and complex clothes is that whereas simple garments tend to be used on a pragmatic thermal basis (being discarded when not needed for warmth), the use of complex clothing is more likely to persist and become a routine or habitual feature of human behaviour. One reason is strictly physiological: when the human body is regularly and more ­completely covered, this creates a more consistently warm micro-environment for the body, which causes impairment of cold tolerance. Another reason why complex clothing tends 191

Ian Gilligan to persist is that, with the skin surface routinely covered, decorative functions are displaced from modifying the naked skin surface to decorating the garments, favouring the emergence of purely cultural motives for wearing clothes. Psychological factors can also come into play: when clothes are worn routinely from infancy, this may engender a sense of shame (or modesty) in relation to the naked body, which encourages the ongoing use of clothes, regardless of any physiological need for thermal insulation.

The Invisible Innovation Palaeolithic clothing is archaeologically almost invisible, although its likely presence may sometimes be inferred, for instance, from the distributions of ornaments in human burials (e.g., Pettitt 2011, 140–142), and it can be rendered partially visible with indirect approaches, such as use-wear analyses on tools (e.g., Hayden 1990; Soffer 2004). An innovative approach utilises genetic studies of human lice (Pediculus humanus) to derive a date for the divergence of head and clothing lice, which may provide an estimate for when modern humans adopted clothing on a regular basis. One lice study has yielded a range of 83,000 to 170,000 years ago, with the older date corresponding to early in the penultimate Ice Age, Marine Isotope Stage (MIS) 6 (Toups et al. 2011, 31). Human skeletal morphology can sometimes provide a clue, as shown by anatomical changes associated with habitual footwear use (Trinkaus & Shang 2008). Another strategy is to utilise data from human physiology and palaeoenvironmental sciences to reconstruct past clothing requirements, based on minimum physiological requirements for human survival. We can then examine the archaeological record for the anticipated technological signatures and other correlates of clothing in varying climatic contexts (e.g., the differing technocomplexes associated with simple and complex clothing) and assess the extent to which these correspond to various proposed markers of modern human behaviour (e.g., Gilligan 2010a, 17–21).

Clothing and Behavioural Modernity The sporadic occurrence of early archaeological signs of behavioural modernity appears to indicate a link with fluctuating environmental conditions, not only in Australia but also in other parts of the world (e.g., Hiscock 1994; d’Errico 2003; Henshilwood and Marean 2003; Hiscock & O’Connor 2006; Zilhão 2007). Specifically, it is suggested here that some of the key components relate to thermal adaptations, notably clothing innovations (Table 15.2). The list includes clothing technologies (e.g., blade-based lithics and bone implements associated with making complex garment assemblages) and also some of the repercussions of wearing clothes, as well as other thermal adaptations (Gilligan 2007b, 104–105). Indeed, these latter features, while less tangible, are often now considered to be more consistent indicators of behavioural modernity compared, for example, to lithic technologies (e.g., Bar-Yosef & Kuhn 1999). The thermal adaptations and clothingrelated features include greater control of fire (e.g., structured hearths), resource specializations (e.g., targeted animal hunting, for hides and fur as well as to meet increased caloric requirements in cold environments), more elaborate artificial shelters, and – in the case of complex clothing – the increased archaeological visibility of personal adornment and symbolism. One consequence of routinely wearing complex clothing is that the important social functions of decorating the human body will be transferred onto the garments and also displaced further afield, becoming more visible archaeologically as symbolic modification of artefacts and the physical surroundings. Whereas decoration of the unclad body may leave little tangible trace in the archaeological record, the decorating of garments and displacement of symbolism onto media external to the body will increase the visibility of these functions in the archaeological record. A prime example is the distribution of thousands of beads on human skeletal remains at the Russian site of Sungir, dating to the Last Glacial Maximum (LGM) between 26,000 and 19,000 years ago, 192

Clothing and Modern Human Behaviour Table  15.2.  Archaeological markers of behavioural modernity and the suggested strength of their association with the development of clothing Strength Strong

Moderate

Archaeological signature of behavioral modernity Range extension to previously unoccupied environments (cold) Site reoccupation and modification (e.g., sheltered sites) Greater control of fire (e.g., stone-lined hearths) Specialised hunting (for hides and meat) New lithic technologies (blades) Tools in novel materials (bone awls and eyed needles) Personal adornment (beads and ornaments) Increased artefact diversity and standardization Geographic / temporal variation in formal tool categories Parietal art (and other external images and representations)

where the beads were evidently sewn onto tailored garments (Bader & Bader 2000, 29; Kuzmin et al. 2004). Similarly, beads have been recovered in Africa at sites dating to cold episodes earlier in the last Ice Age, around 72,000 years ago and perhaps 90,000–100,000 years ago (Henshilwood et al. 2004; Vanhaeren et al. 2006). In Ice Age Europe, the increased prevalence of artworks may simply reflect a shift from symbolic modification of the exposed skin surface onto alternative surfaces such as cave walls (and also into other media such as figurines), after ready access to the skin surface was restricted by its routine concealment by clothes.

Modernity in Pleistocene Australia The very limited archaeological evidence for symbolic behaviour in Pleistocene Australia confirms that a lack of such evidence does not signify any lack of capacity for behavioural modernity (Brumm & Moore 2005, 167–169; O’Connell & Allen 2007, 405; Habgood & Franklin 2008, 214). Significantly, the use of ochre (probably for body decoration, among other uses) is documented virtually from the outset, from around 40,000 years ago (e.g., O’Connor & Fankhauser 2001), but other evidence for decoration or adornment is sparse (e.g., Morse 1993a). One example occurs at Devil’s Lair on the southwest coast, where three bone beads are dated to between 19,000 and 12,000 years ago (Dortch 1984), following the LGM. Rock art is widespread although none is dated reliably to the Pleistocene; a few hand stencils in Tasmanian caves, however, probably date to the terminal Pleistocene (Harris et al. 1988; Cosgrove & Jones 1989). Ethnographically, Australian Aborigines typically wore no clothing and, given that their ancestors probably made the journey from Africa without needing to stray beyond the tropics (Bulbeck 2007), this may have been the case from when humans first arrived on the continent. Simple clothing was worn at times in cooler regions, apparently for exclusively thermal reasons, with its use corresponding to local meteorological indices such as wind chill (Gilligan 2008).The single-layer garments (primarily kangaroo and wallaby skin capes, and sewn possum-fur cloaks) were draped, not fitted, and there is little evidence for “cultural” functions that are linked more to complex clothing (cf. Kamminga 1982, 38). Even in cooler regions such as Tasmania, a complete absence of clothing was the norm throughout most of the year (Gilligan 2007c). The functions of personal adornment and social display were served by decorating the skin surface, mainly with body painting and skin scarification, for which the archaeological signatures are meagre – ochre, as noted, is present from an early date, while small tools used ethnographically for decorative scarification (cicatrices) are sometimes found at archaeological sites (e.g., McNiven 2006, 7–8). 193

Ian Gilligan The limited use of clothing in Aboriginal Australia is mirrored by the limited archaeological evidence for behavioural modernity (Gilligan 2010b, 62–66), with an absence of complex clothing being reflected in the poor archaeological visibility of adornment. Ochre provides the only widespread indicator for symbolism, and this is consistent with body painting, in the typical absence of clothing. Elements of behavioural modernity that relate to the manufacture of clothing (e.g., standardized hide-working lithic technologies and targeted hunting of hide-bearing animal species) are, with the exception of Tasmania, generally absent from the archaeological record. Interestingly, though, bone points that may have served to pierce animal skins make an appearance in cooler regions during the late Pleistocene, sometimes in association with scrapers, for example, at Devil’s Lair (Dortch 1984, 50–64), Cloggs Cave (Flood 1973) and, most conspicuously, in Tasmania.

Modernity in Tasmania Winter wind-chill estimates between −2° C and −8° C point to a heightened need for clothing in late Pleistocene Tasmania, although conditions were such that only simple clothing was required for human survival (Gilligan 2007a; 2007b). This contrasts with comparable mid-latitude regions in the Northern Hemisphere during the LGM, where conditions were generally more severe and complex clothing was required. In Tasmania, marsupial skins were an obvious choice as raw materials for clothing, while the main technological requirement was for some form of scraping implement. Faunal assemblages at cave sites in southwestern Tasmania are indeed dominated by remains of the major fur-bearing species in the area, while the lithic industry is dominated by standardized scrapers.Tasmania stands out also in terms of another archaeological marker of behavioural modernity, namely the production of bone tools. These coincident developments in late Pleistocene Tasmania are not only surprising, given the ethnographic simplicity of the Tasmanian tool kit at the time of European contact, but also unique in Pleistocene Australia in that they constitute a whole constellation of features used to identify behavioural modernity. Furthermore, each can be linked to thermal considerations generally and to clothing requirements in particular.

Resource Specialization Faunal data yield clear evidence for the targeting of a single animal species, the red-necked (or Bennett’s) wallaby (Macropus rufogriseus), the fur of which “would have provided excellent thermal insulation” (Cosgrove 1997, 54). Overall, approximately 70% of faunal remains in the cave and rock shelter sites are those of Bennett’s wallaby (Jones 1990). Their use in the manufacture of clothing is suggested by the frequency distributions of body parts, which appear to reflect separation of the skins to make cloaks (Cosgrove & Allen 2001, 413–418), while a paucity of tail bones suggests that tails were either removed whole or left attached to the rest of the skin (Cosgrove 2004, 60). Similar specialized exploitation of fur-bearing species such as wolves and arctic foxes occurs in Ice Age Europe (e.g., Soffer 1985, 310–327), where comparable patterning in skeletal remains sometimes indicates careful separation of skins from carcasses (e.g., Klein 2009, 673).

Standardized Lithics Distinctive “thumbnail scrapers” begin to appear in Tasmania leading into the LGM, around 28,000 years ago; the earliest dates are from Pallawa Trounta, between 30,000 and 27,000 years ago (Cosgrove 1999, 375). These retouched flake implements dominate the lithic industry and would have been suitable in the preparation of animal skin garments. Thumbnail scrapers disappear from the southwest by the end of the Pleistocene, although they persist into the Holocene at some sites elsewhere on the island (e.g., Moore 2000, 71). These tools probably found multiple 194

Clothing and Modern Human Behaviour

Figure 15.1.  Archaeological evidence for behavioural modernity in late Pleistocene Tasmania and location of sites mentioned in the text (illustration by the author). uses from the outset, and woodworking, for instance, might have become more prominent from the early Holocene as forests spread over the island. Fullagar (1986, 348–350) performed use-wear analyses on thumbnail scrapers from one Ice Age cave site (Kutikina), identifying hide preparation in approximately half the sample.

Bone Tools Together with stone scrapers, bone points begin to appear as temperatures declined, with the earliest dated to between 32,000 and 27,000 years ago at the Warreen site (Cosgrove 1999, 382). Many are shaped into needles, with polished ends consistent with piercing animal skins to make sewn garments (Cosgrove 1993, 167). This function is supported by a use-wear analysis of points from two sites (Bone Cave and Warreen) showing evidence for the piercing of dry skin, although in some cases this could also indicate their use as spearpoints (Webb & Allen 1990, 77–78). One intriguing aspect, however, is the complete disappearance of bone tools from the archaeological record of Tasmania during the Holocene. As Jones (1990, 283–284) suggested, the most likely reason is that warmer conditions led to a reduction in the use of garments. Presumably, these tools failed to acquire any important function other than the sewing of marsupial skins, and their manufacture was abandoned when thermal conditions improved.

The Tasmanian Challenge Collectively, these late Pleistocene Tasmanian developments provide unambiguous evidence for a suite of archaeological markers of behavioural modernity in this most southerly part of Sahul (Figure 15.1), coinciding with significant thermal stresses for the local human population. These 195

Ian Gilligan archaeological signatures of behavioural modernity can be seen as adaptive responses to the local environmental conditions, with the majority – resource intensification, standardized lithics and bone tools – being interpretable as archaeological correlates of the manufacture of clothing (Gilligan 2007b, 107–108).

The Holocene Reversal One seemingly unusual feature of the Tasmanian archaeological record is that these signs of behavioural modernity documented in the numerous cave and rock shelter sites of the southwest are reversed during the early to mid-Holocene, and this aspect is also explicable in terms of thermal contingencies. The natural protection from wind chill afforded by these sheltered sites can account for the otherwise unexpected settlement pattern (Gilligan 2007a), where humans gravitated to a higher latitude (and generally higher altitude) during the LGM, and then abandoned this rugged region when thermal conditions ameliorated. While a few cave sites elsewhere in the region show some human occupation extending into the Holocene (e.g., Cosgrove 1995b, 100), both the resource and technological specializations essentially disappear from the archaeological record.

Tasmania and Europe Parallels between Tasmania and Europe in the late Pleistocene may be striking, but the technological resemblance is more to the European Middle than the Upper Palaeolithic. Because LGM conditions were milder in Tasmania than Europe (e.g., Colhoun 2000), complex clothing was not required, hence an absence of archaeological signs of adornment. Additionally, most of the archaeological signatures of complex clothing – notably blade tools for cutting hides and eyed needles associated with finer sewing – are not seen in Tasmania. Interestingly, the presence of bone awls for piercing animal skins – typical of complex clothing in the Northern Hemisphere – is attributable to the smaller skins available in Tasmania, where a large number of wallaby skins is needed to be sewn together to make a substantial cloak (Gilligan 2007b, 109). During the Holocene, since they needed sewn cloaks neither for warmth nor for decoration,Tasmanians could revert to using single wallaby-skin capes on a purely pragmatic basis.Without any other function for bone points, these tools – along with other markers of behavioural modernity such as the lithic and economic specializations – disappeared from the archaeological record.

Adaptive Modernity The Tasmanian evidence illustrates how archaeological markers of behavioural modernity may be linked to clothing-related developments and hence to human adaptive responses to environmental changes. In this formulation, the concept of behavioural modernity becomes more adaptive than inherent (or, at least, a consequence or epiphenomenon of adaptive processes): it becomes less coupled to anatomical modernity and instead can be connected to environmental conditions (e.g., fluctuating Pleistocene climates). There need be no nebulous, mysteriously delayed emergence of these capacities due to cognitive reorganisation within the human brain or development of language abilities (cf. Klein 2000). What is visible in the archaeological record may be little more than the varying visibility of these capacities, largely consequent and contingent upon the adoption of clothing and its repercussions. Depending on whether simple or complex clothing was required, it may then be anticipated that archaeological markers of behavioural modernity at a global level will manifest variation in concert with climatic fluctuations and the inferred presence of clothing. 196

Clothing and Modern Human Behaviour

An Afrocentric Perspective In Africa, the early Middle Stone Age (MSA) and Late Stone Age (LSA) industries – typified by standardized scraper and blade technologies and facilitating the manufacture of simple and complex clothing, respectively – tend to occur in the cooler regions and during the colder phases of the late Pleistocene (Gilligan 2010a, 42–43). The same appears true of Southwest Asia in the late Middle Palaeolithic and Upper Palaeolithic. Similarly, the earliest archaeological evidence for personal adornment comprises perforated shell beads in northern Africa and the Levant, dated broadly to early cold phases in the last Ice Age (Vanhaeren et al. 2006; Bouzouggar et al. 2007). At the other end of the continent, bone awls for piercing hides first appear in southern Africa dating to cold phases – MIS 5a/b and MIS 4 – between 84,000 and 72,000 years ago (d’Errico & Henshilwood 2007). In northern and southern Africa and Southwest Asia, sporadic production of blade tools had waxed and waned throughout the Middle Pleistocene, beginning from around 400,000  years ago (Bar-Yosef & Kuhn 1999; Gopher et al. 2005). Environmental patterning in blade production becomes especially evident during the last glacial cycle: seemingly precocious southern African industries dominated by blade tools (e.g., Howiesons Poort) appear during the very cold period (MIS 4) around 75,000  years ago, and then these industries disappear from the archaeological record early in MIS 3, when climatic conditions ameliorated. Subsequently, intensified blade production (together with archaeological evidence for adornment and the manufacture of eyed needles) defines the LSA, which began during cold climatic fluctuations late in MIS 3 and became well established during MIS 2 (the LGM).

A Eurocentric Perspective Compared to Africa, mid-latitude Eurasia witnessed a more pronounced proliferation and coalescence of most components of behavioural modernity during the late Pleistocene. Among these, thermal adaptations and clothing-related developments include intensive resource exploitation (specialised hunting of hide- and fur-bearing animals), sustained settlement in new (colder) environments, long-term reoccupation of sites (particularly sheltered cave sites), more sophisticated control of fire, new tool forms and greater artefact diversity and standardization (notably scraper and blade-based technocomplexes, with bone awls and, later, eyed needles) and – last but not least – a dramatic fluorescence of art and other signs of symbolic behaviour (e.g., Vanhaeren & d’Errico 2006). This European “package” not only occurs earlier and in generally colder (i.e., higher latitude) environments than in Africa but coincides closely with local climatic fluctuations and intensified physiological requirements for thermal protection. Some of its elements can be linked to developments in clothing (Gilligan 2010a, 41–47), for which ample direct and indirect evidence exists in the archaeological record (e.g., figurines depicting clothed humans, ibid., 56–59). In Western Europe, the relative frequency of scrapers correlates strongly with colder climatic phases during the Middle and Upper Pleistocene (Monnier 2006), while early blade tool industries first appear during the penultimate Ice Age (MIS 6) (see, e.g., Delagnes & Meignen 2006). The routine use of complex clothing (confirmed by the presence of eyed needles in Upper Palaeolithic assemblages) becomes widely established across mid-latitude Eurasia after 30,000 years ago, accompanied by a “creative explosion” (Pfeiffer 1982; Renfrew 2009) in durable, archaeologically visible signs of decoration, adornment and other forms of artistic expression. Many of these Eurasian developments and their repercussions (including derivative technologies) are sustained across the Pleistocene-Holocene boundary, associated with a decoupling of complex clothing from thermal contingencies as acquired decorative and other functions rendered clothing socially indispensable (see Gilligan 2010a, 26). 197

Ian Gilligan

Tasmania and the End of Modernity The independent emergence and disappearance of the Tasmanian developments renders virtually untenable the concept of modernity as a unilinear, universal cultural and technological trajectory underpinned by uniquely human capacities. Nonetheless, the concept gains credence from the cultural trajectory that dominates the Late Quaternary record in many other parts of the world, notably in western Eurasia but also in parts of Africa, northern Asia and the Americas.This trajectory – illustrated in the European Upper Palaeolithic and the emergence of Neolithic societies – seemingly bears hallmarks of unidirectional change towards more complex forms, analogous to biological evolution. The concept of behavioural modernity is the most recent product of this evolutionary analogy, yet it remains “plagued by the deep contradiction between the processes of local adaptation and supposed universal and absolute human progress” (Porr 2010, 30). The clothing perspective addresses this contradiction, with the distinction between simple and complex clothing implying two differing trajectories for clothing-related signs of behavioural modernity. The archaeological correlates of simple clothing are governed by adaptive considerations (e.g., the late Pleistocene developments and Holocene “devolution” in Tasmania), whereas the correlates of complex clothing (e.g., acquisition of decorative functions) allow for a decoupling from environmental contingencies. In the latter case, a threshold may be crossed where the developments acquire self-sustaining momentum (e.g., elaboration of archaeologically visible symbolic behaviour and further technological specializations). At that inflexion point, the trajectory can change from adaptive fluctuations to a seemingly inherent process resembling “progress”.

Summary and Conclusions The distinction between “simple” and “complex” clothing can help to explain both the parallels and the contrasts between late Pleistocene Tasmania and comparable trends in archaeologically visible components of behavioural modernity seen in Africa and Eurasia. With simple clothing, certain components of behavioural modernity can fluctuate with environmental conditions, whereas the adoption of complex clothing not only amplifies and expands the range of archaeological signals but has a greater tendency to be sustained by acquired psychosocial factors. In Tasmania, the thermal need for clothing was limited to simple garments, even during the LGM. As a consequence, the corresponding archaeological markers of behavioural modernity (e.g., resource specialization, standardized lithics and bone tools) remained coupled closely to local environmental conditions – as was generally the case in Africa and Eurasia prior to the Upper Pleistocene. The challenge to the concept of behavioural modernity posed by the fleeting appearance of some of its elements in Tasmania can be accommodated in this perspective, as can the archaeological trajectories seen in other parts of the world – for example, the very early African developments, the relative paucity of evidence in Australia, and the “creative explosion” in late Pleistocene Europe. The archaeological correlates and consequences of clothing offer a range of intervening variables (technological, demographic, economic, psychosocial) that connect components of modernity with climatic conditions. Furthermore, a transition from simple to complex clothing carries the potential for a “package” of archaeological traits to become decoupled from climatic factors and, in persisting, to seemingly become an inherent aspect of anatomical modernity – but only when the challenge from Tasmania is ignored.

Acknowledgments This chapter derives from a paper presented at the 2009 IPPA conference in Hanoi. The author thanks Peter Bellwood, secretary-general of IPPA, and Miriam Haidle and Alfred Pawlik, convenors 198

Clothing and Modern Human Behaviour of the session on behavioural modernity, for their support and encouragement. Eric Colhoun, Richard Cosgrove and Richard Fullagar provided helpful data (on Tasmanian palaeoclimates, faunal analyses and lithic use-wear findings, respectively), Bob Steadman clarified issues relating to wind-chill estimates, and Yaroslav Kuzmin provided valuable references. David Bulbeck, Colin Groves and Peter White kindly reviewed earlier papers relating to the topic, while Robin Dennell and Martin Porr reviewed this chapter, and it has benefited greatly from their comments.

199

Chapter 16 Patterns of Modernity Taphonomy, Sampling and the Pleistocene Archaeological Record of Sahul

Michelle C. Langley

Introduction Man entered a new continent, adapted his own culture to deal with unfamiliar country, animals and plants, and conversely subjected the environment to the ecological pressures of his own technology. Jones (1968,186) The ‘behavioural modernity debate’ has generated a plethora of publications discussing singlespecies or multiple-species models, ‘bow waves’, innovation hypotheses and cultural skill accumulation and resulted in a debate that has grabbed headlines for the past 20 or more years (recent examples include McBrearty & Brooks 2000; Henshilwood et al. 2002; d’Errico 2003; Mellars 2005, 2010; Mellars et  al. 2007; Bouzouggar et  al. 2007; Powell et  al. 2009; Zilhão et  al. 2010; Shea 2011b). While significant advances in our understanding of the origins and development of our own and neighbouring hominin species’ cognition, social and economic behaviour have occurred, the impact that taphonomic processes have had on the archaeological records under study has not played as prominent a role in these debates as perhaps they should. The trait lists used by archaeologists to identify modern cognition in the archaeological record are dominated by organic-based artefacts and features (e.g., burials, shell and bone beads, mobile art, osseous technologies, evidence for marine, freshwater and terrestrial faunal exploitation) as well as fragile features such as rock art. The use of so much organic-based and archaeologically fragile evidence for identifying the first appearance of modern cognition and its consequent development in the archaeological record indicates that taphonomy should be a key consideration when identifying and describing spatial and temporal patterns. However, this has not always been the case. Several researchers have cautioned about the potential impact that taphonomy and sample size may have on our global interpretations of the early record of modern cognition (e.g., Bednarik 1992; McBrearty & Brooks 2000; d’Errico 2003; Henshilwood & Marean 2003; Wadley 2010). However, the majority has discussed taphonomy only in relation to specific artefacts or features central to the debate: for example, burials (McBrearty & Brooks 2000; d’Errico 2003), pigment 200

Taphonomy, Sampling and the Archaeological Record of Sahul (Bednarik 1992; McBrearty & Brooks 2000; d’Errico 2003), osseous technology (Bednarik 1992; d’Errico 2003; Henshilwood & Marean 2003) and the impact of sampling on faunal assemblages and their subsequent interpretation (Mellars 1989a, 1989b; Marean 1998; Marean & Kim 1998; Marean & Assefa 1999; McBrearty & Brooks 2000; Stiner 2002; d’Errico 2003; Henshilwood & Marean 2003), rather than the impact of these processes on the wider regional patterns of modernity. The potential impact that taphonomic factors may have on global interpretations of the origins and development of modern cognition becomes no more apparent than when focusing on the Pleistocene archaeological record of Sahul. The combined Pleistocene low-sea-level landmass of New Guinea and Australia plays an important role in global debates concerning the origins and spread of modern humans and the characterisation of cultural complexity present among them. As an isolated island continent that was uninhabited by hominin species prior to modern human arrival, Sahul offers a window on the cultural and economic characteristics of early colonising human groups in an unfamiliar environment that was free from inter-specific competition. Additionally, because Sahul was exclusively colonized by so-called modern humans, the relationships between material culture and hominin forms should be unequivocal, that is, Sahul should give a more unambiguous view of modern human variability than in other regions.

Previous Interpretations of Sahul In the past, archaeologists have emphasised differences between the archaeological records of Pleistocene Sahul, Middle Stone Age (MSA) Africa and Upper Palaeolithic Eurasia. Habgood and Franklin (2008, 32) stated that their review of the Sahul archaeological evidence “demonstrates that late Pleistocene Sahul is different to Middle Stone Age Africa, and Middle and Upper Palaeolithic Europe”, while Shea (2011c, 12) included “the elusive character of evidence for modern human behavior in Late Pleistocene and Early Holocene Sahul” as a problem for the traditional definition for ‘behavioural modernity’. Davidson (2003, 61) argued that while the construction of seaworthy watercraft and ground lithic technology implied “that the first Australians had modern cognitive ability; they just did not use it in the way Eurocentric archaeologists would like them to have done”. Similarly, Brumm and Moore (2005, 167, 169) stated that “the Australian record demonstrates that fully modern symbolling humans did not necessarily produce a repetitive package of symbolic traces”, and that “the Australian example effectively highlights that the absence of evidence for repeated patterning in symbolic behaviour cannot by itself be taken as evidence for the absence of behavioural modernity among past people”. Finally, O’Connell and Allen (2007, 404) stated that “in short, it [Sahul’s Pleistocene archaeological record] does not appear to be the product of modern human behaviour as such products are conventionally defined”. More specific description of these differences are given by O’Connell, who states that “the archaeological record of Sahul displays few of the commonly nominated markers of behavioral modernity through its first 25,000 years” and that “in short, this record is much simpler overall than that of the European Upper Paleolithic. . . . ‘Standard indicators’ of that quality become common in Sahul only in the Holocene” (O’Connell 2011, 21; also see O’Connell & Allen 2007). O’Connell (2006, 45) further states that “the bulk of the pre-Holocene Australian archaeological record, including nearly all of it from before the Last Glacial Maximum, is strongly reminiscent of the Middle Paleolithic, the broad archaeological category that includes the Mousterian” and that “assemblages from more than 75 Pleistocene Australian sites dated >20 ka display a narrow array of stone tool production techniques, little evidence of bone technology, and very limited indications of anything that might be characterized as art, ornament or ‘style” (O’Connell 2006, 45; emphasis added). Previous researchers, with few exceptions (Balme et al. 2009), have therefore concluded that Sahul’s Pleistocene archaeological record, particularly that prior to the Last Glacial Maximum 201

Michelle C. Langley (LGM), is largely devoid of evidence for complex (read modern) behaviour and that, as a whole, is much ‘simpler’ than its African and European counterparts and is even ‘strongly reminiscent’ of Middle Palaeolithic Eurasia. But is it? And how much of these perceived differences is the result of taphonomic and archaeological sampling factors in this region? Several commentators note that preservation and sampling may be key factors structuring the present state of knowledge about early cultural complexity and innovation in Sahul (Brumm & Moore 2005; Davidson 2007b; Balme et al. 2009), yet none has systematically investigated these issues, and some have played down their importance altogether (e.g., Habgood & Franklin 2008). Yet accurate depictions of cultural change and diversity in Sahul, and hence the place of the Sahul story in the ongoing debate about a mosaic-like or full-blown emergence of cultural complexity in modern human societies (Klein 1995, 2000; Mellars 1996, 2005; d’Errico et al. 1998; McBrearty & Brooks 2000;Wadley 2001; d’Errico 2003; Zilhão 2007), are dependent on how much and how well archaeological evidence is preserved and recovered. The extreme environments and highly variable climates of Sahul dictate that taphonomy must have played a major role in the survival of archaeological evidence in this region. Therefore, this chapter aims to investigate two main questions: How has the Pleistocene archaeological record in Sahul been impacted by taphonomic factors, and does the archaeological record of Pleistocene Sahul (particularly pre-LGM) resemble that of Middle Palaeolithic Eurasia? These two questions are investigated through the analysis of the impact of taphonomy on the temporal and spatial distribution of evidence for complex behaviour in the archaeological record of Pleistocene Sahul and a comparison of these data with those previously compiled for Middle Palaeolithic Eurasia (see Langley et al. 2008). It concludes by outlining the implications of these results for the interpretation of those patterns previously identified for MSA Africa, and the common observation that the archaeological record of Upper Palaeolithic Europe is significantly more abundant in its evidence for complex behaviours.

The Impact of Taphonomy and Sampling on Sahul’s Record of Pleistocene Complex Cultural Behaviour Data were compiled for this study from published and unpublished sources, yielding 2,096 absolute ages and 464 dated instances of complex behaviour from 223 archaeological sites.This dataset dramatically expands earlier compilations undertaken by Jones (1968; 1973; 1979) and Smith and Sharp (1993). As in these earlier reviews, any site (enclosed or open) containing a Pleistocene archaeological deposit (> 10,000 years BP) was included in the dataset (Figure 16.1). Information for each site was systematically reviewed to determine geographic location, excavation history and assemblage composition and to compile absolute (14C, AMS, OSL, U-series, TL) and relative (ESR, biostratigraphic) ages. Following Smith and Sharp (1993), each site was assigned to one of six depositional contexts: rock shelters/caves, alluvial terraces, coastal dunes, lunettes, wetlands/swamps, or other open sites (e.g., artefact scatters). Sites were present in all six depositional contexts, and all environmental regions (equatorial, tropical, sub-tropical, semi-arid grassland, temperate and desert) as defined by the Australian Bureau of Meteorology (based on annual rainfall) (http://www.bom.gov.au), thereby providing a sample representative of the diversity of depositional contexts across Sahul. Radiocarbon ages were calibrated using OxCal v.4.1 (Bronk Ramsey 2009). Atmospheric ages were calibrated using the calibration datasets SHCal04 (McCormac et al. 2004) for ages between 0 and 11,000 years BP and INTCAL09 (Reimer et al. 2009) for ages between 11,000 and 50,000 years BP with a +41 ± 14 Southern Hemisphere offset (McCormac et al. 2004). Marine ages were calibrated using MARINE09 (Reimer et al. 2009) and the ΔR regional offset recommended by Ulm (2006). To examine the spatial and temporal distribution of evidence for complex behaviours in Sahul, artefacts and features identified as representing ‘complex cultural behaviour’ as defined in Africa 202

Taphonomy, Sampling and the Archaeological Record of Sahul 0

1000km

Rockshelter Coastal Dune Other Open sites Lunette Alluvial Terrace Wetlands and Swamps

Temperate Sub-Tropical Grassland Desert Tropical Equatorial

Figure 16.1.  Distribution of 223 Pleistocene sites identified in Sahul (illustration by the author). and Europe were recorded as separate instances in time and space. Evidence used to identify ‘complex cultural behaviour’ (behavioural modernity) typically includes artefacts and features suggested to indicate abstract thinking, planning depth, innovativeness and the appearance of symbolism (e.g., Chase & Dibble 1987; McBrearty & Brooks, 2000; Wadley 2001; Klein & Edgar 2002; d’Errico 2003; d’Errico et al. 2003; Henshilwood & Marean 2003; Conard 2008). These categories have traditionally included archaeological objects and features indicative of personal ornamentation, social communication, ritual disposal of the dead, complex subsistence behaviours, formalised or complex technologies, long-distance social interaction, deliberate environmental modification, exploitation of new habitats and long-distance water crossings. While the appropriateness of some of these categories has been challenged, they have been employed in all previous studies of complex behaviours, both in Pleistocene Sahul and in Africa and Eurasia (e.g., McBrearty & Brooks 2000; d’Errico 2003; Brumm & Moore 2005; Mellars 2005; Habgood & Franklin 2008), therefore necessitating their use in this study in order to make its findings comparable with previous research. For this study, an archaeological instance of complex behaviour was defined as a single type of artefact or feature indicative of complex (modern) behaviour from a single site in a securely dated context, regardless of the abundance of that artefact type or feature. Thus, a single incised object or a group of 250 fragments of pigment from the same dated context each represents a single instance in the study. Where several distinct instances occur within a site and are dated by two 203

Michelle C. Langley or more distinct dates, each would represent a separate instance. This approach therefore favours diversity of artefact types per dated context as opposed to quantity of artefacts, although the two may be correlated. The date of first colonisation of Sahul is imprecise, and is best estimated as 50,000 ± 10,000 years BP (Hiscock & Wallis 2005). This level of uncertainty, as well as the large error ranges for some dates, makes determining both the earliest instances of complex behaviour and rates of initial change in complexity within Sahul very difficult. To avoid attributing undue precision to early instances of complexity in Sahul, I assigned all individually dated instances to 5,000-year intervals spanning the period from 55,000 to 10,000 years BP. If an artefact or feature was associated with more than one date, or only relative dates were available, a mean for these dates was calculated to determine the appropriate chronological unit.This creates some temporal ambiguity but captures the uncertainty that exists over the age of some instances. Owing to differences in the detail of available information, not all artefact type, site, excavation and absolute date fields could be completed for every site.The excavators of sites with missing data were contacted directly to address these gaps; however, the required data were often not available because it either was not collected at the time of excavation or had subsequently been lost. This meant that not all 223 sites could be included in every analysis undertaken in this study. Sub-samples of various sizes were used for each analysis, which included as many of the 223 sites as possible. Advantages of this dataset over those employed in previous studies are that both sites containing evidence for complex behaviours and sites without archaeological evidence for complex behaviours are included. Additionally, evidence was defined as either symbolic or non-symbolic instances in order to make more direct comparison between instances more clearly indicative of abstract thought and symbolism versus those more clearly representing forward-planning and innovation. This dual system of data exploration allowed both the comparison of the results with those of previous studies and a more rigorous analysis of only those artefacts argued to represent symbolic behaviour (as opposed to all complex behaviours), which is felt to provide a better indication of changes in social complexity. This approach allowed the generation of a more complete picture of the frequency and distribution of archaeological indicators of complexity as well as a better understanding of how complex cultural behaviours are represented in the Pleistocene archaeological record of Sahul. While both organic and inorganic artefacts and features are used to identify complex behaviour in the archaeological record, about two-thirds (59.8%) of the evidence for complex behaviour identified in Pleistocene Sahul is organic based. These include burials, beads, bone, shell and wooden tools and evidence for fishing and shellfishing. This result is the reverse of what one might expect to find since inorganic instances should be less susceptible to post-depositional destructive processes. However, as previously mentioned, the traditional trait lists utilised by researchers consist heavily of organic as well as fragile inorganic (rock art) artefacts and features. This result is therefore consistent with the methods used to investigate complex behaviour in the archaeological record and is further explained when the structure of Sahul’s archaeological record is examined. Sahul’s Pleistocene archaeological record is dominated by enclosed rock shelters and caves (61.8%) rather than open (e.g., lunette, coastal dune, alluvial terrace, wetlands and swamps) contexts. More than twice the amount of evidence (68.9%) is found in enclosed contexts than open contexts. Furthermore, when we focus on the enclosed contexts, it can be seen that the alkaline environment of limestone (versus sandstone) shelters has preferentially preserved the majority of organic based evidence (86.7%, n = 144 instances out of 166 instances in total found in enclosed limestone or sandstone contexts). Thus, there is differential preservation of evidence for complex behaviour between enclosed and open depositional contexts and limestone versus sandstone contexts. Together, these data alone suggest that a substantial amount of evidence for complex behaviours in Pleistocene Sahul may have been lost, so that what we have to study today is a fraction of what was originally deposited. 204

Taphonomy, Sampling and the Archaeological Record of Sahul 180

Pleistocene Sahul

160

Non-Symbolic

No. of Instances

140

Symbolic

120 100 80 60 40

148 Sites

20

464 Instances

0 10 15kya

16 20kya

21 25kya

26 30kya

31 35kya

35

36 40kya

41 45kya

46 50kya

51 55kya

Middle Palaeolithic Eurasia

No. of Instances

30 25 20 15 30 Sites

10

49 Instances

5 0 40 60kya

61 80kya

81 100kya

101 120kya

121 140kya

141 160kya

Figure  16.2.  Comparison of the archaeological record of complex behaviours in Pleistocene Sahul and Middle Palaeolithic Eurasia (illustration by the author). Evidence for symbolism in Sahul includes 128 instances from six categories, incorporating ornamentation, pigment use, notational pieces, rock art, ritual burial and cultural cranial deformation. Although sites are represented back to more than 50,000 years BP, evidence for symbolic behaviour first appears between 41,000 and 45,000 years BP (Figure 16.2). The use of pigment at sites such as Malakunanja II (ochre crayon) (Roberts et al. 1994), Carpenter’s Gap 1 (painted roof fall) (O’Connor 1995; O’Connor & Fankhauser 2001) and Lake Mungo (exotic ochre) (Allen 1972; Bowler 1998) are the first evidence for symbolic activities in Sahul. Pigment use provides the most abundant form of symbolic evidence throughout the Pleistocene. The early use of pigment at Lake Mungo is potentially the most significant of these owing to its use in the WLH3 ritual extended burial and the transport of ochre over circa 200 km to the burial site (Bowler et al. 2003).The world’s earliest cremation is also found at this site, dating to circa 40,000 years BP, together demonstrating the ritual complexity of early populations in southeast Australia (Bowler et al. 1970; 2003). The first evidence for rock art appears early in the record between 36,000 and 40,000 years BP at Carpenter’s Gap 1 (O’Connor & Fankhauser 2001), and it remains one of the main components 205

Michelle C. Langley of the symbolic record throughout the Pleistocene. The decoration or transfer of cultural meanings through this medium therefore contributes one of the oldest symbolic behaviours in Sahul, as well as the most frequently identified in the archaeological record (see Langley & Tac˛on 2010 for an analysis of dated Australian rock art). Burials at Lake Mungo (Allen 1972; Bowler 1998) and personal ornamentation from Riwi (pierced shell beads) (Balme 2000; Balme & Morse 2006), Mandu Mandu Creek (pierced shell beads) (Morse 1993a; Balme & Morse 2006) and Buang Merabak (drilled shark’s tooth) (Leavesley & Allen 1998) first appear between 36,000 and 40,000 years BP; however, these traits contribute only sporadically to the record of complex behaviour until 10,000–15,000 years BP. Evidence for notational pieces appears in just three time slices, once at 31,000–35,000 years BP at Spring Creek (tooth with incised grooves) (Vanderwal & Fullagar 1989), twice at 21,000–25,000 years BP at Devil’s Lair (engraved limestone plaque) (Dortch 1976) and Cave Bay Cave (macropod femur with grooves) (Stockton 1981) and finally a single instance during the last 5,000-year time slice again found at Devil’s Lair (engraved limestone plaque) (Dortch 1976). Finally, intentional cranial deformations along with possible cemeteries at Coobool Creek and Kow Swamp (Brown 1987a; Pardoe 1993) appear in the last 5,000  years of the Pleistocene. The significant increase in the quantity of symbolic evidence in the last 5,000 years of the Pleistocene suggests either that there was an increased need for symbolic communication amongst the populations of Sahul, perhaps caused by increased population density or economic stress as a result of the LGM, or simply that these artefacts and features are better preserved in the terminal Pleistocene. A combination of these two scenarios probably represents the best interpretation of these data. A total of 336 instances of non-symbolic complex behaviours from 16 different categories were identified from 123 Pleistocene sites. Instances appear in all time slices with innovative and complex technologies (waisted axes) comprising both the oldest non-symbolic instances and the oldest instances in the total dataset (51,000–55,000 years BP time-slice) (Groube et al. 1986; Roberts 1997; Summerhayes et al. 2010). The 20,000 years following the appearance of waisted axes sees the appearance of a wide variety of new complex and standardised technologies including bone tools (points, spatulas, awls), shell tools (adzes, containers, flaked shell tools) and dedicated foodprocessing tools such as grindstones (found at sites including Malakunanja II, Nauwalabila 1 and Willandra Lakes). The oldest-known wooden projectile technologies appear a little later at around 25,000 years BP (as indicated by imprints of shafts at Lake Mungo (see Webb et al. 2006). Additional evidence for wooden artefacts along with other perishable technologies such as clothing and textiles are found in rock art, which includes images of elaborate clothing, head dresses, baskets, nets and other textiles (Welch 1996). Their inclusion in these images demonstrates that these items were part of the material culture repertoire of Pleistocene populations dating back to at least circa 20,000 years BP when these images are argued to have been painted (Taçon & Chippindale 1994; Flood 1997). Evidence for forward planning and innovation in subsistence strategies makes up the dominant component of the non-symbolic dataset. These include the exploitation of resources that were difficult to obtain and demanding environments requiring adaptations to subsistence as well as possible risk-reduction strategies. To exploit these environments successfully, colonists developed and undertook a range of activities such as the detoxification of plants (e.g., Macrozamia pits at Cheetup, Smith 1996; taro at Kilu, Gosden 1995), exploiting arboreal or flighted animals including cuscus, bats and a wide range of birds (as evidenced at Buang Merabak, Balof 2 and Batari; Steadman et al. 1999), exploiting deep sea resources including several species of large sharks, tuna and mackerel (as at Balof 2, Leavesley 2007; Kafiavana, White 1972; Buang Merabak, Leavesley 2007), and the transportation of staple flora and fauna such as the cuscus (Phalanger orientalis), rat (Rattus mordax) and bandicoot (Echymipera kalubu) to islands previously devoid of these species (Allen et al. 1989; Fredericksen et al. 1993; Leavesley 206

Taphonomy, Sampling and the Archaeological Record of Sahul & Allen 1998). People also modified their landscape as evidenced by possible fire-stick farming (Turney et al. 2001b) and vegetation clearing (Groube 1989; Denham et al. 2009). The deliberate modification of the landscape together with the introduction of favoured staple species to these landscapes suggests that colonising populations were fully capable of altering their environment to suit their needs. The first evidence for long-distance social interaction or exchange appears as marine shells, shell beads and ochre transported more than 200–300 km at sites including Riwi and Lake Mungo from 36,000 to 40,000 years BP (Allen 1972; Balme 2000; Balme & Morse 2006). Other exotic raw materials were also being exploited and moved long distances from this time, such as Darwin glass (> 50 km) in western Tasmania and obsidian in Melanesia (Allen et  al. 1989; Jones 1995; Cosgrove 1999;Torrence et al. 2004). Although oceangoing watercraft are not directly known from Pleistocene sites, and hence are not included in this analysis, obsidian was transported from quarries on the New Guinea mainland to Melanesian islands more than 23 km away, requiring the manufacture and use of sturdy watercraft around circa 35,000–45,000  years BP (Torrence et al. 2004). Allen and O’Connell (2008) additionally argue that watercraft must have been employed from before 40,000 years BP to obtain the deep-sea fish found in Melanesian sites (including Buang Merabak, Matenkupkum, Kilu), as well as to make journeys of up to 175 km to remote islands such as Manus and the Solomons. Evidence for the non-symbolic traits of behavioural modernity is found in all time slices stretching from 10,000 to 55,000 years BP (Figure 16.2). The temporal distribution of non-symbolic evidence differs from that of the symbolic evidence. While the symbolic evidence appears fairly evenly distributed until the last 5,000-year time slice, the non-symbolic evidence clearly continues to increase from the 51,000- to 55,000-year time slice. In all, the appearance of each single trait, both symbolic and non-symbolic, is different from every other, and it is only when they are combined that it becomes clear that evidence in total is increasing in quantity and diversity. Comparison of the temporal distribution of evidence against the types and number of sites occupied in 5,000-year time slices presents several important observations.The first is that organic evidence appears in the dataset only after the first enclosed site appears in the sample (51,000– 55,000 years BP). The second is that evidence, particularly organic evidence, begins to steadily increase only after limestone shelters appear in the sample (41,000–45,000 years BP). Finally, and probably most significantly, it is not until the 26,000- to 30,000-year BP time slice that more than 10 open and 10 enclosed sites appear in the sample. This result means that the sample size currently available for analysis pre-30,000  years BP is significantly smaller than that for post30,000 years BP. In fact, both the temporal distribution of sites occupied and the distribution of evidence identified easily fit Surovell and Brantingham’s (2007) and Surovell et al.’s (2009) taphonomic degradation curve, and when the number of sites occupied versus the number of instances identified in each 5,000-year time slice is plotted, we find that there is a significant correlation (R2 = 0.9541). Therefore, the abundance of evidence identified in each 5,000-year time slice is tied to the sample size available (the number of sites identified as occupied) for each time slice. Correlations can also be identified when examining the impact of different environmental regions on the spatial distribution of evidence in Sahul. If we examine the distribution of sites occupied prior to the LGM and overlay them onto Hope et al.’s (2004) map of environmental regions during the LGM, we find three key results: 1. No sites are located in the sub-tropics. 2. Sites in climatically unstable regions are mostly located in enclosed contexts (73%). 3. Open sites are largely located in the temperate and grassland regions of the southeast and southwest (60%) or the climatically stable northern equatorial zone (20%). 207

Michelle C. Langley A similar analysis of those sites identified as occupied post LGM and overlayed onto a modern map of environmental regions (Figure 16.1) provided by the Australian Bureau of Meteorology (based on annual rainfall) also finds in three main results: 1. Only 5% of sites are located in sub-tropical and tropical zones. 2. Of Pleistocene open sites, 75% are located in the grassland and temperate regions along with the equatorial north. Here proportions of organic artefacts actually exceed those of inorganic artefacts. 3. While desert regions contain few sites, the proportion of organic to inorganic instances is likewise higher than for tropical and sub-tropical regions. This distribution pattern suggests that past climatic zones have affected the survival of Pleistocene sites, with site (and artefact) survival far more common in the regions (temperate, grassland and equatorial) where preservation conditions are expected to be at their best. These spatial distribution results have implications for Franklin and Habgood’s ‘zones of innovation’. In 2007 Franklin and Habgood (2007, 11) proposed several zones of innovation, which were argued to reflect the early appearance of different kinds of evidence for complex (modern) behaviour – for example, bone tools, the exploitation of Bennett’s wallaby, rock art and longdistance transport or exchange networks in the Tasmania zone. However, when these zones are overlayed with the modern environmental regions, it can be seen that each of these zones correlates with a different environmental region in Sahul (e.g., the Northern Australia zone and the tropical environmental region, the Central Australia zone and the desert environmental region, or the Tasmania zone and the temperate environmental region). It could therefore be argued that the early appearance of different types of artefacts (e.g., bone tools, waisted-axes, shell beads) in each region may be the result of the differing preservation circumstances unique to each environmental region rather than reflecting independent zones of innovation.This interpretation would account for the more abundant organic evidence cited for the Southwestern Australia, Murray-Darling and Tasmania zones of innovation, which coincide with those environmental regions that were shown to have the best conditions for long-term preservation of organic artefacts and features. These results do not imply that peoples occupying each of these regions were not developing new and regionally distinct technologies, only that the role of preservation in their long-term survival to discovery must be considered and accounted for. These results obtained through the analysis of data from 223 Pleistocene sites located throughout Sahul demonstrate that taphonomy and sample size are significant factors in the patterning exhibited in Sahul’s Pleistocene archaeological record. Spatially, this impact was the preferential survival of organic evidence in enclosed limestone contexts or sites located within the southeast, southwest and equatorial north of the continent. The impact on the temporal distribution of evidence is even more apparent, with a significant correlation between the sample size available (number of sites occupied) and the quantity of evidence recorded through time.Various explanations can be offered for this result, including increasing population size, increasing use of marginal landscapes and increasing site survival. A combination of all of these factors appears the most parsimonious explanation.

Pleistocene Sahul as Reminiscent of Middle Palaeolithic Eurasia? Having identified the impact of taphonomic processes on Sahul’s Pleistocene archaeological record, we can now compare it to the Middle Palaeolithic Eurasian and Pleistocene modern human records of Europe and Africa respectively. The apparently sparse evidence for complex cultural behaviours, especially symbolic behaviour, in Sahul has often been used to demonstrate that absence of evidence is not evidence of absence in terms of modern cognition (e.g., Brumm & 208

Taphonomy, Sampling and the Archaeological Record of Sahul Moore 2005; O’Connell 2006, 2011; O’Connell & Allen 2007; Shea 2011b), most frequently with reference to the cognitive capacities of Neanderthals. But are these two archaeological records as similar in character as often stated? In 2008, Langley et al. undertook a review of the evidence for complex behaviour in Middle Palaeolithic Neanderthals. This study identified a pattern of increasing abundance and diversity in the evidence for complex behaviour in this archaeological record towards 40,000 years BP. At the time this review was published, data pertaining to the impact of taphonomic processes or increasing population densities during the Middle Palaeolithic was unavailable, and this continues to be the case in relation to taphonomy, however, recent research by Piercy (2009) indicates that innovation rates may be tied to both population size and density. Despite the slight vertical lengthening in the distribution of evidence for complex behaviours in Middle Palaeolithic Neanderthals created by the use of 20,000-year time slices in contrast to the use of 5,000-year time slices for the Pleistocene Sahul data, both records display a remarkable similarity in the increasing abundance and diversity of evidence identified through time. If the pattern present within Sahul’s Pleistocene archaeological record is the result of both increasing population density and taphonomic factors, does this imply that a similar trend for Middle Palaeolithic Eurasia was also the product of similar demographic and taphonomic factors? Further detailed investigation of the taphonomic processes at work within this record is required before this question can be answered. This observation also begs the question of whether the pattern identified by McBrearty and Brooks (2000) for MSA Africa may also be partially explained by taphonomic factors. When the distribution of different kinds of evidence (e.g., beads, grindstones, shellfishing) in their presented figure 13 (2003, 530) is examined, it becomes apparent that the earliest evidence to be identified in this archaeological record is made up of inorganic artefacts and features (blades, grindstones, points, pigment processing) with organic artefacts and features appearing only later in the record. Furthermore, rock art, arguably the most fragile of evidence used to identify behavioural modernity in the archaeological record, appears only during the terminal Pleistocene and is the last trait to appear in this record. Clearly, intensive investigation of the impact of taphonomy on the temporal distribution of the evidence used to identify complex (modern) behaviour in Africa’s Pleistocene archaeological record is necessary to advance our understanding of the identified pattern. Returning to the comparison of the archaeological records of Pleistocene Sahul and that of Middle Palaeolithic Eurasia, several key observations can be made when these two records are compared in detail (Figure 16.2). First, while the same trend in increasing abundance of evidence can be identified, the Middle Palaeolithic Neanderthal archaeological record includes only 30 sites identified as holding evidence for complex behaviour (constituting 49 instances). These 30 sites and 49 instances cover a period of 120,000 years (one instance per 2,448 years). On the other hand, the archaeological record of Pleistocene Sahul holds 148 sites that contain evidence for complex behaviour (constituting 464 dated instances; one instance per 96 years), and all of these are identified for a period of only 45,000 years – a third of the time span examined for the Middle Palaeolithic. Although the area inhabited by Neanderthals is hard to define definitively, current evidence suggests that it is roughly the same size as Greater Australia, and thus geographic spread cannot explain this difference in result. In terms of the frequency of evidence, 39.9% of Pleistocene sites thus far identified in Sahul have evidence for symbolic behaviour (burials, ornamentation, pigment use, rock art, notational pieces [mobile art], cultural cranial deformation), 55.1% of Pleistocene sites have evidence for non-symbolic complex behaviour (e.g., grindstones, ground-edge axes, hafting, long-distance transport) and 66.3% of Pleistocene sites thus far identified in Sahul hold some form of evidence for complex (‘modern’) behaviour. On average, a Pleistocene site in Sahul held 2.1 instances of complex (either symbolic or non-symbolic) behaviour. Furthermore, in terms of the diversity of traits identified in Sahul’s Pleistocene archaeological record, only one new trait appears in the 209

Michelle C. Langley record after the LGM – cultural cranial deformation (as evidenced, for example, by Kow Swamp 5, ca. 13,000 years BP, and Coobool Creek 65, ca. 12,500 years BP). Every other trait has appeared in the record prior to the LGM – and, once it has appeared within the record generally remains evident up to the Pleistocene-Holocene boundary. These results therefore indicate that evidence for complex behaviour in Sahul is significantly more frequent and diverse than is commonly stated in the literature and is far from being ‘elusive’ in character. Therefore, while I agree with O’Connell (2006; 2011) and O’Connell and Allen (2007) that artefact diversity and density may not be the best source of information for understanding Neanderthal cognition as the relative simplicity of the Neanderthal archaeological record may be in response to low demographic density rather than reflecting cognitive abilities – the comparison of the archaeological records of Middle Palaeolithic Eurasia and Pleistocene Sahul is unjustified. Detailed comparison reveals that the latter is significantly more abundant, diverse and frequent in evidence for complex behaviour than the former, even while having been heavily impacted by taphonomic processes.

Discussion Two questions were posed at the start of this chapter: How has the Pleistocene archaeological record in Sahul been impacted by taphonomic factors, and does the archaeological record of Pleistocene Sahul (particularly pre-LGM) resemble that of Middle Palaeolithic Eurasia? The data presented here demonstrate that the temporal and spatial patterning of complex behaviours (‘behavioural modernity’) in Pleistocene Sahul are not the result of “chronological and geographic patterning” (Habgood and Franklin 2008, 210) but are rather heavily influenced by taphonomic factors. When the archaeological record is compared in detail to that of the Middle Palaeolithic, there are substantial differences in the quantity and diversity of evidence for complex behaviour between the two records. While it is true that Sahul’s lithic industries appear considerably different from those in contemporary modern human African and Eurasian records, this difference alone should not warrant the comparison of the record from Sahul with the European Middle Palaeolithic Neanderthal record. As Balme et al. (2009) have suggested, Sahul’s Pleistocene peoples may have used an extensive organic technology in place of lithic ones, which would not have survived over the time periods under study. The unique Wyrie Swamp finds of boomerangs and barbed spears, along with the various bone and shell technologies, and depictions of baskets, spears, clothing and headdresses in rock art supports this hypothesis. Collectively, the evidence for early complex behaviours in Pleistocene Sahul shows a remarkable degree of flexibility and innovation in symbolism, subsistence and technology. But what about the commonly cited differences between the archaeological record of Pleistocene Sahul and its modern human contemporaries in Africa and Eurasia? To date, no quantitative analysis of either the MSA African or Upper Palaeolithic Eurasian archaeological records has been carried out, and therefore no direct comparison with Sahul’s record can be undertaken. However, the observation of the similarity between the temporal patterning in the appearance of organic and inorganic artefact or feature types identified in Sahul and those in MSA Africa provides an intriguing example of where further quantitative comparisons should be carried out. Furthermore, several observations on the Magdalenian archaeological record indicate that the quantitative differences between Sahul and archaeological records of modern humans elsewhere may not be as striking as commonly believed. While it is true that the mobile art seen in Upper Palaeolithic assemblages do not appear in Sahul’s Pleistocene archaeological record, it has been pointed out that 80% of the carved items recovered from the Périgord come from only four sites (Laugerie-Basse, La Madeleine, Limeuil and Rochereil) and that 48% from only two sites (St Périer 1965; Nougier 1970; Bahn 1977; White 1985). Bahn (1977, 251) comments that “similar figures would emerge from a calculation of this kind in the Pyrenees – hence the remarkable 210

Taphonomy, Sampling and the Archaeological Record of Sahul uniformity of several types of artistic object found in various Pyrenean sites (St Périer 1957)”. He further discusses the possibility that solitary gifted artists may be responsible for the production of a number of iconic artefacts, the best known being the fawn/bird spear throwers found at Mas d’Azil, Bédeilhac, Arudy, Isturitz and Enlène and the spiral decorated rods found at Isturitz, Arudy, Lourdes, Les Harpons, Lespugue and possibly Duruthy (Bahn 1982, 256–257). St Périer (1920, 227) remarked in relation to the spiral decorated rods that the decoration was so particular that the artefacts may have been the “work of a single craftsman or at least of one tribe”. In relation to Cantabrian Magdalenian sites, Straus (1992, 157–160) states: The distribution of mobile art objects is very uneven: some caves are true ‘supersites’ (to borrow Paul Bahn’s term), whereas most sites are poor or lacking in such artifacts . . . three of these supersites (Altamira, El Castillo, and Tito Bustillo) are also among the half-dozen or so richest, most complex cave art sites in the region, but the others have little or no rock art. On the other hand, several major cave art sites that also have significant, excavated cultural deposits are quantitatively very poor in mobile art (notable Ekain and Peña de Candamo). . . . Many other sites have several decorated sagaies or other pieces. . . . At several other Magdalenian sites, there is only one work of mobile art but that solitary object is extraordinary, leading one to wonder under what (deliberate?) circumstances it may have been abandoned. . . . The only figurative art object found in our recent re-excavation of La Riera (in late Magdalenian Level 24) is a small, flat bone fragment intricately engraved with an animal that is difficult to identify with certainty. Even the much larger-scale excavations by Vega del Sella in La Riera yielded no more than a few engraved sagaies and a harpoon. These examples indicate that while the mobile and rock art identified in some Cantabrian, Pyrenean and Périgordian Magdalenian sites is extraordinary in artistic quality, many others contain little or no evidence of these items. Furthermore, it must be remembered that Pleistocene Australians may have had an extensive repertoire of organic artefacts that did not survive to discovery (but was depicted in rock art), as has been suggested by Balme et al. (2009) and others. Clearly, determining the frequency of instances of complex behaviour in the archaeological record of the Magdalenian (and wider European Upper Palaeolithic) and the intensive investigation of the impact of taphonomy on the temporal distribution of the evidence used to identify complex (modern) behaviour in both Eurasia’s and Africa’s Pleistocene archaeological record are necessary to advance our understanding of the identified patterns.

Conclusion Far from simple and unchanging, the Pleistocene societies of Sahul are represented by a diverse range of symbolic and non-symbolic artefacts and features that collectively hold testament to the adaptive and cultural flexibility of modern human colonists that rapidly modified the cultural components they brought from Africa and Asia to suit dramatically different environments, developing innovative new technologies, subsistence strategies and ritual practices in the process. Unlike the archaeological record for Middle Palaeolithic Eurasia, evidence for complex behaviour is a common feature on this southern continent, appearing in two-thirds of sites thus far identified. And even though the sample size for pre-LGM Sahul is significantly smaller than post-LGM Sahul, all but one type of evidence for complex behaviour identified on this continent is evident in the pre-LGM archaeological record. Furthermore, the decoration or transfer of cultural meanings through the use of pigment contribute one of the oldest symbolic behaviours in Sahul, as well as the most frequently identified throughout the circa 50,000-year-long archaeological record. In all, these results argue for rejecting comparisons of Sahul’s Pleistocene archaeological record with that of Middle Palaeolithic Eurasia. Instead, the differences in this southern 211

Michelle C. Langley archaeological record should be interpreted as simply reflecting the variability of modern human adaptation and material culture present during the Late Pleistocene.

Acknowledgments Thanks to all those Australian archaeologists who provided unpublished data for my MPhil thesis which was the basis of this chapter – in particular, Peter White. Many thanks to my MPhil supervisors Sean Ulm and Chris Clarkson, and also to Nick Barton, Ian Lilley and the editors for comments on the draft of this chapter.

212

Chapter 17 Late Pleistocene Colonisation and Adaptation in New Guinea Implications for Modelling Modern Human Behaviour

Glenn R. Summerhayes and Anne Ford

Introduction The islands of New Guinea and the Bismarck Archipelago are a large tropical landmass supporting some of the greatest biodiversity in the world, not to mention in excess of 800 of the earth’s languages.Yet during the late Pleistocene, New Guinea was joined to Australia to form the landmass known as Sahul, with the islands of New Britain and New Ireland remaining as separate entities throughout human occupation. The migration of modern humans to Sahul was part of a global expansion out of Africa, during which these early migrants colonised and adapted to a wide variety of environments and landscapes. This chapter reviews the nature of the initial colonisation for the northern part of Sahul, comprising New Guinea and the Bismarck Archipelago. First occupied between 50,000 and 45,000  years ago, this region witnessed a rapid spread of humanity into the wide range of environments that form mainland New Guinea, and across the water to New Ireland, New Britain and the northern Solomon Islands. Sites from this region will be assessed in terms of both chronological and behavioural data to provide a timeline of colonisation, which can then be used to investigate the nature and speed of how people moved into the new and unfamiliar landscapes of New Guinea. Archaeological research from this region may prove instructive in modelling modern human dispersals from other regions of the world.

Colonisation of Sahul: Entry from the West Modern humans emerging eastwards out of Africa moved along the Southern Dispersal Route more than 60,000  years ago (Lahr & Foley 1998; Mellars 2006c; Petraglia et  al. 2010), passing through South Asia and into Southeast Asia. At this time, the area currently comprising the Malay Peninsula and the western Indonesia islands was joined together to form the continent of Sunda. This period, known as the Marine Isotope Stage 3 (MIS 3), was marked by cold and dry conditions that eventually deteriorated to the Last Glacial Maximum (LGM) (Daniau et al. 2009). Although climates were cooler than today, resulting in lower sea levels during the late Pleistocene (Farrera et al. 1999), migrations into the continent of Sahul from Sunda would have involved a 213

Glenn R. Summerhayes and Anne Ford

Northern Corridor

SUNDA

BORNEO

New Ireland

NEW GUINEA

N New Britain Northern Solomons

Southern Corridor

0

325

650

SAHUL 1300

KMS

SCALE

Figure 17.1.  Possible corridors of entry into Sahul. Sea levels, shown at −66 m, are based on proposed sea level at 45,000 years from ‘Sahul Time’ (illustration by G. Summerhayes and A. Ford). series of water crossings across the Wallacean Archipelago, as part of an island hopping movement. The nature of the movement and specific route taken into Sahul has been widely debated and will not be repeated here. Suffice to say that the ability to use watercraft and undertake island-hopping journeys influenced early modelling of Sahul colonisation to suggest that the first colonisers were relatively adept seafarers specialised in utilising coastal resources (Bowdler 1977). By following coastlines, exploiting shellfish and marine resources, it has been argued that rapid colonisation of new areas could take place. Proponents of this model, Allen and O’Connell (2008; O’Connell et al. 2010) argue a convincing case for the intentional movement of people along Birdsell’s (1977) Northern Route into Sahul (Figure 17.1), as this route allowed for a better chance of survival in terms of being able to complete the return voyages critical for maintaining a colonising population. Taking into account birth and death rates and the availability of mates, they note that “successful colonization of an uninhabited landfall probably requires the near simultaneous arrival of several groups, each including at least 5–10 women of reproductive age . . . the odds that members of one small group cast adrift by chance could later generate a persistent population . . . are slim” (O’Connell et al. 2010, 59). The speed, distances covered and overall success of colonisation suggest that voyaging was deliberate and two way.

Timing the Initial Occupation of New Guinea Testimony to the successful nature of colonisation can be seen in the speed of colonising groups. The evidence suggests that these early colonisers did not waste time in moving to the islands east of New Guinea soon after colonising the mainland. Such movements required maritime technology as the islands of New Britain, New Ireland and the northern Solomons all remained separate entities, free from the mainland and each other. Before exploring the speed of colonisation, we first consider its timing. Occupation in New Guinea and the Bismarck Archipelago began between 50,000 and 45,000 years ago (see Table 17.1 and Figure 17.2; all dates referred to in text are calibrated as per Table 17.1). Evidence for occupation older than 40,000 years comes from a handful of sites. One of the issues with reconstructing modern human migrations to Sahul is that any early sites located in coastal areas are likely to be submerged owing to sea level rises during the Holocene. Exceptions to this are the sites of Bobongara (Groube et al. 1986; Muke 1984), Lachitu (Gorecki et  al. 1991; O’Connor et  al. 2011b), Buang Merabak (Leavesley 2004, 2005, 2007; Leavesley & Allen 1998; Leavesley & Chappell 2004; Leavesley et al. 2002), Matenkupkum (Allen et al. 214

Table 17.1.  Dates of archaeological sites mentioned in text Archaeological sites

Age (BP) uncalibrateda

Mainland Papua New Guinea Huon Peninsula Bobongara Owen Stanley Ranges Kosipe Mission 34,531 + 626 Vilakuav 41,951 + 1571 South Kov 40,298 + 956 Airport Mound 39,836 + 909 North Coast Lachitu 35,410 + 1400* Central Highlands Nombe 25,000 + 550† 14,700 + 180† West Papua Toe Cave 25,920 + 180 BISMARCK ARCHIPELAGO New Britain Kuponana Dari Yombon 35,570 + 480 New Ireland BuangMerabak 40,090 + 550 Matenkupkum 35,410 + 430* Matenbek 20,430 + 180* Admiralty Islands Pamwak 20,900 + ? Solomon Islands Kilu Cave 28,740 + 280*

Lab number

C calibrated 95.4% probability 14

Luminescence

References

> 44,000

Groube et al. 1986; Chappell 2002

Wk 17900 Wk-27072 Wk 23354 Wk 23356

41,110–37,970 48,690–42,970 45,540–42,760 45,170–42,520

Summerhayes et al. 2010 Summerhayes et al. 2010 Summerhayes et al. 2010 Summerhayes et al. 2010

ANU-7610

42,700–37,000

O’Connor et al. 2011a

ANU-2578 ANU-2580

30,850 >–< 17,270

Mountain 1991a

OZG063

31,040–30,350

Clarke et al. 2007; Pasveer 2004

Beta-62319

41,640–39,400

ANUA-15809 ANU-8178 Beta-29007

44,890–43,100 41,100–38,950 24,400–23,440

Leavesley & Chappell 2004 Allen 1994 Allen et al. 1989

?unknown

< 25,860?

Spriggs 2001

ANU-5990

33,365–31,690

Wickler 1990; 2001

34,000–38,000

Torrence et al. 2004 Pavlides & Gosden 1994

215

All terrestrial samples dates calibrated using IntCal09. *Dates on saltwater shells, calibrated using Marine09.†Dates on calcite/flowstone providing upper and lower limits of clay associated with human occupation. Dates include corrections made by Mountain 1991. a

Glenn R. Summerhayes and Anne Ford

Toe Cave Buang Merabak Pamwak

Matenkupkum Matenbek

Lachitu

Kupona na Dari Kilu

Nombe Bobongara

Yombon N

Ivane Valley

900 kms

Figure 17.2.  Late Pleistocene sites of New Guinea and the Bismarck Archipelago (illustration: G. Summerhayes). 1989; Allen et al. 1988; Gosden & Robertson 1991) and Kupona na Dari (Torrence et al. 2004) (Figure 17.2). The Huon Peninsula consists of a series of raised coral terraces which have been lifted out of the sea by a continual process of tectonic uplift (Chappell 1974, 2002; Chappell et al. 1996), thus preserving evidence of past shorelines. At the site of Bobongara (Groube et al. 1986), more than 100 stone artefacts were collected from creek beds either on or above the surface of coral terrace IIIa. Two sections were also excavated at Jo’s Creek, a small stream gully located at the back of coral terrace IIIa. From these excavations, in-situ stone artefacts were found wedged between two volcanic tephras (T2 and T3), which were dated using thermoluminescence (Tephra 2 = 60,500–37,100, Tephra 3 = 59,800–36,200). Whilst a large error range exists with these dates, the authors note that both tephras are likely to be older than 40,000 years old because of effects of moisture in the soil (Groube et al. 1986). This is consistent with new dates obtained of 43,000– 44,000 years ago for coral terrace IIIb (Chappell et al. 1996; Chappell 2002); this terrace underlies IIIa and shows no evidence for the tephras, indicating that they were deposited before the formation of this reef. First occupation of the Huon Peninsula would therefore appear to predate at least 43,000 years ago. The only other mainland Pleistocene New Guinea coastal site is the Lachitu rock shelter, located west of Vanimo on the north coast. Gorecki (Gorecki et al. 1991) surveyed the limestone country from Vanimo west to the Irian Jaya border, noting the presence of raised coral terraces similar to the Huon Peninsula. During the initial fieldwork, several rock shelters were identified and excavated, including Lachitu. A single date on shell indicates occupation from between 43,000 and 37,000 years ago (O’Connor et al. 2011b). Further east, two coastal sites from New Ireland provide evidence for early occupation. The earliest of these is Buang Merabak, located on the central east coast and initially occupied between 45,000 and 43,000  years ago (Leavesley 2004; Leavesley & Chappell 2004). Matenkupkum, located at the southern end of the island (Gosden 1995), has occupation dating from 41,000 to 39,000 years ago (Allen et al. 1989). Both Buang Merabak and Matenkupkum are located close to the present-day shoreline, at a distance of 200 and 50 m respectively (Allen et al. 1988; Leavesley 2004). Although changing sea levels would have affected the distance between these coastal sites and prehistoric shorelines, Allen et al. (1989) argue that owing to the steep contours of the eastern coastline of New Ireland, even at the height of the Last Glacial Maximum (LGM) the shoreline 216

200

0

Late Pleistocene Colonisation and Adaptation in New Guinea

00

00

Kosipe

Vilakuav (AAXF)

Swamp Iva

22

00

00

28

Kosipe Mission (AER)

00

20

00

22

Airport Mound (AAXD)

26

24

Joes Garden (AAXC) South Kov (AAXE)

ne er Riv

1 kilometre

Figure 17.3.  Archaeological sites of the Ivane Valley (illustration by G. Summerhayes). would have been no further than 500 m away from where it currently stands, thus not greatly affecting the distances between these sites and coastal resources. The last coastal site is Kupona na Dari, located on a small hill, approximately 600 m from the current shoreline, on the southeastern corner of the Willaumez Peninsula (Torrence et al. 1999). The dating of this site has been difficult owing to a lack of material available for radiocarbon dating. However, using several different strands of evidence, including fission track dating of obsidian artefacts and optically stimulated luminescence (OSL) dating, initial occupation of the site has been placed at occurring between 34,000 and 38,000 years ago (Torrence et al. 2004). Two locations are non-coastal: the Ivane Valley and Yombon. Both of these sites are vitally important as their early occupation challenges the coastal adaptation model of modern human colonisation. The Ivane Valley is located in the interior mountainous region of Papua New Guinea, at a height of 2,000 m above sea level. Earliest occupation is dated to between 49,000 and 44,000 years ago from four open sites:Vilakuav, Kosipe Mission, South Kov and Airport Mound (Figure 17.3) (Summerhayes et al. 2010). These early dates indicate that the colonisers of Papua New Guinea were moving into the montane regions soon after first colonisation. Even today, the Ivane Valley, although located only 110 km northwest of Papua New Guinea’s capital, Port Moresby, remains a difficult place to access, being surrounded by rugged mountain terrain. Yombon also demonstrates early interior occupation, located 33 km inland from the central south coast of New Britain in rugged limestone karst country (Figure 17.2). This site sits at an altitude of 500 m in a lowland rainforest jungle environment (Pavlides 1999, 2004; Pavlides and Gosden 1994). First occupation of Yombon is dated between 42,000 and 39,000 years ago. The beginning of occupation in other parts of this region is later in time. Evidence for colonisation of the main highland region is sparse and debatable. From the site of Nombe, occupation could have begun anywhere within a stratigraphic unit that dates between 30,800–28,600 years ago and 18,500–17,270 years ago (Mountain 1991a, 3.21). In the northern Solomons, the site of Kilu Cave, located on Buka Island, has evidence for human occupation dating from 33,000 to 32,000 years (Wickler 1990, 139). Occupation of the Admiralties is much later with a tentative date of at least 25,800 years ago at Pamwak Cave, Manus Island (Spriggs 2001, 367). It is difficult 217

Glenn R. Summerhayes and Anne Ford to assess occupation at Pamwak as the excavators have never published the radiocarbon date with lab number and standard deviation. Unfortunately, little archaeological work has been undertaken in the western half of New Guinea.The earliest archaeological evidence for people here dates from 31,040 to 30,350 years ago from Toe Cave, in the Ayamaru region, on the central Bird’s Head of West Papua (see Figure 17.2) (Pasveer 2004).The bones of montane and lowland species of animals have been found at this site. No doubt earlier occupation dates will be found here in the future.

Speed of Colonisation The evidence suggests that the initial colonisation of New Guinea and the Bismarck Archipelago was rapid, with little time separating the earliest dates for occupation of mainland New Guinea and New Ireland to the east. Sea crossings were necessary. Despite lower sea levels during the late Pleistocene, ranging from −56 m between 44,500 and 46,000 years ago (Chappell 2002) to a maximum depth of −130 m between 30,000 and 20,000 years ago (Lambeck and Chappell 2001), both New Britain and New Ireland have always remained separate from the New Guinea mainland. Sea crossings varied. Although New Britain and New Guinea are separated by 95–100 km of water, early colonists would have passed the stepping-stone volcanic island of Umboi, linking the two land masses. Umboi is some 50 km long and 1552 m high. Sea crossings of 50 km from New Guinea (Vitiaz Strait) and 25 km to New Britain (Dampier Strait) would still have been required. Such sea crossings across both the Vitiaz and Dampier Straits are difficult today in a powered dinghy, let alone in canoe. Siassi traders undertook regular voyages across these straits in large outrigger sailing canoes (Harding 1967). From New Britain to New Ireland, only a short crossing of 30 km across the St George Straits was required. This was even shorter if the stepping-stone Duke of York Islands were used, making the maximum crossing 20 km. To the Solomons, the distance of open ocean is much greater. Initial colonisation is normally seen as direct from New Ireland, necessitating a sea crossing of 180 km. However, thought must also be given to movements through the Anir islands, which lie circa 65 km from the south New Ireland coast. From Anir, it is 130 km direct to Buka. Separating Anir and Buka are the Green Islands (Nissan and Pinipel) in which the latter is clearly seen today from Anir’s higher elevations. Nissan is 65 km from Buka and some 60 km from Anir. Nissan is of “probable” Pleistocene age, and although argued by some to have been covered by water in the past, it is 40 m at its highest point (Spriggs 1991, 222). Whichever route was taken, the archaeologically instantaneous appearance of people from Lachitu to Buang Merabak is testimony to the seafaring skills of these first colonisers. Their maritime ability suggests that it is far more likely that initial colonisation of the New Guinea region was part of a deliberate and intentional strategy. But was this strategy by its very nature restricted to coastal and marine environments? As discussed earlier, new data from the interior of New Guinea challenge this conception, with dates from the Ivane Valley either predating or cooccurring with migration across to New Ireland and New Britain. This would appear to suggest that the first colonisers to New Guinea, whilst competent maritime voyagers, were not restricted to solely exploiting these types of environments.

Nature of Settlement and Subsistence Coastal Adaptations Although New Guinea was first occupied during Marine Isotope Stage 3 (MIS 3), a period marked by cold and dry conditions, the environment at these equatorial coastal sites was similar 218

Late Pleistocene Colonisation and Adaptation in New Guinea to today (Farrera et al. 1999, 841). Original models of Sahul colonisation focussed on the maritime abilities of the early colonisers, suggesting that these people were adapted to exploiting coastal and marine environments, thus allowing them to rapidly move along coastlines (Bowdler 1977). For New Guinea, continuities in the marine and vegetation resources located along the coast with that previously encountered in the Malesian zone of Southeast Asia (Paijmans 1976) have been argued to have assisted in the colonisation of these areas as few adjustments to the subsistence economy would have been required (Allen 2000; Groube 1989). The availability of coastal resources, however, may have differed during Sahul’s prehistory because of varying sea levels (O’Connell et al. 2010). Chappell (1993) noted that stable and rising sea levels would have favoured the development of coral reefs, estuaries, coastal wetlands and swamps, areas rich in littoral and coastal resources. In contrast, falling sea levels would have produced less productive hard shorelines (Specht 2005). The fluctuating availability and richness of coastal areas may therefore have directly impacted upon the reliance on a coastal technology. Evidence for assessing early coastal subsistence patterns is restricted to the assemblages of New Ireland and the northern Solomons. There is a paucity of information on Lachitu’s early faunal assemblage, although Gorecki et al. (1991) noted a focus on coastal resources, and a complete lack of faunal materials from the New Britain sites owing to their acidic soils. As a whole, the native fauna of the islands of the Bismarck Archipelago are considered depauperate, with few native mammal species. Analysis of the faunal remains from the initial occupation stage of Buang Merabak shows a low density mixture of both inland and coastal species, including Bare-Backed Bat (Dobsonia anderseni), lizards, fish and shellfish (Leavesley 2004; 2005). This is consistent with a low intensity use of the cave site by small groups of mobile hunter-gatherers, who visited the cave periodically to hunt the D. anderseni that used the cave as a roost but were also utilising nearby coastal resources. The use of D. anderseni is particularly significant as these would have provided good protein sources in high-density localised patches, which could be obtained using simple technology, such as placing barriers across the cave mouth and smoking out the bats (Summerhayes et al. 2009, 732). Leavesley (in Summerhayes et al. 2009, 732) makes an important point that such access to protein was not evenly distributed across the Pleistocene landscape, pointing out its absence in the north of central New Ireland. To venture farther north of this region was to move into areas of effectively lower-density protein resources. The use of coastal resources is more evident at Matenkupkum, which has dense shell middens present. At both Buang Merabak and Matenkupkum, however, similar shell foraging strategies were employed. At both of these sites, large individual specimens of the large-sized shell species Turbo argyrostema predominate (Allen et  al. 1989), with the addition of Chitons also present at Buang Merabak (Leavesley and Allen 1998). Thus, during initial occupation, shellfish were subject to a targeting foraging strategy, where only the largest specimens of a small range of rocky shore and reef taxa were collected. Fish species, which occur in low numbers at Buang Merabak and Matenkupkum, indicate exploitation of the reef zone, with reef fish and tropical sharks dominating. Deep-sea fishing is also evident during the Pleistocene as seen in the presence of pelagic fish at these coastal sites and confirming skills in seafaring. At Kilu Cave, reef species dominate, but pelagic (deep-sea) fish make up 20% of the number of identified specimens (NISP) in the earliest levels (Wickler 2001, 224–226). Developing technologies for pelagic fishing are also found later in time from the Pleistocene cave of Matenbek (see Table 17.1), located in close vicinity to Matenkupkum, where possible fish hook manufacture is evident in the earliest levels dated to between 24,400 and 23,440 years ago (Allen et al. 1989, 55; Smith & Allen 1999, 294). The emphasis on extensive fishing ability is comparable with recent data from East Timor. Pelagic fish comprised approximately half of the fish assemblage from the earliest occupied layers of Jerimalai, between 38,000 and 42,000 years ago (O’Connor et al. 2011b), with fish hook technology developing between 23,000 and 16,000 years ago. Importantly, at Buang Merabak, the earliest evidence so far recorded for personal adornment during the Pleistocene in Papua New Guinea 219

Glenn R. Summerhayes and Anne Ford was also recovered.This consisted of a perforated tiger shark tooth, most likely worn as a pendant, which was recovered from layers dating between 40,000 and 43,000 years ago (Leavesley 2007). In contrast to the faunal data, few plant remains have been identified.The best evidence obtained currently includes taro residues found on stone tools at Kilu from the earliest occupation (Loy et al. 1992). The evidence suggests that the earliest colonists were “small groups of mobile, broad-spectrum foragers” who exploited both maritime and terrestrial resources (Allen et al. 1989, 558–9; Allen & Gosden 1996, 187; Leavesley 2007). A review of the lithic technology present in New Ireland and northern Solomons sites supports this scenario, being dominated by simple flakes manufactured from river cobbles obtained from nearby waterways, with little evidence for retouch or formal shaping of tools (Allen et al. 1989; Leavesley 2004; Leavesley & Allen 1998; Wickler 1990). A size distinction between the different assemblages has been identified, with large flakes produced at Matenkupkum and Matenbek, but only small flakes at Buang Merabak. Allen et al. (1989) have attributed the large flakes at the two southern sites to the local availability of abundant raw material. However, river cobbles are also present in close vicinity to Buang Merabak, within 2.5 km of the cave (Leavesley 2004); therefore it is not clear whether raw material availability is the sole determinant of artefact size in the assemblages. The earliest Pleistocene assemblage from Kilu was also characterised by small, unretouched flakes made from locally obtainable rocks, most of which were crystalline igneous rocks, with lesser amounts of siliceous and fine grained igneous rocks (Wickler 1990). Over time, a shift in raw material usage appears to occur at Buang Merabak, Matenkupkum and Matenbek. During pre-glacial occupation at Matenkupkum and Matenbek, a wide range of raw materials is used, including volcanic and siliceous rocks. Over time, the volcanic rocks dominate. Allen et al. (1989) argue that this is due to the desirable siliceous rocks being depleted from local waterways through overexploitation. Similar changes are noted at Buang Merabak where, during the pre-glacial phase, a siliceous rock, jasper, dominates, constituting 65% of the total assemblage (Leavesley & Allen 1998), before being gradually replaced by other raw materials. By the terminal Pleistocene or early Holocene, volcanics dominate, which again suggests the initial selection of desired raw material types, with shifts towards raw material of lesser quality over time. In summary, the data from New Ireland and the northern Solomons are consistent with small groups of highly mobile foragers initially occupying the island and utilising local resources as a strandlooper strategy (Allen et al. 1989; Allen & Gosden 1996). Although at low intensity, the foraging strategy is targeted at obtaining the highest-value resources available, as evidenced by differential selection of the shellfish, fauna and lithics. Initial lithic technology is characterised as expedient, where locally available raw materials are obtained as river cobbles, struck to produce simple flakes, with minimal further modification prior to use and discard. Can we assess the impact these early populations had on the coastal environments? The data suggest that there was little impact on coastal resources, with any change seen only after 20,000 to 30,000 years of occupation. Over-exploitation of shellfish as seen by a reduction in the size of the available shellfish is not evident till the 24,400 to 20,500-year-old levels. Unfortunately, our knowledge of human impacts on vegetation for these coastal sites is virtually nil because of a lack of palaeoenvironmental reconstructions. This is in stark contrast with the interior occupation. From New Britain, the coastal site of Kupona na Dari presents a contrasting lithic technology compared to the New Ireland sites (see Torrence et al. 2004). Rather than using local raw materials, obsidian was imported into the site from three different sources. Two of these were located to the north on the Willaumez Peninsula, Baki and Gulu, with the third, Mopir, being located to the east. Each of these sources was located between 23 and 39 km away. The disparate location of these sources suggests a high level of mobility during the initial occupation of the area, where people were actively exploring the landscape, identifying and utilising new sources. The technology at Kupona na Dari was based primarily on flakes, which were not only utilised themselves but were also used as cores to produce more flakes. This dual purpose for the flakes means that 220

Late Pleistocene Colonisation and Adaptation in New Guinea excess raw material could be left as quarry sources. By carrying only the minimum usable material, groups were thus creating a highly portable tool kit, light but still effective for use, attributes that should be desired by highly mobile groups. Torrence et al. (2004) argue that the duality of the flakes indicate a high level of foresight and planning, appropriate for colonising peoples who aim to minimise risk in ensuring that they have available tools when required. From mainland New Guinea, the situation is different, although only one coastal site, Bobongara, has been described, and even then only six artefacts were recovered in situ (Groube et al. 1986). Waisted tools were the dominant artefact type collected from both the excavation and surveys, although non-waisted stone artefacts are also present (Groube 1984; Muke 1984). The waisted tools are large, heavy, unifacially flaked pebble tools, primarily made from locally available coarse-grained andesite river cobbles, and with a pronounced indentation on both sides of the tool, forming the waist which is usually attributed to facilitating hafting. All of the waisted tools are heavily patinated as a result of weathering, thus obscuring technological and potential functional data. However, Groube (1989) identifies these artefacts as tools involved in forest clearance, because of the large size and weight of the tools; the blunt edge, which is a function of the nature of the raw material used; macroscopic use wear; and the potential evidence for hafting, including possible hafting breaks. Forest clearance activities include ring barking, felling of small trees, branch trimming and root clearance, where opening up the forest is argued to promote the growth of grassland areas. Manipulating the environment to create new grassland ecotones may be attractive for several reasons including opening up new grass­ land habitat for terrestrial hunting, as well as providing suitable cleared areas for management of economic plant species, which require the removal of the canopy to allow sunlight through for growth. Whilst this argument has been applied to a number of sites with waisted tools present, primarily located in the montane regions of Papua New Guinea, Bobongara is the only coastal site to include large amounts of waisted tools during the late Pleistocene. Because of the early age of this site and its coastal location, it is not clear as to why forest manipulation may have been required when coastal resources would have been readily available. The image of highly mobile broad-spectrum foragers exploiting locally available resources before moving onto new areas would appear to be at odds with the investment of time and energy involved in altering the local landscape.

Interior Adaptations Although MIS 3 had little effect on equatorial coastal locations (Farrera et al. 1999), there was a dramatic drop in temperatures at the higher altitudes. The interior site of Yombon, however, is located only 500 m above sea level, and thus would not have greatly differed in temperatures from today. Recent palaeoenvironmental reconstructions completed at Yombon have shown that during the late Pleistocene the local vegetation consisted of a closed tropical lowland rainforest with mixed forest and grassland species. It was argued that evidence of burning at Yombon may have been humanly induced but not extensive, with burning used to open a gap in the rainforest canopy (Lentfer et al. 2010). What drew people into the interior of New Britain is unknown, although the presence of chert may have been a factor.Yombon is unique in its situation close to rich geological deposits of chert. All of the late Pleistocene artefacts obtained from Yombon were quarried from in situ chert sources, which, again, is a sharp distinction from the New Ireland sites where river cobbles dominate the assemblage. Although only a small assemblage of artefacts dates to the late Pleistocene (a total of 29 artefacts) (Pavlides 1999), Yombon contains evidence for differences in stone artefact production trajectories. One formal tool type is present, a unifacial ovoid scraper, which Pavlides (1999) argues is evidence for the creation of morphological types. Two flakes also appear to have been dislodged from the working edge of a larger tool, indicating possible re-sharpening on site. 221

Glenn R. Summerhayes and Anne Ford Even though chert is available locally, it would appear that efforts were being made to conserve and re-use tools, suggesting the presence of a curated technology. Because of the limited numbers of artefacts, behavioural inferences have to be drawn cautiously from the Yombon assemblage. Overall, however, the assemblage would imply considerable investment in time and energy, in terms of both initial raw material procurement through quarrying primary context raw material and production, as shown by the presence of the ovoid scraper. This investment is further reflected in the re-sharpening of working edges rather than simply discarding them if edges become broken or blunt. Again, this would appear to be a significant contrast to the expedient technology present at the New Ireland sites, where raw material procurement consisted of picking up river cobbles from a secondary context and manufacturing simple flakes for use. Unlike Yombon, the Ivane Valley lies at 2,000 m above sea level, in the mountainous region of mainland New Guinea. Initial occupation dates between 49,000 and 44,000 years ago and is found at locations scattered round the valley (Figure 17.3) (Summerhayes et al. 2010). Moving into this mountainous region would have presented new challenges as it meant experiencing not only a new and unfamiliar landscape but also a novel climate and subsistence resources. Both the montane forest and subalpine grassland regions would have contained new and unfamiliar remnant Gondwanaland flora species. Whilst Gondwanic elements dominate the montane forest, there are similarities between the alpine grasslands of New Guinea and Malesia, suggesting migration of species perhaps during the Quaternary period (Smith 1979). However, as no contemporary archaeological sites have been found within montane or alpine areas in Southeast Asia, it is unclear how much familiarity the first colonists of New Guinea would have had with these species. New faunal species may also have presented new challenges, especially as it is argued that now-extinct large mammal species such as Protemnodons and Diprotodontids may have been present in the montane region up to 15,000 years ago (Mountain 1991a; Flannery et al. 1983). Added to differences in flora, the high altitude montane region would have been significantly colder with temperatures well below that of the coastal areas to which these colonists were purportedly adapted to expedite their progress across the globe.The climate would have been colder than today, although not as cold as during the LGM, when temperatures would have dropped to between −5 o C to −6o C (Farrera et al. 1999). Palaeobotanical research confirms cooler and wetter conditions for the initial period of occupation, with local vegetation initially comprising swamp forest or sub-alpine forest before shifting to upper montane forest with sub-alpine grassland elements after 45,000 years cal BP (Hope 2009). This time period incorporates the bottom cultural layers of the site, Layer 4 (49,000–39,000 years ago) and Layer 3b (39,000–30,000 years ago) (Summerhayes et al. 2009). Occupation continues in Layer 3a (30,000–26,000 years ago), which corresponds to the shift to much colder sub-alpine conditions from MIS 3 to MIS 2. Exploitation of wild pandanus and yam (Dioscorea sp.) is evident from the earliest levels (Summerhayes et al. 2010). Pandanus nuts would have provided these early inhabitants with a rich source of protein and oil (Bourke & Allen 2009). The presence of Dioscorea starch grains on stone tools is all the more significant as it would not have been possible to grow yam at this altitude at this time because of the reduced temperatures. It would have had to have been brought up from warmer lower altitudes, demonstrating the territorial range of these peoples. Highly fragmented burnt bone from a small animal was also recovered from the earliest levels, which may indicate possible hunting; however, identification of species was not possible (see Summerhayes et  al. 2010, 80). Alongside the evidence for subsistence, the stone technology indicates that the earliest inhabitants adapted to locally available raw materials. Ford (2011, 51)  argues that “the use of local raw materials shows innovative and adaptive behaviour where, from initial colonisation, modern humans were moving in and out of the Ivane landscape, making, using and discarding tools on site”. Importantly, different trajectories for stone tool production appeared to be associated with 222

Late Pleistocene Colonisation and Adaptation in New Guinea A

0

1

2

3

B

4

5cm

0

1

2

3

4

5cm

C

0

1

2

3

4

5cm

D

0 1

2 3 4

5cm

Figure 17.4.  Pleistocene stone tools from the Ivane Valley. A: South Kov, Layer 4; B: Joe’s Garden, Layer 3b; C: Airstrip Mound, Layer 4; D: South Kov, Layer 4 (illustration by G. Summerhayes).

the different types of raw materials available locally. Metamorphic rocks appear to have largely been obtained as river cobbles, which were then flaked into larger tool forms, including waisted tools (Figure 17.4). Small siliceous rocks, however, were used to produce a small flake-based technology. As noted earlier in relation to Bobongara, waisted tools have been argued to have been used in forest clearance (Groube 1989) to create grasslands for management of economic plant species, but this remains to be demonstrated. Evidence for other potential forms of forest clearance and land management are found in the pollen record of the Ivane Valley, where Hope (2009) noted an increase in charcoal after 45,000 years ago becoming significant from 41,000 to 38,000 years ago, which he associated with 223

Glenn R. Summerhayes and Anne Ford human firing of the wet montane forest. Its impact would have been to extend open sup-alpine grasslands into the montane forest areas, thus replacing the forest with grasses, Cyathea treeferns and shrubs. The use of burning to create new disturbance gaps or ecotones has been argued by Mountain (1991b) to have provided areas conducive to growing economic plants such as tubers and also attracting animal prey to the new growth inspired in these areas. However, the scale of human impact due to firing of the local environment is difficult to assess. It has been noted that the occurrence of fire was globally higher in MIS 3, and increased burning could have corresponded with warmer phases known as Dansgaard-Oescher (D-O) events from 50,000 to 40,000 years ago during the last glacial cycle (Daniau et al. 2009; Huber et al. 2006). Yet an increase in fire could also be related to the spread across the globe of modern humans. So, from the Ivane Valley we are presented with a mosaic of innovative and flexible adaptations. We have evidence of people bringing in new plant foods from lower altitudes, such as yams to provide carbohydrates, which can supplement locally available foods, such as the protein-rich pandanus nuts. We also have evidence that people are adapting to and using local raw materials for their stone artefact technology. At the same time, there is potential that they are also altering the landscape through firing, which may have been designed to increase the hunting potential of the area. The Ivane Valley is unique along the spine of Papua New Guinea as it is the only valley of its kind at that altitude, being surrounded by steep-sided mountainous terrain and high-energy V-shaped valleys. Even today it is an inaccessible place to get to and involved a 5- to 8-day walk inland for the physically fit from either the northern or southern coastal plains. What drove people into this environment must be related to the calorific returns that fits well with models of foraging collecting strategies and optimal foraging theory, which suggests regular residential mobility to allow use of widely distributed resource patches (see Allen & O’Connell 2008; Denham & Barton 2006). There is another site from the interior that could shed light on early adaptations, although the evidence is meagre. Nombe rock shelter is located at 1,720 m altitude in the central highlands of Papua New Guinea (Figure 17.2). Dating the earliest occupation is problematic, however; as noted earlier, occupation could have begun anytime within a stratigraphic unit (Unit D) that dates between 30,800–28,600 and 18,500–17,270 years ago (Mountain 1991a, 3.21). It would appear that the first humans to occupy Nombe may have been in small groups that visited the area periodically. Both extinct and extant fauna are present in Stratum D, which indicates that montane forest and alpine grassland species were being selected for prey. As Nombe would have been located close to the tree line prior to the LGM, both habitats would have been readily available to the Nombe hunters. Included in these early layers are four species of extinct large herbivorous mammals, including three macropods and one diprotondontid. Only two specimens of the diprotondontid were recovered, making identification difficult, but all three of the macropods have been classified, into Protemnodon tumbuna, Protemnodon nombe and Dendrolagus niobe (Flannery et al. 1983), all of which are browsers that would have inhabited montane rainforest. The relationship between people and the large mammal fauna is difficult to assess; however, if associated, it would suggest that they co-existed for perhaps up to 25,000 years prior to the large mammal extinction. If shown to be associated, this long co-existence would negate any concept of a blitzkrieg-like effect on large fauna in New Guinea following the arrival of humans into the area.

Developments after 30,000 Years Ago Modelling the nature of these early colonisers is difficult because we are assessing a period of time described so aptly by Chris Gosden (1993, 133) as a world without ethnographic parallel.Yet the evidence for small populations of highly mobile foragers led Gosden to infer that continual movement across such expanses did not lend itself to the building of territories or group boundaries, and there was considerable interaction among groups. Changes seen in the archaeological 224

Late Pleistocene Colonisation and Adaptation in New Guinea record some 30,000–20,000 years after initial colonisation suggest the definition of group boundaries as a result of the filling-up of the landscape. Once the colonisation process between Sunda and Sahul ceased, both areas become isolated from each other for the next 40,000 years (Friedlaender et al. 2007; Rasmussen et al. 2011; Summerhayes 2007; van Holst Pellekaan 2013).The same could also be said of interaction between mainland New Guinea, New Britain and New Ireland. There is no hard archaeological evidence for the subsequent movement of people or goods between mainland New Guinea and Island Melanesia for another 30,000–20,000  years. A similar barrier could also have existed between the Bismarck Archipelago and northern Solomons. Once Bougainville was colonised, there is no evidence for any further interactions between it and the Bismarck Archipelago until 3,300 years ago (see Summerhayes 2007). Evidence to support this model of isolation can be seen in the recent work on Melanesian mitochondrial haplogroups P and Q by Friedlaender and his team (Friedlaender et al. 2005a; 2005b).They argue that the mitochondrial haplogroup Q2, which is not found on mainland New Guinea, developed within the Bismarck Archipelago with an estimated coalescence time of 36,000–37,000 years ago (a standard error of 11,000–12,000 years). In short, human populations (or at least the females as mtDNA evidence relates only to female isolation) of the Bismarck Archipelago remained in isolation from populations on mainland New Guinea. However, changes seen in the archaeological record from 25,000 years ago suggest a widening of social interactions, at least in the coastal regions of New Guinea. First, the island of Manus was colonised by 25,000 years ago (Spriggs 2001), probably from the north coast of New Guinea.The colonisation of Manus implies some form of sophisticated water transport, as it involves a substantial water crossing of either 230 km from the north coast of New Guinea, 200 km from Mussau, or 230 km from New Hanover or Lavongai (Irwin 1992, 21). This crossing was not an isolated event. There is a gap of several thousand years before occupation levels again appear at Pamwak with remains of mainland New Guinea animals and nut trees, cuscus (Spilocuscus kraemeri), bandicoot (Echymipera kalubu) and Canarium indicum nuts (Kennedy 2002, 20; Specht 2005, 252–259). Second, obsidian from west New Britain sources has been identified in southern New Ireland archaeological assemblages (Summerhayes & Allen 1993), again suggesting a maritime transport strategy to access and procure obsidian from the sources. Third, there is evidence of the movement of animals from mainland New Guinea to the Bismarck Archipelago. At 25,000 years ago, a shift occurs in the main focus of the mammal diet with evidence for the introduction of Phalanger orientalis (Northern common cuscus) at Buang Merabak (Leavesley 2005) and Matenbek (Gosden 1995). As argued earlier, before the arrival of P. orientalis, New Ireland was relatively depauperate in relation to mammal species; therefore, its addition would have meant an important increase in animal meat available. Allen and Gosden (1996, 188) argued that they were brought to the Bismarck Archipelago as a breeding population or as an accidental by-product, being escaped pets or potential food (Allen et al. 1989, 557). Either way, they would have supplemented previous protein sources of reptiles, birds, bats and seafood. Gosden (1993) argued that the introduction of P. orientalis marks a change from people travelling to obtain resources to people carrying resources with them. Apart from species translocations during this time, we are also seeing evidence for impacts on the local environment. As argued earlier, from the midden deposits of the New Ireland cave sites, the initial colonising strategy that focussed on targeting the large individual specimens of large shellfish species such as Turbo argyrostema changed by 24,400–20,500 years ago when a diminution in size of the shellfish occurs, reflecting over-exploitation (Gosden & Robertson 1991). That this type of environmental impact took so long to manifest within the archaeological record is further testimony to the low population numbers prior to 25,000 years ago. It was argued by Summerhayes (2007) that these changes in the archaeological record were a by-product of population increase, not to mention innovations in human management. The nature of human society did not fundamentally change during the first 30,000–20,000  years since colonisation. The evidence from the few archaeological sequences available shows little change to subsistence patterns. What may have been a critical change is the gradual filling up of 225

Glenn R. Summerhayes and Anne Ford the landscape, which would have varied from the larger island of New Guinea that is rich in land animals, compared to the much smaller islands of the Bismarck Archipelago, which are depauperate in land animals (Summerhayes 2007, 15–16). With population increase, group territories would be expected to slowly develop with defined boundaries, especially as different groups came into more frequent contact (see Summerhayes 2007, fig.  2.5). Evidence for long-distance down-the-line exchange of animals and obsidian appeared at 25,000 years ago, because that was the first time a chain of regular and frequent contacts could have developed between mobile communities. That is, widespread resource distribution requires the development of dependable exchange links with other communities.

Summary Testimony to the adaptive skills of the first colonisers of New Guinea can be seen in their occupation from Lachitu in the west to the east coast of New Ireland in an archaeological instant which would have required competent voyaging skills. Yet, at the same time, the early colonists were not beholden to the coastline, with migration into the interior occurring soon after initial occupation. The need to diversify into other habitats and shift reliance away from coastal economies may have been prompted by unstable coastal environments associated with varying sea levels (O’Connell et al. 2010).This evidence suggests that these first groups were highly mobile explorers, venturing rapidly into unknown habitats from the equatorial coastline to the high-altitude montane valleys. Further evidence for their flexibility can be seen in their adaptation to a diverse range of ­environments and developing technologies (or adapting older technologies) to manage the landscape. In every environment entered, new resources were quickly utilised, from the use of local raw materials for their stone technology to the available faunal and floral resources for their subsistence. Along the coast, a wide range of foods was incorporated into their diet, including bats, shellfish, fish and taro. In the interior, new plant foods encountered such as pandanus were also eaten. From the later sites of Nombe and Toe Cave, the faunal remains indicate the taking of a wide variety of prey from different habitats, such as montane and sub-alpine environments at Nombe. Many of these new resources may have required a process of learning and adaptability, such as the ability to procure chert from primary in-situ contexts at Yombon or the processing of pandanus in the Ivane. However, in addition to exploiting local resources, the colonists also transported key resources if needed, as evidenced by the carrying of carbohydrates in the form of yams into the Ivane Valley and obsidian into Kupona na Dari. Other possible evidence for adapting of the landscape to required use can be seen at Yombon and the Ivane Valley, where burning may have been used to remove canopy and extend open areas. In returning to the question as to whether these early peoples were pre-adapted to a coastal lifeway which enabled them to move from Africa to Sahul, there is clear evidence for a competent maritime technology, as shown by their ability to make substantial sea journeys and access coastal and marine resources. At the same time, it is also clear that at least some groups of colonisers were venturing inland soon after initial colonisation and were able to learn, modify and adapt to resources in these areas. The flexibility and adaptability outlined in the colonising strategies presented here may be a key factor in understanding the rapid and successful migration of modern humans around the globe.

Acknowledgments We wish to thank Mr. Herman Mandui, chief archaeologist of the National Museum and Art Gallery, for facilitating research in Papua New Guinea, and Dr. Matthew Leavesley, Deputy Dean, 226

Late Pleistocene Colonisation and Adaptation in New Guinea University of Papua New Guinea, for continuing the teaching of archaeology in that nation. We would also like to acknowledge Heather Sadler, Matthew Hennessey and Les O’Neill, from the Department of Anthropology and Archaeology at the University of Otago, for assistance with the production of figures.

227

Chapter 18 Modern Humans Spread from Aden to the Antipodes With Passengers and When?

Stephen Oppenheimer

Introduction The use of non-recombining DNA and the phylogeographic approach, based on reconstructed gene-trees, has illuminated the geography of the exit route of anatomically modern humans (AMH), their further travels, and the logic of only one ultimately successful exit from subSaharan Africa – likely via the Yemen, around 72 ka (Oppenheimer 2012a). However, there are competing archaeo-genetic views of multiple exits (Underhill et al. 2001; Rasmussen et al. 2011). Chronometric predictions based solely on genetic data have systematic problems, the most important being calibration and wide confidence intervals, and so need testing against more accurate data from archaeology, palaeoanthropology and the earth sciences. Several recent papers and books offer new climatic, fossil, lithic and genetic evidence for earlier exits of AMH, before the great Toba volcanic eruption of 73 ka (Younger Toba Event [YTE]) (Petraglia et al. 2007, 2011; Oppenheimer 2003, 2009; Rosenberg et al. 2011), and even as long ago as the last interglacial, that is, up to 50,000 years before the YTE (Petraglia et al. 2010, 2011; see also Groucutt & Petraglia, this volume; Liu et al. 2010b; Armitage et al. 2011; Cabrera et al. 2009 [cited as genetic evidence in Petraglia et al. 2010]; Rose et al. 2011). In contrast, others argue for a later movement of modern humans north to Europe and east to India (e.g., Mellars 2006b; 2006c) Besides chronology, there are three basic questions, essential in any reconstruction of the African AMH exit, in the following logical order: 1. How many founding exits of AMH can be seen in the fossil or archaeological record? 2. Which of these are evidenced by genetic continuity into modern regional populations (rather than by evanescent interglacial faunal spreads in the fossil record)? 3. Which route or routes were taken? Given these issues, including the depression of relevant mtDNA haplogroup ages on recalibration, the time is ripe for a focussed review of the evidence both for and against a pre-Toba exit from Africa, and also for a re-ordering of the relative importance of the primary South Asian route to Australia versus the later Levantine-European dispersal. 228

Modern Humans Spread from Aden to the Antipodes 46-50,000 years ago Homo sapiens entered Europe. Most Europeans today can trace their ancestry to mtDNA lines that appeared between 50,000 and 13,000 years ago

20-30,000 years ago Central Asians moved west towards Europe and east towards Beringia

Apporx. 15,000 years ago Humans crossed the Bering land bridge that connected Siberia and Alaska

40,000 years ago Humans from the EastAsian coast moved west along the Silk Road

12,000 years ago A group of humans travelled northward through Egypt and Israel but died out 90,000 years ago

te

u

African origins Over 150,000 years ago modern humans - our mtDNA ancestors lived in Africa

l ro

Approx. 72,000 years ago A group of humans travelled through the southern Arabian peninsula towards India. All non-African people are descended from this group

st a co

40,000 years ago Humans trekked north from Pakistan up the Indus River and into Central Asia.

15-19,000 years ago Artefacts and tools found in Pennsylvania give evidence that humans had migrated into the Americas before the Ice Age

50-60,000 years ago Humans crossed from Timor to Australia Modern humans moved east from India Into Southeast Asia and China

12,500 years ago Evidence of human habitation and artefacts found, Monte Verde, Chile

Figure 18.1.  Narrative map of modern human dispersals, as reconstructed in this review (modified from Oppenheimer 2003; 2009).

Regional Setting There is a growing consensus that the dispersal of AMH from Africa occurred via a single southern exit across the mouth of Red Sea, followed by movement round the coasts of the Indian Ocean initially to Bali, but ultimately to Melanesia, Australia and the Americas (Figure 18.1). The argument for a single exit (e.g., Oppenheimer 2003; 2009) rather than multiple exits (e.g., Lahr & Foley 1994; Underhill et al. 2001; Rasmussen et al. 2011) is mainly based on the finding of single ex-African uniparental lineages (both male and female) representing the entire non-African genetic diversity on those loci. Recent claims for AMH exits eastwards during the last interglacial (Petraglia et al. 2010; Liu et al. 2010b; Armitage et al. 2011; Rose et al. 2011) need to be viewed in this context. The presumption that modern human range extension out of Africa proceeded initially along coastlines has depended largely on the rapidity of this movement, as inferred from genetic phylogeography, since only three key founder mtDNA haplotypes (M, N and R) are indicated. These gave rise to multiple regionally specific branches en route, in effect three starbursts rooted on these haplotypes, seeding all the way from Arabia to Bali (e.g., fig. 6 in Soares et al. 2009). For this speedy beach-travel model, these colonisers must have been able to exploit coastal marine resources. Shellfish, being mainly limited to the intertidal zone, are easily over-exploited, necessitating a continuous extension of foraging along the beach. This model clearly provides an immediate and continuous motive for uni-directional range extension in a linear fashion. Such a systematic pattern of littoral marine exploitation is characteristic of, if not unique to, modern humans (but in one instance, Neanderthals) starting from at least 160 ka (see McBrearty & Stringer 2007). Shell middens and other archaeological evidence from Africa show that early AMH populations were proficient at such exploitation. Putative middens and other evidence of marine exploitation are found at the very earliest occupation sites in coastal Australia and most significantly in Timor at one of the oldest Asian human coastal sites at the Australian threshold (O’Connor 2007b; see also Balme & O’Connor, this volume). It is unlikely that this human skill229

Stephen Oppenheimer A N

N

Chukchi Australian Australian Chinese Piman Uzbek Samoan Korean New Guinea New Guinea New Guinea Italian Georgian German Saami English Crimean Dutch French New Guinea Australian Evenki Buriat Khirgiz Warao Warao Asian Indian Chinese Siberian Inuit Guarani Japanese Japanese

F

M

B 51k P N R 48k 67k 60k U JT 53k

African L3/Other African L

36k

HV M 72k 8L3 83k

G M8

Non African African

African

149k L2’3 103k

L2 190k 6.5 million years

L1bc 162k

L1b

150k

Mkamba Ewondo Bamileke Lisongo Yoruba Yoruba

70k L2a L2c

L1a

Mandenka Effik Effik Ibo Ibo Mbenzele Biaka Biaka Mbenzele Kikuyu Hausa Mbuti Mbuti San San

57k

L1k 200k

D 41k

70k

L1c 138k

L3

C 37k

100k

Toba

50k

Last Ice Age

Today

Figure 18.2.  Illustrative gene tree based on the first mtDNA complete sequence data available in 2000 (from 52 individuals randomly selected around the world). The structure of the tree remains essentially the same today (modified from fig. 0.3 in Oppenheimer 2003; 2009). Note that the chimp-human coalescent date derives from analysis of coding region – that is, it is not assumed from fossil evidence. complex and the ability to cross short stretches of water determined the date of expansion, since these ‘modern’ skills, were practiced by modern humans within Africa for twice as long as outside (McBrearty & Stringer 2007).

Materials and Methods Used in Reviews of Genetic Phylogeography Phylogeography Genetic phylogeography (literally the geography of gene trees; see Richards et al. 2000b for a description of the method of Founder Analysis) is particularly useful for answering topographic questions of route and is the only discipline that reliably informs on routes and numbers of successful human exits. Genetic phylogeography is used extensively for dating migrations, although it is less precise or accurate than disciplines such as archaeology, hominin palaeontology and palaeoclimatology, which are most useful for testing dating regimes and routes suggested by the genetic models. A combined approach should give increased confidence by triangulation. The study of the geographic spread of gene lines involves three related types of inference: first, an unambiguous gene tree (e.g., Figure 18.2) is built up by sequence study (preferably complete sequence [CS]) of non-recombining DNA in representative population samples. This review refers mainly to maternally transmitted mitochondrial DNA (mtDNA). The non-recombining part of the Y chromosome (male-NRY), however, produces very similar topographic results (Oppenheimer 2003; Richards et al. 2006; Hudjashov et al. 2007). Such trees because of their non-recombining DNA possess great detail and geographic specificity, even 70,000 years after the African exit. Second, migrations are inferred from the geographic distribution of individual gene branches. Third, the diversity of clusters of gene lines is used to infer the time depth of 230

Modern Humans Spread from Aden to the Antipodes individual branches and founding events. Accurate dating of Y trees has lagged behind mtDNA analysis as a result of methodological problems, including the cost of complete sequencing, but a preliminary independent estimate for the African/non-African split using CS-NRY is 75 ka (Cruciani et al. 2011). These three forms of inference are combined in Founder Analysis, which is the identification and dating of specific founder lines moving from a source to a target region (Richards et al. 2000).

Linearity of Dating The linearity of the mtDNA molecular clock is a problem, since purifying selection preferentially removes minimally deleterious non-synonymous mutations over time, resulting in non-linearity of calibration for more recent times. Luckily the effect is regular, predictable and measurable and can be corrected for, giving demonstrable independent linearity between the synonymous and non-synonymous sequence data, thus allowing a more-precise complete sequence clock. A major recalibration paper (Soares et al. 2009) used a worldwide phylogeny of more than 2,000 complete mtDNA sequences to measure, predict and correct for this effect, calibrating on the chimp-human split of 6.5 ma, but cross-checking on recent archaeological founder events. For consistency, recalibrated mtDNA lineage dates are exclusively cited as evidence in this review.

Review of Out-of-Africa Models How Many Exits of AMH Ancestors from Africa? The Genetic Evidence Multiregional View versus Recent Replacement There are two contrasting models of how and when our species appeared in Europe and Asia. The multiregional hypothesis postulates that geographically differentiated populations of Homo sapiens arose independently in different regions from a common ancestral population, but avoided speciation by gene flow with neighbouring populations. Replacement models (as proposed in this chapter) maintain that H. sapiens arose in Africa between 100 and 200 ka and then later left Africa and effectively replaced (with acknowledgment of the possibility for some admixture [Oppenheimer 2003, 49]) indigenous populations in Asia and Europe. Limited Archaic Interbreeding outside Africa Contrary to common perception, the replacement model does not depend on the assumption that there was a lack of interfertility, and thus zero gene-flow, between AMH and different indigenous regional human populations. The genetic evidence merely confirms, with at least 95% probability, morphological arguments that AMH appears to have replaced indigenous populations. If one ignores the degree of replacement, however, both multiregional and replacement models turn out to have a point. It is essential to clarify this point in order to lay to rest the misassumption, sometimes made by multiregionalists, that any evidence of archaic admixture (i.e., reticulation) outside Africa, however small, would necessarily disprove the recent replacement model (Wolpoff et al. 2000). It would not: replacement is simply implied by degree, which does not need to be 100%. Interbreeding with Neanderthals The issue of interbreeding is raised by recent claims of a 1–4% introgression of the Neanderthal genome into modern Eurasians (Green et al. 2010). If valid, and this author is convinced there is at least a prima facie case, this evidence is not necessarily at variance with the simple view of apparently complete replacement or with the unambiguous lack of uniparental evidence for Neanderthal introgression amongst the vast number of uniparental mtDNA/Y genomes already 231

Stephen Oppenheimer studied. It should be remembered that the effective population size of the haploid uniparental genomes would be a quarter that of autosomes. The finding of only single African lineages (L3 and M168) in non-African mtDNA and NRY respectively not only excludes the multiregional model at these two loci but also demonstrates a severe founding drift event. This drift would likely have caused the extinction of minority lineages (e.g., of less than the 4% maximum estimated initial introgression among putative uniparental genetic material acquired from predispersal interbreeding). Whether such hypothetical extinction of putative acquired uni-parental Neanderthal lineages among AMH would have occurred following admixture outside Africa (Green’s preferred geographic option) is clearly debatable, but depends on where and how soon after exit admixture occurred and what population sizes were involved. However, it is difficult to imagine any geographic scenario in which the ancestors of East Asians could have simultaneously acquired the same dose of Neanderthal autosomal genes as those of Europeans, unless it was soon after the exit and before East-West divergence.The authors themselves acknowledge, for this to happen, that “gene flow from Neanderthals into the ancestors of non-Africans occurred before the divergence of Eurasian groups from each other” (Green et al. 2010, 710), thus explicitly depending on the single exit hypothesis for AMH, recruited in this case to explain autosomal miscegenation. The current northwest Eurasian distribution of Neanderthals leaves large lacunae in South Asia. Because of the absence of fossil evidence from this region, the possibility cannot be discounted that Neanderthals lived in this region and also Arabia. If so, it may have been in these regions that admixture occurred with immigrant AMH populations, before the differentiation of Europeans from East Asians. Interbreeding with Denisovans in Oceania An alternative admixture scenario that Green and colleagues (2010) did not consider for that publication but, as co-authors, did offer subsequently in the context of the recently sequenced hominin phalanx and molar tooth from a Middle Palaeolithic/Upper Palaeolithic transitional context in Denisova Cave, Southern Siberia (Krause et al. 2010; Reich et al. 2010), is the possibility that AMH interbred with other non-Neanderthal populations, such as H. heidelbergensis, or H. erectus descendants moving from China to Central Asia. In their first report (Krause et al. 2010), the Leipzig group argued from the complete mtDNA sequence of the phalanx that the Denisova lineage had branched off long ago (1.04 Ma, CI: 0.779–1.3 Ma), leaving both the modern human and Neanderthal lineages more closely related to each other (0.493 Ma, CI: 0.3744–0.6121 Ma) than either were to Denisova. In their second publication, they reported another complete mtDNA sequence (from the tooth) which was very close, but not identical to that of the phalanx from the same site, indicating two ‘Denisovan’ individuals, very similar at the mitochondrial locus, and confirming the phylogenetic age estimate, with a slightly reduced error (0.982 Ma CI: 0.7805–1.208 Ma). Morphologically, the Denisova tooth grouped tightly with early Homo specimens (Australopithecines, H. habilis and African H. erectus), all of whose teeth were significantly larger than a more ‘modern’ group containing Neanderthals, H. heidelbergensis and all other European humans (ancient and modern), all AMH, and Chinese H. erectus. Indonesian H. erectus specimens tended to take an intermediate position between the two groups. Another, non-metric, feature of the dental root, namely a massive, splayed lingual root, again grouped the Denisovan tooth with the erectus types. In offline material for two other genetic comparisons, the Leipzig group displayed Venn diagrams of autosomal segmental duplications showing those held in common, compared first between a present-day human (NA18507), a Neanderthal, Denisova and a chimpanzee, and then between a Neanderthal, Denisova and a present-day human. In each case the modern human and Neanderthal shared most segmental duplications with each other and least with Denisova (Reich et al. 2010). These four comparisons, three genetic and two morphological, are all consistent with each other, first in making Denisova the biological outgroup when compared with the relatively 232

Modern Humans Spread from Aden to the Antipodes closely related European human groups (AMH, Neanderthals and H. heidelbergensis), but further in implying that Denisova could be descended from an earlier African population. (The case for a Eurasian origin for the Denisova population is argued by Martinón-Torres et al. 2010.) In the second Denisovan paper, Reich and colleagues (2010) also performed extensive autosomal sequencing and claimed that 4–6% of the provisional autosomal signatures from the Denisova finger bone were found in three Melanesians (two Papuans and a Bougainvillean), but in none of the other modern human groups they tested (African: San, Yoruba and Mbuti; five “Eurasians”: French, Han Chinese, Sardinian, Cambodian and Mongolian; and South American: Karitiana). Given all the genetic and morphological observations summarised thus far, this inclusive or exclusive geographical scenario could be consistent with ancestors of Oceanic AMH mixing, not with North Chinese H. erectus, but with H. erectus in the region of present-day Indonesia. The regional distinction is important, since controversial evidence published by Swisher et al. (1996) suggested that ‘Solo man’, a significantly larger-brained version of Homo erectus from Ngandong and Sambungmacan, Central Java, may have survived to only 27 ± 2 to 53.3 ± 4 thousand years ago.The controversy over sources of error was such that it was revisited by some of the same researchers, resulting in higher but conflicting dates, the youngest of which was a electron spin resonance (ESR)/U-series of 143 ka (+20/−17) (Indriati et al. 2011; see Dennell, chapter 4, this volume).Whether the first dates were valid or not, these controversial reports could bring the evolved descendants of H. erectus much closer to the Upper Pleistocene. Rather surprisingly, in view of the erectine links and the range of their own genetic and morphological evidence, Reich et al. (2010) argued instead for a closer relationship of descent between the Denisova hominin and Neanderthals than either had with AMH. This change in inference was made on the basis of another comparison, namely of autosomal SNPs (single nucleotide polymorphisms) between Denisovans with AMH, chimpanzees and Neanderthals, but was based on the relative degree of sharing of SNPs distinctive to Neanderthals. The results and implications of the latter were quite different, when compared with the four genetic and morphological comparisons mentioned previously, and there is a simple explanation for this anomaly, which they did not test for: Neanderthal-Denisovan admixture. The strength of evidence for their assertion of more recent common descent for the two archaic groups was less convincing than that for the simple Neanderthal intrusion into AMH. This is because in this more complicated three-group interbreeding puzzle they did not test a more likely local explanation for the autosomal SNP associations between Neanderthals and Denisovans.That would be extended direct hybrid interbreeding between the two archaic human groups locally in Central Asia, prior to AMH arrival, as supported by their geographic, physical and temporal overlap, consistent with the extended presence of Neanderthal mtDNA in that part of Central Asia (shown in Okladnikov Cave, near Denisova Cave; Krause et al. 2007b). Such interbreeding would constitute only admixture – not, of course, recent common descent – and might be one explanation for the extraordinary plesiomorphy shown in the recently discovered Late Pleistocene “Mongolanthropus” at Salkhit in Mongolia.With marked superciliary arches, this skullcap fragment “shows multiple similarities with Neanderthals, Chinese Homo erectus, and West/Far East archaic Homo sapiens” (Coppens et al. 2008; but see also Kaifu & Fujita 2012, who mention its dating of 20 ka and see it as within the range of Late Pleistocene AMH). Perhaps the most interesting result in the Reich et al. (2010) paper was from comparisons of autosomal SNPs in Denisovans with a variety of modern Eurasian populations, which revealed that Denisova shared 4–6% of its genetic material with present-day Melanesians in the Pacific, a quarter of the globe away, but not with any other more proximate populations. The possibility of widespread archaic admixture in Oceania was reinforced by Rasmussen et al. (2011) who showed Denisovan admixture in a 100-year-old Australian Aboriginal hair, similar in degree to that in Melanesians. They also postulated two dispersals of African-derived AMH populations into Australia: the first 62,000 to 72,000 years ago, and a later one 25,000 to 38,000 years ago. Unfortunately, their interpretations were influenced by a selective use of autosomal-dating of 233

Stephen Oppenheimer

Figure 18.3.  Map showing a single southern route out of Africa and beachcomber arc route from the Red Sea along the Indo-Pacific coast to Australia, including likely extensions to China, Japan and New Guinea, from MIS 4. Vegetation and sea level shown as at LGM. Overlay: claimed fossil and cultural evidence for AMH in Eurasia during MIS 5 to MIS 4 (see text for dates) (modified from fig. 1.6 in Oppenheimer 2003; 2009). demographic events, unsupported archaeologically. Their analysis and interpretation of ancestral population phylogeny depends not on any genetic phylogeny but, similar to the Green et  al. (2010) paper, on a statistical association test based on only four complete genomes. On the positive side, the Rasmussen et al. (2011) results still extend the influence of ‘Denisovan DNA’ in Oceania more broadly to the Sahul region (as was predicted, Oppenheimer 2012b) and do not falsify the consensus single AMH exit model preferred here. Indeed, there is skeletal evidence that would reverse the morphological order of their “two Australian waves” since the ear­ liest human crania in Australia are modern and gracile, while there is a late appearance of archaic, robust skulls in the Kow Swamp and Coobool Creek sites, radiocarbon-dating to no earlier than 14,300 years ago (Brown 1992). This Antipodean Denisovan puzzle has been extended by a more recent and fascinating third publication (Reich et al. 2011) presenting results for 33 more locations in Asia, Southeast Asia and Oceania. Briefly, all 15 populations tested to the east of Huxley’s Line (Figure  18.3), including one Negrito and one non-Negrito group in the Southern Philippines, Near Oceania and Polynesia, show significant evidence of Denisovan intrusion, up to but no more than that shown for Melanesians (i.e., Near Oceanians) and a now similar level for Aboriginal Australians. Conversely, of 27 Asian and Southeast Asian (SEA) populations west of the Huxley Line, including two Negrito groups, none shows any significant evidence of such intrusion in Reich’s analysis. All admixture thus appears to have been offshore in Wallacea or Sahul since Huxley’s Line demarcated the true eastern limit of the Asian Mainland until 10 ka. More westerly admixture with later local extinction is possible but much less likely. Relevant questions about these findings are how many times, from where, and when? For the first, with the exception of the Philippines, all the results of relative admixture estimates are 234

Modern Humans Spread from Aden to the Antipodes consistent with a single proximate common source of Denisovan admixture: Wallacea, Australia or New Guinea with subsequent fresh external diluting AMH gene flow. The non-Denisovan admixture analysis in Wallacea, Oceania and the Philippines show New Guinea and Australia as tightly correlated, major alternative candidate sources of AMH gene flow for the Pacific region, but Australia as consistently the richer one, overall by 40%. The apparent anomaly could still be compatible with Near Oceania being the primary dispersal source, but with the New Guinea Highlands (Gosden 2010; see Summerhayes and Ford, this volume) having gone through founding event(s), isolation and subsequent drift, a scenario already inferred from shared unique human leukocyte antigen (HLA) markers (Serjeantson & Hill 1989). However,Wallacea (Nusa Tenggarah and/or Moluccas) still seems the most parsimonious primary Denisovan source, from geographic considerations. This speculation is consistent with the plot of individuals’ Denisovan admixture against Near Oceanian (New Guinea) admixture (Reich et al. 2011, fig. S1), which shows tight correlation and high values of relative Denisovan admixture (ranges from 30% to 100%) for Wallacea and Fiji. In contrast, Polynesians, although grouping together with the lower end of the Wallacean distribution with lower values (20%–30%) and still correlating with Near Oceanian admixture, show consistently less Denisovan intrusion than expected from that admixture, likely resulting from later Southeast Asian admixture and drift. The south Philippines are anomalous compared with the other 13 sites, since both samples show 3–4 times more Denisovan intrusion than would be expected from their Near Oceanian (i.e., non-Denisovan) admixture values and no clear correlation with them. This likely indicates a separate admixture event and subsequent dispersal, though the source region could still be in north Wallacea. The similar, asymptotic levels of Denisova intrusion in Melanesians and Australians suggest that the main admixture event occurred in a single source population before the colonisation of Sahul, most likely in Wallacea. Caveats on the above fresh inferences are the paucity of regional samples from Sahul, which could allow for subsequent minor migrations or even fresh admixture events. Interbreeding with Denisovans in China? The ‘exclusivity’ of the Oceanic Denisovan admixture was brought into doubt by Skoglund and Jakobsson (2011), who, by adjusting for founding genetic drift coupled with ascertainment bias, demonstrated an additional ‘Denisova-related’ intrusion centred in East Asia (mainly China but extending to two populations in South America). This result helps to support similar inferences for East Asia published four months before (Abi-Rached et al. 2011, fig. 2E). A discrete third hominin (Denisova-like) admixture event in China would tempt speculation about which hominin was intrusive and a search for morphological evidence of hybrids. There is a dearth of unambiguous archaic candidates in that region dating to less than 100 ka. Xujiayao, Maba and Dali were late large-brained archaic individuals and morphologically intermediate between H. erectus s.s. and AMH, but none dates to later than the terminal Middle Pleistocene (Brown 1999; Kaifu & Fujita 2012). As to evidence of an unexplained morphological change among AMH, the clearest potential candidate change came surprisingly much later, and after the appearance of more conventional Late Pleistocene AMH fossils such as those from Upper Cave 101, Zhoukoudian and Liujiang in China (Brown 1999). This change was apparently the late and sudden appearance of Mongoloid fossil skulls throughout East Asia and the New World, first clearly evidenced in the Asian Early Holocene–Early Neolithic record, with skulls such as Baoji (Brown 1999). These features were possibly anticipated by distinctive, older, ‘broad-cheeked’ AMH fossils such as the Minatogawa 1 skull (ca. 20 ka; Brown 1999) and, more recently, Longlin 1, which dates to the Pleistocene-Holocene transition (Curnoe et al. 2012). “Mongolanthropus” also needs to be mentioned here for morphological and dating reasons, although its face is missing. One possibility is that the onset of Late Glacial and Early Holocene warming allowed populations to expand from their refugia and interact with each other. More likely, hybridisation could have already occurred in Far Eastern LGM refugia (see elaboration of these two complementary expansion235

Stephen Oppenheimer contraction hybridisation scenarios – attributed respectively to Darwin, Hewitt and Arnold, in Stewart & Stringer 2012). Looking for refuge locations, most known LGM hominin archaeological sites in China are north of the Yangtzi in modern Sinodont, northern Mongoloid regions that might fit dental clines (see figs 6.3, 5.4, 5.3 in Oppenheimer 2003), though ascertainment bias is possible. However, either model would still, as mentioned for Neanderthal intrusion, have to assume a subsequent bottleneck period of selection or drift (perhaps during the Younger Dryas refuge period), followed by expansion. This would explain the lack of uniparental evidence of admixture, as with the Neanderthal story, but also the more uniform and gracile suites of cranial features found in modern southern and northern Mongoloid populations. Several predictions of the recent replacement model can be tested by genetic phylogeography, using non-recombining DNA such as mtDNA and the NRY. Recent replacement outside Africa should show older lineages within Africa and younger lineages outside, depending on the degree of replacement. This was first unequivocally demonstrated to be the case (Oppenheimer 2003) 16 years after the famous ‘Mitochondrial Eve’ paper was published (Cann et al. 1987), after knowledge of mtDNA variation had been improved by a set of complete sequence data from 52 individuals of worldwide distribution (Ingman et al. 2000). This author dated the branches of the world mtDNA tree, reconstructed from those first 52 complete sequences, showing that all non-Africans belonged to two sister twigs (M and N), both aged at only around 70 ka, and arising from L3, one of the youngest of more than a dozen branches originating in Africa (see Figure 18.2).This broad conclusion has not been subsequently falsified by analysis of tens of thousands of mtDNA sequences. Single versus Multiple Recent Exit Models for AMH Two further fundamental observations can be made from that worldwide mtDNA tree. First, of the dozen or more L branches of the tree found only in Africa most are, of course, much older than L3, M or N. The genetic antiquity of the deepest L branch point (192 ka; Soares et al. 2009; 190 ka in Figure 18.2 here) is consistent with new dated fossil finds of early AMH in East Africa (164 ka: White et al. 2003; 195 ka: McDougall et al. 2005). Second, as mentioned, the most singular observation is that ‘L3’, the remaining African branch, is unique in that it also encompasses all non-Africans, with two branches originating outside Africa, namely M and N (Oppenheimer 2002, 2003; Metspalu et al. 2004). In effect, this means that the rest of the world was colonised by two descendants of the ancestral haplotype for M and N (a subtype of the L3 root holding a195 nucleotide transition). The latter now dates by the corrected molecular clock to 71,600 (CI: 57,100; 86,600) years ago (Soares et al. 2009). The overall age estimate of L3, including all the African branches, is the same and likely represents the upper age limit for the exit (i.e., L3 could be older but not younger than the exit). However, the confidence intervals on this estimate straddle the date of YTE, and thus prevent a genetic resolution of the issue of whether AMH entered South Asia before or after the Toba eruption. A single mtDNA lineage exit is unambiguous evidence of a single successful exit of AMH from Africa and is paralleled closely in the Male-NRY tree and also clearly reflected in the X chromosome and in several autosomal loci (Oppenheimer 2003; Hudjashov et al. 2007; Richards et al. 2006; Abu-Amero2009). That all non-African descendants, for both mtDNA and Y lines, each share one and the same close African ancestor is in itself evidence of a period of drift in a small isolated founding group. The chances of two or more founding exit episodes separated in time and place each giving rise, by drift, to the same source African lineage in all different parts of the non-African world are negligible, leaving the single successful exit as default (Oppenheimer 2009). Claims for additional earlier AMH exits during MIS 5e, based on fossil or lithic evidence, are not necessarily implausible in practical terms and need to be taken seriously. Given the dating evidence from Skhu¯ l and Qafzeh, which variously approach and significantly exceed 100 ka 236

Modern Humans Spread from Aden to the Antipodes (Stringer et al. 1989) for at least one abortive AMH, exit near MIS 5e means that similar claims for the southern route may well turn out to be demonstrated in the fossil record but not in the mtDNA or NRY of modern populations: if so, such dispersal events must be presumed evanescent. Furthermore, although autosomal evidence, having far more loci and less drift, is much more sensitive than uniparental study and can thus show less than 5% Archaic admixture into AMH, outside Africa, it would be far less specific when comparing AMH with AMH, thus losing that advantage looking for AMH genetic survival from the last interglacial.

Which Route? A single successful African exit for AMH has several clear implications. Because there was only one route of exit, the number of subsequent route options was also decreased. Whichever route was taken initially, however, a multidisciplinary model also has to explain how both Europe and Asia could have been colonised ultimately from the same single exit group. Southern Rather than Northern Exit: Genetic Evidence As far as the genetic evidence is concerned, there are two logical reasons for identifying the southern route across the mouth of the Red Sea. First, of only two primary branches of L3 mtDNA outside Africa (M and N – Figure 18.2), both are represented in East Eurasia, including South Asia (where they have the highest number of autochthonous M and N lineages in the world), while only N is represented in West Eurasia. Second, Levantine and European N representatives are less diverse, more derived and younger than those of Arabia and Southwest Asia (see fuller discussion in Oppenheimer 2003; Richards et al. 2006; Fernandes et al. 2012). Climatic Considerations and Archaeology of the Exit Route An implication of the single genetic exit, southern or northern, is that there must have been great difficulty in establishing a beachhead outside sub-Saharan Africa (SSA) (e.g., the Arabian Gulf; Oppenheimer 2003, 88) during the Upper Pleistocene, presumably as a result of the SaharanArabian arid zone barrier to migration out of SSA (see Figure  18.3). This barrier was freely permeable only during interglacials but may also have been open between 80 and 75 ka during early MIS 5a (see Rosenberg et al. 2011), a period characterised by a moist welcoming refugium, a ‘Gulf Oasis’ (Rose 2010; Fleitmann et al. 2011; see also Petraglia et al. 2011). A Last Interglacial Exit into Southwest Asia? Two full interglacials have occurred during the past 150,000 years, the last, or Eemian, interglacial (MIS 5e) and the most recent, during the Holocene. Each time a diverse flow of African ­savannah fauna occurred from SSA to the Levant and Arabia (Turner 1999). AMH used this climatic window for a northern exodus during MIS 5e, as evidenced by skeletal evidence at Skhu¯ l and Qafzeh. However, this dates to between only 120 and 90 ka (Grün & Stringer 1991). There is presently some archaeo-genetic consensus that this early exit left no descendants in modern populations (contra Petraglia et  al. 2010). On the archaeological side, there is the long gap of up to 40,000 years between the Skhu¯ l and Qafzeh AMH fossils and the reappearance of AMH, signalled by the Early Upper Palaeolithic in West Eurasia after 49 ka, a period that was intercalated by the Neanderthal presence in the Levant (for a recent perspective on dates and causes of these switches, see Shea 2008). On the genetic side there are no mtDNA or Y lineages outside Africa that date even near an MIS 5e exit at 120 ka, or even the less relevant lower bracket of 90 ka, which evidence virtually excludes survival of this very real Levantine AMH colony into the modern ex-African gene pool (Oppenheimer 2003; Richards et al. 2006). An identical issue applies to recent claims that AMH reached Jebel Faya in southern Arabia between the beginning of MIS 5e and 90 ka (Petraglia et al. 2010; Armitage et al. 2011). Here, 237

Stephen Oppenheimer only lithic evidence is available. The tool kit showed affinities to the late Middle Stone Age in northeast Africa. This evidence was used to infer the presence of AMH in the Arabian Peninsula at two different times: an occupation phase during MIS 5e (two optically stimulated luminescence [OSL] dates: 123±10 ka and 127±16 ka, and another later one at 95 ka: OSL: 94.8±13.0 ka; see Armitage et al. 2011; Petraglia et al. 2010). Rose et al. (2011) have also recently reported the finding of a Middle Stone Age Complex of Nubian typology in Dhofar, Oman, dated at 106 ka (OSL), again fossil-free. Except for the absence of fossils, the chronology and implications for these putative, limited appearances of AMH outside Africa during MIS 5e and its disappearance before 90 ka are similar to Skhu¯ l and Qafzeh in that the colonies seem to have died out with no issue. This author feels that it would be hardly surprising if the Skhu¯ l and Qafzeh event via the northern route out of Africa had been duplicated via the southern route out of Africa as well, although given the likely low population densities of the period, it does not necessarily mean genetic survival of such a migration. However, because there is no human skeletal record from Pleistocene Arabia, the makers of the Jebel Faya assemblages are unknown, and the possibility cannot be excluded that these were made by Neanderthals. Given that stone tools are a poor guide to identifying the hominin that made them, the African Middle Stone Age typology of the tools would not necessarily contradict this possibility. Furthermore, if Neanderthals were in Arabia, it provides a plausible scenario for a Neanderthal admixture into an AMH population that was ancestral to both West Eurasians and East Asians, as predicted by the single southern exit hypothesis. Fossil Evidence in East Asia for an Exit during MIS 5e The recent finding of an anterior mandible with two molars at Zhiren Cave (Zhirendong) in South China has been claimed as possible evidence for an expansion of AMH into East Asia during MIS 5e (Liu et al. 2010b). The authors argue that the mandible dates to more than 100 ka (but see Kaifu & Fujita 2012) and shows derived modern human features, distinct from any known late archaic human. However, they also place it close to later Pleistocene archaic humans (contra Kaifu & Fujita 2012). In a comment in Nature, Dennell (2010, 513) agrees with the dates as “100,000–113,000 years old, and possibly older”, on the basis of stratigraphy and faunal associations, but questions the need to postulate interbreeding, at least in relation to the molars: “The latter are small, and would be considered as modern in a Late Pleistocene west Eurasian (postNeanderthal) sample”. Westaway and colleagues (2007a) also recently suggested the possibility of AMH presence in Java during MIS 5e on the basis of one tooth from Punung Cave (reported in Storm et al. 2005) that was small enough to be classified as H. sapiens. This claim is necessarily very tentative until more evidence is available. The Liujiang skull, a Late Pleistocene cave specimen also from South China, has had similar claims for such antiquity. It is clearly anatomically modern, and other items in the breccias in which it was found date to 111–139 ka (Shen et al. 2002), which could put it in the same time frame as the Zhirendong and Punung Cave fossils. These dates are very controversial, however, and the younger dates (all around 67 ka) based on the covering flowstone seem a much safer minimum. Summary of Genetic Arguments against a Genetically Successful Exit during MIS 5e The recalibrated age estimates and confidence intervals (CI) of the out-of-Africa mtDNA branch L3 of 71,600  years ago (CI: 57,100; 86,600) (Soares et  al. 2009) comfortably exclude the date ranges implied by an Asian exit during MIS 5e (as did previous L3 age estimates: Oppenheimer 2003; 2009). Furthermore, given the clear phylogenetic evidence for only one successful exit, it could be either an exit during only MIS 5e or later (e.g., during MIS 5a–4), but not both. Since the genetic dates and the bulk of the archaeological evidence and dates converge on MIS 4 or 238

Modern Humans Spread from Aden to the Antipodes even later, a population that left Africa during the last interglacial and survived substantially into the modern gene pool looks extremely unlikely. So, one has to presume that, like the Skhu¯ l and Qafzeh exit, any Asian exit by AMH during the last interglacial must have been similarly evanescent. This view is consistent with the archaeological record for the terminus of the putative early Africa-Arabia exits in Oman, which are older than the range of L3 age estimates.

Dating Migrations Possible Dates of the Definitive Exit As mentioned above, a general guide to the earliest possible date for the expansion of H. sapiens into and beyond Arabia is the genetic age of lineage L3 that is now 71,600 years ago (CI: 57,100; 86,600) (Soares et al. 2009), if we exclude an exit in the last interglacial (MIS 5e) on the grounds that it lacks a genetic progeny. For the lower limit, N, one of the two daughter lineages of L3, the nearest outside Africa has been dated in South Asia, where it has a predominantly western distribution, to 71,200 years (CI: 55,800; 87,100). However, the ages of the M and N lineages estimated elsewhere are mainly around 60 ka. Branch M is surprisingly young in South Asia where it has a predominantly eastern distribution, dating from 49,400 years (CI: 39,000; 60,200) (Soares et al. 2009). Delayed Migration to the Near East and Europe Since the genetic evidence favours a single southern exit rather than a northern one, it is important to track and date the appearance of northwest Eurasians out of the southern route. The problem remains, however, as to how and when ancestors of northwest Eurasians crossed the arid zone between Southwest Asia and the Levant and Turkey in northwest Eurasia, and why they were delayed. A climate-permissive model can be reconstructed (Oppenheimer 2009) of AMH populations suddenly being able to move into northwest Eurasia from South Asia during Interstadial 14/1349–55 ka ago (see figs. 1.7, 1.8 and 3.4 in Oppenheimer 2003; also Müller et al. 2011). This dating would coincide with the earliest appearance of Upper Palaeolithic technology in the Levant and southeast Europe (calibrated to 47/49 ka and 46 ka respectively, Mellars 2006a) and also with the age of U8, the oldest mtDNA ‘N’ lineage in the Near East and Europe (at 51.3 ka (CI: 44.0–58.8): Soares et al. 2010; see also Oppenheimer 2003; Richards et al. 2006). Dating the arrival of AMH in India and East Asia While there may arguably be evanescent archaeological or physical evidence for AMH in Arabia and southern China soon after the last interglacial, so far there is little fossil or genetic evidence in India or Southeast Asia that dates to the period immediately before the Toba eruption. Instead, there is a congruence of proxy evidence to either just before or soon after this event, the former (pre-Toba) being only cultural and the latter (just post-Toba) now including (in East Asia) several fossils and the recalibrated genetic dates. India and Toba Given the previous suggestion (Oppenheimer 2003) that AMH may have arrived in Southeast Asia before the super-eruption of the Toba volcano (73 ka ago), much archaeological interest has focussed on India, which has the most extensive deposits of younger Toba tuff (YTT) (Jones 2007). Although Middle Palaeolithic stone artefacts have been found in association with several Indian YTT deposits (both above and below), hominin fossils are absent, and whilst H. sapiens might have made these assemblages, other possibilities cannot be discounted. Despite these issues, recent articles (Blinkhorn & Petraglia, this volume; Petraglia et al. 2007; see also Petraglia et al. 2010, 2011; Clarkson et al. 2012) argue from lithic evidence both below and above Toba ash in the Jureru Valley in Southern India for the presence of AMH in India at the time of the YTT event, with survival afterward. Much of the argument impressively detailed 239

Stephen Oppenheimer and analysed by Clarkson et al. (2012; see also Clarkson, this volume), which supports claims that these tools (pre- and post-Toba) were made by AMH, rests on their tight statistical typological associations both with the South African Middle Stone Age, and Palaeolithic AMH cultures in Australia and Southeast Asia, and their differences from North African and Levantine cultures, archaic hominins, very early modern human and later Aurignacian assemblages. This nuanced perspective is obviously more consistent with the southern route of AMH dispersal, as is implicit in Clarkson’s statement: “There are no obvious differences between the Jurreru assemblages and newly reported assemblages from Arabia dating to MIS5a” (Clarkson et  al. 2012, 178). In the absence of hominin fossils, there are the usual problems of trying to match a technology to a particular hominin, but when this work was presented at a meeting in Oxford in 2010, this author was convinced by the statistical and visual associations, as were several seasoned archaeologists, excepting Paul Mellars (reported in Balter 2010). Previously pre- and post-Toba-ash cultural dates for the first site (Petraglia et al. 2007) were both close to the YTE, but in new results on nine paired Indian sites, only the pre-Toba dates were close to the YTE at 74,000 BP or earlier, whereas nearly all of the post-Toba dates were about 55,000 years BP or younger (Balter 2010), thus raising a question mark over their key claim of rapid local human recovery in India post YTE. Indian Genetics and Toba A prediction of the ‘catastrophic’ interpretation of the conjunction of tools and ash in India, clearly not shared by Petraglia and co-workers, is that a deep and wide genetically sterile furrow would have split East from West Asia, with India eventually recovering from a genetic bottleneck either locally or by re-colonisation from either side (Oppenheimer 2009). Evidence for such a furrow was first argued for in the genetic map of Asia (Oppenheimer 2003, 180–184). Also consistent with the possibility of a local Indian bottleneck, M groups in Eastern India round the Bay of Bengal, although numerous, are notably much younger than elsewhere along the Indian Ocean trail (fig. 3, Oppenheimer 2009; see also table 1, Sun et al. 2006; and table 3, Soares et al. 2009). However, dates for haplogroups N and R, which locate further west in South Asia, where less ash fell, are relatively higher (fig. 3, Oppenheimer 2009 and table 3, Soares et al. 2009; but see also figs. 4.3 and 4.5, Oppenheimer 2003), consistent with local drift near the east coast of India and a pre-Toba exit. Fossils and Genes in East and Southeast Asia Inevitably, as in Europe, flaked stone artefacts are vastly more common than fossil hominin specimens. Nevertheless, the few dated AMH fossils from China and Southeast Asia (and even Australia), date from MIS 4–3 and show a much older age bracket (67–40 ka) than the oldest such fossils in Europe (< 35 ka: Trinkaus 2003) (Figure 18.4). However, with the possible exception of the Punung tooth, none predates Toba. The small collection of East Asian dates for the early arrival of AMH are older than their equivalents in Europe (reviewed in Oppenheimer 2012a; 2012b and elsewhere in this volume and need not be discussed here). They are limited by the predominant use of radiocarbon-based ­dating with its low ceiling, the lack of diagnostic fossils even when the cultural dates are secure, and lack of secure direct dating even when the fossils are securely AMH. There is also a problem with the newly recalibrated estimates of genetic ages for M, N and R haplogroups in East Asia, which now cluster at around 60,000 BP or less (Soares et al. 2009), 5,000–15,000 years lower than estimated before (Oppenheimer 2009), rather than the new archaeological estimates of up to 67,000 BP. Confidence intervals of the new genetic dates for all three haplogroups are, however, still sufficiently wide to include 67,000 BP and even Toba. Dating Pleistocene Arrivals of Humans in Sahul and Near Oceania Archaeology, Radiocarbon and Luminescence Dating For Melanesia, the first archaeological evidence of occupation in the upland Ivane Valley, New Guinea, has been radiocarbon dated to between 49,000 and 43,000 cal BP and 44,890 and 240

Modern Humans Spread from Aden to the Antipodes

Figure  18.4.  Map with archaeological dating of early AMH evidence in ISEA and Sahul as mentioned in the text; for 14C calibrated estimates (the rounded mean of the upper and lower figures), see Summerhayes et al. (2010). Two routes to Sahul involved island hopping (modified from fig. 4.1 in Oppenheimer 2003; 2009). 43,100 cal BP to the east, in the Pacific island of New Ireland (Summerhayes et al. 2010; see also Summerhayes & Ford, this volume). For Australia, the carbon ceiling has similarly been increased to around 48,000 years using the ABOX-SC method (Turney et al. 2001a; also in Summerhayes et  al. 2010, table S2). Non-carbon dating techniques of cultural remains suggest substantially earlier human colonisation of Australia by 50–60 ka (Roberts et al. 1994, 2001; Bird et al. 2002, 2004). There is clearly more work required with non-radiocarbon-based techniques. This review is not intended to replace reviews that focussed on that controversy but merely points out that the present ‘oldest non-carbon dates’ are not inconsistent with Australian colonisation by 50–60 ka. Recent sea level curves, concentrating on troughs (Siddall et al. 2003), suggest that the best window of opportunity for such a long crossing over the large Sahul shelf from Timor would have been circa 65 ka when sea levels were 100 m lower, with occupation in the Northern Territory becoming visible on modern littorals only from 60 to 50 ka onwards owing to higher modern sea levels (Oppenheimer 2004). Genetics As mentioned, genetic phylogeography is a better tool for elucidating migration trails than for dating. Resolution on the main descent of Melanesians and Australian Aboriginals from the single exit is clear. The point, made initially in 2003 (Oppenheimer 2003), that Australian and Melanesian mtDNA all belonged to M or N lineages, is consistent with the single exit, and has since been confirmed by complete mtDNA sequencing and NRY work on an expanded set of samples (Macaulay et al. 2005; Hudjashov et al. 2007). Previous mtDNA genetic dates of arrival of the vanguard in the Sahul, estimated by this author, were consistent with a pause of around 10,000 years at Bali before crossing to Australia (Oppenheimer 2009). Combining the Roberts et al. (1994; 2001) colonisation dates for Australia with the recalibrated genetic ages of M, N and R in East Asia would give no time at all for such 241

Stephen Oppenheimer a pause before arriving in Australia, possibly 50–60 ka.This would infer an extremely rapid transit from Africa to Australia, of circa 12,000 years given the recalibrated molecular age of the out of Africa L3 lineage at 71,600 years.This speed of colonisation is however consistent with the phylogenetic observation that the great arc of dispersal was effected entirely by the three primary root founder haplotypes M, N and R, each giving rise to multiple sets of unique regional daughter branches in every sector round the Indian Ocean.

Conclusions Although no hominin fossils have been found, the recent archaeological evidence from India is consistent with a southern exit of AMH from Africa shortly before the Toba eruption of circa 74 ka. However, Neanderthals or other archaic populations cannot be excluded as potential tool makers of their African-type Middle Stone Age assemblages. The preliminary evidence on the dating of the post-Toba cultural layers would suggest a prolonged period of recovery, consistent with a local Indian bottleneck. The parallel phylogeography of both uniparental lineages (mtDNA and male-NRY) overwhelmingly suggests a single exit of AMH from Africa, most likely by the southern route. Recent recalibration and correction for non-linearity of the mtDNA molecular clock reduces most previous genetically dated estimates (based on the age of L3 for the upper limit) of this exit to approximately 72 ka (CI: 57–87 ka), with rapid migration round the Indian Ocean coast to Australia. The lower limit of the estimate could be described by the hypothetical founding ancestor of the two primary non-African clades M and N at 71.5 ka (CI: 57.0; 86.4), or failing that by the age of the most southwesterly clade, N estimated in South Asia at 71.2 ka (CI: 55.8; 87.1), though the ages of haplotypes M and N elsewhere are younger than this. These genetic confidence intervals straddle the Toba event and thus cannot resolve the issue of whether humans dispersed before the Toba eruption, but they do appear to exclude a significant (> 5%) AMH survival from any earlier MIS 5e exit, as claimed from Jebel Faya, Zhirendong and a controversial older dating of the Liujiang skull. As with the recent inference of Neanderthal and Denisovan admixture outside Africa, there is a possibility that uniparental studies, with their known problems of drift, could miss a less than 5% contribution from any other human demic source, including that from an MIS 5 AMH exit. However, given that the autosomal genomic methods recently used to infer archaic admixture were statistical and relatively non-specific and had no method of dating, these methods would themselves be unlikely to resolve this issue any time soon. Palaeoclimatology, palaeoanthropology and archaeology potentially offer firmer measures for dating the definitive AMH exit from Africa, although with no fully convincing evidence older than the Toba eruption. Dated, contextually uncontroversial, diagnostic fossil evidence in China and SEA goes back only to 50 ka. This asymptote is partly an artefact of method, which has been largely radiocarbon-based. Without that limitation of method, Liujiang in Southern China provides a definite AMH skull, and an accurate and contextually conservative minimal date of 67–68 ka, while Australia offers cultural evidence between 50 and 60 ka. The ash fallout from the Toba super-eruption still offers both a stratigraphic approach to dating sites in South and Southeast Asia and an ongoing source of controversy in the absence of any relevant fossil evidence. Diagnostic AMH fossils in Asia dating to the period 74–90 ka would help resolve the issue.

242

Chapter 19 It’s the Thought that Counts Unpacking the Package of Behaviour of the First People of Australia and Its Adjacent Islands

Iain Davidson

Introduction Sahul is composed of Australia, Tasmania and the island of New Guinea, except during interglacial periods of high sea level, and was colonised only by modern humans. Its archaeohistory,1 therefore, provides unique evidence of what is and is not necessary to show the modernity of behaviour. This chapter concentrates on four types of issues about Sahul archaeohistory: historical, theoretical, methodological and empirical. Historically archaeology began in Western Europe and first encountered early humans in that region. As a result, many of the problems tackled as the practice of archaeology spread to other regions derived their perspectives from what seemed important to the archaeohistory of Europe and debates framed there. Theoretical issues are present in assumptions about the nature of modern human behaviour and the cognitive changes during hominin and human evolution from a more ape-like common ancestor more than 3 million years ago to a more modern type of cognition around 100 ± 50 thousand years ago. The cognitive abilities fundamental to the behaviour that got people to Sahul underpinned the use of symbolism too, and this has been a dominant theme in relating people to individuals, groups and environments. Methodological issues concern the most appropriate evidence for writing archaeohistory generally. Molecular genetics have much to say about past relationships, but such studies must be anchored to the evidence of people on the ground. Archaeological evidence of flaked stones, bones and rock art presents its own problems associated with dating and taphonomy, and there are major problems of interpretation and of the role of knowledge of modern people and conditions in that interpretation. Reaction to these historical, theoretical and methodological issues affect the way the empirical record of Australian archaeohistory contributes to understanding the evolution of modern human cognition and behaviour.

243

Iain Davidson

Historical Questions All people who have ever lived in Sahul and the Americas are modern humans (see comparison in Davidson 2013). Earlier hominins who may or may not have exhibited some of the same behaviours as modern humans were present almost everywhere else. The question of modern human origins was originally framed in terms of the apparent opposition between Neanderthals and moderns but their history later proved to be much more complex. Nevertheless, there is still remarkable interest in what happened in the small peninsula of Western Europe at the end of Eurasia and with Neanderthals, who were restricted primarily to that region. The idea that it might be possible to investigate modern human behaviour without it being linked necessarily to Europe and its hominin types occurred rather late in archaeological discussion. The Middle to Upper Palaeolithic transition in southwest France had been known for some time, but Noble and Davidson (1991) linked that transition to the emergence of modern human behaviour in terms of “the greater information flow, planning depth and conceptualisation consequent upon the emergence of language”, a view influenced by a Eurocentric perspective, despite the clear recognition that Australian archaeohistory provided a fundamental challenge (Davidson & Noble 1989; 1992). McBrearty and Brooks (2000) sought to correct the Eurocentric bias by showing important behavioural changes in African archaeohistory similar to and sometimes earlier than in Europe. They mostly used the Eurocentric criteria established in 1991, but systematised the list of traits. Many scholars have sought to measure particular aspects of behaviour against that trait list (e.g., Nowell 2010). Others have shown its limitations (d’Errico 2003 and, for East and Southeast Asia, Haidle & Pawlik 2010), paradoxically reinforcing its apparent importance. Because other hominins preceded modern humans in both Africa and Europe, it is sometimes difficult to be certain that the traits being discussed were really distinctive. Some of the traits occurred before modern humans evolved, some first occurred only long after modern humans, and none appeared within the 50 thousand or so years of emergence, suggesting a complex relationship between modern humans and the archaeological evidence said to characterise their behaviour. Finally, the evidence of Australian archaeology suggests that not all of the traits are necessary to identify modern human behaviour (Davidson 2010c). There was no “package” of traits (cf. Habgood & Franklin 2011), because the most important feature of modern humans is really behavioural flexibility (Veth et al. 2011) and the cognition that allows that. It is, therefore, irrelevant that the traits of the “package” accumulated in Australia gradually (despite the importance attached by Powell et al. 2009). Discussing the presence or absence in Australia of traits important in Africa is letting the tail of the history of archaeology wag the dog of archaeohistory. Historically, the McBrearty and Brooks trait list may have reached the end of its usefulness.

Theoretical Questions The question is, What sort of differences between modern humans and earlier hominins might account for modern humans reaching Australia (and the Americas) while earlier hominins did not? Homo erectus had been in Sunda (the continent that joins the islands of Indonesia to Southeast Asia except during interglacial periods of high sea level) for more than 1.5 million years (Dennell 2009, 165–166). Hominins had to make some sea crossings to get to Flores before a million years ago (Brumm et al. 2010; Morwood & van Oosterzee 2007), but there is no evidence they went further.The options are either nothing changed and people made the several crossings to Australia by a series of accidents or a fundamental change made intentional crossings possible. Australian archaeohistory is important because the best explanation for the difference is a key to an explanation of all modern human behaviour. Understanding the uniqueness of Australian archaeohistory reveals important things about the evolution of all human cognition. 244

The First People of Australia and Its Adjacent Islands Table 19.1.  Problems solved in Australian colonisation Cognitive issue

Practical problem

Conceptualisation of problem Visualisation of solution Implementation of tasks necessary for solution Delay between conceptualisation and achievement

How to obtain fish Cross a water Pelagic fish from deep water barrier remains Making nets or lines Building a (Balme with hooks seaworthy craft 2013) Working Obtain materials from different memory locations (Coolidge & Combine them in Wynn 2009) creative ways Use resultant artefact in another location Continue focus on problem solving despite distractions Carry water One crossing always more than 70 km Both men and women Population large Simulation in enough to grow McArthur, rapidly 1976

Thought for the future

Alternative problem

Empirical support

Theoretical issue

Some time before the first colonisation of Australia there was a fundamental change between hominin and human cognition which made human intentionality possible. By about 50,000 years ago (O’Connell & Allen 2007), modern humans had the cognitive ability to be undertaking routine activities in sea-worthy watercraft (Allen & O’Connell 2008; Davidson & Noble 1992; O’Connell et al. 2008). Earlier hominins must have made sea crossings, because of their longterm presence (with at least two long hiatuses) in Flores (Brumm et al. 2010). In the current state of knowledge of contemporary hominin abilities (Davidson 2001), such crossing(s) seem more likely to have been accidental (e.g., by rafting, as suggested by Smith 2001). Anecdotal evidence from recent tsunamis shows that some such water crossing might have been possible (Morwood & Davidson 2005). But to get to Australia, by any route, at least eight crossings were required (Birdsell 1977); at least one of them had to be about 70 km, and by most routes one was about 90 km. Even for modern people setting out to reach an unseen Australia, there were significant logistical problems. Making watercraft implies modern cognitive ability (see Table 19.1). Previously, Davidson and Noble (1992) emphasised language – communication using symbols – and its impact on mental ability. They supported that argument by showing material evidence of symbol use early in Australian archaeohistory. While that was an important first approximation, it led to an understanding that the evolution of cognition was more complex than the dichotomy we envisaged (see Davidson 2010c). In cognitive terms, separation of tasks in time and space is one of the most indicative features of this achievement. For some parts of the manufacture of the watercraft, mental processes were necessary that were not directly related to stimuli from the interaction of the senses with the external environment (Barnard et  al. 2007). The process was necessarily protracted such that 245

Iain Davidson Table 19.2.  Examples of early indications of emerging complex technology Example

Date

Location

Reference

Heat treatment of tool stone

164 ka

Brown et al. 2009

Paint preparation

100 ka

Preparation of gums and resins for hafting Bow and arrow set technology

70 ka

Hafted projectile point (found embedded in bone of ass) Pitch/resin preparation

50 ka

Pinnacle Point, South Africa Blombos, South Africa Sibudu Cave, South Africa Sibudu Cave, South Africa Umm el Tlel, Syria

50 ka

Königsaue, Germany

64 ka

Henshilwood et al. 2011 Wadley et al. 2009 Lombard & Haidle 2012 Boëda et al. 1999 Koller et al. 2001

Table 19.3.  Evidence of early voyages around the edge of Sahul Evidence

Location

Date

Reference

Water crossing

New Ireland

44 cal ka

Leavesley & Chappell 2004 O’Connor et al. 2011a Leavesley 2006

Substantial numbers Jerimalai in Timor Before 41 cal ka of pelagic fish L’Este Crossed 120 km of Buka at the northern Before 32 cal ka open sea end of Bougainville Crossed 200 km of Manus 25 cal kaa open sea a

Spriggs 2001

The date, and its stratigraphic position, are currently unpublished.

distraction from the solution was possible, and its overall achievement was possible only with extended Working Memory (e.g., Coolidge & Wynn 2009). Examination of the “cognigrams” of artefact manufacture analysed for other and earlier locations (Haidle 2010) shows that even complex earlier toolmaking did not achieve this degree of displacement within the whole chain of actions and function. Some indications of early emerging cognitive complexity can be seen particularly but not exclusively in southern Africa (see Table 19.2). As I have argued previously, such cognitive achievement would account for the deliberate use of watercraft to the north and west of Sahul shortly before its first colonisation, probably between 50 and 45 ka (see also O’Connell et al. 2008; Summerhayes et al. 2010; Summerhayes & Ford, this volume). The deliberate nature of the use of watercraft can be seen in the further voyages around the edges of Sahul (see Table 19.3). Finally, people carried animals in their boats to the east, where the cuscus, Phalanger orientalis, was introduced to New Ireland by 24 cal ka (Gosden 1995). Several Sahul species were also carried westwards to Wallacea during the Holocene (Heinsohn 2001). Our argument of 1992 sought to establish that the cognitive differences between modern humans and earlier hominins stemmed from the capacity to communicate with others about things not in the immediate contingency of the utterance, by using sounds and gestures that were not necessarily determined by the objects of the utterance (Noble & Davidson 1996). People were thus able to speak about memories of past occurrences and to imagine futures on the basis 246

The First People of Australia and Its Adjacent Islands of such knowledge. Out of this arose the conceptualisation necessary to solve problems such as the construction of watercraft. This version of the implications of the first colonisation was improved by more complex modelling of a sequence of conditions in the evolution of human and hominin cognition derived from Barnard’s Interacting Cognitive Systems (ICS) model (Barnard et al. 2007). In the final phase of the sequence, modern humans can think about things for which there is no stimulus from the external environment – and can thus think novel thoughts.The emergence of this internal mental process could be identified as the emergence of reflexive meaning entailing the use of symbols that was at the heart of the earlier argument (Davidson & Noble 1989). Moreover, because the ICS model emphasises the integration of systems that receive inputs from external stimuli, it is much more suitable for understanding the essentially social nature of human mindedness than any model of cognition that is principally based on processes taking place inside the brain. In these arguments the evolution of cognition required more than just genetic changes but also derives from the behavioural contexts (evidenced from archaeology) of such changes. The first Australians had achieved the cognitive abilities of all modern humans, yet they appear to have fallen short of the McBrearty and Brooks’s standards (e.g., Brumm & Moore 2005). The traits do not assemble neatly into a “package” that resembles the behaviour of the first modern humans in Europe, which is a problem with the method derived from European (and African) archaeology rather than for the archaeological record of Australia. The circumstances of the first colonisation of Australia establish the modernity of the cognition of all Australians. All other elements of the so-called package are irrelevant. So why is it so difficult to see this outside Australia?

Fundamental Questions At issue are some very fundamental aspects of the interpretation of archaeology – particularly the interpretation of hominin and human skeletal remains in physical anthropology and attempts to write culture histories derived from unstated assumptions about stone tools.

Physical Anthropology The first question is the validity of classifications of hominin fossils as a framework for understanding the past. This was brought into stark relief by the discovery in 2003 of strange hominin skeletal remains at Liang Bua on Flores (Morwood et al. 2004), one of the Lesser Sunda islands of Indonesia that were always separated from Sunda and Sahul by sea barriers. The straightforward view is that this was a new species, Homo floresiensis (Brown et al. 2004), distantly related to humans, which, if present dating is confirmed, survived on Flores long after the dates for colonisation of Australia (Westaway et al. 2009). The different interpretations of this evidence threw the discipline of physical anthropology into an unresolved crisis. Claims have been made that involved comparisons of one or more characters with all hominins from Australopithecus afarensis (Homo habilis, H. rudolfensis, H. georgicus, H. erectus, H. antecessor), or that described the creatures as the remains of modern humans either with one pathology or another, or with no pathology (see Aiello 2010).The date of the earliest remains at Liang Bua (Westaway et al. 2007b) is earlier than most estimates for the earliest modern humans leaving Africa, making problematic any claim that these are remains of a modern human. A recent study seems to support the hypothesis that H. floresiensis descended from early Indonesian H. erectus by extreme insular dwarfing (Kaifu et al. 2011), which would forestall speculation about the possibility of earlier expansion of hominin species that have never been found outside Africa. Some have argued that the presence of Homo floresiensis in Flores makes it less likely that modern humans passed through that island before they eventually reached Sahul (Balme et al. 2009). 247

Iain Davidson It is possible that the Lesser Sundas were only colonised by modern humans coming later from the east (Davidson 2007a). In support of this argument, the weak genetic signal of pre-agricultural populations in Bali had coalescence ages of 8.2 ka or 12.7 ka, probably involving colonisation from further east in Indonesia or from Melanesia (Karafet et al. 2005). Further support comes from analysis of the presence of Denisovan genetic material in Australia, New Guinea and some of the islands of Southeast Asia (Reich et al. 2011). The unresolved disputes about the skeletal remains from Flores pose a further problem for archaeohistory generally:Are the identifications and the naming of species, and particularly the methods of identifying relationships between such species, good enough to allow the definition of patterns of movement of hominins out of Africa? The Liang Bua remains include the remains of many parts of the skeleton of one individual. If such complete remains are so difficult to interpret, how confident can we be of the groupings, names and relationships applied to other, more fragmentary fossils? A new hominin species has been identified in north East Asia at Denisova in the Altai Mountains of eastern Russia on the basis of its ancient mtDNA (Krause et al. 2010) and its whole genomic DNA (Reich et al. 2010). The remains are a large upper molar and a finger bone from mixed sediments unrelated to the well-dated sediments elsewhere in the cave (see supplementary on-line material [SOM] in Reich et al. 2010) so they are actually undated (but probably less than 125 thousand years old). Such remains are very undiagnostic, so unless skeletal remains can be found that preserve aDNA of Denisovan type, it will be impossible to identify other bones as Denisovans. Some of the less well-classified skeletal remains from China could actually be those of Denisovans. There could even be hitherto unrecognised species of hominin. Recently described anomalous skeletal remains from Mongolia, Southwest China, and Laos (Coppens et al. 2008; Curnoe et al. 2012; Demeter et al. 2012) could be Denisovans or a new species. This possibility and the presence of high proportions of Denisovan DNA among modern people of Sahul (Reich et al. 2011) raises the possibility that the first colonisation of Sahul was from East Asia either through Taiwan and Luzon or on the eastern side of the Sunda Shelf. The difficulties of interpreting skeletal remains have always been apparent in Australia with the conflict between the trihybrid (Birdsell 1967), dihybrid (Thorne 1977b) and single origin theories (Brown 1987a), and the persistence of beliefs in multiregional evolution long after it was rejected in the rest of the scholarly world (e.g., Groves & Lahr 1994). The suggestion that robustness in Australian skeletal remains might be an indication of a close genetic relationship with Javanese Homo erectus has been killed more than once (Brown 1992, 2010; Westaway & Groves 2009). One of the reasons for such confusion in Australia is the determination to give names to groups of fossils as if they represented species as reproductively isolated as living biological species. Clearer insights can be had from representing the evidence in terms of variable interbreeding populations (Pardoe 1991a).

Stone Tool Studies The second question is whether stone tools are a good guide to the connections between behaviours in different parts of the world – again arising from naming practices for types of tools and of assemblages of tools. A deliberately selective handful of artefacts from the finds of Liang Bua, published with the original species description (Morwood et al. 2004), became part of an argument that the tools were too “sophisticated” to have been made by any hominin other than modern humans (Martin et al. 2006). Detailed analysis of the artefacts from Liang Bua (Moore et al. 2009), and of those from the earliest sites in Flores several hundreds of thousands of years older (Brumm et al. 2006), has exposed the error of relying on drawings of a few carefully selected artefacts to pronounce on such matters. Notoriously, Australian archaeologists have been sceptical of typologies over the past 40 years and more (see Holdaway & Stern 2004, 283–315), to the extent that artefacts are now rarely 248

The First People of Australia and Its Adjacent Islands illustrated in published reports (but see Smith 2006). One synthesis of Australian archaeohistory did not mention stone tools at all (Davidson 1999), and the latest synthesis has only eight illustrations of stone artefacts (Hiscock 2008). Interpretations of Australian artefacts can be misleading, particularly those comparing Australian stone industries to the Old World Oldowan, Acheulean, Levallois, Mousterian, Upper Palaeolithic sequence (OALMUP; Davidson 2009). Aspects of Australian stone tool use have been compared with the Oldowan (e.g., Toth 1985); handaxes have been found, particularly in western Queensland (see analysis in Moore 2003) and the Northern Territory, but they are completely unrelated to the Acheulean “tradition” spatially and chronologically (Davidson 2002); a technique that would be called Levallois by experienced analysts (Sonneville-Bordes 1986) was used in Australia (Moore 2003), but some of the artefacts that might be identified as Levallois points elsewhere were made by a blade-making technique on single-platform cores known as horsehoof cores (Binford & O’Connell 1984). The equifinality involved weakens the case that the Levallois and related techniques were the fundamental technologies taken by modern humans from Africa to the rest of the world (see, e.g., Foley & Lahr 1997). The fact that, in 1974, these artefacts were made on horsehoof cores otherwise said to characterise the earliest stone industries of Australia (Bowler et al. 1970) further undermines the use of both “Levallois” points and horsehoof cores as type fossils of early industries. Finally, it was suggested that “blade” industries do not occur in early assemblages from Australia (Mellars 2006b), somewhat confirmed by the details of the assemblage from Puritjarra (Smith 2006). The whole question of blades is more complicated than it is generally said to be in the standard Old World literature: blades do appear throughout the sequence in Australia (Davidson 2003). In particular, at Lake Mungo, before 40 ka, conjoining flakes onto horsehoof cores showed that blades had been removed from the site (Shawcross 1998). The percentage at Puritjarra (3.6% calculated by the author from the data in Smith 2010)2 is very low for an assemblage from the period of modern humans but higher than the proportion at the Middle Pleistocene site of Kapthurin in Africa (Johnson & McBrearty 2010). There has been a long history of trying to fit Australian stone artefacts into the framework established for Europe and, to a lesser extent, Africa, and each attempt either failed or was only partially successful. There are two contradictory explanations for this: either the OALMUP sequence may indeed be a universal aspect of hominin and human progress, and Australians just did not measure up; or the Australian story tells us something fundamental about the flaws of trying to write a narrative of hominin and human archaeohistory based on five basic stages that covered a 2-million-year period. The reason why the second option is more likely is well exemplified by examining Mellars’s (2006b, 798–799) attempt to account for the inconsistency with the OALMUP sequence by (1) a scarcity of suitable raw materials; (2) the lack of necessity for complex tools in economies that emphasised marine resources and did not need to prepare skins for clothing; and (3) founder effects, drift and “progressive loss in the complexity and diversity of culture and technological patterns”. Underlying this argument is an assumption about the continuity of traditions and the probability of convergence on particular forms of stone-flaking product, which goes something like this: • • • •

Archaeologists can recognise patterns in flaking products that vary in time and space. These patterns are meaningful in terms of the intentional actions of the knappers. Flaking products that are close to each other in time show similarities. When there are major changes in flaking products between chronostratigraphic periods, this indicates a cultural change. • Flaking products are generally more similar between adjacent regions than between distant regions, provided that raw material does not vary very much from one region to another. • When flaking products show big differences between adjacent regions, this may indicate a lack of cultural connection between them, provided raw materials are not substantially different. 249

Iain Davidson Table 19.4.  Separate traditions of bifacially flaked points Tradition

Location

Date range

References

Stillbay Szeletian

Southern Africa Central Europe

Tribolo 2009 d’Errico et al. 2011

Solutrean

Western Europe

Clovis

Central United States

Kimberley

Kimberley, Australia

70–80 ka 36.8–42.4 cal ka for 25–75 percentile of dates 21.8–24.3 cal ka for 25–75 percentile of dates 12.7–13.2 cal ka youngest and oldest dates After 1.4 cal ka

d’Errico et al. 2011 Collard et al. 2010 Harrison 2004

• The necessary assumption for all of this to be true is that methods for making such flaking products were handed down from one generation to another. This set of generally unstated assumptions has been dominant from the 19th century in Europe, but the ethnographic record from Australia does not offer strong support for archaeologically defined types being important for people who made and used flaked stone artefacts (Holdaway & Douglass 2012). Some of these propositions seem consistent with the archaeological record from stratified sites, but in the absence of the stratigraphic evidence, inference is much more problematic. Some anomalies can be well illustrated by two different patterns of bifacial flaking. In the first of these cases, bifacial flaking had similar end results in the chronologically separated Acheulean and Mousterian of Acheulean Tradition as well as in Australia very much later. This chronological and spatial separation suggests that the particular outcome may not be the result of tradition at all, but one end product of particular knapping strategies. The Australian evidence suggests that, rather than a continuous tradition, bifacial flaking, the products of which archaeologists call “handaxes”, may have been invented or discovered thousands of times independently. At later dates, bifacially flaked points (Table 19.4) are unlikely to have any connection through a common tradition (for the Solutrean and Clovis, see, e.g., Shott 2013; Straus et al. 2005). The most likely interpretation is that these bifacially flaked points are also examples of people using a range of similar techniques to arrive at a solution to particular needs quite independently. If pieces as apparently complex as large bifacial cores (as in the Acheulean) or smaller bifacial points could be discovered or invented more than once, they may have been invented many, many times, independently, and the same would be true of other “types” such as backed artefacts (contra Mellars 2006b) where the arguments about what is to be included in the comparison can be quite complex (Hiscock & O’Connor 2005). But the propensity of archaeologists to see connection where there is similarity produced narratives that depended almost entirely on unstated and unexamined assumptions. Despite appearances, continuity of cultural tradition is not as strong as it has been assumed to be; given a relatively small number of options in knapping, similar outcomes are to be expected (Moore 2013). It has been misleading to claim that similarities provide an adequate basis to show cultural connection over long distances or long time periods. By the same token, lack of similarity does not licence writing a cultural history in terms of “failure” to achieve the goals of an illusory progress defined inductively from the record of Europe.

What Is to Be Explained? In spite of all reservations about the sources of evidence, the picture emerges of people, Homo sapiens, reaching Sahul, possibly by more than one route, between 50 and 45 ka. By the very fact of their achievement they must all have had modern cognitive abilities. One of those abilities 250

The First People of Australia and Its Adjacent Islands Table 19.5.  Some counts of languages in Sahul Location

Language grouping

Approximate numbers References

80% Australia

Pama-Nyungan

275

20% Australia New Guinea

Non-Pama-Nyungan Trans New Guinea (TNG) Papuan (not TNG) Other languages Austronesian

75 350, maybe 450

New Guinea New Guinea New Guinea

> 250 150

Bowern 2012; Blake 1988; Evans 1988 Bowern 2012; Evans 2005 Pawley 2005 Pawley 2007 Ross 2005 Pawley 2007

was the use of symbols, and this became the dominant feature of their behaviour once on the continent of Sahul. There is a case that rock art was established relatively early and that personal ornaments may have been used too (Balme et al. 2009; Mulvaney 2013). Recent analysis of the early symbolic evidence has shown that symbol use was patterned within regions and different between regions (Habgood & Franklin 2011). One of the functions of archaeohistory is to provide an account of how the conditions recorded historically emerged in an unrecorded past (e.g., Davidson 2010b). There are two questions here. First, what was the variation at the time of the historical record that needs to be accounted for? Second, what processes of introduction, diffusion or independent invention contributed to the historically recorded variation?

Language The small number of first colonists of Sahul presumably spoke one, or at most a few, languages with little variation at the time of their arrival. At the time Europeans arrived, there were more than 1,500 languages in the region and much variation in social and material culture (Table 19.5). The central issue is how that diversity emerged from those small beginnings. The languages of Sahul and adjacent islands, therefore, show enormous diversity but only minor or later regional evidence of introductions from outside. For example, the more recent Austronesian languages (Pawley & Ross 1993) resulted from migration of people from the north who brought with them pottery, pigs, chickens and possibly dogs (e.g., Spriggs 2007), but the majority of the variation seems likely to have resulted from change within the continent. The archaeohistory, therefore, is likely to be related in some way to the generation of that diversity.

Material Culture There is also documentation of variation in many other social and material ways (Table 19.6) (see also Davidson 2013). In each of these cases, variations across the continent can be considered stylistic and hence, at one level, symbolic. Comparisons with and within New Guinea emphasise this point (Davidson 2013). Analysis of the archaeohistory of people on Sahul should show how such diversity could have emerged from presumably less variable origins. A start has been made on such comparisons with syntheses of the earliest rock art of Australia (David et al. 2013; Mulvaney 2013) and of the later diversification and the various social and demographic pressures that may have been involved (Ross 2013; Davidson 2010b). The archaeological record reveals the transfer of materials dating from early in the peopling of Sahul (see Stern 2009;Veth et al. 2011), and at contact there were several examples in Australia of more complex long-distance movement of materials through exchange or trade dating back 251

Iain Davidson Table 19.6.  Selection of material culture items that show variation across Australia Material culture item

Reference

Spearthrowers (woomera or atlatl) Netting and basketry Spears Watercraft Boomerangs Rock art Other features of decorative art Footwear Practices in the disposal of the dead Body decoration Beads and pendants

Davidson 1930 Davidson 1933; Roth 1904 Davidson 1934 Davidson 1935 Davidson 1936b; McAdam 2001 Davidson 1936a; Layton 1992 Davidson 1937 Davidson 1947 Davidson 1948; Roth 1907 Brady 2005; Roth 1910 McAdam 2008; Roth 1910

Note: For a recent summary, see Satterthwaite and Arthur (2005).

1,000 years (Davidson et al. 2005; McBryde 1986; Mulvaney 1976). Trading relationships in New Guinea showed great variety in form and materials (e.g., Hughes 1977; Malinowski 1922;Wiessner and Tumu 1998). In almost all examples, the function of the exchange was not only the provision of goods but also the maintenance of social relations through ritual.

Social Organisation and Ritual There were also variations in the social organisation across the continent (see Keen 2004; 2006) and in rituals associated with initiation that seem to have been changing in their spatial distributions at the time of European disruption. It is inescapable that these variations were structured symbolically. The symbolic and spiritual life of different Aboriginal groups was dominated in most cases by the belief system known as the Dreaming, whatever “subtle . . . variations occurred in different regions” among those beliefs (Stanner 1979 (1962), 115). The appearance of uniformity seems to derive from naming practices among anthropologists (see Wolfe 1991), just as naming has determined how skeletal variation or stone artefact variations are represented. Some published accounts and many informal ones link mythological beliefs with rock engravings (e.g., Layton 1992, esp. ch. 2; Merlan 1989) although it may not be possible to trace the ritual system back more than two millennia (David 2002). Ritual itself is certainly older and used many of the same images as those rituals later associated with the Dreaming (Ross & Davidson 2006).

Subsistence Within the pattern of subsistence practices at contact (see Keen 2004), there was variation dependent on the nature of the available resources, with varying degrees of mobility, aggregation and dispersal. Arguably, the situation was more complex in many parts of Australia, but because of the extent of destruction of Aboriginal society as a result of intrusion of Europeans, subsistence practices have to be reconstructed from fragmentary evidence and inference from archaeology (White 2011). In New Guinea the situation was generally more complex, usually characterised as having agricultural or horticultural systems, but in keeping with the remarks here about the difficulties that arise from unthinking acceptance of named categories, there have always been substantial 252

The First People of Australia and Its Adjacent Islands Table 19.7.  Studies of the subtlety of subsistence of fisher-gatherer-hunters in Australia Subtlety

Region

References

Fire as management tool

Widespread

Propagation of plant fragments after harvesting

Widespread

Use of nets Fish traps Watercraft Hooks Manipulation of eels

Variable Variable Variable Variable Victoria

Bowman & Haberle 2010; Gott 2003; Hallam 1975; Head 1994, 1996; Hope 2009; Latz 1995 Denham et al. 2009; Hallam 1989; O’Connell et al. 1983; Powell 1982; Summerhayes et al. 2010 Satterthwaite 1987 Akerman 1976; Gibbs 2011 Davidson 1935 Lampert & Turnbull 1970; Walters 1988 Lourandos 1980; 1997

numbers of societies that subsist on wild plants, fishing and the meat from wild animals. There is little correlation between such named subsistence types and other measures of social or other complexity (Roscoe 2002). Over the past few decades, there has been a considerable change in the perception of the stereotypes of Aboriginal food-getting (see Table  19.7), with various enhanced subtleties that go beyond the dichotomy of fishing-gathering-hunting and agriculture (e.g., Hynes & Chase 1982; Lourandos 2008; White 2011). Extreme climatic variations could lead to famine and starvation (for Australia, see primary references in Davidson 1990). This was not only a feature of arid lands but has also been documented for the regions of the Murray River and of the Gulf of Carpentaria. Climatic variation also impacted peoples in the Strickland River region of New Guinea where different subsistence strategies resulted in different responses to drought (Minnegal & Dwyer 2007). O’Connell and Allen (2012) argued that much of the variation in Australian archaeology can be understood in terms of the constraints of prey choice, patch choice and other aspects of optimal foraging theory. Two key issues are underestimated in such considerations (see Davidson 2012): How did initial colonists learn about the edible plant species, where to find them and how to cope with the variation in the amount of resource as climatic conditions varied; and how did initial colonists establish their attitudes to food resources and locations through defining symbolic relationships to place and ritualised relationships among the people who acquired food? All the assumptions about the relevance of modern food resource acquisition must be understood in the context of the different roles that men and women have in the food quest (see Bliege Bird & Bird 2008): women gather low-risk resources to maintain networks, men hunt high-risk resources to maintain ritual status and power. It is the symbolic representation of resource options that has a significant impact on responses to variation.

What Sorts of Explanation Are Possible? Population Ups and Downs The apparent concentration of the earliest sites in Australia in refuge zones (Veth 1989) suggests that the first colonists were opportunistic generalists exploiting biogeographic diversity during harsh climatic episodes, consistent with the label “optimal maladaptation” (Rindos & Webb 1992) for people exhausting high-ranked resources and moving on. More recent work emphasises that, after following the savannah to Sahul, people occupied all its environments other than grassland relatively quickly (see Balme et al. 2009) (Table 19.8), reaching Tasmania very soon after the formation of the land bridge to the mainland just before 40,000 years ago (Lambeck & Chappell 253

Iain Davidson Table 19.8.  Early sites in Sahul and their environments Site

Location

Environment

Date

Reference

Ngarrabulgan Bobongora

Cape York Huon Peninsula, New Guinea New Britain

Savannah Tropical rainforests

40 cal ka 44 ka

Tropical rainforests

40.7 cal ka

David et al. 1997 Groube et al. 1986 Pavlides & Gosden 1994 David et al. 2011

Yombon

Shrublands between Arnhem Land, savannah and arid Northern Territory zones Carpenter’s Gap Southern Kimberley Shrublands between and Riwi savannah and arid zones Nawarla Gabarnmang

Lake Mungo

Willandra Lakes of NSW

Puritjarra Devil’s Lair

Central Australia Southwest Western Australia Tasmania

Parmerpar Meethaner

Shrublands between savannah and arid zones Arid zone Savannah (or tall open woodland) Steppe

45.2 cal ka 43 cal ka

42 ka 35 cal ka 48 cal ka 40.7 cal ka

Balme 2000; O’Connor & Fankhauser 2001 Olley et al. 2006 Smith et al. 2001 Turney et al. 2001a Cosgrove 1995b

2001).This rapid colonisation of the continent, not confined to the coast, had implications for the demography of early colonisation. One consequence concerns the demography of the whole occupation of Australia.The growth rate of population over the pre-European period was in the order of 0.0001/yr or 0.01% per year (Davidson 2010b), less if the initial population was higher. We know that Tasmania was occupied almost as soon as the land bridge opened, approximately 10 thousand years after first landfall. Using the same formula we can calculate the total population of Australia at that growth rate: if the initial colonists were 20 people, then the total population of Australia by the time they reached Tasmania would have been only 33 people! Even if the initial population was 1,000, there were only about 500 more in the time it took to reach Tasmania. Something else was going on. It is much more likely that growth rates were often higher (see Davidson 1990), but there were local and regional crashes as a result of environmental hazards or failures to understand the vagaries of Australian plant and animal species and extreme fluctuations in availability of water (see Davidson 1999). Plant species at landfall may have shown some similarities to species available in Sunda (Golson 1971), but these were generally limited in their distribution, and familiarity declined rapidly as people moved away from northern Australia. There is empirical support for population crashes from time-series analysis of radiocarbon databases (Smith et al. 2008) or a less powerful region-by-region site-by-site analysis which shows that early sites were probably occupied discontinuously (Davidson in press).

Demography and Modern Human Behaviour The current fashion for explaining aspects of cultural variation using demographic models got a kick start from Henrich’s (2004) modelling of the implications of Tasmanian material culture impoverishment (e.g., Jones 1977). Henrich argued that because Tasmanians were isolated following the Holocene sea level rise on an island their population lacked sufficient numbers of creative 254

The First People of Australia and Its Adjacent Islands individuals to generate new skills or to pass them on with sufficient accuracy into the cultural system. Henrich’s model might apply to the late emergence of hierarchically structured stonetool-making techniques both among hominins generally and particularly in Australia (Moore 2012), but it would not be enough to account for the convergence on a small number of types and techniques (but see Moore 2011). The demographic model has been extended to the whole history of expansion of modern human behaviour, with impressive results (Powell et al. 2009), particularly in the context of estimates of population history from skyline analysis of mtDNA data (Atkinson et al. 2008). Neither of these arguments deals well with the situation in Australia with high diversity of languages and material culture but generally low population densities. A recent comprehensive overview of hominin and human cultural evolution suggested that such diversity may reflect both ecology and demography but that a key factor was the isolation of small populations (Foley & Lahr 2011) unless mobility strategies spread risk through contact between communities (Nettle 1998). Here is the problem for the application of demographic models to Australia: mobility strategies seem to have been in place through ceremony and shared “Dreamings” (Davidson 2010b; Yengoyan 1968), but these seem likely to be late in the archaeohistory (David 2006). On the other hand, population growth sufficient to generate and maintain the complex artistic, ritual and social lives of the historic period seems inconsistent with the knowledge of the low population at contact and the implied rates of growth over the whole human experience in Australia.

The Importance of Australia for Understanding Human Evolution The archaeohistory of Sahul has some important lessons for understanding the evolution of hominins and humans in other parts of the world. Explanations that work in Sahul should work elsewhere, while those that do not work in Sahul are probably not good explanations. Sahul archaeohistory suggests that conceptualising the emergence of modern human behaviour by generalising from European archaeology and adding inconvenient evidence from Africa presents a distorted picture. The first colonisation of Sahul suggests that there was a real cognitive difference between hominins who lived in Sunda beforehand and those who made it across the seas that separated Sunda from Sahul.The main elements of this cognitive difference were the conceptualisation of a problem without immediate external stimuli, the visualisation of its solution, and the persistence in the task despite the occasions for distraction from it. The skeletal finds from Flores suggest that the archaeohistory of the region is more complex than was envisaged by some simple models. In particular, the finds call into question some of the certainties about identifying species of fossil hominin and make some routes through the islands between Sunda and Sahul more likely than others. Stone artefacts of Australia undermine some of the assumptions about the meanings of the supposed stone tool types identified by people working in the Old World. It may no longer be sufficient to claim that similarity implies connection or that lack of similarity implies “failure” to reach the goals set elsewhere. The enormous diversification of languages, material culture, social organisation, ritual and subsistence across Sahul raises questions about the role of ecology, demography and social organisation in such diversification. At the same time, the appearance of impoverishment in the aspects of Tasmanian culture after isolation may be interpreted in terms of a model of the impact of particular demographic conditions on innovation and transmission of technological skills. While there is some support for this, it would not account for the inventiveness and long traditions in many aspects of Aboriginal life in other parts of Australia. 255

Iain Davidson Finally, the archaeohistory of Sahul is too important to be ignored by theorising that accounts only for hominin and human evolution in other parts of the world.

Acknowledgments For help, advice, discussion and support in various ways I thank Jim Allen, Helen Arthurson, Miriam Haidle, Geoff Irwin, Mark Moore, Frances Morphy, Kimberlee Newman, Jim O’Connell, Sue O’Connor, Nic Peterson, Mike Smith, Matthew Spriggs, Glenn Summerhayes, Sean Ulm, Michael Westaway, Jane Balme, Nikki Stern, Peter Veth and Jo McDonald. I accept sole responsibility for the way I have used their goodwill. Notes 1. I refer to the history constructed from archaeological evidence as “archaeohistory”, as well as using the word for the period in which the only evidence for that history derives from archaeology (Davidson 2010b). The word “prehistory” has become inappropriate because there is too much confusion with dinosaurs. 2. Calculated from substantially whole, unmodified conchoidal flakes from the database 13.2.2_PJ_Lithics_ Dbase-A_Atrributes_Flakes.xls in Smith 2010, which had specified raw material, feather terminations and no signs of core rotation. From this database, I identified blades if the elongation (Length/Breadth) was ≥ 2 and the Parallel Index (Flake Breadth/Platform Breadth: Smith 2006, 385) was ≥ 0.75 and ≤ 1.25.

256

Chapter 20 Essential Questions Modern Humans and the Capacity for Modernity

Martin Porr

Introduction This chapter is aimed at critically discussing some elements guiding more recent discussions of the origins of ‘modern humans’ in the fields of Paleolithic archaeology and paleoanthropology. It argues that the concept of modern humans is informed by an essentialist view of the definition and understanding of human beings in Western thought that has a long philosophical history. In this tradition, humans are foremost separated from the rest of organic life by the fact that they possess specifically “human” mental or cognitive abilities. This view was enhanced by a definition of humanity that was propagated as a reaction to the atrocities that were committed during the 1930s and 1940s in the name of racism. Ernst Mayr, the main proponent of this movement, basically regarded an inclusion of physical or anatomical features as unethical in the definition of our own species. These views have been given a new dimension more recently through molecular studies of the genetic makeup of present human populations, which seem increasingly to point toward an origin of our species in sub-Saharan Africa. Consequently, it is implied that so-called genetically modern humans also are endowed with the capacity for behavioral modernity and this ultimately explains their success in colonizing the world and as a species. These developments, however, seem very much contradicted by discussions within molecular biology itself, where it is increasingly accepted that the links between genome and organism are far from straightforward. As a consequence, the notion of cognitive and genetic capacity for modern thought is in constant danger of becoming a virtual construct that can be filled with a whole range of features and abilities that reflect more the attitudes of the respective researcher than the universal cognitive endowment of our species. As paleoanthropology and Paleolithic archaeology have been mostly influenced by adaptationist models of behavior in the past few decades, ‘modern humans’ appear mostly as products of the optimization of extractive behaviors and technologies (Clark and Willermet 1997; Henshilwood 2007; Henshilwood & Marean 2003; Mellars 2006c; Mellars & Stringer 1989a; Shea 2011b). The consequence is a deterministic, linear, and teleological narrative of modern human evolution, which nevertheless struggles to be coherently linked to the accepted mechanisms of organic (Darwinian) evolution. 257

Martin Porr The reasons for this can be related to the contradiction between local mechanisms of variability and selection and the concept of modern thought (behavioral modernity) that is constructed as if it transcends the variation of organic life and remains essentially unchanged from the moment of its creation.

Historical Legacies Despite the fierce debates surrounding the origins of modern humans in the past few decades, it seems that most authors have at least since the 1980s shifted toward a definition that is based on behavioral/cognitive aspects alone (see, e.g., Clark & Willermet 1997; Mellars & Stringer 1989a). The current most widely accepted view has shifted the identity of the human organism from recognizable external characteristics to internal ones that are expressed (and potentially detectable) through behavior and, by extension, material culture. In a sense, this appears to show resemblance to Carolus Linnaeus’s taxonomic treatment (first published in 1735) of Homo sapiens for which he “abandoned his usual practice of providing a [morphological] diagnosis for each taxon” (Schwartz & Tattersall 2010; 95) and included only the remark nosce te ipsum (know yourself) as a description. Alfred Russel Wallace (e.g., 1889) was also struggling to find an adequate place for “the moral nature and mental faculties”, which he regarded to be specific to the “civilized races” of humanity, within the Darwinian explanatory framework. Famously, he consequently argued that the “moral, intellectual, or spiritual faculties” could not have not been “derived by the action of the same general laws as his physical structure” (see, e.g., Ingold 2004). These remarks point to the deep historical legacies that continue to inform more recent discussions as well as to shape some fundamental issues related to the definition of human beings in relation to differences from and similarities to animals. Aristotle already combined in his definition of humans morphological or anatomical aspects (bipedalism, freeing of the hands from locomotion) and (as one might say) behavioral/cognitive ones, the ability to think and reason. The latter he regarded as humanity’s “substantial being” or “defining character”, which was ultimately a consequence of a human’s essentially divine endowment (Schwartz & Tattersall 2010, 94). During the 19th century and before human fossil specimens were discovered and accepted, Johann Friedrich Blumenbach and Johann Wolfgang von Goethe also struggled with the morphological variability expressed by living humans. Despite their disagreements over the importance of certain anatomical features, they both concluded that “the most important attribute” to separate “man” from “the brutes” must be “reason”. Upon discovering similarities between a juvenile human and a chimpanzee, Edward Tyson had reached the same conclusion already in 1699 (Schwartz & Tattersall 2010, 95). The introduction of Darwin’s theory of natural selection presented a challenge not only to the established view of the divine nature of human beings (and their equally divine origins). It also needed to come to terms with the material or, rather, organic basis of human intellectual abilities and the processes that led to their phylogenetic development. It is not necessary here to discuss Darwin’s own struggle, leading him from The origins of species (first published 1859) to The descent of man (first published 1871) (see, e.g., Landau 1984; Ruse 2008). However, it is important to note that he felt compelled to write his second major work also because he was deeply unhappy about the treatment of the ‘higher faculties’, which included, for example, tool use, complex language, reason, mathematics, and aesthetics, by his colleagues (Cain 2009, xvi). As mentioned, Wallace argued that these abilities could not have been produced by the interplay of variability and natural selection alone. He concluded in the end that the only adequate cause for explaining “the spiritual nature of man” has to be “the unseen universe of Spirit” itself (Wallace 1889) and that there seems to be evidence of a power that has guided the action of evolution in definite directions and for special ends (Cain 2009, xvii). In the decades after the publication of The Origins, this was not a minority opinion. In contrast, this was the most widely accepted view of the origins 258

Modern Humans and the Capacity for Modernity of fully modern or civilized human beings and was also shared by Charles Lyell, one of Darwin’s idols (Cain 2009, xvii). Charles Darwin was deeply frustrated by this development. In The descent of man and The expression of emotions in man and animals, he not only set out to demonstrate morphological similarities between humans and various animals but also ventured to show that aspects of behavior, expressions and, by extension, cognition are also shared. His aim was clearly to demonstrate that even the so-called higher faculties in humans differ only in degree and not in kind from those of animals. In the two volumes just mentioned, he set out to depict “animals as far more sophisticated (that is, endowed with increasingly human-like qualities) than most people usually acknowledged” at that time, and he further presented “human beings as carriers of features which were simply extensions of those found in animals” (Cain 2009, xix). Darwin also constructed a concentric and layered view of human beings where the ‘higher faculties’ surround more primitive layers of behavior and ultimately a core of basic features that are shared by all biological organisms.While all humans and animals are assumed to be united within the grand scheme of evolutionary ascent, the individual organisms (humans or animals) remain immutable representatives of their respective stage of development, defined by their hereditary endowment. As Darwin (2004, 689) himself expressed in the last sentence in The Decent of Man: “Man still bears in his bodily frame the indelible stamp of his lowly origin”. Despite Darwin’s contempt for slavery and his conviction that his theory of the ‘descent of man’ stressed the unity of mankind, these crucial elements of his work planted the seeds for social Darwinism and a racist view of human diversity. The events of the Second World War during which the concepts of race and racial competition played a pivotal role in justifying unimaginable atrocities had a profound impact on the notion of what it means to be a human being. Most well known is probably Ernst Mayr’s rejection of the multitude of genus and species names that had been put forward for fossil specimens until the 1950s (e.g., Sinanthropus pekinensis, Pithecanthropus erectus, Plesianthropus transvaalensis, Eoanthropus dawsoni) (e.g., Mayr 1963). Mayr argued that the genus name Homo had to be given to all hominids that possessed bipedal locomotion. He envisaged human evolution as a single, non-diversifying continuum of change that included only three time-successive species. He strictly opposed the possibility that two human species could have existed at the same time, because humans were no longer subjected to the biological mechanisms that produced the gradual acquisition of new traits as a consequence of new adaptations. In this he followed the argument of T. Dobzhansky, who had put forward the view that the presence of culture “removed all hominids from evolutionary processes that would otherwise lead to divergent speciation” (Schwartz & Tattersall 2010, 96). Mayr’s approach shifted the focus from the species to the genus level of taxonomy, but it especially gave the element of ‘culture’ a much more central and significant position, which nevertheless remained very much undefined. While Mayr equated the genus with bipedal locomotor behavior, he left the species Homo sapiens neither morphologically defined nor diagnosed. For an expert systematist, this seems an extraordinary step. Of particular importance for the discussion here is that Homo sapiens, the most recent of the three time-successive species, was now very much defined by the presence of cognitive or mental abilities alone – unlike Neanderthals and H. erectus, which were defined morphologically. Mayr followed a strategy that echoed the definitions of theorists before Darwin with their exclusive focus on ‘reasoning’ and ‘higher faculties’ (Ingold 2004). The identity of Homo sapiens was to be found in the inside of the organism, in capabilities that were located in the brain: “If groups of apparently disparate morphology are more or less universally agreed on to be members of the same species, it is scientifically ludicrous (and racist) to attach biological, systematic, and thus evolutionary meaning to the differences between them” (Schwartz & Tattersall 2010, 97). Effectively, human beings were defined by a consensus, by recognition (and assumption) of human capacity and potential. The consequence is that the notion of human cognitive potential is now beyond the grasp of natural evolutionary processes and in contradiction to classic Darwinian population thinking. It is no surprise that theorists in archaeology have been struggling to accommodate it in their models without inconsistencies and contradictions. 259

Martin Porr

Human Revolutions: Open and Disguised Together with a second edited book (Mellars 1990), the landmark volume The human revolution (Mellars and Stringer 1989a) was the result of a conference on the ‘origins and dispersal of modern humans’ held in Cambridge in 1987. The most important issue that was structuring The human revolution was the assertion that ‘biological’ and ‘behavioral’ changes were not correlated in the archaeological record. Despite the seemingly various controversial issues between authors in the volume, there seems to have been a widespread agreement that ‘modern human behavior’ originated with “behavioral changes which more or less correlate with the conventional ‘Middle-Upper Paleolithic transition’ in many parts of the Old World, and which are often seen as signaling the arrival of behavior closely comparable to modern hunter-gatherers, in all its essentials” (Mellars & Stringer 1989b, 1). While consequently ‘modernity’ is equated with the behavior and supposedly cognitive abilities of living hunter-gatherers, the character of the ‘essentials’ that are mentioned was not explicitly explored and defined. Mellars and Stringer (1989b, 12) identify as the “ultimate dilemma in studying the earlier phases of human cultural development” the “danger of applying completely circular arguments which equate the expression of culture with the capacity for culture – arguments which tend to assume that because a particular form of cultural and behavioral expression is not reflected in the particular period, the mental (or physical) capacities for it were lacking in the population involved”. In this case, now, the discussion is framed as a discourse of ‘presence/ absence’, which is applied to both material markers and the ‘capacity for culture’. Interestingly, the “well-documented grave offerings found in association with the human burials at both Qafzeh and Skhu¯l” (Mellars & Stringer 1989b, 8) are not further discussed as possible markers for the presence of language or ‘culture’. This example very much reflects what is regarded as the most important factor in the development and, hence, the recognition of ‘modern human behavior’: the structure of the (lithic) subsistence technology. It also reflects the bias that seems to be inherent in these discussions and which can be related to a definition of ‘modernity’ along Western modern values. Greater efficiency in extractive technologies is regarded as a marker for advanced cognitive abilities. The volume The human revolution was followed by a large number of further edited books as well as papers and other contributions on the same subject (see references in, for example, Henshilwood & Marean 2003; McBrearty & Brooks 2000; Shea 2011b). Only one should be mentioned here, because it deals specifically with “conceptual issues in modern human origins research” (Clark & Willermet 1997). In this case it is fascinating to see that the volume indeed presents a large number of controversial issues that are, however, related to either morphological aspects of human evolution or issues connected with the chronologies of Middle and Upper Paleolithic technologies. At the same time, the book contains the underlying consensus of a cognitive identity of modern humans that is independent of anatomical variation. Proponents of the ‘Out of Africa’ as well as the multiregional view of recent human evolution could both easily subscribe to the following statement: “The archaeological record does not always reflect people’s behavioral capacities. . . . human populations are modern when they behave in recognizably modern ways, no matter what they look like” (Wolpoff & Caspari 1997b, 44). It seems furthermore that both camps would even agree that ‘recognizably modern ways’ are largely equivalent with the occurrence of material expressions that can be found in the European Upper Paleolithic record (e.g., blade technology, burins, systematically constructed hearths). While the link between morphology and cognitive capacity or identity is rejected on a general level, the link between these two factors for single individual humans in the past is not questioned. The capacity for modernity is constructed as an innate feature of individual human beings. In the following discussions it was especially the idea of the ‘human revolution’ itself that was scrutinized as well as the connected idea of an Upper Paleolithic package that related to ‘modern human behavior’. In a highly influential article, McBrearty and Brooks (2000) argued that elements of modern behavior did not appear in a revolutionary fashion but were gradually assembled during the Middle and early Late Pleistocene in Africa. They consequently concluded that 260

Modern Humans and the Capacity for Modernity “the origin of our species is linked with the appearance of Middle Stone Age technology [in Africa] at 250–300 ka” (McBrearty & Brooks 2000, 453). The core of their argument was that the most important archaeological items that have been used to define a ‘human revolution’ between 40,000 and 50,000 years ago rather occurred in the African archaeological record “tens of thousands of years earlier” and “at sites that are widely separated in space and time” (McBrearty and Brooks 2000, 453). The authors argue that “the transition to fully modern human behavior was not the result of a biological or cultural revolution, but the fitful expansion of a shared body of knowledge, and the application of novel solutions on an ‘as needed’ basis” (McBrearty & Brooks 2000, 531). Ironically, while they seemingly reject the notion of a ‘human revolution’ in archaeological terms, they implicitly propose a biological and cognitive one, which created the “cognitive equipment” that allowed the subsequent building of the “complex content of human cultures”. This factor – the capacity for culture – leads them to extend the species designation Homo sapiens to include individuals that were previously termed Homo helmei (McBrearty & Brooks 2000, 529–532). The isolated occurrences of instances of so-called modern behaviors therefore become reflections of a common cognitive heritage that is equally shared by all humans in Africa from at least 250,000 years before today (see also McBrearty 2007). As outlined, this view seems to replicate Mayr’s and Dobzhansky’s version of the post–World War II definition of Homo sapiens and also re-establishes a link between morphology and cognitive capacity. In arguing that the changes that are visible in the African archaeological record are explicitly not related to biological evolutionary processes, they also implicitly follow Dobzhansky’s view that human history began with the origin of modern humanity, which is supposedly qualitatively different from the course of biological evolution. Interestingly, similar elements can be traced in Mellars’s (2006, 9384–9385) contributions, who specifically leaves it open if crucial developments were caused by either “a further genetic mutation involving cognitive capacities” or “some major shift in the adaptive and selective pressures to which human populations were subjected”. The underlying assumption remains that all features that might be used to identify the cognitive capacity for modern thought were universally valuable and advantageous and were consequently shared by all modern humans who originated in and eventually made the successful migration out of Africa. Consequently, with the origins of the former element at some unknown point in time, a development is set in motion that fits into a discourse of a progressive and cumulative cognitive unfolding of humanity that achieves increasing insight into and control over reality, a development that eventually culminates in modern Western thinking and science (e.g., Ingold 2000, 373–391; Landau 1984, 99–127). While most of these elements are only implicitly contained in the arguments outlined thus far, they are more explicitly elaborated in Renfrew’s discussion of the ‘Sapient Paradox’, which is also mentioned by Mellars (2006, 9385). It deals with the question why it took modern humans such a long time to develop complex behaviors for which they had already the relevant capacities (Renfrew 1996). The paradox is, of course, only a paradox when it is assumed that the dormant and genetically inherited capacities do indeed have universal value and are generally advantageous across different contexts and environments. The philosophical implications of this assumption cannot be properly discussed here (see, e.g., Landau 1984), but the contained element of arbitrariness can already be discerned in the observation that Renfrew sees the breakthrough to modernity in the previously mentioned sense, basically only with the beginning of writing – and not with the ‘Upper Palaeolithic revolution’ (see also Renfrew 2008). In contrast, Richard Klein has argued that the origin of behavioral modernity can be related to “a fortuitous mutation that promoted the fully modern brain” between 45,000 and 50,000 years ago (Klein 2008, 271). This “big bang” of human culture and consciousness also immediately enabled Upper Paleolithic material technologies as well as art and other items of symbolic expressions (e.g., burials, ornaments) (Klein & Edgar 2002; Porr 2010). The essence behind these developments was “the fully modern ability to invent and manipulate culture”. This ability, furthermore, translated into an “evolutionary advantage . . ., because it permitted its possessors to 261

Martin Porr extract far more energy from nature and to invest it into society [and] allowed the kind of rapidly spoken phonemic language that is inseparable from culture as we know it today [and] people to conceive and model complex natural and social circumstances entirely with their minds” (Klein & Edgar 2002, 24). Klein clearly stresses here the importance of culture as the extrasomatic means of adaptation and that ‘culture’ provides a superior strategy in terms of adaptive speed and flexibility. But his insistence on the fact that these means were indeed created fortuitously echoes the ideas of Wallace and others, who were certain that the modern mind could not have been produced by evolutionary forces alone but must have been a product of divine intervention. In that sense, Klein’s argument is strangely arbitrary in its application and ignorance of evolutionary theory and principles. It also needs to be stressed that in Klein’s framework the essence of cultural, symbolic and linguistic behaviors are all fixed in the genetic code of each individual. Currently, Klein’s position seems to be a minority opinion and most authors apparently do subscribe to a more gradual development of modern human behaviors (see, e.g., Henshilwood & Marean 2003; McBrearty & Brooks 2000). In a series of papers, Shea has recently argued for a replacement of the idea of ‘behavioral modernity’ or ‘modern behavior’ with the concept of ‘behavioral variability’ (Shea 2011a; 2011b; 2011c; 2011d). Shea (2011c, 131) makes the case that the spatial and temporal variability that is observable in lithic Middle Stone Age technology of Eastern Africa between circa 280,000 and 6,000 years ago does not represent “a steady accumulation of novel core technologies. . . . Instead one sees a persistent pattern of wide technological variability”. Overall, modern human behavior can be characterized not by a fixed collection of material expressions but rather by an ability to be particularly strategic and flexible in the employment of material culture and behaviors. The aim of all behavior of humans is supposedly an increase in the efficiency in the extraction of energy from different environments. Consequently, Shea suggests that the main tool to understand modern human behavioral variability should be ‘strategic modeling’ with a particular emphasis on ‘behavioral ecology’. The latter factors and overall framework have, of course, been central elements of the research into human cultural evolution as well as evolutionary biology for decades (Bettinger 1991; Binford 1962, 2001; Kelly 1995; Winterhalder and Smith 1992). The strange situation consequently arises that Shea argues on the one hand for an abandonment of behavioral modernity as an analytical concept and at the same time assumes that the ability for behavioral variability (as an innate characteristic) came into existence only with the origins of anatomically modern humans. This is the core of his argument of ‘Homo sapiens is as Homo sapiens was’ (Shea 2011b), and as I have tried to show, this conviction is widely shared among current authors. Therefore, as he does not extend the idea of behavioral variability to earlier populations (and sees it as the defining characteristic of recent humans that is fixed in their genetic makeup), the essentialism that he throws out of the front door immediately comes in through the back door.

Modern Humans and Adaptationist Explanations This last observation is a reflection of the fact that most recent approaches to modern human origins adopt a largely adaptationist position toward evolutionary causalities and respective explanations, which have been most explicitly put forward by proponents of evolutionary psychology (e.g., Barkow et al. 1996). It needs to be stressed that while these assumptions seem mostly unquestioned within Paleolithic archaeology and paleoanthropology, this is not the case in the wider field of evolutionary biology. Critiques of an adaptationist orientation in evolutionary biology were most forcefully expressed in a series of papers by Lewontin and Gould (esp. Gould 2002; Gould & Lewontin 1979; Pigliucci & Kaplan 2000). These have since then generated a veritable discussion about the role and power of adaptation, selection and other forces in evolutionary contexts as well as the respective epistemological and methodological consequences (Godfrey-Smith 1999; Lewens 2009; Resnik 1997). 262

Modern Humans and the Capacity for Modernity It is of course not possible to fully discuss these issues in relation to recent cognitive human evolution, but it should be noted that some of the fundamental assumptions on which present discussions are resting are in fact far from unproblematic. For example, this applies to the assumed centrality and priority of adaptive mechanisms as an explanation for human behavior as well as the structure, usefulness and applicability of optimization models in this context (Andrews et al. 2002; Godfrey-Smith 2000; Harris 2010; Orzack and Sober 2000; Sansom 2003). The latter are of central importance to current explanatory frameworks within Paleolithic archaeology and human evolution, often in the form of optimal foraging theory (Bettinger 1991; Binford 2001; Gamble 1986; Kelly 1995; Winterhalder & Smith 1992). However, the role, character and importance of adaptive mechanisms are rarely critically discussed in these contexts as well as the resulting methodological and epistemological considerations. Certain critiques that have been raised against different forms of adaptationism clearly apply to approaches to so-called modern human origins as well. Godfrey-Smith (2000) and Stegmann (2005) have criticized the way that adaptationist explanations in evolutionary biology often concentrate on adaptations in organisms, not because they provide the quantitatively most important factors but because they present the highest explanatory value within a Darwinian framework. Consequently, the choices that are made in this direction rather more reflect cultural choices based on their perceived value for the scientific world view than are based on the actual primacy of adaptive causes (Stegmann 2005, 289–290). In this context, the importance of optimality considerations and modeling has to be stressed as a central tool in the development and testing of any adaptationist scenario, both in human evolution studies and in evolutionary biology (Orzack and Sober 1994, 2000; Parker & Maynard Smith 1990). These are facing the same problematic issues in that they suffer from an inherent lack of context dependency and historicity, because they are based on the assumption that the processes that are underlying the models have universal applicability. The specifics, potentials and constraints of both the environmental and organismic configurations involved in each evolutionary process necessitate a constant critical adjustment of what optimality and adaptation might mean at a given point in time and if these concepts in fact can be productively applied. These tensions, however, are mostly absent in current views of modern human origins. It is this overall understanding of the dispersal of modern humans that leads to the view that the archaeological evidence from Australia (as well as East and Southeast Asia) is seen as impoverished and a “challenge to prehistorians [because of] the strikingly simple appearance of the earliest stone-tool industries” (Mellars & Stringer, 1989a, 11). It also leads to Gowlett’s remark of “fully modern man, sitting at the head of 40,000 years of occupation of modern man, making stone tools that could come out of the African or European Lower Paleolithic . . . it leaves us a little baffled” (Gowlett, quoted in Brumm and Moore 2005, 157). Both views are the product of the assumption that humans can be characterized by a predetermined cognitive potential and capacity that is defined by the standards of the European Upper Paleolithic record and, ultimately, the values of modern Western societies (e.g., enhanced efficiency in the extraction of energy from the environment and technological sophistication). As such, they do not help in understanding global human life ways and variability, but rather contribute to reducing them to either fully realized or failed reflections of assumed universal human characteristics or potentials.

Conclusion: Absolute or Relative Capacity The notion of an innate ‘capacity’ for modern human behavior is problematic at a variety of levels, in particular when it is used in the sense of a biological potential (with a genetic basis) that relates to a specific cognitive capacity and allows individuals to develop modern behaviors and to manipulate symbols. It seems to me that this notion has not been subjected to proper critical scrutiny, in relation to both its problematic continuing reference to the Western European Upper Paleolithic evidence and its value in the context of biological mechanisms in general. The 263

Martin Porr former issue especially is not simply a product of the research history of Paleolithic archaeology: the explicit and implicit adherence to notions of material complexity and increasing efficiency in the extraction of energy from the environment as measures of behavioral modernity are in fact products of the strongly adaptationist Darwinian explanatory frameworks that are employed to explain exactly those features. At the end product of this process stands the Western scientific world view that has a privileged access to the causal mechanisms of reality and is supposedly the product of the mechanisms of human evolution itself, which gradually rewarded this deeper understanding of causal relationships and their most efficient use and manipulation. The previously mentioned curious and ambiguous treatment of so-called symbolic expressions of human behavior in the Pleistocene archaeological record is only a reflection of this much more fundamental issue related to the relationship between Western science and cultural diversity. The concern with ‘modern human origins’ and the interpretation of global human variability are two sides of the same coin. Both have a long way to go to critically and productively engage with the issues that have been mentioned in this chapter. I would argue that approaches that recognize humans as socially constituted dynamic beings that are also reflective and cognitively active can overcome the problems that have been discussed. This perspective does not mean that genetic influences, adaptation and optimization processes can be ignored. Rather they have to be seen as one influence within dynamic sets of relationships. Human beings develop and grow through these relationships, which provide both potentials and constraints.What humans are and can do is not a reflection of internal essences of human nature but is a product of situated growth, reflections, interactions and negotiations.

264

References

Abbie, A. A. 1951. The Australian Aborigine. Oceania 22, 91–100. Abbie, A. A. 1969. The original Australians. Wellington: Reed. Abi-Rached, L. et al. 2011.The shaping of modern human immune systems by multiregional admixture with archaic humans. Science 334, 89–94. Abu-Amero, K. et al. 2007. Eurasian and African mitochondrial DNA influences in the Saudi Arabian population. BMC Evolutionary Biology 7, 32. Abu-Amero, K. et  al. 2008. Mitochondrial DNA structure in the Arabian Peninsula. BMC Evolutionary Biology 8, 45. Abu-Amero, K. et al. 2009. Saudi Arabian Y-chromosome diversity and its relationship with nearby regions. BMC Genetics 10, 59. Aczel, A. D. 2007. The Jesuit and the skull:Teilhard de Chardin, evolution, and the search for Peking Man. New York: Riverhead Books. Adcock, G. et  al. 2001. Mitochondrial DNA sequences in ancient Australians: implications for modern human origins. PNAS 98, 2, 537–542. Aiello, L. C. 2010. Five years of Homo floresiensis. American Journal of Physical Anthropology 142, 2, 167–79. Akerman, K. 1976. Fishing with stone traps on the Dampierland Peninsula, Western Australia. Mankind 10, 3, 182. Allchin, B., A. Goudie & K. Hedge 1978. The prehistory and palaeogeography of the Great Indian Desert. London: Academic Press. Allen, H. 1972. Where the crow flies backwards: man and land in the Darling Basin. Unpublished PhD thesis, Australian National University, Canberra. Allen, H. 1998. Reinterpreting the 1969–1972 Willandra Lakes archaeological surveys. Archaeology in Oceania 33, 207–220. Allen, J. 1989. Excavations at Bone Cave, South Central Tasmania, January–February 1989. Australian Archaeology 28, 105–106. Allen, J. 1994. Radiocarbon determinations, luminescence dating and Australian archaeology. Antiquity 68, 339–43. Allen, J. (ed.) 1996a. Report of the Southern Forests Archaeological Project: site description, stratigraphies and chronologies. Volume 1. Melbourne: School of Archaeology, La Trobe University. Allen, J. 1996b. Warreen Cave. In: J. Allen (ed.), Report of the Southern Forests Archaeological Project, Vol. 1: site descriptions, stratigraphies, and chronologies, Melbourne: School of Archaeology, La Trobe University, 135–168.

265

References Allen, J. 2000. From beach to beach: the development of maritime economies in prehistoric Melanesia. In: S. O’Connor & P.Veth (eds.), East of Wallace’s Line: studies of past and present maritime culture of the Indo-Pacific region. Rotterdam: A.A. Balkema, 139–176. Allen, J. 2010. The curious history of the Talgai skull. Bulletin of the History of Archaeology 20, 4–12. Allen, J. & S. Holdaway 1994. The contamination of Pleistocene radiocarbon determinations in Australia. Antiquity 69, 101–112. Allen, J. & C. Gosden 1996. Spheres of interaction and integration: modelling the culture history of the Bismarck Archipelago. In: J. Davidson et al. (eds.), Oceanic culture history: essays in honour of Roger Green. Dunedin North: New Zealand Journal of Archaeology Special Publications, 183–97. Allen, J. & 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. Allen, J. & J. F. O’Connell 2008. Getting from Sunda to Sahul. In: G. Clarke et al. (eds.), Islands of inquiry: colonization, seafaring and the archaeology of maritime landscapes. Canberra: ANU E Press, 31–46. Allen, J., C. Gosden & J. P. White 1989. Human Pleistocene adaptations in the tropical island Pacific: recent evidence from New Ireland, a Greater Australian outlier. Antiquity 63, 548–561. Allen, J. et al. 1988. Pleistocene dates for the human occupation of New Ireland, northern Melanesia. Nature 331, 707–709. Ambrose, S. H. 2002. Small things remembered: origins of early microlithic industries in sub-Saharan Africa. Archeological Papers of the American Anthropological Association 12, 9–30. Ambrose, S. 2010. Coevolution of composite-tool technology, constructive memory, and language. Implications for the evolution of modern human behavior. Current Anthropology 51, Supplement 1, 135–147. Ambrose, S. H. & K. G. Lorenz 1990. Social and ecological models for the Middle Stone Age in southern Africa. In: P. Mellars (ed.), The emergence of modern humans. Edinburgh: Edinburgh University Press, 3–33. Anderson, D. D. 1990. Lang Rongrien rockshelter: a Pleistocene, Early Holocene archaeological site from Krabi, Southwestern Thailand. Pennsylvania: University of Pennsylvania Museum of Archaeology. Anderson, D. 2005. The use of caves in Peninsular Thailand in the Late Pleistocene and Early and Middle Holocene. Asian Perspectives 44, 1, 137–153. Anderson, A. & S. O’Connor 2008. Indo-Pacific migration and colonization: introduction. Asian Perspectives 47, 1, 2–11. Anderson, P. et  al. 1993. Traces et fonction: les gestes retrouvés. Liège: Service de Préhistoire, Université de Liège. Andrefsky,W. 2008. Lithic technology: measures of production, use, and curation. Cambridge: Cambridge University Press. Andrews, J. E. et al. 1998. Do stable isotope data from calcrete record monsoonal climate variation in the Thar Desert of India? Quaternary Research 50, 240–251. Andrews, P. W., S. W. Gangestad & D. Matthews 2002. Adaptationism – How to carry out an exaptationist program. Behavioral and Brain Sciences 25, 489–553. Anell, B. 1960. Hunting and trapping methods in Australia and Oceania. Studia Ethnographica Upsaliensia XVIII. Lund: Håkan Ohlssons Boktryckeri. Anon. 1873. The caves of Mount Sobis. Sarawak Gazette 3, 68, 59–60. Ansyori, M. M. 2010. Fauna from the oldest occupation layer in Song Terus Cave, Eastern Java, Indonesia: biochronological significance of Terus Layer. Unpublished Master’s Thesis, Muséum national d’Histoire naturelle. Anton, D. 1984. Aspects of geomorphological evolution: paleosols and dunes in Saudi Arabia. In: A. R. Jado & J. G. Zötl (eds.), Quaternary period in Saudi Arabia. II. Sedimentological, hydrogeological, hydrochemical, geomorphological, and climatological investigations of Western Saudi Arabia.Vienna: Springer, 275–296. Antón, S. C. 2003. Natural history of Homo erectus. Yearbook of Physical Anthropology 46, 126–170. Aplin, K. P. 2006. Ten million years of rodent evolution in Australasia: phylogenetic evidence and a speculative historical biogeography. In: J. R Merrick et al. (eds.), Evolution and biogeography of Australasian vertebrates. Sydney: Auscipub Pty Ltd., 707–744. Appenzeller, T. 2012. Eastern odyssey. Nature 485, 24–26. Armitage, S. J. et al. 2011. The southern route ‘out of Africa’: evidence for an early expansion of modern humans into Arabia. Science 331, 453–456. Athreya, S. 2007. Was Homo heidelbergensis in South Asia? A test using the Narmada fossil from central India. In: M. D. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia: interdisciplinary studies in archaeology, biological anthropology, linguistics and genetics. Dordrecht: Springer, 137–170.

266

References Athreya, S. 2010. South Asia as a geographic crossroad: patterns and predictions of hominin morphology in Pleistocene India. In: C. J. Norton & D. R. Braun (eds.), Asian paleoanthropology. Dordrecht: Springer, 129–141. Atkinson, Q., R. D. Gray & A. J. Drummond 2008. mtDNA variation predicts population size in humans and reveals a major Southern Asian chapter in human prehistory. Molecular Biology and Evolution 25, 2, 468–474. Bader, O. N. & N. O. Bader 2000. Upper Palaeolithic site Sunghir. In: T. I. Alexeeva & N. O. Bader (eds.), Homo Sungirensis. Upper Palaeolithic man. Ecological and evolutionary aspects of the investigation. Moscow: Scientific World, 21–29. Bacon, A.-M. et al. 2006. New palaeontological assemblage, sedimentological and chronological data from the Pleistocene Ma U’Oi cave (northern Vietnam). Palaeogeography, Palaeoclimatology, Palaeoecology 230, 280–298. Bacon, A.-M. et al. 2008. The Late Pleistocene Duoi U’Oi cave in northern Vietnam: palaeontology, sedimentology, taphonomy and palaeoenvironments. Quaternary Science Reviews 27, 1627–1654. Bahn, P. G. 1977. Seasonal migration in S. W. France during the late glacial period. Journal of Archaeological Science 4, 245–257. Bahn, P. G. 1982. Inter-site and inter-regional links during the Upper Palaeolithic: the Pyrenean evidence. Oxford Journal of Archaeology 1, 3, 247–268. Bailey, R. C. et al. 1989. Hunting and gathering in tropical rain forest: is it possible? American Anthropologist 91, 1, 59–82. Balme, J. 1980. An analysis of charred bone from Devil’s Lair, Western Australia. Archaeology & Physical Anthropology in Oceania 15, 81–85. Balme, J. 1995. 30,000 years of fishery in western New South Wales. Archaeology in Oceania 30, 1–21. Balme, J. 2000. Excavations revealing 40,000 years of occupation at Mimbi Caves, south central Kimberley, Western Australia. Australian Archaeology 51, 1–5. Balme, J. 2013. Of boats and string: the maritime colonisation of Australia. Quaternary International 285, 68–75. Balme, J. & K. Morse 2006. Shell beads and social behaviour in Pleistocene Australia. Antiquity 80, 799–811. Balme, J., D. Merrilees & J. Porter 1978. Late quaternary mammal remains, spanning about 30,000 years, from excavations in Devil’s Lair, Western Australia. Journal of the Royal Society of Western Australia 61, 33–65. Balme, J. et al. 2009. Symbolic behaviour and the peopling of the southern arc route to Australia. Quaternary International 202, 59–68. Balter, M. 2010. Of two minds about Toba’s impact. Science 327, 5970, 1187–1188. Ba Maw 1995. The first discovery of the early man’s fossilized maxillary bone fragment in Myanmar. East Asian Tertiary Quaternary Newsletter 16, 72–79. Barham, L. S. 2002. Systematic pigment use in the Middle Pleistocene of South-Central Africa. Current Anthropology 43, 1, 181–190. Barham, L. 2007. Modern is as modern does? Technological trends and thresholds in the south-central Africa record. In: P. Mellars et al. (eds.), Rethinking the human revolution. Cambridge: University Press, 165–176. Barham, L. & P. Mitchell 2008. The first Africans. Cambridge World Archaeology. Cambridge: Cambridge University Press. Barik, S. S. et al. 2008. Detailed mtDNA genotypes permit a reassessment of the settlement and population structure of the Andaman Islands. American Journal of Physical Anthropology 136, 1, 19–27. Barker, G. 2005. The archaeology of foraging and farming at Niah Cave, Sarawak. Asian Perspectives 44, 1, 90–106. Barker, G. (ed.) 2013. Rainforest foraging and farming in Southeast Asia: the archaeology of the Niah Caves, Sarawak. Cambridge: McDonald Institute for Archaeological Research. Barker, G. et al. 2002. Prehistoric foragers and farmers in Southeast Asia: renewed investigations at Niah Cave, Sarawak. Proceedings of the Prehistoric Society 68, 147–64. Barker, G. et al. 2007.The ‘human revolution’ in lowland tropical Southeast Asia: the antiquity and behavior of anatomically modern humans at Niah Cave (Sarawak, Borneo). Journal of Human Evolution 52, 243–261. Barker, G. et al. 2011. Foraging-farming transitions at the Niah Caves, Sarawak, Borneo. Antiquity 85, 328, 492–509. Barkow, J. H., L. Cosmides & J. Tooby 1996. The adapted mind: evolutionary psychology and the generation of culture. New York: Oxford University Press. Barnabas, S.,Y. Shouche & C. G Suresh 2006. High-Resolution mtDNA Studies of the Indian population: implications for Palaeolithic settlement of the Indian Subcontinent. Annals of Human Genetics 70, 1, 42–58.

267

References Barnard, P. J. et  al. 2007. Differentiation in cognitive and emotional meanings: an evolutionary analysis. Cognition & Emotion 21, 6, 1155–1183. Barnard Davis, J. 1874. On the osteology and peculiarities of the Tasmanians, a race of man recently become extinct. Haarlem: De Erven Loosjes. Barton, H. 2005. The case for rainforest foragers: the starch record at Niah Cave, Sarawak. Asian Perspectives 44, 1, 56–72. Barton, C. M., G. A. Clark & A. E. Cohen 1994. Art as information: explaining Upper Palaeolithic art in western Europe. World Archaeology 26, 2, 185–207. Barton, H.A. et al. 2009. Composite hunting technologies from the Terminal Pleistocene and Early Holocene, Niah Cave, Borneo. Journal of Archaeological Science 36, 1708–1714. Bartstra, G.-J. 1983. Some remarks upon fossil man from Java, his age, and his tools. Bydragen to de Taal-Landand Volkenkunde 139, 421–434. Bartstra, G.-J. 1984. Dating the Pacitanian: some thoughts. Courier Forschungsinstitut Senckenberg 69, 253–258. Bartstra, G.-J. 1985. Sangiran, the stone implements of Ngebung, and the Palaeolithic of Java. Modern Quaternary Research in SE Asia 9, 99–113. Bar-Yosef, O. 2002. The Upper Paleolithic revolution. Annual Review of Anthropology 31, 363–393. Bar-Yosef, O. 2007. The dispersal of modern humans in Eurasia: a cultural interpretation. In: P. Mellars et al. (eds.), Rethinking the human revolution. Cambridge: McDonald Institute for Archaeological Research, 207–218. Bar-Yosef, O. & J. Callander 1999. The woman from Tabun: Garrod’s doubts in historical perspective. Journal of Human Evolution 37, 879–885. Bar-Yosef, O. & S. L. Kuhn 1999. The big deal about blades. Laminar technologies and human evolution. American Anthropologist 101, 2, 322–338. Bar-Yosef, O. et al. 1996. The dating of the Upper Palaeolithic layers in Kebara cave, Mount Carmel. Journal of Archaeological Science 23, 297–306. Basilia, P. A. 2011. Morphological and technological analysis of the cut shell beads from Ille Cave, El Nido, Palawan. Unpublished master’s thesis, University of the Philippines, Archaeological Studies Program. Bautista, A. & J. de Vos 2001. Archaeological exploration and excavation at Solana, Cagayan. Unpublished manuscript. National Museum Manila. Beaton, J. M. 1983. Does intensification account for changes in the Australia Holocene archaeological record? Archaeology in Oceania 18, 94–97. Beaton, J. M. & G. L. Walsh 1977. Che-ka-ra. Mankind 11, 46–48. Bednarik, R. G. 1992. Palaeoart and archaeological myths. Cambridge Archaeological Journal 2, 1, 27–57. Bednarik, R. G. 1998. Microscopic analysis of ‘engraved plaques’ and other objects from Devil’s Lair. Journal of the Royal Society of Western Australia 81, 165–175. Bellwood, P. S. 1979. Man’s conquest of the Pacific: the prehistory of Southeast Asia and Oceania. Oxford: Oxford University Press. Bellwood, P. S. 1997. Prehistory of the Indo-Malaysian Archipelago. 2nd ed.. Honolulu: University of Hawaii Press. Bellwood, P. S. 2005. First farmers.The origins of agricultural societies. Oxford: Blackwell. Bellwood, P. S. 2010. Comment on Donahue and Denham, farming and language in Island Southeast Asia: reframing Austronesian History. Current Anthropology 51, 2, 240–241. Bellwood, P. et al. 1998. 35,000 years of prehistory in the northern Moluccas. In: G.-J. Bartstra (ed.), Bird’s Head approaches: Irian Jaya studies, a programme for interdisciplinary research. Rotterdam: A. A. Balkema, 15, 233–275. Bennett, E. L., A. J. Nyaoi & J. Sompud 1997. Hornbills buceros spp. and culture in northern Borneo: can they continue to co-exist? Biological Conservation 82, 1, 41–46. Bergh, G. D. van den 1999.The Late Neogene elephantoid-bearing faunas of Indonesia and their palaeozoogeographic implications; a study of the terrestrial faunal succession of Sulawesi, Flores and Java, including evidence for early hominid dispersal east of Wallace’s Line. Scripta Geologica 117, 1–419. Bergh, G. D. van den, J. de Vos & P.Y. Sondaar 2001.The late Quaternary palaeogeography of mammal evolution in the Indonesian Archipelago. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 385–408. Bergh, G. D. van den et al. 2009.The Liang Bua faunal remains: a 95-kyr sequence from Flores, East Indonesia. Journal of Human Evolution 57, 5, 527–537. Bergman, C. A. & C. B. Stringer 1989. Fifty years after: Egbert, an early Upper Palaeolithic juvenile from KsarAkil, Lebanon. Paléorient 15, 99–112.

268

References Berry, J. A. & A. W. D. Robertson 1909. Preliminary communication on fifty-three Tasmanian crania, fortytwo of which are now recorded for the first time. Proceedings of the Royal Society of Victoria 22, 47–58. Berry, J. A. & A. W. D. Robertson 1910. The place in nature of the Tasmanian Aboriginal as deduced from a study of his calvarium. Pt. I. His relations to the anthropoid apes, Pithecanthropus, Homo primigenius, Homo fossilis, and Homo sapiens. Proceedings of the Royal Society of Edinburgh 31, 41–69. Berry, J. A. & A. W. D. Robertson 1914. The place in nature of the Tasmanian Aboriginal as deduced from a study of his calvarium. Pt. II. His relation to the Australian Aboriginal. Proceedings of the Royal Society of Edinburgh 34, 144–189. Berry, J. A., W. D. Robertson & K. S. Cross 1910. A biometrical study of the relative degree of purity of race of the Tasmanian, Australian and Papuan. Proceedings of the Royal Society of Edinburgh 31, 17–40. Bethge, P. et al. 2009. Activity patterns and sharing of time and space of platypuses, Ornithorhynchus anatinus, in a subalpine Tasmanian lake. Journal of Mammalogy 90, 6, 1350–1356. Bettinger, R. L. 1991. Hunter-gatherers: archaeological and evolutionary theory. New York: Plenum Press. Bettinger, R. L. & J. Eerkens 1999. Point typologies, cultural transmission, and the spread of bow-and-arrow technology in the Prehistoric Great Basin. American Antiquity 64, 231–242. Beyer, H. O. 1947. Outline review of Philippine archaeology by islands and province. Philippine Journal of Science 77, 3–4. Beyin, A. 2006. The Bab al Mandab vs the Nile-Levant: an appraisal of the two dispersal routes for early modern humans out of Africa. African Archaeological Review 23, 5–30. Beyries, S. (ed.) 1988. Industries lithiques: tracéologie et technologie. BAR International Series 411. Oxford: British Archaeological Reports. Bhopal, R. 2007. The beautiful skull and Blumenbach’s errors. British Medical Journal 335, 1308–1309. Binford, L. R. 1962. Archaeology as anthropology. American Antiquity 28, 217–225. Binford, L. 1980.Willow smoke and dogs’ tails: hunter-gatherer settlement and archaeological site formation. American Antiquity 45, 4–20. Binford, L. R. 2001. Constructing frames of reference: an analytical method for archaeological theory building using hunter-gatherer and environmental data sets. Berkeley: University of California Press. Binford, L. R. & J. F. O’Connell 1984. An Alyawara day: the stone quarry. Journal of Anthropological Research 40, 406–32. Bird, M. I. et al. 2002. Radiocarbon analysis of the early archaeological site of Nauwalabila I, Arnhem Land, Australia: implications for sample suitability and stratigraphic integrity. Quaternary Science Reviews 21, 1061–1075. Bird, M. I., G. Hope & D. Taylor 2004. Populating PEP II: the dispersal of humans and agriculture through Austral-Asia and Oceania. Quaternary International 118–119, 145–163. Bird, M., D. Taylor & C. Hunt 2005. Palaeoenvironments of insular Southeast Asia during the last glacial period: a savanna corridor in Sundaland? Quaternary Science Reviews 24, 2228–2242. Bird, M. et  al. 2007. A long record of environmental change from bat guano deposits in Makangit Cave, Palawan, Philippines. Earth and Environmental Science Transactions of the Royal Society of Edinburgh 98, 59–69. Birdsell, J. B. 1949.The racial origin of the extinct Tasmanians. Records of the Queen Victoria Museum, Launceston 2, 105–122. Birdsell, J. B. 1967. Preliminary data on the trihybrid origin of the Australian Aborigines. Archaeology and Physical Anthropology in Oceania 2, 100–155. Birdsell, J. B. 1977. The recalibration of a paradigm for the first peopling of greater Australia. In: J. Allen, J. Golson & R. Jones (eds.), Sunda and Sahul: prehistoric studies in Southeast Asia, Melanesia and Australia. London: Academic Press, 113–167. Black, D. 1925. Asia and the dispersal of primates. Bulletin of the Geological Society of China 4, 133–183. Blake, B. J. 1988. Redefining Pama-Nyungan: towards a prehistory of Australian languages. Aboriginal Linguistics 1,1–90. Blasco, R. 2008. Human consumption of tortoises at Level IV of Bolomor Cave (Valencia, Spain). Journal of Archaeological Science 35, 2839–2848. Blasco, R. & J. Férnandez Peris 2009. Middle Pleistocene bird consumption at Level XI of Bolomor Cave (Valencia, Spain). Journal of Archaeological Sciences 36, 2213–2223. Bliege Bird, R. & D. W. Bird 2008. Why women hunt: risk and contemporary foraging in a Western Desert Aboriginal community. Current Anthropology 49, 4, 655–693. Boëda, E. 1995. Levallois: a volumetric construction, methods, a technique. In: H. Dibble & O. Bar-Yosef (eds.), The definition and interpretation of Levallois technology. Madison, Wisc.: Prehistory Press, 41–69.

269

References Boëda, E. et al. 1999. A Levallois point embedded in the vertebra of a wild ass (Equus africanus): hafting, projectiles and Mousterian hunting weapons. Antiquity 73, 280, 394–402. Boivin, N. et al. 2013. Human dispersal across diverse environments of Asia during the Upper Pleistocene. Quaternary International 300, 32–47. Bondoc, N. H. 1979. A re-investigation of the Espinosa archaeological sites: Cagayan and Kalinga-Apayao. Manila: National Museum of the Philippines. Boswell, P. G. H. 1935. Human remains from Kanam and Kanjera, Kenya Colony. Nature 135, 371. Bourke, R. M. & B. Allen 2009.Village food production systems. In: R. M. Bourke & T. Harwood (eds.), Food and agriculture in New Guinea. Canberra: ANU E Press, 193–259. Bouzouggar, A. et al. 2007. 82,000-year-old shell beads from North Africa and implications for the origins of modern human behavior. PNAS 104, 9964–9969. Bowdler, S. 1977.The coastal colonisation of Australia. In: J. Allen et al. (eds.), Sunda and Sahul: prehistoric studies in Southeast Asia, Melanesia and Australia. London: Academic Press, 205–246. Bowdler, S. 1984. Hunter Hill, Hunter Island. Canberra: ANU. Bowdler, S. 1990. Peopling Australasia: the ‘coastal colonisation’ hypothesis reconsidered. In: P. Mellars (ed.), The emergence of modern humans: an archaeological perspective. Edinburgh: Edinburgh University Press, 327–343. Bowdler, S. 1992. Homo sapiens in Southeast Asia and the Antipodes: archaeological versus biological interpretations. In:T. Akazawa et al. (eds.), The evolution and dispersal of modern humans in Asia.Tokyo: HokusenSha, 560–589. Bowdler, S. 1993.Views of the past in Australian prehistory. In: M. Spriggs et al. (eds.), A community of culture: the people and prehistory of the Pacific. Canberra: ANU, 123–138. Bowern, C. 2012. in http://pamanyungan.sites.yale.edu/master-list-australianlanguages-v12; accessed 13 July 2012. Bowler, J. M. 1998. Willandra Lakes revisited: environmental framework for human occupation. Archaeology in Oceania 33, 3, 120–155. Bowler, J. M. & A. G. Thorne 1976. Human remains from Lake Mungo: discovery and excavation of Lake Mungo III. In: R. L. Kirk & A. G.Thorne (eds.), The origins of the Australians. Canberra: Australian Institute of Aboriginal Studies, 127–138. Bowler, J. M. & D. M. Price 1998. Luminescence dates and stratigraphic analyses at Lake Mungo: review and new perspectives. Archaeology in Oceania 33, 156–168. Bowler, J. M. & J. W. Magee 2000. Redating Australia’s oldest human remains: a sceptic’s view. Journal of Human Evolution 38, 719–726. Bowler, J. M. et al. 1970. Pleistocene human remains from Australia: a living site and cremation from Lake Mungo, New South Wales. World Archaeology 2, 39–60. Bowler, J. M. et al. 2003. New ages for human occupation and climatic change at Lake Mungo, Australia. Nature 421, 837–840. Bowman, D. M. J. S & S. G. Haberle 2010. Paradise burnt: how colonizing humans transform landscapes with fire. PNAS 107, 50, 21234–21235. Boyd, R. & P. Richerson 1985. Culture and the evolutionary process. Chicago: Chicago University Press. Brady, D. 2005. Australian Aboriginal body scarification and drainage basins. Armidale: University of New England. Brain, C. K. 1981. The hunters or the hunted: an introduction to African cave taphonomy. Chicago: Chicago University Press. Brandt, S. A. & N. Carder 1987. Pastoral rock art in the Horn of Africa: making sense of udder chaos. World Archaeology 19, 194–213. Brantingham, P. J., S. L. Kuhn & K.W. Kerry 2004. The Early Upper Palaeolithic beyond Western Europe. London: University of California Press. Bräuer, G., M. Collard & C. Stringer 2004.On the reliability of recent tests of the Out of Africa hypothesis for modern human origins. The Anatomical Record Part A 279A, 701–707. Breuil, H. & L. Kozlowski 1931. Étude de stratigraphie paléolithique dans le nord de la France, la Bélgique et 1’Angleterre – la vallée de la Somme. L’Anthropologie 41, 449–488. Bricker, H. M. 2007. Hallam Leonard Movius Jr. 1907–1987: a biographical memoir. Washington, D. C.: National Academy of Sciences. Bronk Ramsey, C. 2009. Bayesian analysis of radiocarbon dates. Radiocarbon 51, 337–360.

270

References Brooks, A. S. et al. 2006. Projectile technologies of the African MSA: implications for modern human origins. In: E. Hovers and S. L. Kuhn (eds.), Evolution and stability in the Middle Palaeolithic and Middle Stone Age. New York: Springer, 233–256. Brosius, J. P. 1991. Foraging in tropical rain forests: the case of the Penan of Sarawak, East Malaysia (Borneo). Human Ecology 19, 2, 123–150. Brothwell, D. R. 1960. Upper Pleistocene human skull from Niah caves, Sarawak. Sarawak Museums Journal (new series) 15–16, 323–349. Brown, P. 1981. Artificial cranial deformation: a component in the variation in Pleistocene Australian Aboriginal crania. Archaeology in Oceania 16, 156–167. Brown, P. 1987a. Pleistocene homogeneity and Holocene size reduction: the Australian human skeletal evidence. Archaeology in Oceania 22, 41–71. Brown, S. 1987b. Toward a prehistory of the Hamersley Plateau, Northwest Australia. Canberra: ANU. Brown, P. 1989. Coobool Creek: a morphological and metrical analysis of the crania, mandibles and dentitions of a prehistoric Australian human population. Terra Australis 13. Canberra: Department of Prehistory, Research School of Pacific Studies, Australian National University. Brown, P. 1992. Recent human evolution in East Asia and Australasia. Philosophical Transactions of the Royal Society of London B 337, 235–242. Brown, S. 1993. Mannalargenna Cave: a Pleistocene site in Bass Strait. In: M. Smith et al. (eds.), Sahul in review. Canberra: ANU, 258–271. Brown, P. 1994. A flawed vision: sex and robusticity on King Island. Australian Archaeology, 38, 1–7. Brown, P. 1997. Australian palaeoanthropology. In: F. Spencer (ed.), History of physical anthropology: an encyclopaedia. New York: Garland Publishing, 138–145. Brown, P. 1999. The first modern East Asians? Another look at Upper Cave 101, Liujiang and Minatogawa 1. In: K. Omoto (ed.), Interdisciplinary perspectives on the origins of the Japanese. Kyoto: International Research Center for Japanese Studies, 105–130. Brown, P. 2000. Australian Pleistocene variation and the sex of Lake Mungo 3. Journal of Human Evolution 38, 743–749. Brown, P. 2010. Nacurrie 1: mark of ancient Java, or a caring mother’s hands, in terminal Pleistocene Australia? Journal of Human Evolution 59, 168–187. Brown, P. et al. 2004. A new small bodied hominin from the Late Pleistocene of Flores, Indonesia. Nature 431, 1055–1061. Brown, K. S. et al. 2009. Fire as an engineering tool of early modern humans. Science 325, 859–862. Brumm, A. & M. W. Moore 2005. Symbolic revolutions and the Australian archaeological record. Cambridge Archaeological Journal 15, 157–175. Brumm, A. et al. 2006. Early stone technology on Flores and its implications for Homo floresiensis. Nature 441, 624–28. Brumm, A. et al. 2010. Evidence for hominins on Flores, Indonesia, by more than one million years ago. Nature 464, 748–752. Büchner, L. W. G. 1912a. An investigation of fifty-two Tasmanian crania by Klaatsch’s craniotrigonometrical methods. Proceedings of the Royal Society of Victoria 25, 122–134. Büchner, L. W. G. 1912b. A study of the prognathism of the Australian Aboriginal. Proceedings of the Royal Society of Victoria 25, 135–150. Büchner, L. W. G. 1914. A study of the curvature of the Tasmanian Aboriginal crania. Proceedings of the Royal Society of Edinburgh 34, 128–143. Bulbeck, D. 1982. A re-evaluation of possible evolutionary processes in Southeast Asia since the late Pleistocene. Bulletin of the Indo-Pacific Prehistory Association 3, 1–21. Bulbeck, D. 2007. Where river meets sea. A parsimonious model for Homo sapiens colonization of the Indian Ocean Rim and Sahul. Current Anthropology 48, 2, 315–321. Büller, J. 1988. Handling, hafting and ochre stains. In: S. Beyries (ed.), Industries lithiques: traceologie et technologie. Oxford: British Archaeological Reports, 5–46. Burke, A. 2004. The ecology of Neanderthals: preface. International Journal of Osteoarchaeology 14, 155–161. Busk, G. 1879–1880. Note on the collection of bones from caves in Borneo, referred to in Mr. Everett’s report on the ‘Exploration of the Bornean Caves in 1878–9’. Proceedings of the Royal Society of London 30, 319–321.

271

References Cabrera, V. et al. 2009. The Arabian peninsula: gate for human migrations out of Africa or cul-de-sac? A mitochondrial DNA phylogeographic perspective. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia. Dordrecht: Springer, 79–87. Cadenas, A. M. et  al. 2008. Y-chromosome diversity characterizes the Gulf of Oman. European Journal of Human Genetics 16, 374–386. Cahen, D., L. H. Keeley & F. van Noten 1979. Stone tools, toolkits and human behavior in prehistory. Current Anthropology 20, 4, 661–682. Cain, J. 2009. Introduction. In J. Cain and S. Messenger (eds.), Charles Darwin: the expression of the emotions in man and animals. London: Penguin Books, xi–xxxiv. Calaby, J. 1983. Red-necked wallaby. In R. Strahan (ed.), The Australian Museum complete book of Australian mammals. Sydney: Angus and Robertson, 239–241. Cameron, D., R. Patnaik & A. Sahni 2004.The phylogenetic significance of the Middle Pleistocene Narmada cranium from Central India. International Journal of Osteoarchaeology 14, 419–447. Campbell, J. B. 2000. The Chillagoe and Laura Laser-AMS dating project. In: G. K. Ward & C. Tuniz (eds.), Advances in dating Australian rock-markings: papers from the First Australian Rock-Picture Dating Workshop. Melbourne: Australian Rock Art Research Association, 80–83. Campbell, J. B. & M. Mardaga-Campbell 1993. From Micro- to Nano-stratigraphy. Linking vertical and horizontal dating of archaeological deposits with the direct dating of rock art at ‘The Walkunders’, Chillagoe (north Queensland, Australia). In: J. Steinbring et al. (eds.), Time and space: dating and spatial considerations in rock art research. Melbourne: Australian Rock Art Research Association, 57–63. Campbell, J. et al. 1996. Dating of rock surface accretions with Aboriginal paintings and engravings in North Queensland. In: S. Ulm et al. (eds.), Australian archaeology ’95: proceedings of the 1995 Australian Archaeological Association Annual Conference. St Lucia: Anthropology Museum, University of Queensland, 231–239. Campo, E. van 1986. Monsoon fluctuations in two 20,000-Yr B.P. oxygen-isotope/pollen records off Southwest India. Quaternary Research 26, 376–388. Cane, S. 2001.The Great Flood: eustatic change and cultural change in Australia during the Late Pleistocene and Holocene. In: A. Anderson et al. (eds.), Histories of old ages: essays in honour of Rhys Jones. Canberra: Pandanus Books, ANU, 141–165. Cann, R. L. 2001. Genetic clues to dispersal in human populations: retracing the past from the present. Science 291, 1742–1748. Cann, R. L., M. Stoneking & A. C. Wilson 1987. Mitochondrial DNA and human evolution. Nature 325, 31–36. Carlos, J. 2010. AMS radiocarbon dating of Canarium hirsutum nuts from Ille. Test Pit 17, 8. Castel, J.-C. et al. 2002. Solutrean animal resource exploitation at Combe Saunière (Dordogne, France). In: M. Maltby (ed.), Integrating zooarchaeology. Oxford: Oxbow, 138–152. Caton-Thompson, G. & F. W. Gardner 1934. The Desert Faiyum. London: Royal Anthropological Institute. Cˇerný, V. et al. 2008. Regional differences in the distribution of sub-Saharan, West Eurasian and south Asian mtDNA lineages in Yemen. American Journal of Physical Anthropology 136, 128–137. Cˇerný, V   . et  al. 2011. Internal diversification of Mitochondrial Haplogroup R0a reveals post-Last Glacial Maximum demographic expansions in south Arabia. Molecular Biology and Evolution 28, 71–78. Chalmers, A. F. 1976. What is this thing called science? An assessment of the nature and status of science and its methods. St Lucia: University of Queensland Press. Chaloupka, G. 1993. Journey in time: the world’s longest continuing art tradition. Sydney: Reed. Chandrasekar, A. et al. 2009. Updating phylogeny of mitochondrial DNA macrohaplogroup M in India: dispersal of Modern Human in South Asian Corridor. PLoS ONE 4, 10, e7447. Chappell, J. 1974. Geology of coral terraces, Huon Peninsula, New Guinea: a study of Quaternary tectonic movements and sea-level changes. Geological Society of America Bulletin 85, 553–570. Chappell, J. 1993. Late Pleistocene coasts and human migrations in the Austral region. In: M. Spriggs et al. (eds.), A community of culture: the people and prehistory of the Pacific. Canberra: ANU, 43–48. Chappell, J. 2002. Sea level changes forced ice breakouts in the last glacial cycle: new results from coral terraces. Quaternary Science Reviews 21, 1229–1240. Chappell, J., J. Head & J. Magee 1996. Beyond the radiocarbon limit in Australian archaeology and Quaternary research. Antiquity 70, 543–552. Chappell, J. et al. 1996. Reconciliation of late Quaternary sea levels derived from coral terraces at Huon Peninsula with deep sea oxygen isotope records. Earth and Planetary Science Letters 141, 227–236. Chardin, T. de. 1941. Early man in China. Institut de Géo-Biologie, Pékin, 1–100.

272

References Charoenwongsa, P. 1988. The current status of prehistoric research in Thailand. In: P. Charoenwongsa & B. Bronson (eds.), Prehistoric studies: the Stone and Metal Ages in Thailand. Bangkok: Thai Antiquity Working Group, 17–42. Chase, P. G. & H. L. Dibble 1987. Middle Paleolithic symbolism: A review of current evidence and interpretations. Journal of Anthropological Archaeology 6, 263–296. Chauhan, P. 2009. The Lower Paleolithic of the Indian subcontinent. Evolutionary Anthropology 18, 2, 62–78. Chazine, J.-M. 2005. Rock art, burials, and habitations: caves in East Kalimantan. Asian Perspectives 44, 1, 219–230. Ciochon, R. 2009. The mystery ape of Pleistocene Asia. Nature 459, 910–911. Ciochon, R. et al. 1996. Dated co-occurrence of Homo erectus and Gigantopithecus from Tham Khuyen Cave, Vietnam. PNAS 93, 7, 3016–3020. Clark, J. D. 1958. The natural fracture of pebbles from the Batoka Gorge, Northern Rhodesia. Proceedings of the Prehistoric Society 24, 64–77. Clark, J. G. D. 1968. Australian stone age. In: K. Jazdzewski (ed.), Liber Iosepho Kostrzewski octogenario a veneratoribus dicatus. Warsaw: Ossolineum, Polish Academy of Sciences, 17–28. Clark, J. G. D. 1969. World prehistory. Cambridge: Cambridge University Press. Clark, J. D. 1988. The Middle Stone Age of East Africa and the beginnings of regional identity. Journal of World Prehistory 2, 235–305. Clark, G. A. & C. M. Willermet (eds.) 1997. Conceptual issues in modern human origins research. New York: Aldine de Gruyter. Clark, J. D. et al. 1984. A Middle Stone Age occupation site at Porc Epic Cave, Dire Dawa (East-Central Ethiopia). African Archaeological Review 2, 37–71. Clark, J. D. et al. 2003. Stratigraphic, chronological and behavioural contexts of Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423, 747–751. Clarke, S. J. et al. 2007. The geochronological potential of isoleucine epimerisation in cassowary and megapode eggshells from archaeological sites. Journal of Archaeological Science 34, 1051–1063. Clark, P. U. et al. 2009. The last glacial maximum. Science 325, 710–714. Clarkson, C. 2010. Regional diversity within the core technology of the Howiesons Poort Techno-Complex. In: S. Lycett & P. Chauhan (eds.), New perspectives on old stones: analytical approaches to Palaeolithic technologies. New York: Springer, 43–60. Clarkson, C. & S. O’Connor 2006. An introduction to stone artifact analysis. In: J. Balme and A. Paterson (eds.), Archaeology in practice: a student guide to archaeological analyses. Malden, MA: Blackwell Publishing, 159–206. Clarkson, C., S. Jones & C. Harris 2012. Continuity and change in the lithic industries of the Jurreru Valley, India, before and after the Toba eruption. Quaternary International 258, 165–179. Clarkson, C. et al. 2009.The oldest and longest enduring microlithic sequence in India: 35 000 years of modern human occupation and change at the Jwalapuram Locality 9 rockshelter. Antiquity 83, 320, 326–348. Clarkson, C. et al. submitted. On the stoney trail of modern humans: new phylogenetic evidence for the colonization of Eurasia and Oceania. Clason, A. T. 1976. A preliminary note about the animal remains from Leang I Cave, South Sulawesi, Indonesia. Modern Quaternary Research in Southeast Asia 2, 53–67. CLIMANZ 1983. Proceedings of the First Climanz Conference: a symposium of results and discussion concerned with Late Quaternary climatic history of Australia, New Zealand and surrounding seas. Canberra: ANU. Clottes, J. 2008. Cave art. London: Phaidon Press. Colani, M. 1927. L’âge de la pierre dans la province de Hoa Binh,Tonkin. Hanoi: Imprimé d’Extrême-Orient. Colbert, E. H. 1943. Pleistocene vertebrates collected in Burma by the American Southeast Asiatic expedition for early man. Transactions of the American Philosophical Society 32, 395–428. Cole, F. 2012. Communities of the dead. Practice as an indicator of group identity in the Neolithic and Metal Age Burial Caves at Niah North Borneo. Unpublished PhD, University of Cambridge, Cambridge. Cole, N. & A. Watchman 2005. AMS dating of rock art in the Laura region, Cape York Peninsula, Australia – protocols and results of recent research. Antiquity 79, 661–678. Cole, N., A. Watchman & M. J. Morwood 1995. Chronology of Laura rock art. In: M. J. Morwood & D. R. Hobbs (eds.), Quinkan prehistory: the archaeology of Aboriginal art in S. E. Cape York Peninsula, Australia. St. Lucia: University of Queensland, 147–159. Colhoun, E. A. 2000. Vegetation and climate change during the Last Interglacial–Glacial cycle in western Tasmania, Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 155, 195–209.

273

References Collard, M. & N. Franchino 2002. Pairwise difference analysis in modern human origins research. Journal of Human Evolution 43, 323–352. Collard, M. et  al. 2010. Spatiotemporal dynamics of the Clovis-Folsom transition. Journal of Archaeological Science 37, 10, 2513–2519. Collings, H. D. 1938. Pleistocene site in the Malay Peninsula. Nature 142, 575–576. Coolidge, F. L. & T. Wynn 2009. The rise of Homo sapiens. The evolution of modern thinking. Oxford: Blackwell. Conard, N. J. 2008. A critical view of the evidence for a southern African origin of behavioural modernity. South African Archaeological Society, Goodwin Series 10, 175–179. Conard, N. J. & J. Richter (eds.) 2011. Neanderthal lifeways, subsistence and technology: one hundred and fifty years of Neanderthal study. Dordrecht: Springer. Conkey, M.W. 1978. Style and information in cultural evolution: toward a predictive model for the Paleolithic. In: C. L. Redman et al. (eds.), Social archaeology. New York: Academic Press, 61–85. Conkey, M. 1980. The identification of prehistoric hunter-gatherer aggregation sites: the case of Altamira. Current Anthropology 21, 609–630. Conkey, M. W. 1982. Boundedness in art and society. In: I. Hodder (ed.), Symbolic and structural archaeology. Cambridge: Cambridge University Press, 115–128. Cook, J. 1784. A voyage to the Pacific Ocean. Undertaken, by the command of His Majesty, for making discoveries in the northern hemisphere.Vol. 1. London: G. Nicol and T. Cadell. Coon, C. 1962. The Origin of Races. New York: Knopf. Coppens, Y. et  al. 2008. Discovery of an archaic Homo sapiens skullcap in Northeast Mongolia. Comptes Rendus Palévolution 7, 51–60. Corbey, R. & W. Roebroeks 1999. Periodization and double standards in the study of the Palaeolithic. Analecta praehistorica Leidensia 31, 77–86. Cosgrove, R. 1993. Hunters on the edge of the Tasmanian ice. In: G. Burenhult et al. (eds.), The first humans. Human origins and history to 10,000 BC. St Lucia: University of Queensland Press, 166–167. Cosgrove, R. 1995a. The illusion of riches: scale, resolution and explanation in Tasmanian Pleistocene human behaviour. Oxford: British Archaeological Reports. Cosgrove, R. 1995b. Late Pleistocene behavioural variation and time trends: the case from Tasmania. Archaeology in Oceania 30, 83–104. Cosgrove, R. 1997.Tasmanian Aborigines through the last Ice Age. In: E. K.Webb (ed.), Windows on meteorology. Australian perspective. Collingwood: CSIRO Publishing, 46–58. Cosgrove, R. 1999. Forty-two degrees south. The archaeology of late Pleistocene Tasmania. Journal of World Prehistory 13, 4, 357–402. Cosgrove, R. 2000. Lithic raw material distribution in Tasmania. In: A. Anderson & T. Murray (eds.), Australian archaeologist: collected papers in honour of Jim Allen. Canberra: Coombs Academic Publishing, ANU, 187–200. Cosgrove, R. 2004. Ice-age hunters of Tasmania. Nature Australia 28, 3, 59–63. Cosgrove, R. 2006. The archaeology of global expansion 300,000–8,000 years ago: the archaeology of late Pleistocene Australia and the Pacific. In: S. L. Elias (ed.), Encyclopaedia of Quaternary sciences. Oxford/ Amsterdam: Elsevier, 118–129. Cosgrove, R. & R. Jones 1989. Judds Cavern: a subterranean Aboriginal painting site, southern Tasmania. Rock Art Research 6, 96–104. Cosgrove, R. & J. Allen 2001. Prey choice and hunting strategies in the Late Pleistocene: evidence from southwest Tasmania. In: A. Anderson et al. (eds.), Histories of old ages: essays in honour of Rhys Jones. Canberra: Pandanus Books, ANU, 397–429. Cosgrove, R. & A. Pike-Tay 2004.The Middle Palaeolithic and late Pleistocene Tasmania hunting behaviour: a reconsideration of the attributes of modern human behaviour. International Journal of Osteoarchaeology 14, 321–332. Cosgrove, R., J. Allen & B. Marshall 1990. Palaeoecology and Pleistocene human occupation in south central Tasmania. Antiquity 64, 242, 59–78. Cosgrove, R. et  al. 2010. Overdone overkill  – the archaeological perspective on Tasmanian megafaunal extinctions. Journal of Archaeological Science 37, 2486–2503. Costa, A. G. 2012. Were there stone-tipped armatures in the South Asia Middle Paleolithic? Quaternary International 269, 22–30.

274

References Cranbrook, Earl of 2000. Northern Borneo environments of the past 40,000 years: archaeological evidence. Sarawak Museum Journal 76, 61–110. Cranbrook, Earl of 2010. Late Quaternary turnover in mammals in Borneo: the zooarchaeological record. Biodiversity Conservation 19, 373–391. Cranbrook, Earl of & P. J. Piper 2007a. Sarawak through the ice ages to the present time: environmental change and human impacts on the past and present distribution of mammals. In: R. B. Stuebing et al. (eds.), Proceedings of the Regional Conference: biodiversity conservation in tropical planted forests in South East Asia. Kuching: Forest Department, Sarawak Forest Corporation & Grand Perfect Sdn. Bhd., 75–92. Cranbrook, Earl of & P. J. Piper 2007b. The Javan rhinoceros Rhinoceros sondaicus in Borneo. Raffles Bulletin of Zoology 55, 1, 229–232. Crassard, R. 2007. Apport de la technologie lithique à la definition de la préhistoire du Hadramawt, dans le context du Yémen et de l’Arabie du Sud. Unpublished PhD thesis, Université Paris 1 Panthéon-Sorbonne, Paris. Crassard, R. 2008. La préhistoire du Yémen: diffusions et diversités locales, à travers l’étude d’industries lithiques du Hadramawt. Oxford: Archaeopress. Crassard, R. 2009. The Middle Paleolithic of Arabia: the view from the Hadramawt Region,Yemen. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 151–168. Crassard, R. & C. Thiébaut 2011. Levallois points production from eastern Yemen and some comparisons with assemblages from East-Africa, Europe and the Levant. In: J. M. Le Tensorer et al. (eds.), The Lower and Middle Palaeolithic in the Middle East and neighbouring regions. Liège: Université de Liège, 131–142. Cremaschi, M., S. Di Lernia & E. A. A. Garcea 1998. Some insights on the Aterian in the Libyan Sahara: chronology, environment, and archaeology. African Archaeological Review 15, 261–286. Crowe, I. 2000. The quest for food  – its role in human evolution and migration. Stroud and Charleston, S.C.: Tempus Publishing. Cruciani, F. et al. 2011. A revised root for the human Y chromosomal phylogenetic tree: the origin of patrilineal diversity in Africa. American Journal of Human Genetics 88, 814–818. Cucchi,T., K. Dobney & M. Fujita 2009. New insights into pig taxonomy, domestication and human dispersal in Island South East Asia through molar shape analysis: the Sus remains from Niah, and Lobang Kudih caves in Sarawak. International Journal of Osteoarchaeology 19, 4, 508–530. Curnoe, D. 2011. A 150-year conundrum: cranial robusticity and its bearing on the origin of Aboriginal Australians. International Journal of Evolutionary Biology, 18 pages, doi:10.4061/2011/632484. Curnoe, D. & A. Thorne 2006. The question of cranial robusticity. Before Farming 2, 1–13. Curnoe, D. et al. 2012. Human remains from the Pleistocene-Holocene transition of Southwest China suggest a complex evolutionary history for East Asians. PLoS One 7, e31918. Daniau, A. L., S. P. Harrison & P. J. Bartlein 2009. Fire regimes during the Last Glacial. Quaternary Science Reviews 29, 21–22, 2918–2930. Darwin, C. 1839. Journal and remarks. 1832–1836. In: R. Fitz-Roy, Narrative of the surveying voyages of His Majesty’s Ships Adventure and Beagle, between the years 1826 and 1836… Volume III. London: Henry Colburn. Darwin, C. 1859. On the origin of species by means of natural selection, or the preservation of favoured races in the struggle for life. London: John Murray. Darwin, C. 1871. The descent of man. Reprinted 1998. Amherst: Prometheus Books. Darwin, C. 2004. The descent of man, and selection in relation to sex. London: Penguin Classics. Darwin, J. A. 2007. After Tamerlane: the rise and fall of global empires, 1400–2000. London: Penguin. David, B. 1991. Fern Cave, rock art and social formations: rock art regionalisation and demographic models in southeastern Cape York Peninsula. Archaeology in Oceania 26, 41–57. David, B. 1993. Nurrabullgin Cave preliminary results from a pre-37,000  year old rockshelter, north Queensland. Archaeology in Oceania 28, 50–54. David, B. 2002. Landscapes, rock-art and the dreaming. An archaeology of preunderstanding. London: Leicester University Press. David, B. 2006. Archaeology and the dreaming: toward an archaeology of ontology. In: I. Lilley (ed.), Archaeology of Oceania: Australia and the Pacific Islands. Carlton: Blackwell Publishing, 48–68. David, B. et al. 1997. New optical and radiocarbon dates from Ngarrabuligan Cave, a Pleistocene archaeological site in Australia: implications for the comparability of time clocks and for the human colonization of Australia. Antiquity 71, 183–87. David, B. et al. 2011. Nawarla Gabarnmang, a 45,180±910 Cal Bp site in Jawoyn Country, Southwest Arnhem Land Plateau. Australian Archaeology 73, 73–77.

275

References David, B. et al. 2013. How old are Australia’s pictographs? A review of rock art dating. Journal of Archaeological Science 40, 1, 3–10. Davidson, D. S. 1930. Australian spear-thrower traits and their derivation. Journal of the Polynesian Society 43, 41–72. Davidson, D. S. 1933. Australian netting and basketry techniques. Journal of the Polynesian Society 42, 257–99. Davidson, D. S. 1934. Australian spear traits and their derivations. Journal of the Polynesian Society 43, 41–72, 143–162. Davidson, D. S. 1935.The chronology of Australian watercraft. Journal of the Polynesian Society 44, 1–16, 69–84, 137–152, 192–207. Davidson, D. S. 1936a. Aboriginal Australian and Tasmanian rock carvings and paintings. American Philosophical Society Memoirs 5, 1–151. Davidson, D. S. 1936b. Australian throwing-sticks, throwing-clubs, and boomerangs. American Anthropologist 38, 1, 76–100. Davidson, D. S. 1937.A preliminary consideration of Aboriginal Australian decorative art. American Philosophical Society Memoirs 9. Davidson, D. S. 1947. Footwear of the Australian Aborigines: environmental vs cultural determinism. Southwestern Journal of Anthropology 3, 2, 114–123. Davidson, D. S. 1948. Disposal of the dead in Western Australia. Proceedings of the American Philosophical Society 93, 1, 71–97. Davidson, I. 1990. Prehistoric Australian demography. In: B. Meehan & N. White (eds.), Hunter-gatherer demography: past and present. Sydney: Oceania Monographs, 41–58. Davidson, I. 1999. First people becoming Australian. Anthropologie 37, 1, 125–141. Davidson, I. 2001. The requirements for human colonisation of Australia. In: I. Metcalfe et al. (eds.), Faunal and floral migration and evolution in SE Asia-Australia. Lisse: Swets & Zeitlinger, 399–408. Davidson, I. 2002. The ‘finished artefact fallacy’: Acheulean handaxes and language origins. In: A. Wray (ed.), Transitions to language. Oxford: Oxford University Press, 180–203. Davidson, I. 2003. ‘Part and parcel’ – Blade industries and modern human behaviour. Australian Archaeology 57, 54–63. Davidson, I. 2007a. ‘As large as you need and as small as you can’ – Implications of the brain size of Homo floresiensis. In: A. Schalley & D. Khlentzos (eds.), Mental states: evolution, function, nature. Amsterdam: John Benjamins, 35–42. Davidson, I. 2007b. Tasmanian Aborigines and the origins of language. In: J. Mulvaney & H. Tyndale-Biscoe (eds.), Rediscovering Recherche Bay. Canberra: Academy of the Social Sciences in Australia, 69–85. Davidson, I. 2009. Comment on Bar-Yosef and Van Peer. Current Anthropology 50, 1, 119–120. Davidson, I. 2010a. The archaeology of cognitive evolution. Wiley Interdisciplinary Reviews: Cognitive Science 1, 2, 214–229. Davidson, I. 2010b. Australian archaeology as a historical science. Journal of Australian Studies 34, 3, 377–398. Davidson, I. 2010c.The colonization of Australia and its adjacent islands and the evolution of modern cognition. Current Anthropology 51, s1, s177-s189. Davidson, I. 2012. Four questions about foraging models and the process of colonisation. Australian Archaeology 74, 19–20. Davidson, I. 2013. Peopling the last new worlds: the first colonization of Sahul and the Americas. Quaternary International 285, 1–29. Davidson, I. in press. Hunter-gatherers in Australia: deep histories of continuity and change. In:V. Cummings et al. (eds.), Oxford handbook of the archaeology and anthropology of hunter-gatherers. Oxford: Oxford University Press. Davidson, I. & W. Noble 1989. The archaeology of perception: traces of depiction and language. Current Anthropology 30, 2, 125–55. Davidson, I. & W. Noble 1992. Why the first colonization of the Australian region is the earliest evidence of modern human behaviour. Archaeology in Oceania 27, 135–142. Davidson, I. et  al. 2005. Archaeology in another country: exchange and symbols in North West Central Queensland. In: I. Macfarlane et al. (eds.), Many exchanges: archaeology, history, community and the work of Isabel Mcbryde. Canberra: Aboriginal History Inc., 101–128. Deacon, H. J. 1989. Late Pleistocene palaeoecology and archaeology in the southern Cape, South Africa. In: P. Mellars & C. B. Stringer (eds.), The human revolution: behavioural and biological perspectives on the origins of modern humans. Edinburgh: Edinburgh University Press, 547–564.

276

References Deacon, H. J. 2000. Modern human emergence: an African archaeological perspective. In: P. V. Tobias et al. (eds.), Humanity from African naissance to coming millennia. Florence: Florence University Press, 217–226. Deacon, H. & J. Deacon 1999. Human beginnings in South Africa. Cape Town: David Philip Publishers. Delagnes, A. & L. Meignen 2006. Diversity of lithic production systems during the Middle Paleolithic in France. Are there any chronological trends? In: E. Hovers & S. L. Kuhn (eds.), Transitions before the transition. Evolution and stability in the Middle Paleolithic and Middle Stone Age. New York: Springer, 85–107. Delagnes, A. et  al. 2012. Inland human settlement in southern Arabia 55,000  years ago. New evidence from the Wadi Surdud Middle Paleolithic site complex, western Yemen. Journal of Human evolution 63, 452–474. Delagnes, A. et al. in press. The Middle Paleolithic assemblage of Shi’batDihya 1 (Wadi Surdud site complex, Yemen). Journal of Human Evolution. Demeter, F. et al. 2004. An archaic Homo molar from northern Vietnam. Current Anthropology 45, 4, 535–541. Demeter, F. et al. 2005. Discovery of a second human molar and cranium fragment in the late Middle to Late Pleistocene cave of Ma U’Oi (Northern Vietnam). Journal of Human Evolution 48, 393–402. Demeter, F. et  al. 2012. Anatomically modern human in Southeast Asia (Laos) by 46 ka. PNAS 109, 36, 14375–14380. Denham, T. & H. Barton 2006. The emergence of agriculture in New Guinea: continuity from pre-existing foraging practices. In: D. J. Kennett & B. Winterhalder (eds.), Behavioral ecology and transition to agriculture. Berkeley: University of California Press, 237–264. Denham, T., R. Fullagar & L. Head 2009. Plant exploitation on Sahul: from colonisation to the emergence of regional specialisation during the Holocene. Quaternary International 202, 1–2, 29–40. Dennell, R. W. 1990. Progressive gradualism, imperialism, and academic fashion: Lower Palaeolithic archaeology in the twentieth century. Antiquity 64, 549–558. Dennell, R. W. 1995. The early stone age of Pakistan: a methodological review. Man and Environment: 20, 1, 21–28. Dennell, R.W. 2001. From Sangiran to Olduvai, 1937–1960: the quest for ‘centres’ of hominid origins in Asia and Africa. In: R. Corbey & W. Roebroeks (eds.), Studying human origins: disciplinary history and epistemology. Amsterdam: Amsterdam University Press, 45–66. Dennell, R. W. 2009. The Palaeolithic settlement of Asia. Cambridge: Cambridge University Press. Dennell, R. W. 2010. Early Homo sapiens in China. Nature 468, 512–513. Dennell, R. W. 2013. Hominins, deserts, and the colonisation and settlement of continental Asia. Quaternary International 300, 13–21. Dennell, R. W. & H. Rendell 1991. De Terra and Paterson, and the Soan flake industry: a perspective from the Soan Valley, Pakistan. Man and Environment 16, 2, 91–99. Dennell, R. W. & M. D. Petraglia 2012. The dispersal of Homo sapiens across southern Asia: how early, how often, how complex? Quaternary Science Reviews 47, 15–22. Dennell, R. W. et al. 1992. A 45,000-year-old open-Palaeolithic site at Riwat, northern Pakistan. Journal of Field Archaeology 19, 1, 17–33. Dennell, R. W. et al. 2011. Hominin variability, climatic instability and population demography in Middle Pleistocene Europe. Quaternary Science Reviews 30, 1511–1524. Deraniyagala, S. U. 1992. The prehistory of Sri Lanka: an ecological perspective. Colombo: Department of Archaeological Survey, Government of Sri Lanka. d’Errico, F. 2003. The invisible frontier. A multiple species model for the origin of behavioral modernity. Evolutionary Anthropology 12, 4, 188–202. d’Errico, F. 2008. Le rouge et le noir: implications of early pigment use in Africa, the Near East and Europe for the origin of cultural modernity. South African Archaeological Society Goodwin Series 10, 168–174. d’Errico, F. & C. S. Henshilwood 2007. Additional evidence for bone technology in the southern African Middle Stone Age. Journal of Human Evolution 52, 2, 142–163. d’Errico, F. & M. Vanhaeren 2007. Evolution or revolution? New evidence for the origin of symbolic behaviour. In: P. Mellars et al. (eds.), Rethinking the human revolution. Cambridge: McDonald Institute for Archaeological Research, 275–286. d’Errico, F., C. Henshilwood & P. Nilssen, 2001. An engraved bone fragment from c. 70,000-year-old Middle Stone Age levels at Blombos Cave, South Africa: implications for the origin of symbolism and language. Antiquity 75, 309–318. d’Errico, F. et al. 1998. Neanderthal acculturation in Western Europe? A critical review of the evidence and its interpretations. Current Anthropology 39, S1-S44.

277

References d’Errico, F. et al. 2003. Archaeological evidence for the emergence of language, symbolism, and music – an alternative multidisciplinary perspective. Journal of World Prehistory 17, 1–70. d’Errico, F. et al. 2011. Pacea geo-referenced radiocarbon database. PaleoAnthropology 2011, 1–12. Détroit, F. 2002. Origine et évolution des Homo sapiens en Asie du Sud-Est: Descriptions et analyses morphométriques de nouveaux fossiles. Unpublished PhD thesis, Muséum national d’Histoire naturelle, Paris. Détroit, F. et  al. 2004. Upper Pleistocene Homo sapiens from the Tabon Cave (Palawan, The Philippines): description and dating of new discoveries. Comptes Rendues Palévolution 3, 705–712. Dhir, R. P. et al. 2010. Multiple episodes of aggradation and calcrete formation in Late Quaternary aeolian sands, Central Thar Desert, Rajasthan, India. Journal of Asian Earth Sciences 37, 1, 10–16. Diamond, J. & P. Bellwood 2003. Farmers and their languages: the first expansions. Science 300, 597–603. Díaz-Andreu, M. 2007. A world history of nineteenth-century archaeology: nationalism, colonialism, and the past. Oxford: Oxford University Press. Diez-Martín, F. et al. 2009. The Middle to Later Stone Age technological transition in East Africa: new data from Mumba Rockshelter Bed V (Tanzania) and their implications for the origin of modern human behavior. Journal of African Archaeology 7, 147–173. Dinnis, R.,A. Pawlik & C. Gaillard 2009. Bladelet cores as weapon tips? Hafting residue identification and microwear analysis of three carinated burins from Les Vachons. Journal of Archaeological Science 36, 1922–1934. Dizon, E. Z. & A. Pawlik 2010.The Lower Palaeolithic record in the Philippines. Quaternary International 223–224, 444–450. Dizon, E. et al. 2002. Notes on the morphology and age of the Tabon Cave Homo sapiens. Current Anthropology 43, 660–666. Donlon, D. 1994. Aboriginal skeletal collections and research in physical anthropology: an historical perspective. Australian Archaeology 39, 73–82. Donohue, M. & T. Denham 2010. Farming and language in Island Southeast Asia: reframing Austronesian History. Current Anthropology 51, 2, 223–256. Dortch, C. E. 1976. Two engraved stone plaques of late Pleistocene age from Devil’s Lair, Western Australia. Archaeology and Physical Anthropology in Oceania 3, 33–40. Dortch, C. E. 1979. Australia’s oldest known ornaments. Antiquity 53, 39–43. Dortch, C. E. 1984. Devil’s Lair, a study in prehistory. Perth: Western Australian Museum. Dortch, J. 2004. Palaeo-environmental change and the persistence of human occupation in south-western Australian forests. Oxford: British Archaeological Reports, International Series. Archaeopress. Dortch, C. E. & J. Dortch 1996. Review of Devil’s Lair artifact classification and radiocarbon chronology. Australian Archaeology 43, 28–32. Dortch, J. & R. Wright 2010. Identifying palaeo-environments and changes in Aboriginal subsistence from dual-patterned faunal assemblages, south-western Australia. Journal of Archaeological Science 37, 1053–1064. Dortch, J., J. Balme & J. Ogilvie 2012. Aboriginal responses for late Quaternary environmental change in a Mediterranean-type region: zoo archaeological evidence from south-western Australia. Quaternary International 264, 121–134. Drake, N. A., P. Breeze & A. G. Parker 2013. Palaeoclimate in the Saharan and Arabian deserts during the Middle Palaeolithic and the potential for hominin dispersals. Quaternary International 300, 48–61. Drake, N. A. et al. 2011. Ancient watercourses and biogeography of the Sahara explain the peopling of the desert. PNAS 108, 458–462. Dubois, E. 1921. The proto-Australian fossil man of Wadjak, Java. Koninklijke Nederlandsche Akademie van Wetenschappen Proceedings 23, 2, 1013–1051. Duckworth,W. L. H. 1894.A critical study of the collection of crania of Aboriginal Australians in the Cambridge University Museum. Journal of the Anthropological Institute of Great Britain and Ireland 23, 284–314. Duckworth, W. L. H. 1902. Craniological notes on the Aborigines of Tasmania. Journal of the Anthropological Institute of Great Britain and Ireland 32, 177–181. Dunlap, T. R. 1997. Remaking the land: the acclimatization movement and Anglo ideas of Nature. Journal of World History 8, 2, 303–319. Eberl, B. 1930. Die Eiszeitenfolge im nördlichen Alpenvorlande. Augsburg: Benno Filser. Edens, C. A. 2001. A bladelet industry in southwestern Saudi Arabia. Arabian Archaeology and Epigraphy 12, 137–142. Elkin, A. P. 1949. The origin and interpretation of petroglyphs in south-east Australia. Oceania 20, 119–157. Endicott, K. & P. Bellwood 1991. The possibility of independent foraging in the rain forest of peninsular Malaysia. Human Ecology 19, 2, 151–185.

278

References Endicott, P. & S.Y. W. Ho 2008. A Bayesian evaluation of human mitochondrial substitution rates. American Journal of Human Genetics 82, 142–146. Endicott, P., M. Metspalu & T. Kivisild 2007. Genetic evidence on modern human dispersals in South Asia: Y chromosome and mitochondrial DNA perspectives: the world through the eyes of two haploid genomes. In: M. D. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia. Amsterdam: Springer, 229–244. Endicott, P. et al. 2009. Evaluating the mitochondrial timescale of human evolution. Trends in Ecology and Evolution 24, 515–521. Erlandson, J. M. 2001. The archaeology of aquatic adaptations: paradigms for a new millennium. Journal of Archaeological Research 9, 4, 287–350. Erdbrink, D. P. 1954. Mesolithic remains of the Sampung stage in Java: some remarks and additions. Southwestern Journal of Anthropology 10, 294–303. Es, L. J. C. van 1930. The prehistoric remains in the Sampoeng cave, residency of Ponorogo, Java. Proceedings of the Fourth Pacific Science Congress, Java, 1929. Volume III: Biological papers, 329–340. Esposito, M. et al. 2002. U-series dating of fossil teeth and carbonates from Snake Cave, Thailand. Journal of Archaeological Science 29, 341–349. Esseltyn, J. A. et al. 2010. Integrating phylogenetic and taxonomic evidence illuminates complex biogeographic patterns along Huxley’s modification of Wallace’s Line. Journal of Biogeography 37, 2054–2066. Etheridge, R. 1890. Has man a geological history in Australia? Proceedings of the Linnaean Society of New South Wales, 2nd series, 5, 262–265. Evans, N. 1988. Arguments for Pama-Nyungan as a genetic subgroup, with particular reference to initial laminalization. Aboriginal Linguistics 1, 91–110. Evans, I. H. N. 1990 (orig. published 1923). Among primitive people in Borneo: a description of the lives, habits and customs of the piratical head-hunters of North Borneo, with an account of objects of prehistoric antiquity discovered on the Island. Singapore: Oxford University Press. Evans, N. 2005. Australian languages reconsidered: a review of Dixon (2002). Oceanic Linguistics 44, 242–286. Farrera, I. et al. 1999. Tropical climates at the Last Glacial Maximum: a new synthesis of terrestrial palaeoclimate data. I.Vegetation, lake-levels and geochemistry. Climate Dynamics 15, 823–856. Faure, H., R. C. Walter & D. R. Grant 2002. The coastal oasis: Ice Age springs on emerged continental shelves. Global and Planetary Change 33, 47–56. Fenner, F. J. 1939.The Australian Aboriginal skull: its non-metric morphological characters. Transactions of the Royal Society of South Australia 63, 248–306. Fenton, M. et al. 2000. Aplanktonic zones in the Red Sea. Marine Micropaleontology 40, 277–294. Fernandes, C. A. 2009. Bayesian coalescent inference from Mitochondrial DNA variation of the colonization time of Arabia by the Hamadryas baboon (Papio hamadryas hamadryas). In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 89–102. Fernandes, C. A., E. J. Rohling & M. Siddall 2006. Absence of post-Miocene Red Sea land bridges: biogeographic implications. Journal of Biogeography 33, 961–966. Fernandes,V. et al. 2012. The Arabian cradle: mitochondrial relicts of the first steps along the southern route out of Africa. American Journal of Human Genetics 90, 347–355. Field, J. S. & M. M. Lahr 2005. Assessment of the southern dispersal: GIS based analyses of potential routes at Oxygen Isotope Stage 4. Journal of World Prehistory 19, 1–45. Field, J. S., M. D. Petraglia & M. M. Lahr 2007. The southern dispersal hypothesis and the South Asian archaeological record: examination of dispersal routes through GIS analysis. Journal of Anthropological Archaeology 26, 1, 88–108. Finlayson, C. & J. S. Carrión 2007. Rapid ecological turnover and its impact on Neanderthal and other human populations. Trends in Ecology & Evolution 22, 4, 213–222. Fischer, A. et al. 1984. Macro and micro wear traces on lithic projectile points: experimental results and prehistoric examples. Journal of Danish Archaeology 3, 19–46. Flannery, T. 1988. Origins of the Australo-Pacific land mammal fauna. Australian Zoological Reviews 1, 15–24. Flannery, T. 1995. Mammals of the South-West Pacific and Moluccan Islands. Chatswood: Reed Books, Australian Museum. Flannery, T. & J. P. White 1991. Animal translocation. Zoogeography of New Ireland mammals. National Geographic Research and Exploration 7, 96–113.

279

References Flannery, T., M.-J. Mountain & K. Aplin 1983. Quaternary kangaroos (Macropodidae: Marsupialia) from Nombe rockshelter, Papua New Guinea, with comments on the nature of megafaunal extinction in the New Guinea Highlands. Proceedings of the Linnean Society of New South Wales 107, 2, 75–97. Fleitmann, D. et al. 2011. Holocene and Pleistocene pluvial periods in Yemen, southern Arabia. Quaternary Science Reviews 30, 783–787. Flood, J. 1973. Pleistocene human occupation and extinct fauna in Cloggs Cave, Buchan, south-east Australia. Nature 246, 5431, 303. Flood, J. 1995. Archaeology of the dreamtime. Sydney: Collins. Flood, J. 1997. Rock art of the dreamtime. Sydney: Angus and Robertson. Flower, W. H. 1878. The native races of the Pacific Ocean. Proceedings of the Royal Institute of Great Britain 8, 602–652. Foley, R. & M. M. Lahr 1997. Mode 3 technologies and the evolution of modern humans. Cambridge Archaeological Journal 7, 3–36. Foley, R. & M. M. Lahr 2003. On stony ground: lithic technology, human evolution, and the emergence of culture. Evolutionary Anthropology 12, 109–122. Foley, R. & M. M. Lahr 2011. The evolution of the diversity of cultures. Philosophical Transactions of the Royal Society B: Biological Sciences 366, 1567, 1080–1089. Ford, A. 2011. Learning the lithic landscape: using raw material sources to investigate Pleistocene colonisation in the Ivane Valley, Papua New Guinea. Archaeology in Oceania 46, 42–53. Foster, J. B. 1964. The evolution of mammals on islands. Nature 202, 234–235. Fox, R. 1970. The Tabon Caves: archaeological explorations on Palawan Island, Philippines. Manila: National Museum of the Philippines. Fox, R. B. 1971. Ancient man and Pleistocene fauna in Cagayan Valley, Northern Luzon Philippines: a progress report. Unpublished manuscript of the National Museum of the Philippines, Manila. Fox, R. B. 1978. The Philippine Palaeolithic. In: F. Ikawa-Smith (ed.), Early Palaeolithic in South and East Asia. Paris: Moutton, 59–85. Fox, R. B. & J. Peralta 1974. Preliminary report on the Palaeolithic archaeology of Cagayan, Philippines, and the Cabalwanian Industry. Proceedings of the first regional seminar on Southeast Asian prehistory and archaeology (July–August 1972). Manila, 100–147. Franklin, N. R. 1984. Of macropods and men: an analysis of the Simple Figurative Styles. Unpublished BA (hons) thesis, Department of Anthropology, University of Sydney. Franklin, N. R. 1989. Research with style: a case study from Australian rock art. In: S. J. Shennan (ed.), Archaeological approaches to cultural identity. London: Unwin Hyman, 278–290. Franklin, N. R. 1996. Rock art in Australia and New Guinea 1990–1994: research, management and conservation. In: P. J. Bahn & A. Fossati (eds.), Rock arts studies: news of the World I. Oxford: Oxbow Books, 145–161. Franklin, N. R. 2004. Explorations of variability in Australian prehistoric rock engravings. Oxford: British Archaeological Reports. Franklin, N. R. 2007a. Aboriginal engravings in the southwest of Western Australia: analysis of the Kybra site. Records of the Western Australian Museum 24, 65–79. Franklin, N. R. 2007b. Discontinuous dreaming networks: analyses of variability in Australian pre-historic petroglyphs. Rock Art Research 24, 79–104. Franklin, N. R. & P. J. Habgood 2007. Modern human behaviour and Pleistocene Sahul in review. Australian Archaeology 65, 1–16. Freedman, L. 1964. Metrical features of Aboriginal crania from coastal New South Wales, Australia. Records of the Australian Museum 26, 309–326. Fredericksen, C., M. Spriggs & W. Ambrose 1993. Pamwak rockshelter: a Pleistocene site on Manus Island, Papua New Guinea. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene archaeology in Australia, New Guinea and Island Melanesia. Canberra: ANU, 144–152. Freeman, D. 1979. Severed heads that germinate. In: R. J. Hook (ed.), Fantasy and symbol: studies in anthropological interpretation. London: Academic Press, 233–246. Frenzen, J. 1984. G. H. R. von Koenigswald and Asia – an obituary. Asian Perspectives 25, 43–51. Friedlaender, J. S. et al. 2005a. Expanding Southwest Pacific mitochondrial haplogroups P and Q. Molecular Biology and Evolution 22, 6, 1506–1517.

280

References Friedlaender, J. S. et al. 2005b. Mitochondrial genetic diversity and its determinants in Island Melanesia. In: A. Pawley et al. (eds.), Papuan pasts: studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, 693–716. Friedlaender, J. S. et al. 2007. Melanesian mtDNA complexity. PLoS Online 2, e248. Fullagar, R. 1986. Use-wear and residues on stone tools. Functional analysis and its application to two southeastern Australian archaeological assemblages. PhD thesis, Department of Archaeology, La Trobe University. Fullagar, R. & J. Field 1997. Pleistocene seed grinding implements from the Australian arid zone. Antiquity 71, 300–307. Gagge, A. P., A. C. Burton & H. C. Bazett 1941. A practical system of units for the description of the heat exchange of man with his environment. Science 94, 2445, 428–430. Gale, S. J. & P. G. Hoare 1992. Quaternary sediments. London: Belhaven. Gamble, C. 1982. Interaction and alliance in Palaeolithic society. Man n.s.17, 92–107. Gamble, C. S. 1986. The Palaeolithic settlement of Europe. Cambridge: Cambridge University Press. Gamble, C. S. 1992. Archaeology, history and the uttermost ends of the earth – Tasmania, Tierra del Fuego and the Cape. Antiquity 66, 712–720. Gamble, C. S. 2007.Origins and revolutions: human identity in earliest prehistory. Cambridge: Cambridge University Press. Gamble, C. & G. Marshall 2001. The shape of handaxes, the structure of the Acheulian world. In: S. Milliken & J. Cook (eds.), A very remote period indeed: papers on the Palaeolithic presented to Derek Roe. Oxford: Oxbow Books, 19–27. Gamble, C. S. & E. K. Gittins 2004. Social archaeology and origins research. In: R. Preucel & M. Meskell (eds.), Social archaeology: a companion. Oxford: Blackwell, 96–118. Gamble, C. S. et al. 2004. Climate change and evolving human diversity in Europe during the last glacial. Philosophical Transactions of the Royal Society, Biological Sciences 359, 243–254. Gao, X. et al. 2010. Revisiting the origin of modern humans in China and its implications for global human evolution. Science China (Earth Series) 53, 12, 1927–1940. Gardner, R. & H. Martingell 1990. Microlithic sites and their paleoenvironmental setting, Southeast India: a reevaluation. Geoarchaeology 5, 1, 1–13. Garrard, A., C. P. D Harvey & V. R. Switsur 1981. Environment and settlement during the Upper Pleistocene and Holocene at Jubba in the Great Nefud, Northern Arabia. Atlal 5, 137–148. Garrod, D. A. E. & D. M. Bate 1937. The stone age of Mount Carmel,Vol. I. Oxford: Clarendon Press. Garvey, J. 2006. Preliminary zooarchaeological interpretations from Kutikina Cave, south-west Tasmania. Australian Aboriginal Studies 1, 58–63. Garvey, J. 2010. Economic anatomy of the Bennett’s wallaby (Macropus rufogriseus): implications for understanding human hunting strategies in late Pleistocene Tasmania. Quaternary International 211, 144–256. Garvey, J. 2011. Bennett’s wallaby (Macropus rufogriseus) marrow quality vs quantity: evaluating human decisionmaking and seasonal occupation in late Pleistocene Tasmania. Journal of Archaeological Science 38, 763–783. Gathorne-Hardy, F. J. et al. 2002. Quaternary rainforest refugia in South-East Asia: using termites (Isoptera) as indicators. Biological Journal of the Linnean Society 75, 453–466. Gaudzinski-Windheuser, S. & L. Niven 2009. Hominin subsistence patterns during the Middle and Late Paleolithic in Northwestern Europe. In: J.-J. Hublin & M. P. Richards (eds.), The evolution of hominin diets: integrating approaches to the study of Palaeolithic subsistence. Leipzig: Springer, 99–111. Geneste, J.-M. et al. 2010. Earliest evidence for ground-edge axes: 35,400 ± 410 cal BP from Jawoyn Country, Arnhem Land. Australian Archaeology 71, 66–69. Ghirotto, S., L. Penso-Dolfin & G. Barbujani 2011. Genomic evidence for an African expansion of anatomically modern humans by a southern route. Human Biology 83, 477–489. Gibbs, M. 2011. An Aboriginal fish trap on the Swan Coastal Plain: the Barragup Mungah. Records of the Western Australian Museum Supplement 79, 4–15. Gijn, A. L. van 1989. The wear and tear of flint. Leiden: Leiden University. Gilbertson, D. et al. 2005. Past human activity and geomorphological change in a guano-rich tropical cave mouth: initial interpretations of the Late Quaternary succession in the Great Cave of Niah, Sarawak. Asian Perspectives 44, 1, 16–41. Gilbertson, D. et  al. 2013. The cave entrance sequences and environmental change. In: G. Barker (ed.), Rainforest foraging and farming in Southeast Asia: the archaeology of the Niah Caves, Sarawak. Cambridge: McDonald Institute for Archaeological Research, 71–134.

281

References Gilganic, L. A. et al. 2012. New ages for Middle and Later Stone Age deposits at Mumba rockshelter,Tanzania: optically stimulated luminescence dating of quartz and feldspar grains. Journal of Human Evolution 62, 4, 533–547. Gillespie, R. 1998. Alternative timescales: a critical review of Willandra Lakes dating. Archaeology in Oceania 33, 169–182. Gillespie, R. 2002. Dating the first Australians. Radiocarbon 44, 455–472. Gillespie, R. & R. G. Roberts 2000. On the reliability of age estimates for human remains at Lake Mungo. Journal of Human Evolution 38, 727–732. Gilligan, I. 2007a. Resisting the cold in Ice Age Tasmania: thermal environment and settlement strategies. Antiquity 81, 555–568. Gilligan, I. 2007b. Clothing and modern human behaviour: prehistoric Tasmania as a case study. Archaeology in Oceania 42, 102–111. Gilligan, I. 2007c. Another Tasmanian paradox. Clothing and thermal adaptations in Aboriginal Australia. Oxford: Archaeopress. Gilligan, I. 2008. Clothing and climate in Aboriginal Australia. Current Anthropology 49, 3, 487–495. Gilligan, I. 2010a. The prehistoric development of clothing: Archaeological implications of a thermal model. Journal of Archaeological Method and Theory 17, 15–80. Gilligan, I. 2010b. Clothing and modern human behaviour in Australia. Bulletin of the Indo-Pacific Prehistoric Association 30, 54–69. Glover, I. & R. Lampert 1969. Puntutjarpa rockshelter excavation by R.A. Gould: a critical review. Archaeology and Physical Anthropology in Oceania 4, 222–228. Godfrey-Smith, P. 1999. Adaptationism and the power of selection. Biology and Philosophy 14, 181–194. Godfrey-Smith, P. 2000. Three kinds of adaptationism. In S. H. Orzack and E. Sober (eds.), Optimality and adaptationism. Cambridge: Cambridge University Press, 335–357. Gollan, K. 1984. The Australian dingo: in the shadow of man. In: M. Archer & C. Clayton (eds.), Vertebrate zoogeography and evolution in Australasia. Perth: Hesperian Press, 921–927. Golson, J. 1971. Australian Aboriginal food plants: some ecological and culture-historical implications. In: D. J. Mulvaney & J. Golson (eds.), Aboriginal man and environment in Australia. Canberra: ANU Press, 196–238. Gonder, M. K. et al. 2007. Whole-mtDNA genome sequence analysis of ancient African lineages. Molecular Biology and Evolution 24, 3, 757–768. Gopher, A. et al. 2005. Qesem Cave. An Amudian site in central Israel. Journal of the Israel Prehistoric Society 35, 69–92. Gorecki, P., M. Mabin & J. Campbell 1991. Archaeology and geomorphology of the Vanimo Coast, Papua New Guinea: preliminary results. Archaeology in Oceania 26, 3, 119–122. Gorman, C. F. 1970. Excavations at Spirit Cave, Northern Thailand. Asian Perspectives 13, 79–107. Gorman, C. F. 1971. The Hoabinhian and after: subsistence patterns in Southeast Asia during the Late Pleistocene and early recent periods. World Archaeology 2, 3, 300–320. Gosden, C. 1993. Understanding the settlement of the Pacific Islands in the Pleistocene. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene archaeology in Australia and Island Melanesia. Canberra: ANU, 131–136. Gosden, C. 1995. Arboriculture and agriculture in coastal Papua New Guinea. Antiquity 69, Special Number 265, 807–817. Gosden, C. 2010. When humans arrived in the New Guinea highlands. Science 330, 41–42. Gosden, C. & N. Robertson 1991. Models for Matenkupkum: interpreting a Late Pleistocene site from southern New Ireland, Papua New Guinea. In: J. Allen & C. Gosden (eds.), Report of the Lapita Homeland Project. Canberra: ANU, 20–45. Gott, B. 2003. Fire-making in Tasmania: absence of evidence is not evidence of absence. Current Anthropology 43, 4, 650–656. Gould, R. 1969. Puntutjarpa rockshelter: A reply to Messrs Glover and Lampert. Archaeology and Physical Anthropology in Oceania 4, 229–237. Gould, R. A. 1980. Living archaeology. Cambridge: Cambridge University Press. Gould, S. J. 1996. The mismeasure of man. 2nd ed. New York: W. W. Norton. Gould, S. J. 2002. The structure of evolutionary theory. Cambridge, Mass.: Harvard University Press. Gould, S. J. & R. Lewontin 1979.The spandrels of San Marco and the Panglossian paradigm: a critique of the adaptationist programme. Proceedings of the Royal Society London B 205, 581–598.

282

References Green, R. E. et  al. 2008. A complete Neandertal mitochondrial genome sequence determined by highthroughput sequencing. Cell 134, 416–426. Green, R. E. et al. 2010. A draft sequence of the Neandertal genome. Science 328, 710–722. Groube, L. 1984. Waisted axes of Asia, Melanesia and Australia. In: G. K. Ward (ed.), Archaeology at ANZAAS, Canberra. Canberra: ANU, 168–177. Groube, L. 1989. The taming of the rainforests: a model for late Pleistocene forest exploitation in New Guinea. In: D. R. Harris & G. C. Hillman (eds.), Foraging and farming: the evolution of plant exploitation. London: Unwin Hyman, 292–304. Groube, L. et al. 1986. A 40,000 year old occupation site at Huon Peninsula, Papua New Guinea. Nature 324, 6096, 453–455. Groucutt, H. & M. D. Petraglia 2012.The prehistory of the Arabian Peninsula: deserts, dispersals and demography. Evolutionary Anthropology 21, 113–125. Groves, C. P. & M. M. Lahr 1994. A bush not a ladder: speciation and replacement in human evolution. Perspectives in Human Biology 4, 1–11. Grün, R. & C. B. Stringer 1991. Electron spin resonance dating and the evolution of modern humans. Archaeometry 33, 153–199. Grün, R. & C. Stringer 2000. Tabun revisited: revised ESR chronology and new ESR and U-series analyses of dental material from Tabun C1. Journal of Human Evolution 39, 6, 601–612. Grün, R. & A. Thorne 1997. Dating the Ngandong humans. Science 276, 1575. Grün, R., H. P. Schwarcz & J. M. Chadam 1988. ESR dating of tooth enamel: coupled correction for U-uptake and U-series disequilibrium. Nuclear Tracks and Radiation Measurements 14, 237–241. Grün, R. et al. 2000. Age of the Lake Mungo 3 skeleton, reply to Bowler & McGee and to Gillespie & Roberts. Journal of Human Evolution 38, 733–741. Grün, R. et al. 2005. U-Series and ESR analyses of bones and teeth relating to the human burials from Skhul. Journal of Human Evolution 49, 316–334. Grün, R. et al. 2010. ESR and U-series analysis of faunal material from Cuddie Springs, NSW, Australia: implications for the timing of the extinction of the Australian megafauna. Quaternary Science Reviews 29, 596–610. Grün, R. et al. 2011. Stratigraphy and chronology of the WLH 50 human remains, Willandra Lakes World Heritage Area, Australia. Journal of Human Evolution 60, 597–604. Güleç, E. 2007. Early Upper Paleolithic human dental remains from Ucagizli Cave (Hatay, Turkey). American Journal of Physical Anthropology 132, S44, 122. Gunn, R. G. 1995. Regional patterning in the Aboriginal rock art of Central Australia: a preliminary report. Rock Art Research 12, 117–128. Gunn, R. G. 1997. Rock art, occupation and myth: the correspondence of symbolic and archaeological sites within Arrernte rock art complexes of central Australia. Rock Art Research 14, 125–136. Gunn, R. G. 2000. Central Australian rock art: a second report. Rock Art Research 17, 111–126. Gunn, R. G. 2007. The interpretation of handedness in Australian Aboriginal rock art. Rock Art Research 24, 199–208. Gunn, R. G. & R. L. Whear 2008. A singular beeswax representation of Namarrkon, the Lightning Man, from western Arnhem Land. Australian Aboriginal Studies 2, 54–69. Haberle, S. G. & B. David 2004. Climates of change: human dimensions of Holocene environmental change in low latitudes of the PEPII transect. Quaternary International 118–119, 165–179. Habgood, P. J. 1986. The origin of the Australians: A multivariate approach. Archaeology in Oceania 21, 130–137. Habgood, P. J. 1991. Aboriginal fossil hominids: evolution and migrations. In: R. Foley (ed.), The origins of human behaviour. London: Unwin Hyman, 97–113. Habgood, P. J. & N. R. Franklin 2008.The revolution that didn’t arrive: a review of Pleistocene Sahul. Journal of Human Evolution 55, 2, 187–222. Habgood, P. J. & N. R. Franklin 2011. Explanations for patterning in the ‘package of traits’ of modern human behaviour within Sahul. Bulletin of the Indo-Pacific Prehistory Association 30, 14–27. Haidle, M. 2008. Kognitive und kulturelle Evolution. Erwägen Wissen Ethik (EWE) 19, 2, 149–159. Haidle, M. N. 2010. Working memory capacity and the evolution of modern cognitive potential: implications from animal and early human tool use. Current Anthropology 51, S1, S149-S166. Haidle, M. & A. Pawlik 2009. Missing types: overcoming the typology dilemma of lithic archaeology in Southeast Asia. Bulletin of the Indo-Pacific Prehistoric Association 29, 2–5.

283

References Haidle, M. & A. Pawlik 2010. Pleistocene modernity: an exclusively Afro-European issue? Bulletin of the IndoPacific Prehistoric Association 30, 3–8. Hale, H. M. & N. B. Tindale 1930. Notes on some human remains in the lower Murray Valley, South Australia. Records of the South Australian Museum 4, 145–218. Hallam, S. 1975. Fire and hearth: a study of Aboriginal usage and European usurpation in South-Western Australia. Canberra: Australian Institute of Aboriginal Studies. Hallam, S. 1989. Plant usage and management in South-West Australian societies. In: D. R. Harris & G. C. Hillman (eds.), Foraging and farming. London: Unwin-Hyman, 136–150. Hammond, M. 1988.The shadow man paradigm in paleoanthropology, 1911–1945. In: G.W. Stocking Jr. (ed.), Bones, bodies, behavior: essays on biological anthropology. Madison: University of Wisconsin Press, 117–137. Harding,T. 1967. Across the Vitiaz Straits: a study of a New Guinea trade system. Seattle: University of Washington Press. Harper, W. R. & A. Clarke 1898. Notes on the measurements of the Tasmanian crania in the Tasmanian Museum. Papers and Proceedings of the Royal Society of Tasmania for 1897, 97–110. Harris, E. E. 2010. Nonadaptive processes in primate and human evolution. Yearbook of Physical Anthropology 53, 13–45. Harris, S., D. Ranson & S. Brown 1988. Maxwell River archaeological survey 1986. Australian Archaeology 27, 89–97. Harrison, R. 2004. Kimberley points and colonial preference: new insights into the chronology of pressure flaked point forms from the Southeast Kimberley, Western Australia. Archaeology in Oceania 39, 1, 1–11. Harrisson, T. 1937. Savage civilisation. London: Gollancz and Left Book Club. Harrisson, T. 1943. Living among cannibals. London: Harrap. Harrisson, T. 1958. The caves of Niah: a history of prehistory. Sarawak Museum Journal 8 (n.s. 12), 549–595. Harrisson, T. 1959a. World within: a Borneo story. London: The Cresset Press. Harrisson, T. 1959b. New archaeological and ethnological results from Niah Caves, Sarawak. Man 59, 1–8. Harrisson, T. 1959c. Radiocarbon datings from Niah: a note. Sarawak Museum Journal (n. s.) 13–14, 136–142. Harrisson, T. 1961. Britain revisited. London:Victor Gollancz. Harrisson,T. 1972.The Borneo Stone Age in the light of a recent research. Sarawak Museum Journal 20, 40–41, 385–412. Harrold, F. B. 1989. Mousterian, Châtelperronian and early Aurignacian in Western Europe: continuity or discontinuity. In: P. Mellars & C. B. Stringer (eds.), The human revolution: behavioural and biological perspectives on the origins of modern humans. Edinburgh: Edinburgh University Press, 677–713. Haslam, M. et al. 2010.The 74,000 BP Toba super-eruption and southern Indian hominins: archaeology, lithic technology and environments at Jwalapuram Locality 3. Journal of Archaeological Science 37, 3370–3384. Haslam, M. et  al. 2011. Late Acheulean hominins at the Marine Isotope Stage 6/5e transition in northcentral India. Quaternary Research 75, 3, 670–682. Haslam, M. et al. 2012. Dhaba: A southern Indian Middle Palaeolithic occupation surface sealed by the 74 ka Toba eruption: further evidence from Jwalapuram Locality 22. Quaternary International 258, 148–164. Hautala, S. J. et al. 2001.Velocity structure and transport of the Indonesian Throughflow in the major straits restricting flow into the Indian Ocean. Journal of Geophysical Research 106, C9, 19527–19546. Hawkes, J., C. Hawkes & H. de Terra 1934. Palaeolithic human industries in the northwest Punjab and Kashmir and their geological significance. Memoirs of the Connecticut Academy of Arts and Sciences 8, 1–15. Hawks, J. et al. 2000. An Australasian test of the recent African origin theory using the WLH-50 calvarium. Journal of Human Evolution 39, 1–22. Hayden, B. 1987. Alliances and ritual ecstasy: human responses to resource stress. Journal for the Scientific Study of Religion 26, 81–91. Hayden, B. 1990. The right rub. Hide working in high ranking households. In: B. Gräslund et al. (eds.), The interpretative possibilities of microwear studies. Uppsala: Societas Archaeologica Upsaliensis, 89–102. Head, L. 1994. Landscapes socialised by fire: post-contact changes in Aboriginal fire use in Northern Australia, and implications for prehistory. Archaeology in Oceania 29, 172–181. Head, L. 1996. Rethinking the prehistory of hunter-gatherers, fire and vegetation change in Northern Australia. The Holocene 6, 4, 481–487. Heaney, L. R. 1985. Zoogeographic evidence for middle and late Pleistocene land bridges to the Philippine islands. Modern Quaternary Research in Southeast Asia 9, 127–144. Heaney, L. R. 1986. Biogeography of mammals in S. E. Asia. Biological Journal of the Linnaean Society 28, 127–165.

284

References Heaney, L. R. 1991. A synopsis of climate and vegetational change in Southeast Asia. Climate Change 19, 53–61. Heaney, L. R. 1993. Biodiversity patterns and the conservation of mammals of the Philippines. Asia Life Sciences 2, 2, 261–274. Heaney, L. R. 2002. Island life along Wallace’s Line: biogeography and patterns of endemism in the Philippines and Indonesia. In: E. Wikramanakye et al. (eds.), Terrestrial ecoregions of the IndoPacific: A conservation assessment. Washington, D.C.: Island Press, 28–30. Heaney, L. R., P. J. Piper & A. S. B. Mijares 2011. The first fossil record of endemic murid rodents from the Philippines: A Late Pleistocene cave fauna from northern Luzon. Proceedings of the Biological Society of Washington 124, 3, 234–247. Heekeren, H. R. van 1948. Stone age discoveries in Siam. Proceedings of the Prehistoric Society 14, 24–32. Heekeren, H. R. van 1972. The stone age of Indonesia. 2nd ed.. The Hague: Martinus Nijhoff. Heimann, J. 1997. The most offending soul alive: Tom Harrisson and his remarkable life. Honolulu: University of Hawai’i Press. Heinsohn, T. E. 2001. Human influences on vertebrate zoogeography: animal translocation and biological invasions across and to the east of Wallace’s Line. In: I. Metcalfe et al. (eds.), Faunal and floral migration and evolution in SE Asia-Australia. Lisse: Swets & Zeitlinger, 153–170. Henrich, J. 2004. Demography and cultural evolution: how adaptive cultural processes can produce maladaptive losses – the Tasmanian case. American Antiquity 69, 2, 197–214. Henry, A. G., A. S. Brooks & D. R. Piperno 2011. Microfossils in calculus demonstrate consumption of plants and cooked foods in Neanderthal diets (Shanidar III, Iraq; Spy I and II, Belgium). PNAS 108, 486–491. Henshilwood, C. S. 2007. Fully symbolic sapiens behaviour: innovations in the Middle Stone Age at Blombos Cave, South Africa. In: P. Mellars et al. (eds.), Rethinking the human revolution: new behavioural and biological perspectives on the origin and dispersal of modern humans. Cambridge: McDonald Institute for Archaeological Research, 123–132. Henshilwood, C. S. 2008. Winds of change: palaeoenvironments, material culture and human behaviour in the Late Pleistocene (~77 ka–48 ka ago) in the Western Cape Province, South Africa. South African Archaeological Society Goodwin Series 10, 35–51. Henshilwood, C. S. & C. W. Marean 2003. The origin of modern human behavior: a critique of the models and their test implications. Current Anthropology 44, 5, 627–651. Henshilwood, C. S., F. d’Errico & I. Watts 2009. Engraved ochres from the Middle Stone Age levels at Blombos Cave, South Africa. Journal of Human Evolution 57, 1, 27–47. Henshilwood C. S. et al. 2002. Emergence of modern human behavior: Middle Stone Age engravings from South Africa. Science 295, 1278–1280. Henshilwood, C. S. et al. 2004. Middle Stone Age shell beads from South Africa. Science 304, 404. Henshilwood, C. S. et  al. 2011. A 100,000-year-old ochre-processing workshop at Blombos Cave, South Africa. Science 334, 219–222. Hewlett, B. S. et al. 2011. Social learning among Congo Basin hunter-gatherers. Philosophical Transactions of the Royal Society B, Biological Sciences 366, 1168–1178. Higham, T. F. G. et al. 2009. Radiocarbon dating of charcoal from tropical sequences: results from the Niah Great Cave, Sarawak, and their broader implications. Journal of Quaternary Science 24, 2, 189–197. Higham, T. et al. 2010. Chronology of the Grotte du Renne (France) and implications for the context of ornaments and human remains within the Châtelperronian. PNAS 107, 47, 20234–20239. Hilbert, Y., J. Rose & R. Roberts 2012. Late Palaeolithic core-reduction strategies in Dhofar, Oman. Proceedings of the Seminar for Arabian Studies 42, 101–118. Hill, K. R. et al. 2011. Co-residence patterns in hunter-gatherer societies show unique human social structure. Science 331, 1286–1289. Hillgruber, F. 2007. Das mittelpaläolithische Werkzeugspektrum der Fundstelle Neandertal. Archäologisches Korrespondenzblatt 37, 335–346. Hiscock, P. 1994. Technological responses to risk in Holocene Australia. Journal of World Prehistory 8, 3, 267–292. Hiscock, P. 2002. Pattern and context in the Holocene proliferation of backed blades in Australia. In: R. G. Elston & S. L. Kuhn (eds.), Thinking small: global perspectives on micro-lithization. Archaeological Papers of the American Anthropological Association (AP3A) 12. Arlington, VA: American Anthropological Association, 163–177.

285

References Hiscock, P. 2005. Artefacts on Aru: Evaluating the technological sequences. In: S. O’Connor, M. Spriggs & P. Veth (eds.), The archaeology of the Aru Islands, Eastern Indonesia. Terra Australis 23. Canberra: Pandanus Books, 205–254. Hiscock, P. 2006. Blunt and to the point: changing technological strategies in Holocene Australia. In: I. Lilley (ed.), Archaeology of Oceania: Australia and the Pacific Islands. Oxford: Blackwell, 69–95. Hiscock, P. 2008. Archaeology of ancient Australia. London/New York: Routledge, Taylor & Francis Group. Hiscock, P. 2009. Prehistory of ancient Australia. London: Routledge. Hiscock, P. & V. Attenbrow 1998. Early Holocene backed blades from Australia. Archaeology in Oceania 33, 49–63. Hiscock, P. & H. Allen 2000. Assemblage variability in the Willandra Lakes. Archaeology in Oceania 35, 97–103. Hiscock, P. & S. O’Connor 2005. Arid paradises or dangerous landscapes: a review of explanations for Paleolithic assemblage change in Arid Australia and Africa. In: P.Veth et al. (eds.), Desert peoples: archaeological perspectives. Oxford: Blackwell, 58–77. Hiscock, P. & L. Wallis 2005. Pleistocene settlement of deserts from an Australian perspective. In: P.Veth et al. (eds.), Desert peoples: archaeological perspectives. Oxford: Blackwell, 34–57. Hiscock, P. & S. O’Connor 2006. An Australian perspective on modern behaviour and artefact assemblages. Before Farming 2, 1–10. Hiscock, P., C. Clarkson & A. Mackay 2011. Big debates over little tools: ongoing disputes over microliths on three continents. World Archaeology 43, 4, 660–671. Hoáng Xuân Chinh 1991. Faunal and cultural changes from Pleistocene to Holocene in Vietnam. Bulletin of the Indo-Pacific Prehistory Association 10, 74–78. Holcombe, C. 2011. A history of East Asia. Cambridge: Cambridge University Press. Holdaway, S. 2004. Continuity and change: an investigation of the flaked stone artefacts from the Pleistocene deposits at Bone Cave, Southwest Tasmania, Australia. Melbourne: Archaeology Program, School of Historical and European Studies, La Trobe University. Holdaway, S. J. & N. Porch 1995. Cyclical patterns in the Pleistocene human occupation of south-west Tasmania. Archaeology in Oceania 30, 74–82. Holdaway, S. J. & N. Porch 1996. Dates as data: an alternative approach to the construction of chronologies for Pleistocene sites in southwest Tasmania. In: J. Allen (ed.), Report of the Southern Forests Archaeological Project: site descriptions, stratigraphies and chronologies. Melbourne: La Trobe University, 251–277. Holdaway, S. & R. Cosgrove 1997. Archaeological attributes of behaviour: difference or variability? Endeavour 21, 2, 66–71. Holdaway, S. J. & N. Stern 2004. A record in stone. The study of Australia’s flaked stone artefacts. Canberra/ Melbourne: Aboriginal Studies Press/Museum Victoria. Holdaway, S. J. & M. Douglass 2012. A twenty-first century archaeology of stone artefacts. Journal of Archaeological Method and Theory 19, 1, 101–131. Hooijer, D. A. 1963. Further ‘Hell’ mammals from Niah. Sarawak Museum Journal 12 (n.s. 21–22), 196–200. Hope, J. 1993. Pleistocene archaeological sites in the central Murray-Darling Basin. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene archaeology in Australia, New Guinea and Island Melanesia. Canberra: ANU, 183–196. Hope, G. 2004. History of vegetation and habitat change in the Austral-Asian region. Quaternary International 118–119, 103–126. Hope, G. 2009. Environmental change and fire in the Owen Stanley Ranges, Papua New Guinea. Quaternary Science Reviews 28, 23–24, 2261–2276. Hope, G. et  al. 2004. History of vegetation and habitat change in the Austral-Asian region. Quaternary International 118–119, 103–112. Hou, Y. et  al. 2000. Mid-Pleistocene Acheulean-Iike stone technology of the Bose Basin, South China. Science 287, 1622–1666. Hovers, E. 2009. The lithic assemblages of Qafzeh. Oxford: Oxford University Press. Howell, F. C. 1999. Paleo-demes, species clades, and extinctions in the Pleistocene hominin record. Journal of Anthropological Research 55, 191–243. Hrdlicˇka, A. 1928. Catalogue of human crania in the United States National Museum collections. Proceedings of the National Museum of Science 71, 24, 1–140. Huang, W. 1989. Bifaces in China. Human Evolution 4, 1, 87–92.

286

References Huber, C. et al. 2006. Isotope calibrated Greenland temperature record over Marine Isotope Stage 3 and its relation to CH4. Earth and Planetary Science Letters 243, 504–519. Hublin, J.-J. 2009. The origin of Neandertals. PNAS 106, 16022–16027. Hublin, J.-J. & W. Roebroeks 2009. Ebb and flow or regional extinctions? On the character of Neandertal occupation of northern environments. C.R. Palévol 8, 503–509. Hudjashov, G. et al. 2007. Revealing the prehistoric settlement of Australia by Y chromosome and mtDNA analysis. PNAS 104, 21, 8726–8730. Hughes, I. 1977. New Guinea stone axe trade: the geography and ecology of traffic in the interior. Canberra: ANU. Hunt, C. O. & G. M. Coles 1988. The application of palynofacies analysis to geoarchaeology. In: E. A. Slater & J. O. Tate (eds.), Science and archaeology. Oxford: Archaeopress, 473–484. Hunt, C. O. & R. Premathilake 2012. Early Holocene vegetation and climate history from Sarawak, Malaysian Borneo. Quaternary International 249, 105–119. Hunt, C. O. & G. Rushworth 2005a. Airfall sedimentation and pollen taphonomy in the West Mouth of the Great Cave, Niah. Journal of Archaeological Science 32, 465–473. Hunt, C. O. & G. Rushworth 2005b. Cultivation and human impact at 6000 cal yr BP in tropical lowland forest at Niah, Sarawak, Malaysian Borneo. Quaternary Research 64, 460–468. Hunt, C. O., D. D. Gilbertson & G. Rushworth 2007. Modern humans in Sarawak, Malaysian Borneo, during Oxygen Isotope Stage 3: palaeoenvironmental evidence from the Great Cave of Niah. Journal of Archaeological Science 34, 1953–1969. Hunt, C. O., D. Gilbertson & G. Rushworth 2012. A 50,000-year record of late Pleistocene tropical vegetation and human impact in lowland Borneo. Quaternary Science Reviews 37, 61–80. Hutchison, C. S. 2005. Geology of North-West Borneo: Sarawak, Brunei and Sabah. Amsterdam: Elsevier Science. Hutterer, K. 1977. Reinterpreting the Southeast Asian Paleolithic. In: J. Allen et al. (eds.), Sunda and Sahul: Prehistoric studies in Southeast Asia, Melanesia and Australia. New York: Academic Press, 31–71. Hynes, R. A. & A. K. Chase 1982. Plants, sites and domiculture: Aboriginal influence upon plant communities in Cape York Peninsula. Archaeology in Oceania 17, 1, 38–50. Indriati, E. et al. 2011. The age of the 20 Meter Solo River Terrace, Java, Indonesia and the survival of Homo erectus in Asia. PLoS ONE 6, 1–10, e21562. Ingman, M. et al. 2000. Mitochondrial genome variation and the origin of modern humans. Nature 408, 708–713. Ingold, T. 2000. The Perception of the Environment. Essays in Livelihood, Dwelling and Skill. London: Routledge. Ingold, T. 2004. Beyond biology and culture. The meaning of evolution in a relational world. Social Anthropology 12, 2, 209–221. Inizan, M. L. & L. Ortlieb 1987. Préhistoire dans la région de Shabwa au Yemen du Sud (R. D. P. Yemen). Paléorient 13, 5–22. Irwin, G. J. 1992. The prehistoric exploration and colonization of the Pacific. Cambridge: Cambridge University Press. Jacobs, Z. & R. G. Roberts 2008.Testing times: old and new chronologies for the Howieson’s Poort and Still Bay industries in environmental context. South African Archaeological Society Goodwin Series 10, 9–34. Jacobs, Z. et al. 2008. Ages for the Middle Stone Age of Southern Africa: implications for human behaviour and dispersals. Science 322, 733–735. Jacobs, Z. et al. 2011. Single-grain OSL chronologies for Middle Palaeolithic deposits at El Mnasra and El Harhoura 2, Morocco: Implications for Late Pleistocene human–environment interactions along the Atlantic coast of northwest Africa. Journal of Human Evolution 62, 3, 377–394. Jagher, R. 2009.The Central Oman Paleolithic Survey: Recent research in southern Arabia and reflection on the prehistoric evidence. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 139–150. Jain, M. et al. 2005. Quaternary alluvial stratigraphical development in a desert setting: a case study from the Luni River basin, Thar Desert of western India. Fluvial Sedimentology 7, 349–371. James, H. 2007.The emergence of modern human behavior in South Asia: A review of the current evidence and discussion of its possible implications. In: M. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia. Dordrecht: Springer, 201–227. James, H. 2011. Becoming human: The emergence of modern human behaviour in South Asia. Unpublished PhD Thesis, University of Cambridge, Cambridge.

287

References James, H. & M. D. Petraglia 2005. Modern human origins and the evolution of behavior in the later Pleistocene record of South Asia. Current Anthropology 46, S3–S27. James, H. & M. D. Petraglia 2009. The Lower to Middle Palaeolithic transition in South Asia and its implications for hominin cognition and dispersals. In: M. Camps & P. Chauhan (eds.), Sourcebook of Paleolithic transitions. Dordrecht: Springer, 255–264. Jansa, S.A., F. K. Barker & L. R. Heaney 2006.The pattern and timing of diversification of Philippine endemic rodents: evidence from mitochondrial and nuclear gene sequences. Systematic Biology 55, 1, 73–88. Jin, C. et al. 2009. The Homo sapiens cave hominin site of Mulan Mountain, Jiangzhou District, Chongzuo, Guangxi with emphasis on its age. Chinese Science Bulletin 54, 3848–3856. Jochim, M. A. 1983. Palaeolithic cave art in ecological perspective. In: G. Bailey (ed.), Hunter gatherer economy in prehistory. Cambridge: Cambridge University Press, 212–219. Johns, T. & I. Kubo 1988. A survey of traditional methods employed for the detoxification of plant foods. Journal of Ethnobiology 8, 1, 81–129. Johnson, C. R. & S. McBrearty 2010. 500,000 Year old blades from the Kapthurin Formation, Kenya. Journal of Human Evolution 58, 2, 193–200. Johnson, K. R. et al. 2006. East Asian summer monsoon variability during Marine Isotope Stage 5 based on speleothem δ18O records from Wanxiang Cave, central China. Palaeogeography Palaeoclimatology Palaeoecology 236, 5–19. Johnston, H. 1993. Pleistocene shell middens of the Willandra Lakes. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene archaeology in Australia, New Guinea and Island Melanesia. Canberra: ANU, 197–203. Jones, R. 1964–1965. Archaeological reconnaissance in Tasmania, summer 1963/1964. Oceania 35, 191–201. Jones, R. 1966. A speculative archaeological sequence for northwest Tasmania. Records of the Queen Victoria Museum 25, 1–12. Jones, R. 1968.The geographical background to the arrival of man in Australia and Tasmania. Archaeology and Physical Anthropology in Oceania 3, 186–215. Jones, R. 1971. Rocky Cape and the Problem of the Tasmanians. Unpublished PhD thesis, University of Sydney. Jones, R. 1973. Emerging picture of Pleistocene Australians. Nature 246, 78–81. Jones, R. 1977. The Tasmanian paradox. In: R.V. S. Wright (ed.), Stone tools as cultural markers: change, evolution and complexity Canberra: A.I.A.S., 189–204. Jones, R. 1979.The fifth continent: problems concerning the human colonization of Australia. Annual Review of Anthropology 8, 445–466. Jones, R. 1985. Archaeological conclusions. In: R. Jones (ed.), Archaeological research in Kakadu National Park. Special Publication 13. Canberra: Australian National Parks and Wildlife Service, 291–298. Jones, R. 1990. From Kakadu to Kutikina. The southern continent at 18 000 years ago. In: C. Gamble & O. Soffer (eds.), The world at 18 000 BP.Volume 2. Low latitudes. London: Unwin Hyman, 264–295. Jones, R. 1995. Tasmanian archaeology: establishing the sequences. Annual Review of Anthropology 24, 423–446. Jones, S. 2007. The Toba supervolcanic eruption: tephra-fall deposits in India and palaeoanthropological implications. In: M. D. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia. Hamburg: Springer, 173–200. Jones, R. & I. Johnson 1985. Deaf Adder Gorge: Lindner Site Nauwalabila. In: R. Jones (ed.), Archaeological research in Kakadu National Park, Special Publication 13. Canberra: Australian National Parks and Wildlife Service, 165–227. Jones, R. et al. 1988. An archaeological reconnaissance of karst caves within the Southern Forests region of Tasmania, September 1987. Australian Archaeology 26, 1–23. Juel Jensen, H. 1988. Functional analysis of prehistoric flint tools by high power microscopy: a review of West European literature. Journal of World Prehistory 2, 1, 53–88. Juyal, N. et al. 2006. Continental record of the southwest monsoon during the last 130 ka: evidence from the southern margin of the Thar Desert, India. Quaternary Science Reviews 25, 19–20, 2632–2650. Kaars, S. van der & M. A. C. Dam 1995. A 135,000-year record of vegetational and climatic change from the Bandung area, West-Java, Indonesia. Palaeogeography, Palaeoclimatology, Palaeoecology 117, 55–72. Kaars, S. van der et al. 2000. A Late Quaternary palaeoecological record from the Banda Sea, Indonesia: patterns of vegetation, climate and biomass burning in Indonesia and northern Australia. Palaeogeography, Palaeoclimatology, Palaeoecology 155, 135–153. Kaifu, Y. & M. Fujita 2012. Fossil record of early modern humans in East Asia. Quaternary International 248, 2–11.

288

References Kaifu,Y. et al. 2011. Craniofacial morphology of Homo floresiensis: description, taxonomic affinities and evolutionary implications. Journal of Human Evolution 61, 644–682. Kamminga, J. 1979. The nature of use-polish and abrasive smoothing on stone tools. In: B. Hayden (ed.), Lithic use wear analysis. New York: Academic Press, 143–157. Kamminga, J. 1982. Over the edge. Functional analysis of Australian stone tools. St. Lucia: University of Queensland Press. Kamminga, J. 2007. Reinterpretation of the Hoabinhian at Sai Yok, Thailand. In: D. P. Zuraina Majid & M. Saidin (eds.), Archaeological heritage of Malaysia No. 1. Kuala Lumpur: Universiti Sains Malaysia, Centre for Archaeological Research in Malaysia, 38–58. Karafet, T. M. et al. 2005. Balinese Y-chromosome perspective on the peopling of Indonesia: genetic contributions from Pre-Neolithic Hunter-Gatherers, Austronesian farmers, and Indian traders. Human Biology 77, 1, 93–114. Keates, S. G. 1998. A discussion of the evidence for early hominids on Java and Flores. Modern Quaternary Research in Southeast Asia 15, 171–191. Keeley, L. H. 1974. Technique and methodology in microwear studies: a critical review. World Archaeology 5, 323–336. Keeley, L. H. 1980. Experimental determination of stone tool uses. Chicago: University of Chicago Press. Keeley, L. H. 1982. Hafting and retooling: effects in the archaeological record. American Antiquity 47, 798–809. Keen, I. 2004. Aboriginal economy and society. Australia at the threshold of colonisation. Melbourne: Oxford University Press. Keen, I. 2006. Constraints on the development of enduring inequalities in Late Holocene Australia. Current Anthropology 47, 1, 7–38. Kefous, K. 1977. We have a fish with ears, and wonder if it’s valuable. Unpublished honours thesis, ANU, Canberra. Keinan, A. & D. Reich 2010. Can a sex-biased human demography account for the reduced effective population size of chromosome x in non-Africans? Molecular Biology and Evolution 27, 2312–2321. Kellock,W. L. & P. A. Parsons 1970.Variation of minor non-metrical cranial variants in Australian Aborigines. American Journal of Physical Anthropology 32, 409–422. Kelly, R. L. 1995. The foraging spectrum. diversity in hunter-gatherer lifeways. Washington, D.C.: Smithsonian Institution. Kennedy, J. 2002. Manus from the beginning: an archaeological overview. In: C. Kaufmann et  al. (eds.), Admiralty art: art from the South Seas. Zürich: Museum Rietberg, 17–28. Kennedy, K. A. R. 1979. The deep skull of Niah: an assessment of twenty years of speculation concerning its evolutionary significance. Asian Perspectives 20, 32–50. Kennedy, K. A. R. 2000. God-apes and fossil men: palaeoanthropology in South Asia. Ann Arbor: University of Michigan Press. Kennedy, K. A. R. & S. U. Deraniyagala 1989. Fossil remains of 28,000-year-old hominids from Sri Lanka. Current Anthropology 30, 3, 394–399. Kennedy, K. A. R. et al. 1991. Is the Narmada hominid an Indian Homo erectus? American Journal of Physical Anthropology 86, 4, 475–496. Kent, P. E. 1942. The Pleistocene beds of Kanam and Kanjera, Karivondo, Kenya. Geological Magazine 79, 117–132. Kerwin, D. 2010. Aboriginal Dreaming paths and trading routes: the colonisation of the Australian economic landscape. Brighton: Sussex Academic Press. King,V. T. 2007. Iban art, sexual selection and severed heads: weaving, sculpture, tattooing and other arts of the Iban of Borneo. Bijdragen Tot de Taal- Land- en Volkenkunde 163, 1, 177–180. Kirkpatrick, J. B. & M. Fowler 1998. Locating likely glacial forest refugia in Tasmania using palynological and ecological information to test alternative climatic models. Biological Conservation 85, 171–182. Kivisild,T. et al. 1999. Deep common ancestry of Indian and western-Eurasian mitochondrial DNA lineages. Current Biology 9, 22, 1331–1334. Kivisild, T. et al. 2003. The genetic heritage of the earliest settlers persists both in Indian tribal and caste populations. American Journal of Human Genetics 72, 2, 313–332. Kivisild, T. et al. 2004. Ethiopian mitochondrial DNA heritage: tracking gene flow across and around the gate of tears. American Journal of Human Genetics 75, 752–770.

289

References Klein, R. G. 1992. The archeology of modern human origins. Evolutionary Anthropology 1, 1, 5–14. Klein, R. G. 1995. Anatomy, behaviour and modern human origins. Journal of World Prehistory 9, 167–198. Klein, R. G. 2000. Archaeology and the evolution of human behavior. Evolutionary Anthropology 9, 12–36. Klein, R. G. 2008. Out of Africa and the evolution of modern behaviour. Evolutionary Anthropology 17, 267–281. Klein, R. G. 2009. The human career: human biological and cultural origins.3rd edition. Chicago: University Chicago Press. Klein, R. G & B. Edgar 2002. The dawn of human culture. A bold new theory on what sparked the ‘Big Bang’ of human consciousness. New York: John Wiley & Sons. Koenigswald, G. H. R. von 1933. Ein neuer Urmensch aus dem Diluvium Javas. Centralblatt für Mineralogie, Geologie und Paläontologie (B) 1933, 29–42. Koenigswald, G. H. R. von 1936. Early Palaeolithic stone implements from Java. Bulletin of the Raffles Museum (Singapore) B1, 52–60. Koenigswald, G. H. R. von 1958. Preliminary report on a newly discovered Stone Age culture from Northern Luzon, Philippine Islands. Asian Perspectives 2, 69–71. Koenigswald, G. H. R. von & A. K. Ghosh 1973. Stone implements from the Trinil beds of Sangiran, Central Java. Proceedings of the Koninklijke Nederlandse Akademie van Wetanschappen B76, 1, 1–17. Koenigswald, G. H. R. & F. Weidenreich 1939. The relationship between Pithecanthropus and Sinanthropus. Nature 144, 926–929. Koller, J., U. Baumer & D. Mania 2001. High-tech in the Middle Palaeolithic: Neandertal-manfactured pitch identified. European Journal of Archaeology 4, 385–397. Korisettar, R. 2007. Toward developing a basin model for Paleolithic settlement of the Indian subcontinent: geodynamics, monsoon dynamics, habitat diversity and dispersal routes. In: M. D. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia. Dordrecht: Springer, 69–96. Krause, J. et al. 2007a. The derived FOXP2 variant of modern humans was shared with Neandertals. Current Biology 17, 21, 1908–1912. Krause, J. et al. 2007b. Neanderthals in central Asia and Siberia. Nature 449, 902–904. Krause, J. et al. 2010. The complete mitochondrial DNA genome of an unknown hominin from southern Siberia. Nature 464, 894–897. Kress, J. 1979. Tom Harrison, North Borneo, and Palawan: a preliminary assessment. Asian Perspectives 20, 1, 75–86. Kress, J. 1980. The ceramics from Pilanduk Cave and Sa’gung rockshelter, Quezon Municipality, Palawan Island, the Philippines. Asian Perspectives 21, 1, 58–85. Krigbaum, J. 2005. Reconstructing human subsistence in the West Mouth (Niah Cave, Sarawak) burial series using stable isotopes of carbon. Asian Perspectives 44, 1, 73–89. Krigbaum, J. & I. Datan 1999. The deep skull of Niah. Borneo 5, 1, 13–17. Krigbaum, J. & I. Datan 2005. The Deep Skull and associated human remains from Niah Cave. In: D. P. Zuraina Majid (ed.), The Perak Man and other prehistoric skeletons of Malaysia. Pulau Pinag: Penerbit Universiti Sains Malaysia, 131–154. Krogman, W. M. 1932. The morphological characters of the Australian skull. Journal of Anatomy 66, 283–231. Kuhn, S. L. & M. C. Stiner 2007. Body ornamentation as information technology: towards an understanding of the significance of early beads. In: P. Mellars et al. (eds.), Rethinking the human revolution. Cambridge: McDonald Institute for Archaeological Research, 45–54. Kuhn, S. L. et al. 2001. Ornaments of the earliest Upper Paleolithic: new insights from the Levant. PNAS 98, 7641–7646. Kuhn, S. L. et al. 2009. The Early Upper Paleolithic occupations at Üçag˘ ızlı Cave (Hatay, Turkey). Journal of Human Evolution 56, 87–113. Kuhnt,W. et al. 2004. Neogene history of the Indonesian Throughflow. In: P. Clift et al. (eds.), Continent-ocean interactions in the East Asian marginal seas. Washington, D.C.: American Geophysical Union, 299–320. Kuzmin,Y.V. et al. 2004. AMS 14C age of the Upper Palaeolithic skeletons from Sungir site, Central Russian Plain. Nuclear Instruments and Methods in Physics Research B 223–224, 731–734. Lahr, M. M. 1996. The evolution of modern human diversity: a study of cranial variation. Cambridge: Cambridge University Press. Lahr, M. M. & R. Foley 1994. Multiple dispersals and modern human origins. Evolutionary Anthropology 3, 48–60. Lahr, M. M. & R. A. Foley 1998. Towards a theory of modern humans origins: geography, demography, and diversity in recent human evolution. Yearbook of Physical Anthropology 41, 37–176.

290

References Lambeck, K. & J. Chappell 2001. Sea level change through the last glacial cycle. Science 292, 679–686. Lambeck, K., T. M. Esat & E.-K. Potter 2002. Links between climate and sea levels for the past three million years. Nature 419, 199–206. Lambeck, K. et al. 2011. Sea level and shoreline reconstructions for the Red Sea: isostatic and tectonic considerations and implications for hominin migration out of Africa. Quaternary Science Reviews 30, 3542–3574. Lampert, R. & G. Turnbull 1970. The manufacture of shell fish hooks on the south coast of N.S.W. Mankind 7, 4, 312–313. Landau, M. 1984. Human evolution as narrative. American Scientist 72, 262–268. Landau, M. 1993. Narratives of human evolution. New Haven:Yale University Press. Langley, M. 2009. Material culture and behaviour in Pleistocene Sahul: examining the archaeological representation of Pleistocene behavioural modernity in Sahul. Unpublished MPhil Thesis, School of Social Science, University of Queensland. Langley, M. C. & P. S. C. Tac˛on 2010. The age of Australian rock art: a review. Australian Archaeology 71, 70–73. Langley, M. C., C. Clarkson & S. Ulm 2008. Behavioural complexity in Eurasian Neanderthal populations: a chronological examination of the archaeological evidence. Cambridge Archaeological Journal 18, 289–307. Langley, M. C., C. Clarkson & 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. Larnach, S. L. & N. W. G. Macintosh 1966. The craniology of the Aborigines of coastal New South Wales. Sydney: University of Sydney. Larnach, S. L. & N. W. G. Macintosh 1970. The craniology of the Aborigines of Queensland. Sydney: University of Sydney. Latz, P. 1995. Bushfires and bushtucker. Alice Springs: Institute of Aboriginal Development. Layton, R. 1992. Australian rock art. A new synthesis. Melbourne: Cambridge University Press. Leakey, L. S. B. 1934. Adam’s ancestors. London: Methuen. Leakey, L. S. B. 1951. Olduvai Gorge. Cambridge: Cambridge University Press. Leavesley, M. G. 2004. Trees to the sky: prehistoric hunting in New Ireland, Papua New Guinea. Unpublished PhD thesis, Australian National University, Canberra. Leavesley, M. 2005. Prehistoric hunting strategies in New Ireland, Papua New Guinea: the evidence of the Cuscus (Phalanger orientalis) remains from Buang Merabak cave. Asian Perspectives 44, 207–218. Leavesley, M. 2006. Late Pleistocene complexities in the Bismarck Archipelago. In: I. Lilley (ed.), Archaeoogy of Oceania: Australia and the Pacific Islands. Carlton: Blackwell, 189–204. Leavesley, M. G. 2007. A shark-tooth ornament from Pleistocene Sahul. Antiquity 8, 308–315. Leavesley, M. G. & J. Allen 1998. Dates, disturbance and artefact distributions: another analysis of Buang Merabak, a Pleistocene site on New Ireland, Papua New Guinea. Archaeology in Oceania 33, 2, 63–82. Leavesley, M. G. & J. Chappell 2004. Buang Merabak: additional early radiocarbon evidence of the colonisation of the Bismarck Archipelago, Papua New Guinea. Antiquity 78, 301 (project gallery at http:// antiquity.ac.uk/ProjGall/leavesley/index.html; accessed 30 October 2011). Leavesley, M. G. et  al. 2002. Buang Merabak: early evidence for human occupation in the Bismarck Archipelago, Papua New Guinea. Australian Archaeology 54, 55–57. Lee L. 2008. Quantitative palaeozoology. Cambridge: Cambridge University Press. Leinders, J. J. M. et al. 1985. The age of the hominid-bearing deposits of Java: State of the art. Geologie en Mijnbouw 64, 167–173. Leng Jian & C. L. Shannon 2000. Rethinking early paleolithic typologies in China and India. Journal of East Asian Archaeology 2, 1–2, 9–35. Lentfer, C., C. Pavlides & J. Specht 2010. Natural and human impacts in a 35,000-year vegetation history in central New Britain, Papua New Guinea. Quaternary Science Reviews 29, 3750–3767. Leplongeon, A. & D. Pleurdeau 2011.The Upper Palaeolithic lithic industry of Nazlet Khater 4 (Egypt): implications for the Stone Age/Palaeolithic of Northeastern Africa. African Archaeological Review 28, 3, 213–236. Leuschner, D. C. & F. Sirocko 2003. Orbital insolation forcing of the Indian Monsoon – a motor for global climate changes? Palaeogeography, Palaeoclimatology, Palaeoecology 197, 1–2, 83–95. Levi-Sala, I. 1996. A study of microscopic polish on flint implements. British Archaeological Reports Series 629. Oxford: British Archaeological Reports. Lewens, T. 2009. Seven types of adaptationism. Biology and Philosophy 24, 161–182. Lewis, H. et al. 2008.Terminal Pleistocene to mid-Holocene occupation and an early cremation burial at Ille Cave, Philippines. Antiquity 82, 316, 318–335.

291

References Lindly, J. M. & G. A. Clark 1990. Symbolism and modern human origins. Current Anthropology 31, 233–261. Liu, H. et al. 2006. A geographically explicit genetic model of worldwide human-settlement history. American Journal of Human Genetics 79, 230–237. Liu,W. et al. 2010a. Huanglong Cave: a late Pleistocene human fossil site in Hubei Province, China. Quaternary International 211, 29–41. Liu,W. et al. 2010b. Human remains from Zhirendong, South China, and modern human emergence in East Asia. PNAS 107, 19201–19206. Lloyd-Smith, L. 2009. Chronologies of the dead: later prehistoric burial practice at the Niah Caves, Sarawak. Unpublished PhD dissertation, University of Cambridge, Cambridge. Lombard, M. 2005a. Evidence of hunting and hafting during the Middle Stone Age at Sibudu Cava, KwaZulu-Natal, South Africa: a multianalytical approach. Journal of Human Evolution 48, 279–300. Lombard, M. 2005b. A method for identifying Stone Age hunting tools. South African Archaeological Bulletin 60, 182, 115–120. Lombard, M. & M. N. Haidle 2012. Thinking a bow-and-arrow set: cognitive implications of Middle Stone Age bow and stone-tipped arrow technology. Cambridge Archaeological Journal 22, 2, 237–264. Lombard, M. & J. Pargeter 2008. Hunting with Howiesons Poort segments: pilot experimental study and the functional interpretation of archaeological tools. Journal of Archaeological Science 35, 2523–2531. Long,V.T., J. de Vos & R. S. Ciochon 1996.The fossil mammal fauna of the Lang Trang caves,Vietnam, compared with Southeast Asian fossil and recent mammal faunas: the geographical implications. Bulletin of the Indo-Pacific Prehistory Association 14, 101–109. Lorblanchet, M. & R. Jones 1979. Les premieres fouilles à Dampier (Australié occidentale), et leur place dans l’ensemble Australien. Bulletin de la Société Préhistorique Française 76, 463–487. Lothrup, S. M. 1928. The Indians of Tierra del Fuego. New York: Museum of the American Indian, Heye Foundation (reprint: Buenos Aires: Zagier & Urrity, 2002). Lourandos, H. 1980. Change or stability? Hydraulics, hunter-gatherers and population in temperate Australia. World Archaeology 11, 3, 245–264. Lourandos, H. 1997. Continent of hunter-gatherers: new perspectives in Australian archaeology. Cambridge: Cambridge University Press. Lourandos, H. 2008. Constructing ‘hunter-gatherers’, constructing ‘prehistory’. Australian Archaeology 67, 69–78. Louys, J. & A. Turner 2012. Environment, preferred habitats and potential refugia for Pleistocene Homo in Southeast Asia. Comptes Rendues Palévolution 11, 203–211. Lovell, J. 2006. The Great Wall: China against the world, 1000 BC–2000 AD. London: Atlantic Books. Loy, T., M. Spriggs & S. Wickler 1992. Direct evidence for human use of plants 28,000 years ago: starch residues on stone tools from the northern Solomon Islands. Antiquity 66, 898–912. Luis, J. R. et al. 2004. The Levant versus the Horn of Africa: evidence for bidirectional corridors of human migrations. American Journal of Human Genetics 74, 532–544. Lukacs, J. R. 2007. Interpreting biological diversity in South Asian prehistory: early Holocene population affinities and subsistence adaptations. In: M. D. Petraglia & B. Allchin (eds.), The evolution and history of human populations in South Asia. Dordrecht: Springer, 271–296. Lycett, S. & C. Bae 2010. The Movius Line controversy: the state of the debate. World Archaeology 42, 521–544. Lycett, S. & C. Norton 2010. A demographic model for Palaeolithic technological evolution:The case of East Asia and the Movius Line. Quaternary International 211, 55–65. Lycett, S. J., N. von Cramon-Taubadel & R. A. Foley 2006. A crossbeam co-ordinate caliper for the morphometric analysis of lithic nuclei: a description, test and empirical examples of application. Journal of Archaeological Science 33, 6, 847–861. MacArthur, R. H. & E. O. Wilson 1967. The theory of island biogeography. Princeton: Princeton University Press. Maca-Meyer, N. et al. 2001. Major genomic mitochondrial lineages delineate early human expansions. BMC Genetics 2, 1, 13. Macaulay,V. et al. 2005. Single, rapid coastal settlement of Asia revealed by analysis of complete mitochondrial genomes. Science 308, 1034–1036. MacDonald, H. 2005. Human remains: episodes in human dissection. Melbourne: Melbourne University Press. Macintosh, N. W. G. 1965. The physical aspect of man in Australia. In: R. M. Berndt & C. H. Berndt (eds.), Aboriginal man in Australia: essays in honour of Emeritus Professor A. P. Elkin. Sydney: Angus & Robertson, 29–70.

292

References Macintosh, N. W. G. 1967. Fossil man in Australia. Australian Journal of Science 4, 86–98. Macintosh, N.W. G. & B. C.W. Barker 1965. The osteology of Aboriginal man in Tasmania. University of Sydney: Sydney. McIntosh, P. D. et al. 2009. Late Quaternary erosion events in lowland and mid-altitude Tasmania in relation to climate change and first human arrival. Quaternary Science Reviews 28, 850–872. Mackay, A. & A. Welz 2008. Engraved ochre from a Middle Stone Age context at Klein Kliphuis in the Western Cape of South Africa. Journal of Archaeological Science 35, 6, 1521–1532. Mackintosh, A. N. et al. 2006. Exposure dating and glacial reconstruction at Mt. Field, Tasmania, Australia, identifies MIS 3 and MIS 2 glacial advances and climatic variability. Journal of Quaternary Science 21, 4, 363–376. Madulid, D. 1998. Bases of biodiversity. In: G. Magdaraog (ed.), Environmental and Natural resources atlas of the Philippines. Manila: SEAICE, 42–67. Maher, L. A. 2009. The Late Pleistocene of Arabia in relation to the Levant. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 187–202. Mahony, D. J. 1943a. The Keilor Skull: geological evidence of antiquity. Memoirs of the National Museum 13, 79–81. Mahony, D. J. 1943b. The problem of antiquity of man in Australia. Memoirs of the National Museum 13, 7–56. Maji, S., S. Krithika & T. S.Vasulu 2008. Distribution of mitochondrial DNA macrohaplogroup N in India with special reference to haplogroup R and its sub-haplogroup U. International Journal of Human Genetics 8, 1/2, 85. Malinowski, B. 1922. The argonauts of the Western Pacific. New York: Dutton. Manalo, K. 2011. Preliminary identification of cut mark morphology on animal bones: methods and applications. Unpublished Master’s Thesis, Archaeological Studies Program, University of the Philippines. Manica, A., F. Prugnolle & F. Balloux 2005. Geography is a better determinant of human genetic differentiation than ethnicity. Human Genetics 118, 3, 366–371. Marean, C. W. 1998. A critique of the evidence for scavenging by Neanderthals and early modern humans: New data from Koben Cave (Zagros Mousterian) and Die Kelders (South Africa Middle Stone Age). Journal of Human Evolution 35, 111–136. Marean, C. W. & S.Y. Kim 1998. Mousterian large mammal remains from Kobeh Cave. Current Anthropology, Supp. 39, 79–113. Marean, C. W. & Z. Assefa 1999. Zooarchaeology evidence for the faunal exploitation behavior of Neanderthals and early modern humans. Evolutionary Anthropology 8, 22–37. Marean, C. W. et al. 2007. Early human use of marine resources and pigment in South Africa during the Middle Pleistocene. Nature 44, 905–909. Marshall, B., 1992. Late Pleistocene human exploitation of the platypus in southern Tasmania. In: M. Augee (Ed.), Platypus and Echidnas. Royal Zoological Society of New South Wales, Sydney, 268–276. Marston, D. 2005. The Pacific War companion: from Pearl Harbour to Hiroshima. Oxford: Osprey Publishing. Martin, R. D. et al. 2006. Flores hominid: new species or microcephalic dwarf? Anatomical Record Part A: Discoveries in Molecular, Cellular, and Evolutionary Biology 288A, 11, 1123–1145. Martinón-Torres, M., R. Dennell & J. M. Bermudez de Castro 2010.The Denisova hominin need not be an African story. Journal of Human Evolution 60, 251–255. Marwick, B. 2009. Biogeography of Middle Pleistocene hominins in mainland Southeast Asia: a review of current evidence. Quaternary International 202, 51–58. Matisso-Smith, E. 2007. Animal translocations, genetic variation and the human settlement of the Pacific. In: J. Friedlander (ed.), Genes, language and culture history in the Southwest Pacific. Oxford: Oxford University Press, 10–35. Matsumua, H. & S. Pookajorn 2005. A morphometric analysis of the Late Pleistocene human skeleton from the Moh Khiew cave in Thailand. Journal of Comparative Human Biology 56, 93–118. Matthew, W. D. 1915. Climate and evolution. Annals of the New York Academy of Science 24, 171–318. Maynard, L. 1979. The archaeology of Australian Aboriginal art. In: S. M. Mead (ed.), Exploring the visual art of Oceania. Honolulu: University Press of Hawaii, 83–110. Mayr, E. 1963. The taxonomic evaluation of fossil hominids. In S. L. Washburn (ed.), Classification and human evolution. Chicago: Aldine, 332–346. Mazel, A. D. 1989. Changing social relations in the Thukela Basin, Natal 7000–2000 BP. South African Archaeological Society Goodwin Series 6, 33–41.

293

References Mazel, A. D. 2009. Unsettled times: shaded polychrome paintings and hunter-gatherer history in the southeastern mountains of southern Africa. Southern African Humanities 21, 85–115. Mazza, P. P. A. et al. 2006. A new Palaeolithic discovery: tar-hafted stone tools in a European Mid-Pleistocene bone-bearing bed. Journal of Archaeological Science 33, 1310–1318. McAdam, L. E. S. 2001. Boomerangs and drainage basins: variation in Aboriginal material culture in the region between Lake Eyre and the Gulf of Carpentaria. Armidale: University of New England. McAdam, L. E. S. 2008. Beads across Australia: an ethnographic and archaeological view of the patterning of Aboriginal ornaments. Armidale: University of New England. McArthur, N. 1976. Computer simulations of small populations. Australian Archaeology 4, 53–57. McBrearty, S. 1993. Lithic artifacts from Abu Dhabi’s western region. Tribulus: Bulletin of the Emirates Natural History Group 3, 13–14. McBrearty, S. 1999. Earliest stone tools from the Emirate of Abu Dhabi, United Arab Emirates. In: P. J. Whybrow & A. Hill (eds.), Fossil vertebrates of Arabia. New Haven:Yale University Press, 373–388. McBrearty, S. 2007. Down with the revolution. In: P. Mellars et al. (eds.), Rethinking the human revolution: new behavioural and biological perspectives on the origin and dispersal of modern humans. Cambridge: McDonald Institute for Archaeological Research, 133–151. McBrearty, S. & A. S. Brooks 2000. The revolution that wasn’t: a new interpretation of the origin of modern human behavior. Journal of Human Evolution 39, 5, 453–563. McBrearty, S. & C. Stringer 2007. The coast in colour. Nature 449, 793–794. McBryde, I. 1986. Artefacts, language and social interaction: a case study from South-Eastern Australia. In: G. N. Bailey & P. Callow (eds.), Stone-Age prehistory. Studies in memory of Charles McBurney. Cambridge: Cambridge University Press, 77–93. McBurney, C. B. M. 1950. The geographical study of the older Palaeolithic stages in Europe. Proceedings of the Prehistoric Society 11, 163–183. McClure, H. A. 1976. Radiocarbon chronology of Late Quaternary lakes in the Arabian Desert. Nature 263, 755–756. McClure, H. A. 1994. A new Arabian stone tool assemblage and notes on the Aterian industry of North Africa. Arabian Archaeology and Epigraphy 5, 1–16. McCormac, F. G. et al. 2004. SHCAL04 southern hemisphere calibration 0–11.0cal kyr BP. Radiocarbon 46, 1087–1092. McDermott, F. et al. 1993. Mass spectrometric dates for Israeli Neanderthal/early modern sites. Nature 363, 252–255. McDonald, J. 2005. Archaic faces to headdresses: the changing role of rock art across the arid zone. In: P.Veth et al. (eds.), Desert peoples: archaeological perspectives. Oxford: Blackwell, 116–142. McDonald, J. 2008. Dreamtime superhighway: Sydney Basin rock art and prehistoric information exchange. Canberra: ANU E Press. McDougall, I., F. H. Brown & J. G. Fleagle 2005. Stratigraphic placement and age of modern humans from Kibish, Ethiopia. Nature 433, 733–736. McEvoy, B. P. et al. 2011. Human population dispersal ‘Out of Africa’ estimated from linkage disequilibrium and allele frequencies of SNPs. Genome Research, doi:10.1101/gr.119636.110. McGowan, A. et al. 1993. Photogrammetric recording of Pleistocene cave paintings in southwest Tasmania. In: M. A. Smith et  al. (eds.), Sahul in review: Pleistocene archaeology in Australia, New Guinea and Island Melanesia. Canberra: ANU, 225–232. McIntosh P. D. et al. 2009. Late Quaternary erosion events in lowland and mid-altitude Tasmania in relation to climate change and first human arrival. Quaternary Science Reviews 28, 850–872. McLaren, S. J. et al. 2009. First evidence for episodic flooding events in the arid interior of central Saudi Arabia over the last 60 ka. Journal of Quaternary Science 24, 198–297. McNiven, I. J. 2006. Dauan 4 and the emergence of ethnographically-known social arrangements across the Torres Strait during the last 600–800 years. Australian Archaeology 62, 1–12. Mehlman, M. J. 1979. Mumba-Höhle revisited: the relevance of a forgotten excavation to some current issues in East African prehistory. World Archaeology 11, 1, 80–94. Meijaard, E. & C. P. Groves 2006. The geography of mammals and rivers in mainland Southeast Asia. In: S. H. Lehman & J. C. Fleagle (eds.), Primate biogeography. New York: Springer, 305–329. Meiri, S. et  al. 2008. Mammals of Borneo  – small size on a large island. Journal of Biogeography 35, 6, 1087–1094.

294

References Mellars, P. 1989a. Technological changes across the Middle-Upper Paleolithic transition: economic, social, and cognitive perspectives. In: P. Mellars & C. Stringer (eds.), The human revolution: behavioral and biological perspectives on the origins of modern humans. Princeton: Princeton University Press, 338–365. Mellars, P. 1989b. Major issues in the origin of modern humans. Current Anthropology 30, 349–385. Mellars, P. (ed.) 1990. The emergence of modern humans: an archaeological perspective. Ithaca: Cornell University Press. Mellars, P. 1995. Symbolism, language, and the Neanderthal mind. In: P. Mellars & K. Gibson (eds.), Modelling the human mind. Cambridge: MacDonald Institute for Archaeological Research, 15–32. Mellars, P. 1996. The Neanderthal legacy: an archaeological perspective from Western Europe. Princeton: Princeton University Press. Mellars, P. 2005. The impossible coincidence. A single-species model for the origins of modern human behaviour in Europe. Evolutionary Anthropology 14, 12–27. Mellars, P. 2006a. A new radiocarbon revolution and the dispersal of modern humans in Eurasia. Nature 439, 931–935. Mellars, P. 2006b. Going east: new genetic and archaeological perspectives on the modern human colonization of Eurasia. Science 313, 796–800. Mellars, P. 2006c. Why did modern human populations disperse from Africa ca. 60,000 years ago? A new model. PNAS 103, 25, 9381–9386. Mellars, P. 2010. Neanderthal symbolism and ornament manufacture: the bursting of a bubble? PNAS 107, 47, 20147–20148. Mellars, P. & C. Stringer (eds.) 1989a. The human revolution. Behavioural and biological perspectives in the origins of modern humans. Edinburgh: Edinburgh University Press. Mellars, P. & C. Stringer 1989b. Introduction. In P. Mellars and C. Stringer (eds.), The human revolution: behavioural and biological perspectives in the origins of modern humans. Edinburgh: Edinburgh University Press, 1–14. Mellars, P. & J. Tixier 1989. Radiocarbon-accelerator dating of Ksar ‘Aqil (Lebanon) and the chronology of the Upper Palaeolithic sequence in the Middle East. Antiquity 63, 61–68. Mellars, P. et al. (eds.) 2007. Rethinking the human revolution: new behavioural and biological perspectives on the origin and dispersal of modern humans. Cambridge: McDonald Institute for Archaeological Research. Mercier, H. & H.Valladas 2003. Reassessment of TL age estimates of burnt flints from the Paleolithic site of Tabun Cave, Israel. Journal of Human Evolution 45, 401–409. Mercier, N. et al. 1993. Thermoluminescence data for the Mousterian burial site of Es-Skhul, Mt. Carmel. Journal of Archaeological Science 20, 2, 169–174. Mercier, N. et al. 1999. Thermoluminescence dating of a Middle Palaeolithic occupation at Sodmein Cave, Red Sea Mountains (Egypt). Journal of Archaeological Science 26, 1339–1345. Merlan, F. 1989.The interpretive framework of Wardaman rock art: a preliminary report. Australian Aboriginal Studies 1989/2, 14–24. Metcalf, P. 1991. A Borneo journey into death: Berawan eschatology from its rituals. Kuala Lumpur: S. Abdul Majeed. Metspalu, M. et  al. 2004. Most of the extant mtDNA boundaries in South and Southwest Asia were likely shaped during the initial settlement of Eurasia by anatomically modern humans. BMC Genetics 5, 1, 26. Miller-Antonio, S. et al. 2004. Lithic raw material use at the late Middle Pleistocene site of Panxian Dadong. Asian Perspectives 43, 2, 314–332. Mijares, A. S. B. 2002. The Minori Cave expedient lithic technology. Diliman, Quezon City: University of the Philippines Press. Mijares, A. S. B. 2004. Lithic analysis of recently excavated Tabon Cave flakes. Proceedings of the Society of Philippine Archaeologists 2, 15–19. Mijares, A. S. B. 2007. The Late Pleistocene to Early Holocene foragers of Northern Luzon. Bulletin of the Indo-Pacific Prehistoric Association 28, 99–107. Mijares, A. S. B. 2008. The Peñablanca flake tools: an unchanging technology? Hukay 12, 13–34. Mijares, A. S. B. et al. 2010. New evidence for a 67,000-year-old human presence at Callao Cave, Luzon, Philippines. Journal of Human Evolution 59, 123–132. Minnegal, M. & P. D. Dwyer 2007. Foragers, farmers and fishers: responses to environmental perturbation. Journal of Political Economy 14, 34–57.

295

References Misra,V. N. 1995a. Evolution of environment and culture in the Rajasthan Desert during the Late Quaternary. In: E. Johnson (ed.), Ancient peoples and landscapes. Lubbock: Museum of Texas Tech University, 77–104. Misra, V. N. 1995b. Geoarchaeology of the Thar desert, northwest India. In: S. Wadia, R. Korisettar & V. S. Kale (eds.), Quaternary environments and geoarchaeology of India. Bangalore: Memoirs Geological Society of India, 210–230. Misra,V. N. 2001. Prehistoric human colonization of India. Journal of Biosciences 26, 491–531. Mithen, S. 1993.Technology and society during the Middle Pleistocene: hominid group size, social learning and industrial variability. Cambridge Archaeological Journal 4, 3–32. Molengraaff, G. A. F. & M.Weber 1921. On the relation between the Pleistocene glacial period and the origin of the Sunda Sea (Java and South China-Sea), and its influence on the distribution of coral reefs and on the land- and freshwater fauna. Netherlands Academy of Arts and Sciences (KNAW) Proceedings 23, 1, 395–439. Monnier, G. F. 2006.The Lower/Middle Paleolithic periodization in western Europe. An evaluation. Current Anthropology 47, 5, 709–744. Mooney, S. D. et al. 2011. Late Quaternary fire regimes of Australasia. Quaternary Science Reviews 30, 28–46. Moore, M. W. 2000. Lithic technology in Tasmania. Archaeology in Oceania 35, 2, 57–73. Moore, M. W. 2003. Australian Aboriginal biface reduction technique on the Georgina Rive, Camooweal, Queensland. Australian Archaeology 56, 22–34. Moore, W. 2005. The knife man. London: Bantam. Moore, M.W. 2011.The design space of stone flaking: implications for cognitive evolution. World Archaeology 43, 4, 702–715. Moore, M. W. 2013. Simple stone flaking in Australasia: patterns and implications. Quaternary International 285, 140–149. Moore, M. W. & A. Brumm 2007. Stone artifacts and hominins in island Southeast Asia: new insights from Flores, eastern Indonesia. Journal of Human Evolution 52, 85–102. Moore, M. W. et al. 2009. Continuities in stone flaking technology at Liang Bua, Flores, Indonesia. Journal of Human Evolution 57, 5, 503–526. Morant, G. M. 1927. A study of the Australian and Tasmanian skulls, based on previously published measurements. Biometrika 19, 417–440. Morant, G. M. 1939. Note on Dr J. Wunderly’s survey of Tasmanian crania. Biometrika 30, 338–340. Morlote, D. et al. 2011. The Soliga, an isolated tribe from Southern India: genetic diversity and phylogenetic affinities. Journal of Human Genetics 56, 4, 258–269. Morphy, H. 1981. Art of northern Australia. In: Aboriginal Australia (Exhibition Catalogue). Sydney: Australian Gallery Directors Council Ltd., 53–65. Morse, K. 1988. Mandu Mandu Creek Rockshelter: Pleistocene human coastal occupation of North West Cape, Western Australia. Archaeology in Oceania 23, 81–88. Morse, K. 1993a. Shell beads from Mandu Mandu Creek rock-shelter, Cape Range peninsula, Western Australia, dated before 30,000 b.p. Antiquity 67, 877–883. Morse, K. 1993b. New radiocarbon dates from North West Cape, Western Australia: a preliminary report. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene sites in Australia, New Guinea and Island Melanesia. Canberra: ANU, 155–163. Morse, K. 1999. Coastwatch: Pleistocene resource use on the Cape Range Peninsula. In: J. Hall & I. McNiven (eds.), Australian Coastal Archaeology. Canberra: ANH Publications, R.S.P.A.S., Australian National University, 73–78. Morwood, M. J. 1979. Art and stone: towards a prehistory of Central Western Queensland. Unpublished PhD thesis, Department of Prehistory and Anthropology, Australian National University. Morwood, M. J. 2002. Visions from the past: the archaeology of Australian Aboriginal art. Sydney: Allen and Unwin. Morwood, M. J. & P.Trezise 1989. Edge-ground axes in Pleistocene greater Australia: new evidence from S.E. Cape York Peninsula. Queensland Archaeological Research 6, 77–90. Morwood, M. J. & D. R. Hobbs 1995. Antiquity 69, Special Number 265, 747–768. Morwood, M. J. & S. Jung 1995. Excavations at Magnificent Gallery. In: M. J. Morwood & D. R. Hobbs (eds.), Quinkan prehistory: the archaeology of Aboriginal art in S. E. Cape York Peninsula, Australia. St. Lucia: Anthropology Museum, University of Queensland, 92–100. Morwood, M. J. & I. Davidson 2005. Out of ancient disasters, forebears may have colonised new lands. Sydney Morning Herald. Sydney, 17. Morwood, M. J. & P. van Oosterzee 2007. The discovery of the Hobbit: the scientific breakthrough that changed the face of human history. Melbourne: Random House.

296

References Morwood M. J., D. R. Hobbs & D. M. Price 1995. Excavations at Sandy Creek 1 and 2. In: M. J. Morwood & D. R. Hobbs (eds.), Quinkan prehistory: the archaeology of Aboriginal art in S. E. Cape York Peninsula, Australia. St. Lucia: Anthropology Museum, University of Queensland, 71–91. Morwood, M. J. et al. 2003. Revised age for Mojokerto 1, an early Homo erectus cranium from East Java, Indonesia. Australian Archaeology 57, 1–4. Morwood, M. J. et al. 2004. Archaeology and age of a new hominin from Flores in eastern Indonesia. Nature 431, 1087–1091. Morwood, M. J. et al. 2005. Further evidence for small-bodied hominins from the Late Pleistocene of Flores, Indonesia. Nature 437, 1012–1017. Morwood, M. J. et al. 2008. Climate, people and faunal succession on Java, Indonesia: evidence from Song Gupuh. Journal of Archaeological Science 35, 7, 1776–1789. Morwood, M. J. et al. 2009. Preface: research at Liang Bua, Flores, Indonesia. Journal of Human Evolution 57, 5, 437–449. Moser, J. 2001. Hoabinhian: Geographie und Chronologie eines steinzeitlichen Technokomplexes in Südostasien. Cologne: Linden Soft. Mountain, M.-J. 1991a. Highland New Guinea hunter-gatherers: the evidence of Nombe Rockshelter, Simbu with emphasis on the Pleistocene. Unpublished PhD thesis, Australian National University, Canberra. Mountain, M.-J. 1991b. Landscape use and environmental management of tropical rainforest by pre-agricultural hunter-gatherers in northern Sahulland. Indo-Pacific Prehistory Association Bulletin 11, 54–68. Movius, H. L. 1943. The stone age of Burma. Transactions of the American Philosophical Society 32, 341–393. Movius, H. L., Jr. 1944. Early man and Pleistocene stratigraphy in Southern and Eastern Asia. Cambridge, Mass.: Peabody Museum. Movius, H. L. 1948. The Lower Palaeolithic cultures of southern and eastern Asia. Transactions of the American Philosophical Society 38, 4, 329–420. Movius, H. L. 1978. Southern and eastern Asia: Conclusions. In: I. K. Ikawa-Smith (ed.), Early Palaeolithic in South and East Asia. The Hague: Mouton, 351–355. Mudar, K. & D.  Anderson 2007. New evidence for Southeast Asian Pleistocene foraging economies: faunal remains from the early levels of Lang Rongrien rockshelter, Krabi, Thailand. Asian Perspectives 46, 2, 298–334. Muke, J. D. 1984. The Huon discoveries: a preliminary report on the stone artefacts and a comparative analysis of the distribution of waisted tools in Greater New Guinea. Unpublished Bachelor of Letters thesis, Australian National University, Canberra. Müller, U. C. et al. 2011.The role of climate in the spread of modern humans into Europe. Quaternary Science Reviews 30, 273–279. Mulvaney, D. J. 1964. The Pleistocene colonization of Australia. Antiquity 38, 263–267. Mulvaney, D. J. 1976. ‘The chain of connection’: the material evidence. In: N. Peterson (ed.), Tribes and boundaries in Australia. Canberra: Australian Institute of Aboriginal Studies, 72–94. Mulvaney, D. J. 1981. Origins. In: Aboriginal Australia (Exhibition Catalogue). Sydney: Australian Gallery Directors Council Ltd., 15–27. Mulvaney, K. 2013. Iconic imagery: Pleistocene rock art development across Northern Australia. Quaternary International 285, 99–110. Mulvaney, D. J. & E. B. Joyce 1965. Archaeological and geomorphological investigations on Mt Moffatt Station, Queensland, Australia. Proceedings of the Prehistoric Society 31, 147–212. Mulvaney, J. & J. Kamminga 1999. Prehistory of Australia. St. Leonards: Allen & Unwin. Murty, M. L. K. 1979. Recent research on the Upper Palaeolithic phase in India. Journal of Field Archaeology 6, 3, 301–320. Musser, G. G. 1981.The giant rat of Flores and its relatives east of Borneo and Bali. Bulletin of the American Museum of Natural History 169, 69–175. Narr, K. 1966. Die frühe und mittlere Altsteinzeit Süd- und Ostasiens. Bern: Francke. Neri, L. A. 2002. Obsidian sourcing at Huluga open site: an evidence of exchange? Unpublished Master’s thesis, University of the Philippines, Archaeological Studies Program. Neri, L. A. 2005. Obsidian from the Huluga site and its implications. Hukay 7, 55–67. Nettle, D. 1998. Explaining global patterns of language diversity. Journal of Anthropological Archaeology 17, 4, 354–374. Newsome, J. & J. R. Flenley 1998. Late Quaternary vegetational history of the Central Highlands of Sumatra II. Palaeopalynology and vegetational history. Journal of Biogeography 15, 555–578.

297

References Nguyen Lang Cuong 1992. A reconsideration of the chronology of hominid fossils from Vietnam. In: T. Akazawa et al. (eds.), The evolution and dispersal of modern humans in Asia. Tokyo: Hokusen-Sha, 321–335. Nicolaisen, I. 2003. Pillars of faith: souls, fertility, and mortuary rituals among the Punan Bah of Central Borneo – a preliminary analysis. In: W. D. Wilder (ed.), Journeys of the soul: anthropological studies of death, burial and reburial practices in Borneo. Borneo Research Council Monograph Series no. 7, 121–174. Noble, W. & I. Davidson 1991. The evolutionary emergence of modern human behaviour: language and its archaeology. Man: Journal of the Royal Anthropological Institute 26, 223–253. Noble, W. & I. Davidson, 1996. Human evolution, language and mind. A psychological and archaeological inquiry. Cambridge: Cambridge University Press. Norton, C. & J. J. H. Jin 2009. The evolution of modern human behaviour in East Asia: current evidence. Evolutionary Anthropology 18, 247–260. Norton, C. J. et al. 2006. Middle Pleistocene handaxes from the Korean Peninsula. Journal of Human Evolution 51, 527–536. Nougier, L.-R. 1970. L’économie préhistorique.‘Que sais-je?’ No. 1397. Paris: PUF. Nowell, A. 2010. Defining behavioral modernity in the context of Neandertal and anatomically modern human populations. Annual Review of Anthropology 39, 1, 437–452. Oakley, K. & M. Leakey 1937. Report on excavations at Jaywick Sands, Essex (1934) with some observations on the Clactonian industry and on the fauna and geological significance of the Clacton Channel. Proceedings of the Prehistoric Society 3, 216–260. Oakley, K. P. & H. De Vries 1959. Radiocarbon dating of the Piltdown skull and jaw. Nature 184, 224–226. Ochoa, J. 2009. Terrestrial vertebrates from Ille Cave, northern Palawan, Philippines: subsistence and palaeoecology in the terminal Pleistocene to the Holocene. Unpublished Master’s Thesis, University of the Philippines. Ochoa, J. & P. J. Piper in press. Holocene large mammal extinctions in Palawan Island, Philippines. In G. Monks (ed.), Climate change, human response and zooarchaeology.Vertebrate paleobiology and paleoanthropology series, New York: Springer. O’Connell, J. F. 2006. How did modern humans displace Neanderthals? Insights from hunter-gatherer ethnography and archaeology. In: N. J. Conard (ed.), When Neanderthals and modern humans met. Tübingen: Kerns, 43–64. O’Connell, J. F. 2011. Comment on J. J. Shea, Homo sapiens is as Homo sapiens was: behavioral variability versus ‘behavioral modernity’ in Paleolithic archaeology. Current Anthropology 52, 1, 20–21. O’Connell, J. F. & J. Allen 1998. When did humans first arrive in Greater Australia and why is it important to know? Evolutionary Anthropology 6, 4, 132–146. O’Connell, J. F. & J. Allen 2004. Dating the colonization of Sahul (Pleistocene Australia–New Guinea): a review of recent research. Journal of Archaeological Science 31, 835–853. O’Connell, J. F. & J. Allen 2007. Pre-LGM Sahul (Pleistocene Australia-New Guinea) and the archaeology of early modern humans. In: P. Mellars et al. (eds.), Rethinking the human revolution: new behavioural and biological perspectives on the origin and dispersal of modern humans. Cambridge: McDonald Institute for Archaeological Research, 395–410. O’Connell, J. F. & J. Allen 2012. The restaurant at the end of the universe: modelling the colonisation of Sahul. Australian Archaeology 74, 5–31. O’Connell, J. F., P. K. Latz & P. Barnett 1983. Traditional and modern plant use among the Alyawarra of Central Australia. Economic Botany 37, 1, 80–109. O’Connell, J. F., J. Allen & J. Hawkes 2010. Pleistocene Sahul and the origins of seafaring. In: A. Anderson et al. (eds.), The global origins and development of seafaring. Cambridge: McDonald Institute for Archaeological Research, 57–68. O’Connor, S. 1995. Carpenter’s Gap Rockshelter 1: 40,000 years of Aboriginal occupation in the Napier Ranges, Kimberley, W. A. Australian Archaeology 40, 58–59. O’Connor, S. 1999. 30,000 years of Aboriginal occupation: Kimberley, North West Australia. Canberra: ANU. O’Connor, S. 2005. Carpenter’s Gap Rockshelter 1: 40,000 years of Aboriginal occupation in the Napier Ranges, Kimberley, W. A. Australian Archaeology 40, 58–59. O’Connor, S. 2007a. Revisiting the past: changing interpretations of settlement, subsistence, and demography in the Kimberley region. In: M. Donaldson & K. Kenneally (eds.), Rock art of the Kimberley: proceedings of the Kimberley Society Rock Art Seminar. Perth: Kimberley Society, 57–80. O’Connor, S. 2007b. New evidence from East Timor contributes to our understanding of earliest modern human colonisation east of the Sunda shelf. Antiquity 81, 313, 518–535.

298

References O’Connor, S. 2012. Out of Asia. Australian Science 33, 1, 16–19. O’Connor, S. & B. Fankhauser 2001. Art at 40,000 BP? One step closer: an ochre covered rock from Carpenter’s Gap Shelter 1, Kimberley region, Western Australia. In: A. Anderson et al. (eds.), Histories of old ages: essays in honour of Rhys Jones. Canberra: Pandanus Books, 287–310. O’Connor, S. & D. Bulbeck in press. Homo sapiens societies in Indonesia and Southeastern Asia. In: V. Cummings, P. Jordan & M. Zvelebil (eds.), The Oxford handbook of the archaeology and anthropology of hunter-gatherers. Oxford: Oxford University Press. O’Connor, S., P. Veth & N. Hubbard 1993. Changing interpretations of postglacial human subsistence and demography in Sahul. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene sites in Australia, New Guinea and Island Melanesia. Canberra: ANU, 95–105. O’Connor, S., P. Veth & C. Campbell 1998. Serpent’s Glen rockshelter: report of the first Pleistocene-aged occupation sequence from the Western Desert. Australian Archaeology 46, 12–22. O’Connor, S., M. Spriggs & P. Veth 2002. Direct dating of shell beads from Lene Hara Cave, East Timor. Australian Archaeology 55, 18–21. O’Connor, S. et al. 2005. Liang Lemdubu, a Pleistocene cave site in the Aru Islands. In: S. O’Connor, M. Spriggs & P. Veth (eds.), The archaeology of the Aru Islands, Eastern Indonesia. Terra Australis 23. Canberra: Pandanus Books, 171–204. O’Connor, S. et al. 2010. Cave archaeology and sampling issues in the tropics:A case study from Lene Hara Cave, a 42,000 year old occupation site in East Timor, Island Southeast Asia. Australian Archaeology 71, 29–40. O’Connor, S. et al. 2011a. Pelagic fishing at 42,000 BP and the maritime skills of modern humans. Science 334, 1117–1121. O’Connor, S. et al. 2011b.The power of paradigms: examining the evidential basis for early to mid-Holocene pigs and pottery in Melanesia. Journal of Pacific Archaeology 2, 2, 1–25. Odell, G. H. 1981. The mechanics of use-breakage of stone tools: some testable hypotheses. Journal of Field Archaeology 8, 197–209. Oliver, W. L. R. & L. R. Heaney 1996. Biodiversity and conservation in the Philippines. International Zoo News 45, 3, 329–337. Olley, J. M. et  al. 2006. Single-grain optical dating of grave-infill associated with human burials at Lake Mungo, Australia. Quaternary Science Reviews 25, 19–20, 2469–2474. Olsen, J. W. & R. Ciochon 1990. A review of evidence for postulated Middle Pleistocene occupation in Viet Nam. Journal of Human Evolution 19, 761–768. Ono, R., S. Soegondho & M.Yoneda 2009. Changing marine exploitation during late Pleistocene in Northern Wallacea: shell remains from Leang Sarru rockshelter in Talaud. Asian Perspectives 48, 2, 318–341. Oota, H. et al. 2005. Recent origin and cultural reversion of a hunter–gatherer group. PLOS Biology 3, 3, 536–542. Oppenheimer, S. 2002. The first exodus: journey out of Africa. Geographic Magazine 74, 32–35. Oppenheimer, S. 2003. Out of Eden:The peopling of the world. London: Robinson. Oppenheimer, S. 2004. Out of Eden:The peopling of the world. Revised paperback edition. London: Robinson. Oppenheimer, S. 2009. The great arc of dispersal of modern humans: Africa to Australia. Quaternary International 202, 1–2, 2–13. Oppenheimer, S. 2012a. A single southern exit of modern humans from Africa: before or after Toba? Quaternary International 258, 88–99. Oppenheimer, S. 2012b. Out-of-Africa, the peopling of continents and islands: tracing uniparental gene trees across the map. Philosophical Transactions of the Royal Society B, Biological Sciences 367, 1590, 770–784. Oppenoorth,W. E. F. 1932. Homo (Javanthropus) soloensis: een Pleistocene Mensch van Java. Scientific Proceedings of the Mining Company of the Dutch East Indies 20, 49–75. Orzack, S. H. & E. Sober 1994. Optimality models and the test of adaptationism. American Naturalist 143, 361–380. Orzack, S. H. & E. Sober (eds.) 2000. Optimality and adaptationism. Cambridge: Cambridge University Press. Osborne, A. H. et al. 2008. A humid corridor across the Sahara for the migration of early modern humans out of Africa 120,000 years ago. PNAS 105, 16444–16447. Owen, R. 1853. Descriptive catalogue of the osteological series contained in the Museum of the Royal College of Surgeons of England.Vol. II. Mammalia Placentiala. London: Taylor and Frances. Oysten, B. 1996.Thermoluminescence age determinations for the Mungo III human burial, Lake Mungo, Southeastern Australia. Quaternary Science Reviews 15, 739–747.

299

References Paijmans, K. 1976. New Guinea vegetation. Canberra: Australian National University Press. Pardoe, C. 1986. Population genetics and population size in prehistoric Tasmania. Australian Archaeology 22, 1–6. Pardoe, C. 1988.The cemetery as symbol.The distribution of prehistoric Aboriginal burial grounds in southeastern Australia. Archaeology in Oceania 23, 1–16. Pardoe, C. 1991a. Competing paradigms and ancient human remains: the state of the discipline. Archaeology in Oceania 26, 79–85. Pardoe, C. 1991b. Isolation and evolution in Tasmania. Current Anthropology 32, 1, 1–21. Pardoe, C. 1993.The Pleistocene is still with us: analytical constraints and possibilities for the study of ancient human remains in archaeology. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene sites in Australia, New Guinea and Island Melanesia. Canberra: ANU, 81–94. Pardoe, C. 1995. Riverine, biological and cultural evolution in southeastern Australia. Antiquity 69, Special Number 265, 696–713. Pardoe, C. 2004. Australian biological anthropology for archaeologists. In: T. Murray (ed.), Archaeology from Australia. Melbourne: Australian Scholarly Publishing, 131–150. Pardoe, C. 2006. Becoming Australian: evolutionary processes and biological variation from ancient to modern times. Before Farming 1, 1–21. Parker, A. G. 2009. Pleistocene climate change in Arabia: developing a framework for hominin dispersal over the last 350 ka. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 39–49. Parker, G. A. & J. Maynard Smith 1990. Optimality theory in evolutionary biology. Nature 348, 27–33. Parton, A. et al. 2010. An early MIS 3 wet phase at palaeolake Aqabah: preliminary interpretation of the multi-proxy record. Proceedings of the Seminar for Arabian Studies 40, 267–276. Parton, A. et al. in press. An early MIS 3 pluvial phase in southeast Arabia: climatic and archaeological implications. Quaternary International. Pasveer, J. 2004. The Djief hunters: 26,000 years of rainforest exploitation on the Bird’s Head of Papua, Indonesia. London: A. A. Balkema Publishers. Paterson, T. T. 1937. Studies in the Palaeolithic succession in England: No. 1. The Barnham sequence. Proceedings of the Prehistoric Society 3, 87–135. Patnaik, R. 2000. Paleoenvironment of Homo erectus bearing Pleistocene Narmada Valley deposits of India. Proceedings of the XVI Indian Colloquium on Micropaleontology and Stratigraphy (Bulletin of the Oil and Natural Gas Corporation Ltd.) 37, 2, 247–249. Patnaik, R. et al. 2009. New geochronological, palaeoclimatological and Palaeolithic data from the Narmada Valley hominin locality, central India. Journal of Human Evolution 56, 114–133. Patole-Edoumba, E. 2002. L’industrie lithique préhistorique de débitage des Philippines de la fin du Pleistocène à l’Holocène moyen. Unpublished PhD Thesis, University of Aix-Marseille I. Patole-Edoumba, E., 2006. Caractérisation des premières industries de pierre taillée des Philippines. Aseanie 17, 61–93. Patole-Edoumba, E. 2009. A typo-technological definition of Tabonian industries. Bulletin of the Indo-Pacific Prehistoric Association 29, 21–25. Pavlides, C. 1999. The story of Imlo: the organisation of flaked stone technologies from the lowland tropical rainforest of West New Britain, Papua New Guinea. Unpublished PhD Thesis, La Trobe University, Melbourne. Pavlides, C. 2004. From Misisil Cave to Eliva Hamlet: rediscovering the Pleistocene in Interior West New Britain. Records of the Australian Museum, Supplement 29, 97–108. Pavlides, C. & C. Gosden 1994. 35,000-year-old sites in the rainforests of West New Britain. Antiquity 68, 604–610. Pawley, A. 1999. Chasing rainbows: implications of the rapid dispersal of Austronesian languages for subgrouping and reconstruction. In: E. Zeitoun & P. J. Li (eds.), Selected papers from the Eighth International Conference on Austronesian Linguistics. Taipei: Institute of Linguistics, Academia Sinica, 95–138. Pawley, A. 2005. The chequered career of the Trans New Guinea Phylum: recent historical research and its implications. In: A. Pawley et al. (eds.), Papuan pasts: investigations into the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, Australian National University, 67–108. Pawley, A. 2007. Recent research on the historical relationships of the Papual languages, or, what does linguistics say about the prehistory of Melanesia? In: J. S. Friedlaender (ed.), Genes, language, and culture history in the Southwest Pacific. Oxford: Oxford University Press, 36–58.

300

References Pawley, A. & M. Ross 1993. Austronesian historical linguistics and culture history. Annual Review of Anthropology 22, 425–459. Pawlik, A. 1992. Mikrogebrauchsspurenanalyse, Methoden  – Forschungsstand  – Neue Ergebnisse. Tübingen: Archaeologica Venatoria. Pawlik,A. 1995. Die mikroskopische Analyse von Steingeräten. Experimente – Auswertungsmethoden – Artefaktanalysen. Tübingen: Archaeologica Venatoria. Pawlik, A. 2002. Is there an Early Palaeolithic in the Philippines? New approaches for lithic analysis in the Philippines. In: M. Horrocks & P. Sheppard (eds.), Australasian connections and new directions: proceedings of the 7th Australasian Archaeometry Conference. Auckland: University of Auckland, 255–270. Pawlik, A. 2004. The Palaeolithic site of Arubo 1 in Central Luzon, Philippines. Bulletin of the Indo-Pacific Prehistory Association 24, 3–12. Pawlik, A. 2009a. Is the functional approach helpful to overcome the typology dilemma of lithic archaeology in Southeast Asia? Bulletin of the Indo-Pacific Prehistory Association 29, 6–14. Pawlik, A. 2009b. Morphological and functional analysis of lithic assemblages from the Pleistocene layers of Ille Cave, Dewil Valley, El Nido, Palawan. A progress report. Unpublished manuscript submitted to OVCRD, University of the Philippines. Pawlik, A. F. 2010. Have we overlooked something? Hafting traces and indications of modern traits in the Philippine Palaeolithic. Bulletin of the Indo-Pacific Prehistory Association 30, 35–53. Pawlik, A. & W. P. Ronquillo 2003. The Palaeolithic in the Philippines. Lithic Technology 28, 2, 79–93. Pawlik, A. & J. Thissen 2011. Hafted armatures and multi-component tool design at the Micoquian site of Inden-Altdorf, Germany. Journal of Archaeological Science 38, 1699–1708. Pearce R. H. & M. Barbetti 1981. A 38,000 year old archaeological site at Upper Swan, Western Australia. Archaeology in Oceania 16, 173–178. Peer, P. van 1998. The Nile corridor and the Out-of-Africa model. Current Anthropology 39, S115-S140. Peer, P. van, P. M.Vermeersch & E. Paulissen 2010. Chert quarrying, lithic technology and a modern human burial at the Middle Palaeolithic site of Taramsa 1, Upper Egypt. Leuven: Leuven University Press. Peer, P. van et al. 1996. Palaeolithic sequence of Sodmein cave, Red Sea mountains, Egypt. In: G. Pwiti (ed.), Aspects of African archaeology. Harare: University of Zimbabwe, 149–156. Penck, A. & E. Brückner 1909. Die Alpen im Eiszeitalter. Leipzig: Tauchnitz. Peng Shulin n.d. Study of the pebble tool culture in south China. Unpublished manuscript, Museum of Guangxi, China. Perera, N. et al. 2011. People of the ancient rainforest: Late Pleistocene foragers at the Batadomba-lena rockshelter, Sri Lanka. Journal of Human Evolution 61, 254–269. Peresani, M. et al. 2011. Late Neandertals and the intentional removal of feathers as evidenced from bird bone taphonomy at Fumane Cave 44 ky B.P., Italy. PNAS 108, 3888–3893. Peterson, N. 1986. Australian territorial organisation: a band perspective. Oceania Monograph 30. Sydney: University of Sydney. Petit-Maire, P. et al. 2010. A vast Eemian palaeolake in southern Jordan (29°N). Global and Planetary Change 72, 368–373. Petraglia, M. D. 2005. Hominin responses to Pleistocene environmental change in Arabia and South Asia. Geological Society, London, Special Publications 247, 1, 305. Petraglia, M. D. 2011. Trailblazers across Arabia. Nature 470, 50–51. Petraglia, M. D. & J. I. Rose (eds.) 2009. The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer. Petraglia, M. D. et al. 2007.Middle Paleolithic assemblages from the Indian subcontinent before and after the Toba super-eruption. Science 317, 114–116. Petraglia, M. D. et al. 2009. Population increase and environmental deterioration correspond with microlithic innovations in South Asia ca. 35,000 years ago. PNAS 106, 30, 12261–12266. Petraglia, M. D. et al. 2010. Out of Africa: new hypotheses and evidence for the dispersal of Homo sapiens along the Indian Ocean rim. Annals of Human Biology 37, 288–311. Petraglia, M. D. et al. 2011. Middle Paleolithic occupation on a Marine Isotope Stage 5 lakeshore in the Nefud Desert, Saudi Arabia. Quaternary Science Reviews 30, 1555–1559. Petraglia, M. D. et al. 2012a. Hominin evolutionary history in the Arabian Desert and the Thar Desert. In: L. Mol & T. Sternberg (eds.), Changing deserts: integrating people and their environment. Cambridge: White Horse Press, 61–82.

301

References Petraglia, M. D. et al. 2012b.The Toba volcanic super-eruption, environmental change, and hominin occupation history in India over the last 140,000 years. Quaternary International 258, 119–134. Petrow, S. 1997. The last man: the mutilation of William Lanne in 1869 and its aftermath. Aboriginal History 21, 90–112. Pettitt, P. 2011. The Palaeolithic origins of human burial. London: Routledge. Pfeiffer, J. E. 1982. The creative explosion. An inquiry into the origins of art and religion. New York: Harper & Row. Piercy, I. L. 2009. The demography of Eurasian Neanderthals: implications for cultural complexity and transmission. Unpublished BA (Hons) Thesis, School of Social Science, University of Queensland. Pigliucci, M. & J. Kaplan 2000. The fall and rise of Dr Pangloss: adaptationism and the spandrels paper 20 years later. Trends in Ecology and Evolution 15, 66–70. Pike-Tay, A. & R. Cosgrove 2002. From reindeer to wallaby: recovering patterns of seasonality, mobility, and prey selection in the Paleolithic Old World. Journal of Archaeological Method and Theory 9, 101–146. Pike-Tay, A. et al. 2008. Systematic seasonal land use by late Pleistocene Tasmanian Aborigines. Journal of Archaeological Science 35, 2532–2544. Piper, P. J. & A. S. B. Mijares 2007. A preliminary report on a Late Pleistocene animal bone assemblage from Callao Cave, Peñablanca, northern Luzon, Philippines. Unpublished report, Archaeological Studies Program, University of the Philippines/National Museum of the Philippines. Piper, P. J. & R. J. Rabett 2009a. Hunting in a tropical rainforest: evidence from the Late Pleistocene and Early Holocene at Lobang Hangus, Niah Cave, Borneo. International Journal of Osteoarchaeology 19, 4, 551–565. Piper, P. J. & R. J. Rabett 2009b. Disentangling the Harrisson Archive to interpret the spatial and temporal distribution of vertebrate remains at Niah Cave, Borneo. International Journal of Osteoarchaeology 19, 4, 464–475. Piper, P. J., R. J. Rabett & G. Barker in prep. The vertebrate fauna. In: G. Barker et al. (eds.), Rainforest foraging and farming in Island Southeast Asia: the archaeology and environmental history of the Niah Caves, Sarawak. Cambridge: McDonald Institute for Archaeological Research. Piper, P. J. et al. 2008. The first evidence for the past presence of the tiger Panthera tigris (L.) on the island of Palawan, Philippines: extinction in an island population. Palaeoclimatology, Palaeogeography, Palaeoecology 264, 123–127. Piper, P. J. et al. 2011. Palaeozoology of Palawan Island, Philippines. Quaternary International 233, 142–158. Pleurdeau, D. 2005. Human technical behavior in the African Middle Stone Age: the lithic assemblage of Porc-Epic Cave (Dire Dawa, Ethiopia). African Archaeological Review 22, 4, 177–197. Plicht, J. van der, A. van der Wijk & G. J. Bartstra 1989. Uranium and thorium in fossil bones: Activity ratios and dating. Applied Geochemistry 4, 339–342. Pookajorn, S. 1988. Archaeological research of the Hoabinhian Culture or technocomplex and its comparison with ethnoarchaeology of the Phi Tong Luang, a hunter-gatherer group of Thailand. Tübingen: Archaeologica Venatoria. Pookajorn, S. 1996. Human activities and environmental changes during the Late Pleistocene to Middle Holocene in South Thailand and Southeast Asia. In: L. G. Straus et al. (eds.), Humans at the end of the Ice Age: the archaeology of the Pleistocene-Holocene transition. New York: Plenum Press, 201–213. Pope, G. C. 1989. Bamboo and human evolution. Natural History 98, 10, 49–57. Pope, G. C. & S. G. Keates 1994. The evolution of human cognition and cultural capacity. In: R. Corruchini & S. Ciochon (eds.), Integrative paths to the past: paleoanthropological advances in honor of F. Clark Howell. Upper Saddle River, N.J.: Prentice Hall, 531–567. Pope, G. et  al. 1986. Earliest radiometrically dated artefacts from SE Asia. Current Anthropology 27, 3, 275–279. Porch, N. & J. Allen 1995. Tasmania: archaeological and palaeoecological perspectives. Antiquity 69, Special Number 265, 714–732. Porr, M. 2010. Identifying behavioural modernity. Lessons from Sahul. Bulletin of the Indo-Pacific Prehistory Association 30, 28–34. Powell, J. M. 1982. Plant resources and palaeobotanical evidence for plant use in the Papua New Guinea Highlands. Archaeology in Oceania 17, 1, 28–37. Powell, A., S. Shennan & M. G. Thomas 2009. Late Pleistocene demography and the appearance of modern human behavior. Science 324, 1298–1301. Preusser, F., D. Radies & A. Matter 2002. A 160,000-year record of dune development and atmospheric circulation in Southern Arabia. Science 296, 2018–2020.

302

References Pritchard, P., R. Rabett & P. Piper 2009. Distinguishing species of geoemydid and trionychid turtles from shell fragments: evidence from the Pleistocene at Niah Cave, Sarawak, Borneo. International Journal of Osteoarchaeology 19, 4, 531–550. Prugnolle, F., A. Manica & F. Balloux 2005. Geography predicts neutral genetic diversity of human populations Current Biology 15, 5, R159–160. Przywolnik, K. 2005. Long-term transitions in hunter-gatherers of coastal northwestern Australia. In: P.Veth, M. Smith & P. Hiscock (eds.), Desert peoples: archaeological perspectives. Oxford: Blackwell, 177–205. Pulleine, R. 1929.The Tasmanians and their stone culture. Australasian Association for the Advancement of Science 19, 294–314. Pyatt, B. et al. 2010. Analytical examination of animal remains from Borneo: the painting of bone and shell. Journal of Archaeological Science 37, 2102–2105. Quintana-Murci, L. L. et al. 1999. Genetic evidence of an early exit of Homo sapiens sapiens from Africa through eastern Africa. Nature Genetics 23, 437–444. Rabett, R. 2012. Human adaptation in the Asian Palaeolithic: hominin dispersal and behaviour during the Late Quaternary. Cambridge : Cambridge University Press. Rabett, R. & G. Barker 2007. Through the looking glass: new evidence on the presence and behaviour of Late Pleistocene humans at Niah Cave, Sarawak, Borneo. In: P. Mellars et al. (eds.), Rethinking the human revolution. Cambridge: McDonald Institute for Archaeological Research, 383–394. Rabett, R. & G. Barker 2010. Late Pleistocene and Early Holocene forager mobility in Southeast Asia. In: B. Bellina et al. (eds.), 50 years of archaeology in Southeast Asia: essays in honour of Ian Glover. Bangkok: River Books, 66–77. Rabett, R. & P. J. Piper 2012.The emergence of bone technologies at the end of the Pleistocene in Island and Mainland Southeast Asia and their regional implications. Cambridge Archaeological Journal 22, 1, 37–56. Rabett, R. J., P. J. Piper & G. Barker 2006. Bones from hell: preliminary results of new work on the Harrisson faunal assemblage from the deepest part of Niah Cave, Sarawak. In: E. A. Bacus et al. (eds.), Uncovering Southeast Asia’s past. Singapore: University of Singapore Press, 46–59. Rabett, R. et al. 2009. The Tràng An project: Late- to Post-Pleistocene settlement of the lower Song Hong valley, North Vietnam. Journal of the Royal Asiatic Society 19, 1, 83–109. Rabett, R. et  al. 2011. Inland shell midden site-formation: investigation into a late Pleistocene to early Holocene midden from Tràng An, northern Vietnam, Quaternary International 239, 153–169. Rabinovich, R. & E.Tchernov 1995. Chronological, paleoecological and taphonomical aspects of the Middle Palaeolithic site of Qafzeh, Israel. In: H. Buitenhuis & H.-P. Uerpmann (eds.), Archaeozoology of the Near East. Leiden: Backhuys Publishers, 5–44. Rabinovich, R. & E. Hovers 2004. Faunal assemblages from Amud Cave: preliminary results and interpretations. International Journal of Osteoarchaeology 14, 287–306. Rasmussen, M. et al. 2011. An Aboriginal genome reveals separate human dispersals into Asia. Science 334, 94–98. Reich, D. et al. 2010. Genetic history of an archaic hominin group from Denisova Cave in Siberia. Nature 468, 1053–1060. Reich, D. et al. 2011. Denisova admixture and the first modern human dispersals into Southeast Asia and Oceania. American Journal of Human Genetics 89, 516–528. Reimer, P. J. et al. 2009. IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal BP. Radiocarbon 51, 1111–1150. Relethford, J. H. 2001. Ancient DNA and the origin of modern humans. PNAS 98, 2, 390–391. Relethford, J. H. 2008. Genetic evidence and the modern human origins debate. Heredity 100, 6, 555–563. Rendell, H. & R.W. Dennell 1987.The dating of an upper pleistocene archaeological site at Riwat, northern Pakistan. Geoarchaeology 1, 6–12. Rendell, H. R., R. W. Dennell & M. Halim 1989. Pleistocene and Palaeolithic investigations in the Soan Valley, Northern Pakistan. Oxford: Archaeopress. Renfrew, C. 1996.The sapient behaviour paradox: How to test for potential? In P. Mellars and K. Gibson (eds.), Modelling the early human mind. Cambridge: McDonald Institute for Archaeological Research, 11–15. Renfrew, C. 2008. Neuroscience, evolution and the sapient paradox: The factuality of value and the sacred. Philosophical Transactions of the Royal Society B 363, 2041–2047. Renfrew, C. 2009. Situating the creative explosion. Universal or local? In: C. Renfrew & I. Morley (eds.), Becoming human. Innovation in prehistoric material and spiritual culture. Cambridge: Cambridge University Press, 74–94.

303

References Resnik, D. 1997. Adaptationism: Hypothesis or heuristic? Biology and Philosophy 12, 39–50. Reynolds, T. E. G. 1993. Problems in the stone age of South-East Asia. Proceedings of the Prehistoric Society 59, 1–16. Richards, M. P. & E. Trinkaus 2009. Isotopic evidence for the diets of European Neanderthals and early modern humans. PNAS 106, 16034–16039. Richards, M. et al. 2000a. Neanderthal diet at Vindija and Neanderthal predation: the evidence from stable isotopes. PNAS 97, 13, 7663–7666. Richards, M. et al. 2000b. Tracing European founder lineages in the Near Eastern mtDNA pool. American Journal of Human Genetics 67, 5, 1251–1276. Richards, M. et al. 2001. Stable isotope evidence for increasing dietary breadth in the European mid-Upper Paleolithic. PNAS 98, 11, 6528–6532. Richards, M. et al. 2003. Extensive female-mediated gene flow from Sub-Saharan African into Near Eastern Arab populations. American Journal of Human Genetics 72, 1058–1064. Richards, M. et al. 2006. A model for the dispersal of modern humans out of Africa. In: H.-J. Bandelt, V. Macaulay & M. Richards (eds.), Human mitochondrial DNA and the evolution of Homo sapiens. Hamburg: Springer Verlag, 223–263. Richter, C. et al. 2008. Collapse of a new living species of giant clam in the Red Sea. Current Biology 18, 1349–1354. Rightmire, G. P. 2009. Middle and later Pleistocene hominins in Africa and Southwest Asia. PNAS 106, 16046–16050. Rindos, D. & E. Webb 1992. Modelling the initial human colonisation of Australia: perfect adaptation, cultural variability, and cultural change. Proceedings of the Australiasian Society for Human Biology 5, 441–454. Rink, W. J. et al. 2001a. Age of the Mousterian Industry at Hayonim Cave, Northern Israel, using electron spin resonance and 230Th/234U methods. Journal of Archaeological Science 31, 953–964. Rink, W. J. et al. 2001b. Electron spin resonance (ESR) and thermal ionization mass spectroscopic (TIMS) 230Th/234U dating of teeth in Middle Paleolithic layers at Amud Cave, Israel. Geochronology 16, 701–717. Roberts, R. G. 1997. Luminescence dating in archaeology: from origins to optical. Radiation Measurements 27, 5/6, 819–892. Roberts, R. G. & R. Jones 2001. Chronologies of carbon and of silica: evidence concerning the dating of the earliest human presence in Northern Australia. In: P.V. Tobias et al. (eds.), Humanity from African naissance to coming millennia. Firenze: Florence University Press, 239–248. Roberts, R. G., R. Jones & M. A. Smith 1990. Thermoluminescence dating of a 50,000-year-old human occupation site in northern Australia. Nature 345, 153–156. Roberts, R. G. et al. 1994.The human colonisation of Australia: optical dates of 53,000 and 60,000 years bracket human arrival at Deaf Adder Gorge, Northern Territory. Quaternary Science Reviews 13, 5–7, 575–583. Roberts, R. G. et al. 1996. Preliminary luminescence dating for archaeological sediments on the Nullarbor Plain, South Australia. Australian Archaeology 42, 7–16. Roberts, R. G. et al. 1997. Luminescence dating of rock art and mud-wasp nests in northern Australia. Nature 387, 696–699. Roberts, R. G. et al. 1998. Single-aliquot and single-grain optical dating confirm thermoluminescence age estimates at Malakunanja II rock shelter in northern Australia. Ancient TL 16, 1, 19–24. Roberts, R. G. et al. 2001. New ages for the last Australian megafauna: continent-wide extinction about 46,000 years ago. Science 292, 1888–1889. Robertson, G. B. & J. R. Prescott 2006. Luminescence dating at the archaeological and human burial site at Roonka, South Australia. Quaternary Science Reviews 25, 19–20, 2586–2593. Roebroeks, W. 2008. Time for the Middle to Upper Paleolithic transition in Europe. Journal of Human Evolution 55, 918–926. Roebroeks, W. & A. Verpoorte 2009. A ‘language-free’ explanation for differences between the European Middle and Upper Paleolithic Record. In: R. Botha & C. Knight (eds.), The cradle of language. New York: Oxford University Press, 150–166. Roebroeks, W. & P. Villa 2011. On the earliest evidence for habitual use of fire in Europe. PNAS 108, 5209–5214. Roebroeks, W., J.-J. Hublin & K. MacDonald 2011. Continuities and discontinuities in Neandertal presence: a closer look at Northwestern Europe. In: N. M. Ashton et al. (eds.), The ancient human occupation of Britain, developments in Quaternary science. Oxford/Amsterdam: Elsevier, 113–123.

304

References Ronquillo,W. P. 1981. The technological and functional analyses of lithic flake tools from Rabel Cave, Northern Luzon, Philippines. Manila: National Museum of the Philippines. Ronquillo, W. P. et al. 1993. The 1992 re-excavation of the Balobok Rockshelter, Sanga Sanga, Tawi Tawi Province, Philippines. Journal of the Historiographical Institute 18, 1–40. Roscoe, P. 2002. The hunters and gatherers of New Guinea. Current Anthropology 43, 1, 153–162. Rose, J. E. 1994. The question of Upper Pleistocene connections between East Africa and South Arabia. Current Anthropology 45, 4, 551–555. Rose, J. I. 2006. Among Arabian sands: defining the Paleolithic of southern Arabia. PhD dissertation, Southern Methodist University, Dallas. Rose, J. I. 2007. The Arabian corridor migration model: archaeological evidence for hominin dispersals into Oman during the Middle and Upper Pleistocene. Proceedings of the Seminar for Arabian Studies 37, 219–237. Rose, J. I. 2010. New light on human prehistory in the Arabo-Persian Gulf Oasis. Current Anthropology 51, 6, 849–883. Rose, J. I. et al. 2011. The Nubian complex of Dhofar, Oman: an African Middle Stone Age industry in southern Arabia. PloS ONE 6, e28239. Rosenberg,T. M. et al. 2011. Humid periods in southern Arabia: windows of opportunity for modern human dispersal. Geology 39, 1115–1118. Rosenberg,T. M. et al. 2012. Late Pleistocene palaeolake in the interior of Oman: a potential key area for the dispersal of anatomically modern humans out-of-Africa? Journal of Quaternary Science 27, 13–16. Rosenfeld, A. 1981. Excavations at the Early Man Shelter. In: A. Rosenfeld, D. Horton & J. Winter (eds.), Early man in North Queensland: art and archaeology in the Laura Area. Canberra: ANU, 5–34. Ross, J. 1997. Painted relationships: an archaeological analysis of a distinctive anthropomorphic rock art motif in Northwest Central Queensland. Unpublished BA (hons) thesis, Department of Archaeology and Palaeoanthropology, University of New England, Armidale. Ross, M. 2005. Pronouns as a preliminary diagnostic for grouping Papuan languages. In: A. Pawley et al. (eds.), Papuan pasts: investigations into the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, Australian National University, 15–65. Ross, J. 2013. A continent of nations: the emergence of regionally distinct rock art styles across Australia. Quaternary International 285, 161–171. Ross, J. & I. Davidson 2006. Rock art and ritual: an archaeological analysis of rock art in Arid Central Australia. Journal of Archaeological Method and Theory 13, 4, 305–341. Ross, J. & M. Smith 2009. An engraved ‘archaic face’ in the Northeastern Simpson Desert. Australian Archaeology 69, 68–70. Rossano, M. J. 2009. Ritual behaviour and the origins of modern cognition. Cambridge Archaeological Journal 19, 2, 243–256. Roth, W. E. 1904. Domestic implements, arts, and manufactures. Brisbane: Home Secretary’s Department. Roth, W. E. 1907. Burial ceremonies, and disposal of the dead. North Queensland Ethnography. Bulletin No. 9. Records of the Australian Museum 6, 5, 365–403. Roth, W. E. 1910. Decoration, deformation, and clothing. North Queensland Ethnography. Bulletin No. 15. Records of the Australian Museum 8, 1, 20–54. Rottman, G. L. 2005. Japanese army in World War II: conquest of the Pacific 1941–42. Oxford: Osprey Publishing. Ruse, M. 2008. Charles Darwin. Oxford: Blackwell. Russell, M. D. 1985. The supraorbital torus: “a most remarkable peculiarity”. Current Anthropology 26, 337–360. Said, E. 1978. Orientalism: western conceptions of the Orient. New York/London: Pantheon Books/Routledge and Kegan Paul. Sali, S. 1985. The Upper Palaeolithic culture at Patne, District Jalgaon, Maharashtra. In: V. N. Misra & P. Bellwood (eds.), Recent advances in Indo-Pacific prehistory: proceedings of the international symposium held at Poona, December 19–21, 1978. Canberra: Indo-Pacific Prehistory Association, 137–158. Sali, S. A. 1989. The Upper Palaeolithic and Mesolithic cultures of Maharashtra. Pune: Deccan College Post Graduate and Research Institute. Sandford, S. & W. J. Arkell 1933. Palaeolithic man and the Nile Valley in Nubia and Upper Egypt. Chicago: University of Chicago.

305

References Sansom, R. 2003. Constraining the adaptationism debate. Biology and Philosophy 18, 493–512. Santa Luca, A. P. 1980. The Ngangdong fossil hominids: a comparative study of a Far Eastern Homo erectus group. New Haven:Yale University. Sather, C. 2003. Transformation of self and community in Saribas Iban death rituals. In: W. D. Wilder (ed.), Journeys of the soul: anthropological studies of death, burial and reburial practices in Borneo. Kota Kinabalu: Borneo Research Council, 175–247. Sathiamurth, E. & H. K.Voris 2006. Maps of Holocene sea transgression and submerged lakes on the Sunda Shelf. Natural History Journal of Chulalongkorn University, Supplement 2, 1–43. Satterthwaite, L. D. 1986. Aboriginal Australian net hunting. Mankind 16, 31–48. Satterthwaite, L. 1987. Socioeconomic implications of Australian Aboriginal net hunting. Man: Journal of the Royal Anthropological Institute 22, 613–636. Satterthwaite, L. & B. Arthur 2005. Technology and material culture. In: B. Arthur & F. Morphy (eds.), Macquarie atlas of Indigenous Australia. Macquarie University: Macquarie Dictionary, 48–65. Sauer, C. O. 1962. Seashore – primitive home of man? Proceedings of the American Philosophical Society 106, 1, 41–47. Saurin, E. 1969. Les recherches préhistoriques au Cambodge, Laos et Vietnam. Asian Perspectives 12, 12–41. Scally,  A. & R. Durbin 2012. Revising the human mutation rate: implications for understanding human evolution. Nature Reviews Genetics 13, 745–753. Scerri, E. 2012. A new stone tool assemblage revisited: reconsidering the ‘Aterian’ in Arabia. Proceedings of the Seminar for Arabian Studies 42, 357–370. Schepartz, L. A., S. Miller-Antonio & D. A. Bakken 2000. Upland resources and the early Palaeolithic occupation of Southern China,Vietnam, Laos, Thailand and Burma. World Archaeology 32, 1, 1–13. Schick, K. D. 1994. The Movius Line reconsidered: perspectives on the Earlier Paleolithic of Eastern Asia. In: R. Corruchini & S. Ciochon (eds.), Integrative paths to the past: paleoanthropological advances in honor of F. Clark Howell. Upper Saddle River, N.J.: Prentice Hall, 569–594. Schick, K. D. & D. Zhuan 1993. Early Paleolithic of China and Eastern Asia. Evolutionary Anthropology 2, 1, 22–35. Schillaci, M. A. 2008. Human cranial diversity and evidence for an ancient lineage of modern humans. Journal of Human Evolution 54, 814–826. Schiller, A. 2001. Talking heads: capturing Dayak deathways on film. American Ethnologist 28, 1, 32–55. Schmidt, P. 2008. Characterization and geological provenance of Jasper that was used for debitages in the archaeological site of Tabon Cave, Philippines. Hukay 14, 3–12. Scholander, P. F. et  al. 1958. Cold adaptation in Australian Aborigines. Journal of Applied Physiology 13, 2, 211–218. Schrire, C. 1982. The Alligator Rivers: prehistory and ecology in western Arnhem Land. Canberra: ANU. Schultz, E. & J. W. Whitney 1986. Upper Pleistocene and Holocene lakes in the An Nafud, Saudi Arabia. Hydrobiologica 143, 175–190. Schwarcz, H. P. & W. J. Rink 1988. Progress in ESR and U-series chronology of the Levantine Paleolithic. In: T. Akazawa et al. (eds.) Neanderthals and modern humans in the Levant. New York: Plenum Press, 57–67. Schwarcz, H. P. et  al. 1988. ESR dates for the hominid burial site of Qafzeh in Israel. Journal of Human Evolution 17, 733–737. Schwartz, J. H. & I. Tattersall 2000. What constitutes Homo erectus? Acta Anthropologica Sinica Supplement 19, 18–22. Schwartz, J. H. & I. Tattersall 2003. The human fossil record. Volume Two: Craniodental morphology of genus Homo (Africa and Asia). Hoboken, N.J.: Wiley-Liss. Schwartz, J. H. & I. Tattersall 2010. Fossil evidence for the origin of Homo sapiens. Yearbook of Physical Anthropology 53, 94–121. Schwartz, J. H. et al. 1994. A diverse hominoid fauna from the late Middle Pleistocene breccia cave of Tham Khuyen, Socialist Republic of Vietnam. Anthropological Papers of the American Museum of Natural History 73, 2–11. Schwartz, J. H. et  al. 1995. A review of the Pleistocene hominoid fauna of the Socialist Republic of Vietnam (excluding Hylobatidae). Anthropological Papers of the American Museum of Natural History 76, 2–24. Scott-Jackson, J., W. Scott-Jackson & J. I. Rose 2009. Paleolithic stone tool assemblages from Sharjah and Ras al Khaimah in the United Arab Emirates. In: M. D. Petraglia & J. I. Rose (eds.) The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 125–138.

306

References Sellato, B. 1994. Review of Through central Borneo – an account of two years’ travel in the land of the head-hunters between the years 1913 and 1917 by C. S. Lumholtz (Singapore: Oxford University Press 1991  [1920]). L’Homme 130, 212–214. Sémah, A.-M. & F. Sémah 2012. The rain forest in Java through the Quaternary and its relationships with humans (adaptation, exploitation and impact on the forest). Quaternary International, 249, 120–128. Sémah, F. et al. 1992. Did they also make stone tools? Journal of Human Evolution 23, 439–446. Sémah, F. et al. 2004. The significance of the Punung karstic area (eastern Java) for the chronology of the Javanese Palaeolithic, with special reference to the Song Terus cave. In: S. Keates et al. (eds.), Quaternary Research in Indonesia. Rotterdam: A. A. Balkema, 45–61. Semenov, S. A. 1964. Prehistoric technology. Bath: Adams and Dart. Serjeantson, S. W. & A.V. S. Hill 1989. The colonisation of the Pacific: the genetic evidence. In: A.V. S. Hill & S. W. Serjeantson (eds.), The colonisation of the Pacific: a genetic trail. Oxford: Clarendon Press, 286–294. Shang, H. & E. Trinkaus 2010. The early modern human from Tianyuan Cave. College Station: Texas A&M University Press. Shang, H. et  al. 2010. An early modern human from Tianyuan Cave, Zhoukoudian, China. PNAS 104, 6573–6578. Shawcross, W. 1998. Archaeological excavations at Mungo. Archaeology in Oceania 33, 183–200. Shea, J. J. 2003. The Middle Palaeolithic of the East Mediterranean Levant. Journal of World Prehistory 17, 313–94. Shea, J. J. 2008. Transitions or turnovers? Climatically-forced extinctions of Homo sapiens and Neanderthals in the East Mediterranean Levant. Quaternary Science Reviews 27, 2253–2270. Shea, J. J. 2011a. Behavioral modernity – Not again. A reply to Porr. Current Anthropology 52, 4, 583–584. Shea, J. J. 2011b. Homo sapiens is as Homo sapiens was. Current Anthropology 52, 1, 1–35. Shea, J. J. 2011c. Refuting a myth about human origins. American Scientist 99, 128–135. Shea, J. J. 2011d. Stone tool analysis and human origins research: Some advice from Uncle Screwtape. Evolutionary Anthropology 20, 48–53. Shea, J. J. 2011e. The archaeology of an illusion: the Middle-Upper Paleolithic transition in the Levant. In: J. M. Le Tensorer et al. (eds.), The Lower and Middle Palaeolithic in the Middle East and neighbouring regions. Liège: Études et recherches archéologiques de l’Université de Liège, 169–182. Shea, J. J. & M. L. Sisk 2010. Projectile technology and Homo sapiens dispersal from Africa to western Eurasia. Paleoanthropology 2010, 100–122. Shen G. et al. 2002. U-series dating of Liujiang hominid site in Guanxi, Southern China. Journal of Human Evolution 43, 817–829. Shennan, S. 2001. Demography and cultural innovation: a model and its implications for the emergence of modern human culture. Cambridge Archaeological Journal 11, 5–16. Sherratt, A. J. 2002. Darwin among the archaeologists: the John Evans nexus and the Borneo Caves. Antiquity 76, 151–157. Shen, G. & V. Michel 2007. Chronological position of modern Homo sapiens sites in China based on U-series dating. L’Anthropologie 111, 157–165. Shen, G. et al. 2001a. U-series dating of hominid site Ganqian Cave at Tubo, Liujiang, Guangxi in South China. Acta Anthropologica Sinica 20, 3, 238–244. Shen, G. et  al. 2001b. U-series dating of Bailiangdong site in Liuzhou, Guangxi, South China. Journal of Stratigraphy 25, 2, 89–106. Shen, G. et al. 2002. U-series dating of Liujiang hominid site in Guangxi, southern China. Journal of Human Evolution 43, 817–829. Shen, G. et al. 2007. Mass spectrometric U-series dating of Laibin hominid site in Guangxi, southern China. Journal of Archaeological Science 34, 2109–2114. Sheppard, P. J. 1997. Characterisation of cherts from sites in southwest Tasmania. Archaeology in Oceania 32, 47–53. Shoocongdej, R. 2006. Late Pleistocene activities at the Tham Lod rockshelter in highland Pang Mapha, Mae Hong Son province, northwestern Thailand. In: E. A. Bacus et al. (eds.), Uncovering Southeast Asia’s past. Singapore: National University of Singapore Press, 22–37. Shott, M. J. 2013. Human colonization and Late Pleistocene lithic industries of the Americas. Quaternary International 285, 150–160. Shutler, R. & M. Mathisen, 1979. Pleistocene studies in the Cagayan Valley of Northern Luzon, Philippines. Journal of the Hong Kong Archaeological Society 8, 105–114.

307

References Shuttler, R. et al. 2004. AMS radiocarbon dates on bone from cave sites in southeast Java, Indonesia, including Wadjak. Modern Quaternary Research in SE Asia 18, 89–94. Siddall, M. et al. 2003.Sea-level fluctuations during the last glacial cycle. Nature 423, 853–858. Sim, R. 1998. The archaeology of isolation? Prehistoric occupation in the Furneaux Group of Islands, Bass Strait, Tasmania. Unpublished PhD Thesis, Department of Archaeology and Natural History, Australian National University, Canberra. Simanjuntak, T. & F. Sémah 1996. A new insight into the Sangiran flake industry. Bulletin of the Indo-Pacific Prehistory Association 14, 22–26. Simanjuntak, T. & I. N. Asikin 2004. Early Holocene human settlement in eastern Java. Bulletin of the IndoPacific Prehistory Association 24, 13–19. Simanjuntak, H. T., B. Prasetyo & R. Handini (eds.) 2001. Sangiran: man, culture, and environment in Pleistocene times. National Research Centre of Archaeology, EFEO dan Yayasan Obor Indonesia. Singer, R. & J. Wymer 1982. The Middle Stone Age at Klasies River Mouth in South Africa. Chicago: University of Chicago Press. Singhvi, A. K., S. U. Deraniyagala & D. Sengupta 1986.Thermoluminescence dating of Quaternary red-sand beds: a case study of coastal dunes in Sri Lanka. Earth and Planetary Science Letters 80, 139–144. Singhvi, A. K. et al. 2010. A ~200 ka record of climatic change and dune activity in the Thar Desert, India. Quaternary Science Reviews 29, 23–24, 3095–3105. Sisk, M. L. & J. J. Shea 2011.The African origin of complex projectile technology: an analysis using tip crosssectional area and perimeter. International Journal of Evolutionary Biology 2011, doi: 10.4061/2011/968012. Skeat,W.W. 1902.The wild tribes of the Malay Peninsula. Journal of the Anthropological Institute of Great Britain and Ireland 32, 124–141. Skoglund, P. & M. Jakobsson 2011. Archaic human ancestry in East Asia. PNAS 108, 18301–18306. Slack, M. J. et  al. 2004. New Pleistocene ages for backed artefact technology in Australia. Archaeology in Oceania 39, 131–137. Smith, S. A. 1918. The fossil human skull found at Talgai, Queensland. Philosophical Transactions of the Royal Society of London. Series B. Containing Papers of a Biological Character 208, 351–387. Smith, J. M. B. 1979. Origins, affinities and distribution of the high altitude flora. In: P. van Royen (ed.), The Alpine flora of Papua New Guinea, Volume 1.Vaduz: J. Cramer, 133–151. Smith, B. 1985. European Vision and the South Pacific. 2nd ed. New Haven:Yale University Press. Smith, M. A. 1989. The case for a resident population in the Central Australian Ranges during full glacial aridity. Archaeology in Oceania 24, 93–105. Smith, M. 1996. Revisiting Pleistocene Macrozamia. Australian Archaeology 42, 52–53. Smith, J. M. B. 2001. Did early hominids cross sea gaps on natural rafts? In: I. Metcalfe et al. (eds.), Faunal and floral migration and evolution in SE Asia-Australia. Lisse: Swets & Zeitlinger, 409–416. Smith, M. A. 2006. Characterising Late Pleistocene and Holocene stone artefact assemblages from Puritjarra rock shelter: a long sequence from the Australian desert. Records of the Australian Museum 58, 371–410. Smith, M.A. 2010. Puritjarra rock shelter: final report on excavations, 1986–90.Technical Reports Series 1. Canberra: National Museum of Australia Press. Smith, M. A. & N. D. Sharp 1993. A revised bibliography of Pleistocene archaeological sites in Australia, New Guinea and Island Melanesia. In: M. A. Smith et al. (eds.), Sahul in review: Pleistocene sites in Australia, New Guinea and Island Melanesia. Canberra: ANU, 283–312. Smith, A. & J. Allen 1999. Pleistocene shell technologies: evidence from Island Melanesia. In: J. Hall & I. McNiven (eds.), Australian coastal archaeology. Canberra: ANU, 291–297. Smith, M., J. R. Prescott & M. J. Head 1997. Comparison of 14C and luminescence chronologies at Puritjarra rock shelter, central Australia. Quaternary Science Reviews 16, 299–320. Smith, M. A., B. Fankhauser & M. Jercher 1998. The changing provenance of red ochre at Puritjarra rock shelter, Central Australia: Late Pleistocene to present. Proceedings of the Prehistoric Society 64, 275–292. Smith, M. A. et al. 2001. New abox AMS-14C ages remove dating anomalies at Puritjarra Rock Shelter. Australian Archaeology 53, 45–47. Smith, M. A. et al. 2008. Human-environment interactions in Australian drylands: exploratory time-series analysis of archaeological records. The Holocene 18, 389–401. Soares, P. et al. 2009. Correcting for purifying selection: an improved human mitochondrial molecular clock. American Journal of Human Genetics 84, 740–759. Soares, P. et al. 2010. The archaeogenetics of Europe. Current Biology 20, R174–R183.

308

References Soejono, R. P. 1984. Prehistoric Indonesia. In P. van de Velde (ed.), Prehistoric Indonesia: a reader. Dordrecht: Foris, 49–78. Soffer, O. 1985. The Upper Paleolithic of the Central Russian Plain. Orlando, Fla.: Academic Press. Soffer, O. 2004. Recovering perishable technologies through use wear on tools. Preliminary evidence for Upper Paleolithic weaving and net making. Current Anthropology 45, 3, 407–425. Söhnge, P. G., D. J. Visser & C. van Riet Lowe 1937. The geology and archaeology of the Vaal river basin. Memoirs of the Geological Survey of South Africa, Pretoria 35, 1–184. Solecki, R. S. 1975. Shanidar IV, a Neanderthal flower burial in northern Iraq. Science 190, 880–881. Solecki, R. S. 1977. The implications of the Shanidar cave Neanderthal flower burial. Annals of the New York Academy of Science 293, 114–124. Solheim, W. G., II 1970. Prehistoric archaeology in Eastern Mainland Southeast Asia and the Philippines. Asian Perspectives 13, 47–58. Solheim, W. 1983. Archaeological research in Sarawak, past and future. Sarawak Museum Journal 32 (n.s. 53), 35–58. Sollas, W. T. 1911. Ancient hunters and their modern representatives. London: Macmillan. Sonakia, A. 1985. Early Homo from the Narmada Valley, India. In: E. Delson (ed.), Ancestors: the hard evidence. New York: Alan Liss Inc., 334–338. Sonakia, A. & H. de Lumley 2006. Narmada Homo erectus  – a possible ancestor of the modern Indian. Comptes Rendues de Palévolution 5, 353–357. Sondaar, P.Y. 1977. Insularity and its effect on mammal evolution. In: M. N. Hecht et al. (eds.), Major patterns in vertebrate evolution. New York: Plenum Publishing Corporation, 671–707. Sondaar, P. Y. 1984. Faunal evolution and the mammalian biostratigraphy of Java. Courier Forschungsinstitut Senckenberg 69, 219–235. Sonneville-Bordes, D. de 1986. Technique Levallois dans L’australie du Sud-Est. Arqueologia 13, 71–75. Soressi, M. & F. d’Errico 2007. Pigments, gravures, parures: les comportements symboliques controversés des Néandertaliens. In: B. Vandermeersch & B. Maureille (eds.), Les Néandertaliens. Biologie et cultures. Paris: Éditions du CTHS, 297–309. Soriano, S., P. Villa & L. Wadley 2007. Blade technology and tool forms in the Middle Stone Age of South Africa: the Howiesons Poort and post-Howiesons Poort at Rose Cottage Cave. Journal of Archaeological Science 34, 681–703. Specht, J. 2005. Revisiting the Bismarcks: some alternative views. In: A. Pawley et al. (eds.), Papuan pasts: studies in the cultural, linguistic and biological history of the Papuan speaking peoples. Canberra: Pacific Linguistics, ANU, 235–88. Speth, J. D. & E. Tchernov 2001. Neanderthal hunting and meat processing in the Near East: evidence from Kebara Cave (Israel). In: C. B. Stanford & H. T. Bunn (eds.), Meat-eating and human evolution. Oxford: Oxford University Press, 52–72. Spriggs, M. 1991. Nissan, the Island in the middle. Summary report on excavations at the north end of the Solomons and the south end of the Bismarcks. In: J. Allen & C. Gosden (eds.), Report of the Lapita Homeland Project. Canberra: ANU, 222–243. Spriggs, M. 1997. The Island Melanesians. Oxford: Blackwell. Spriggs, M. 2001. How AMS dating changed my life. In: M. Jones & P. Sheppard (eds.), Australasian connections and new directions: proceedings of the 7th Australasian Archaeometry Conference. Auckland: Department of Anthropology, University of Auckland, 365–374. Spriggs, M. 2007. The Neolithic and Austronesian expansion within Island Southeast Asia and into the Pacific. In: S. Chiu & C. Sand (eds.), From Southeast Asia to the Pacific. Archaeological Perspectives on the Austronesian Expansion and the Lapita Cultural Complex. Taipei: Academia Sinica, 104–140. Stanner, W. E. H. 1979 (1962). Religion, totemism and symbolism. In: W. E. H. Stanner, White man got no dreaming. Essays, 1938–1973. Canberra: Australian National University Press, 106–143. Steadman, R. G. 1995. Comments on ‘Wind Chill Errors’. Part 1. Bulletin of the American Meteorological Society 76, 9, 1628–1630. Steadman, D.W., J. P.White & J. Allen 1999. Prehistoric birds from New Ireland, Papua New Guinea: extinctions on a large Melanesian island. PNAS 96, 5, 2563–2568. Stefan,V. H. & E. Trinkaus 1998. Discrete trait and dental morphometric affinities of the Tabun C ­mandible. Journal of Human Evolution 34, 443–468. Stegmann, U. 2005. Die Adaptionismus-Debatte. In U. Krohs and G. Toepfer (eds.), Philosophie die Biologie. Eine Einführung. Frankfurt am Main: Suhrkamp, 287–303.

309

References Stephens, M. et al. 2005. Micromorphology of cave sediments in the humid tropics: Niah cave, Sarawak. Asian Perspectives 44, 42–55. Stephens, M. et al. in press. The sedimentology of the West Mouth lithofacies. In: G. Barker et al. (eds.), Archaeological investigations in the Niah Caves, Sarawak, 1954–2004. Cambridge: McDonald Institute for Archaeological Research. Stern, N. 2009. The archaeological signature of behavioural modernity: a perspective from the southern periphery of the modern human range. In: J. J. Shea & D. E. Lieberman (eds.), Transitions in prehistory: essays in honor of Ofer Bar-Yosef. Oxford: Oxbow/David Brown for the American School of Prehistoric Research, 258–288. Stern, N. & B. Marshall 1993. Excavations at Mackintosh 90/1 in western Tasmania: a discussion of stratigraphy, chronology and site formation. Archaeology in Oceania 28, 8–17. Stevenson, J. et al. 2010. Paoay Lake, northern Luzon, the Philippines: a record of Holocene environmental change. Global Change Biology 16, 1672–1688. Stewart, J. R. & C. B. Stringer 2012. Human evolution out of Africa: the role of refugia and climate change. Science 335, 1317–1321. Stimpson, C. M. 2009. Raptor and owl bone from Niah Caves, Sarawak: identifications and morphological variation in the humerus and tarsometatarsus of selected raptors. International Journal of Osteoarchaeology 19, 4, 476–490. Stimpson, C. M. 2011. Late Quaternary environments and human impact in northern Borneo: the evidence from the birds (Aves) and bats (Mammalia: Chiroptera) from the archaeology of the Great Cave of Niah, Sarawak. Unpublished PhD. thesis, Department of Archaeology, University of Cambridge. Stimpson, C. M. 2012. Local scale, proxy evidence for the presence of closed canopy forest in North-western Borneo in the late Pleistocene: bones of Strategy I bats from the archaeological record of the Great Cave of Niah, Sarawak. Palaeogeography, Palaeoclimatology, Palaeoecology 331–332, 136–149. Stiner, M. C. 2002. Carnivory, coevolution, and the geographic spread of the genus. Homo. Journal of Archaeological Research 10, 1, 1–63. Stiner, M. 2006. Middle Palaeolithic subsistence economy in the Mediterranean region. In: E. Hovers & S. L. Kuhn (eds.), Transitions before the transition: evolution and stability in the Middle Paleolithic and Middle Stone Age. New York: Springer, 213–231. Stiner, M. C., N. D. Munro & T. A. Surovell 2000. The tortoise and the hare. Small-game use, the broadspectrum revolution, and Paleolithic demography. Current Anthropology 41, 1, 39–73. Stockton, J. 1981. Radiocarbon dates for archaeological sites in Tasmania. Australian Archaeology 12, 97–101. Stone, T. & M. L. Cupper 2003. Last Glacial Maximum ages for robust humans at Kow Swamp, southern Australia. Journal of Human Evolution, 45, 99–111. Storm, P. 2001a. The evolution of humans in Australasia from an environmental perspective. Palaeogeography, Palaeoclimatology, Palaeoecology 171, 363–383. Storm, P. 2001b. An environmental approach to the fate of Homo erectus in Australasia. In: L. Barham & K. Robson-Brown (eds.), Human roots: Africa and Asia in the Middle Pleistocene. Bristol:Western Academic and Specialist Press Ltd., 203–215. Storm, P. & A. Nelson 1992. The many faces of Wadjak Man. Archaeology in Oceania 27, 7–46. Storm, P. et al. 2005. Late Pleistocene Homo sapiens in a tropical rainforest fauna in East Java. Journal of Human Evolution 49, 536–545. St Périer, R.-S. de 1920. Les migrations des tribus Magdaléniennes des Pyrénées. Revue Anthropologique 30, 136–141. St Périer, R.-S. de 1957. A propos d’une gravure inedited de Massat (Ariége). Essai d’Interprétation de certaines oeuvres d’art pyrénéennes. In: Mélanges Pittard offerts au Professeur Eugéne Pittard ar ses collègues et ses amis en l’honneur de son 90e anniversaire, 5 juin 1957. Brive: Musée et Institut d’Ethnographie de Genéve, Imprimerie Chastrusse, 315–329. St Périer, R.-S. de 1965. Inventaire de l’art mobilier paleolithique du Perigord. Périgueux: Societé Historique et Archeologique du Périord, 139–159. Straus, L. G. 1992. Iberia before the Iberians.The Stone Age prehistory of Cantabrian Spain. Albuquerque: University of New Mexico Press. Straus, L. G., D. J. Meltzer & T. Goebel 2005. Ice Age Atlantis? Exploring the Solutrean-Clovis ‘connection’. World Archaeology 37, 4, 507–532. Stringer, C. 1998. A metrical study of the WLH-50 calvaria. Journal of Human Evolution 34, 327–332.

310

References Stringer, C. 2001. Modern human origins  – distinguishing the models. African Archaeological Review 18, 67–75. Stringer, C. 2002. Modern human origins: progress and prospects. Philosophical Transactions of the Royal Society of London B 357, 563–579. Stringer, C. B. & P. Andrews 1988. Genetic and fossil evidence for the origin of modern humans. Science 239, 1263–1268. Stringer, C. et al. 1989. ESR dates for the burial site of Es Skhul in Israel. Nature 338, 756–758. Su, B. et al. 1999.Y-Chromosome evidence for a northward migration of modern humans into Eastern Asia during the last ice age. American Journal of Human Genetics 65, 6, 1718–1724. Subhavan,V. 2009. Homo erectus in Thailand: a comparative analysis of fossils uncovered from the site of Doi Ta Ka (Loality 1) Lampang Province, Northern Thailand. Paper presented at the 19th Congress of the Indo-Pacific Prehistory Association, Hanoi 2009. Summerhayes, G. R. 2007. Island Melanesian pasts – a view from archaeology. In: J. Friedlaender (ed.), Genes, language and culture history in the Southwest Pacific. New York: Oxford University Press, 10–35. Summerhayes, G. & J. Allen 1993. The transport of Mopir obsidian to late Pleistocene New Ireland. Archaeology in Oceania 28, 144–148. Summerhayes, G. R., M. Leavesley & A. Fairbairn 2009. Impact of human colonisation on the landscape – a view from the Western Pacific. Pacific Science 63, 4, 725–745. Summerhayes, G. R. et  al. 2010. Human adaptation and plant use in Highland New Guinea 49,000 to 44,000 years ago. Science 330, 6000, 78–81. Sun, X. & X. Li 1999. A pollen record of the last 37 ka in deep sea core 17940 from the northern slope of the South China Sea. Marine Geology 156, 227–244. Sun, X. et al. 2000. The vegetation and climate at the last glaciation on the emerged continental shelf of the South China Sea. Palaeogeography, Palaeoclimatology, Palaeoecology 160, 301–316. Sun, X. et al. 2003. Deep-sea pollen from the South China Sea: Pleistocene indicators of East Asian monsoon. Marine Geology 201, 97–118. Sun, C. et al. 2006. The dazzling array of basal branches in the mtDNA macrohaplogroup M from India as inferred from complete genomes. Molecular Biology and Evolution 23, 3, 683–690. Surovell, T. A. & P. J. Brantingham 2007. A note on the use of temporal frequency distributions in studies of prehistoric demography. Journal of Archaeological Science 34, 1868–1877. Surovell, T. A. et  al. 2009. Correcting temporal frequency distributions for taphonomic bias. Journal of Archaeological Science 36, 1715–1724. Swisher, C. C., III., et al. 1994. Age of the earliest known hominids in Java, Indonesia. Science 263, 1118–1121. Swisher, C. C., III, et al. 1996. Latest Homo erectus of Java: potential contemporaneity with Homo sapiens in Southeast Asia. Science 274, 1870–1874. Szabó, K. 2005. Technique and practice: shell-working in the Western Pacific and Island Southeast Asia. Unpublished PhD dissertation, Australian National University. Szabó, K. & J. Amesbury 2012. Molluscs in a world of islands: the use of shellfish as a food resource in the tropical island Asia-Pacific region. Quaternary International 239, 8–18. Szabó, K., M. Swete Kelly & A. Peñalosa 2004. Preliminary results from excavations in the eastern mouth of Ille Cave, northern Palawan. In: V. Paz (ed.), Southeast Asian archaeology: Wilhelm G. Solheim II Festschrift. Diliman: University of the Philippines Press, 209–224. Szabó, K., A. Brumm & P. Bellwood 2007. Shell artifact production at 32,000–28,000 BP in island Southeast Asia. Current Anthropology 48, 701–723. Szabó, K., P. J. Piper & G. Barker 2008. Sailing between worlds: the symbolism of death in northwest Borneo. In: G. Clark et al. (eds.), Islands of inquiry: colonisation, seafaring and the archaeology of maritime landscapes. Canberra: ANU E-press, 149–170. Taçon, P. S. C. 1994. Socialising the landscape: the long-term implications of signs, symbols and marks on the land. Archaeology in Oceania 29, 117–129. Taçon, P. S. C. 2008. Rainbow colour and power among the Waanyi of northwest Queensland. Cambridge Archaeological Journal 18, 163–176. Taçon, P. S. C. & C. Chippindale 1994. Australia’s ancient warriors: changing depictions of fighting in the rock art of Arnhem Land, N.T. Cambridge Archaeological Journal 4, 211–248. Taçon, P. S. C. et al. 2006. Wollemi petroglyphs, N.S.W., Australia: An unusual assemblage with rare motifs. Rock Art Research 23, 227–238.

311

References Taçon, P. S. C. et  al. 2010. Eagle’ s Reach: a focal point for past and present social identity within the Northern Blue Mountains World Heritage Area, Australia. In: I. Domingo Sanz et al. (eds.), Archaeologies of art. Walnut Creek, Calif.: Left Coast Press, 195–214. Takashi,T. 2012. MIS3 edge-ground axes and the arrival of the first Homo sapiens in the Japanese archipelago. Quaternary International 248, 70–78. Tanudirjo, D. 2001. Islands in between: prehistory of the northeastern Indonesian archipelago. Unpublished PhD thesis, Australian National University, Canberra. Teodosio, S. F. R. 2006. A functional analysis of the Arubo stone tools. Unpublished master’s thesis, University of the Philippines, Archaeological Studies Program, Manila. Terra, H. de 1937. The Siwaliks of India and early man. In: G. G. MacCurdy (ed.), Early man. Philadelphia: Lippincott, 257–268. Terra, H. de 1938. Preliminary report on recent geological and archaeological discoveries relating to early man in Southeast Asia. PNAS 24, 10, 407–413. Terra, H. de 1943a. The Pleistocene of Burma. Transactions of the American Philosophical Society 32, 271–339. Terra, H. de 1943b. Pleistocene geology and early man in Java. Transactions of the American Philosophical Society 32, 437–464. Terra, H. de & T. de Chardin 1936. Observations on the Upper Siwalik formations and later Pleistocene deposits in India. Proceedings of the American Philosophical Society 76, 791–822. Terra, H. de & T. Paterson 1939. Studies on the Ice Age in India and associated human cultures. Washington, D.C.: Carnegie Institute. Terra, H. de & H. L. Movius 1943. Research on early man in Burma. Transactions of the American Philosophical Society 32, 271–393. Thangaraj, K. et al. 2005. Reconstructing the origin of Andaman Islanders. Science 308, 996. Thieme, H. 1997. Lower Palaeolithic hunting spears from Germany. Nature 385, 807–810. Thorne, A. G. 1975. Kow Swamp and Lake Mungo: toward an osteology of early man in Australia. Unpublished PhD thesis, University of Sydney. Thorne, A. G. 1976. Morphological contrasts in Pleistocene Australians. In: R. L. Kirk & A. G. Thorne (eds.), The origin of the Australians. Canberra: Australian Institute of Aboriginal Studies, 95–112. Thorne, A. G. 1977a. Separation or reconciliation? Biological clues to the development of Australian society. In: J. Allen et al. (eds.), Sunda and Sahul. London: Academic Press, 187–204. Thorne, A. G. 1977b. Morphological contrasts in Pleistocene Australians. In: R. L. Kirk & A. G.Thorne (eds.), The origin of the Australians. Canberra: Australian Institute of Aboriginal Studies, 95–112. Thorne, A. G. & M. H. Wolpoff 1981. Regional continuity in Australasian Pleistocene hominid evolution. American Journal of Physical Anthropology 55, 337–349. Thorne, A. & R. Sim 1994. The gracile male skeleton from late Pleistocene King Island, Australia. Australian Archaeology 38, 8–10. Thorne, A. & D. Curnoe 2000. Sex and significance of Lake Mungo 3: reply to Brown. Journal of Human Evolution 39, 587–600. Thorne, A. et al. 1999. Australia’s oldest human remains: age of the Lake Mungo 3 skeleton. Journal of Human Evolution 36, 591–612. Tindale, N. B. 1937. Relationship of the extinct Kangaroo Island culture and cultures of Australia, Tasmania and Malaya. Records of the South Australian Museum 6, 39–60. Tindale, N. B. 1957. Culture succession in south eastern Australia from late Pleistocene to the present. Records of the South Australian Museum 13, 1–49. Tindale, N. 1974. Aboriginal tribes of Australia. Their terrain, environmental controls, distribution, limits, and proper names. Canberra: ANU Press. Tobias, P. V. 1992. Piltdown: an appraisal of the case against Sir Arthur Keith. Current Anthropology 33, 243–293. Topinard, P. 1868. Étude sur les Tasmaniens. Mémoires de la Société d’Anthropologie de Paris, 1ère série, 3, 307–329. Topinard, P. 1872. Étude sur les races indigènes de l’Australie. Bulletins de la Société d’anthropologie de Paris, 2nd série, 7, 211–327. Torrence, R. et al. 1999. A Pleistocene landscape in West New Britain, PNG. Australian Archaeology 49, 44–45. Torrence, R. et al. 2004. Pleistocene colonization of the Bismark Archipelago: new evidence from West New Britain. Archaeology in Oceania 39, 3, 101–130.

312

References Toth, N., 1985. The Oldowan reassessed: a close look at early stone artifacts. Journal of Archaeological Science 12, 101–120. Tougard, C. et al. 1996. Extension of the geographic distribution of the giant panda (Ailuropoda) and search for the reasons for its progressive disappearance in Southeast Asia during the latest Middle Pleistocene. Comptes Rendus de l’Académie des Sciences Paris IIa 323, 973–979. Tougard, C. et  al. 1998. Discovery of a Homo sp. tooth associated with a mammalian cave fauna of Late Middle Pleistocene age, Northern Thailand. Journal of Human Evolution 35, 47–54. Toups, M. A. et al. 2011. Origin of clothing lice indicates early clothing use by anatomically modern humans in Africa. Molecular Biology and Evolution 28, 1, 29–32. Tribolo, C. et al. 2009. Thermoluminescence dating of a Stillbay-Howiesons Poort sequence at Diepkloof Rock Shelter (Western Cape, South Africa). Journal of Archaeological Science 36, 730–739. Tringham, R. et al. 1974. Experimentation in the formation of edge damage: a new approach to lithic analysis. Journal of Field Archaeology 1, 171–196. Trinkaus, E. & H. Shang 2008. Anatomical evidence for the antiquity of human footwear: Tianyuan and Sunghir. Journal of Archaeological Science 35, 1928–1933. Trinkaus, E. et  al. 2003. An early modern human from the Pes¸tera cu Oase, Romania. PNAS 100, 20, 11231–11236. Tschirch, A. & G. Glimmann 1896. Archiv der Pharmazie 234, 8, 585–589. Tulang, C. S. 2000. A functional analysis of the reported small flake and blade’ industry from Duyong Cave, Palawan, Philippines. Unpublished manuscript. Archaeological Studies Program, University of the Philippines. Turnbull, P. 1991. ‘Ramsay’s regime’: The Australian Museum and the procurement of Aboriginal bodies, c. 1874–1900. Aboriginal History 15, 108–121. Turnbull, P. 1994.To what strange uses: the procurement and use of Aboriginal peoples’ bodies in early colonial Australia. Voices: the Quarterly Journal of the National Library of Australia 4, 5–20. Turner, W. 1908. The craniology, racial affinities and descent of the Aborigines of Tasmania. Transactions of the Royal Society of Edinburgh 46, 365–403. Turner, W. 1910. The Aborigines of Tasmania. Pt II. The skeleton. Transactions of the Royal Society of Edinburgh 47, 411–454. Turner, A. 1999. Assessing earliest human settlement of Eurasia: late Pliocene dispersions from Africa. Antiquity 73, 363–370. Turney, C. S. M. et al. 2001a. Early human occupation at Devil’s Lair, Southwestern Australia 50,000 years ago, Quaternary Research 55, 3–13. Turney, C. S. M. et al. 2001b. Redating the onset of burning at Lynch’s Crater (North Queensland): implications for human settlement in Australia. Journal of Quaternary Science 16, 8, 767–771. Tylor, E. B. 1894. On the Tasmanians as representatives of Palaeolithic Man. Journal of the Royal Anthropological Society of Great Britain and Ireland 23, 141–152. Tylor, E. B. 1895. On the occurrence of ground stone implements of Australian type in Tasmania. Journal of the Anthropological Institute of Great Britain and Ireland 24, 335–340. Ulm, S. 2006. Australian marine reservoir effects: a guide to ΔR values. Australian Archaeology 63, 57–60. Underhill, P. et al. 2001.The phylogeography of Y chromosome binary haplotypes and the origins of modern human populations. Annals of Human Genetics 65, 43–62. Unrath, G. et al. 1986. An evaluation of use-wear studies: a multi-analyst approach. Early Man News 9/10/11, 117–175. Urquhart, D. 1989. Biology. In: E. Hall (ed.), People and caribou in the Northwest Territories. Yellowknife: Department of Renewable Resources, Government of the Northwest Territories, 67–73. Usik,V. I. et al. 2013. Nubian Complex reduction strategies in Dhofar, southern Oman. Quaternary International 300, 244–266. Valladas, H. et al. 1987. Thermoluminescence dates for the Neanderthal burial site at Kebara in Israel. Nature 330, 159–160. Valladas, H. et al. 1988.Thermoluminescence dating of Mousterian ‘Proto-Cro-Magnon’ remains from Israel and the origin of modern man. Nature 331, 614–616. Valladas, H. et al. 1999. TL dates for the Neanderthal sites of the Amud cave, Israel. Journal of Archaeological Science 26, 259–268. Vandermeersch, B. 2006. Ce que nous apprennent les premières sepultures. Comptes Rendus Palévolution 5, 1–2, 161–167.

313

References Vanderwal, R. & R. Fullagar 1989. Engraved Diprotodon tooth from the Spring Creek locality, Victoria. Archaeology in Oceania 24, 1, 13–16. Vanhaeren, M. 2005. Speaking with beads:The evolutionary significance of personal ornaments. In: F. d’Errico & L. Backwell (eds.), From tools to symbols. Johannesburg: Witwatersrand University Press, 525–553. Vanhaeren, M. & F. d’Errico 2006. Aurignacian ethno-linguistic geography of Europe revealed by personal ornaments. Journal of Archaeological Science 33, 8, 1105–1128. Vanhaeren, M. et al. 2006. Middle Paleolithic shell beads in Israel and Algeria. Science 312, 1785–1788. van Holst Pellekaan, S. 2013. Genetic evidence for the colonization of Australia. Quaternary International 285, 44–58. Vaughan, P. 1985. Use-wear analysis of flaked stone tools. Tucson: University of Arizona Press. Vermeersch, P. M. 2001. ‘Out of Africa’ from an Egyptian point of view. Quaternary International 75, 103–112. Vermeersch, P. M. et  al. 1998. A Middle Palaeolithic burial of a modern human at Taramsa Hill, Egypt. Antiquity 72, 475–484. Verpoorte, A. 2006. Neanderthal energetics and spatial behaviour. Before Farming 2006/3, 1–6. Veth, P. M. 1989. Islands in the Interior: a model for the colonisation of Australia’s arid zone. Archaeology in Oceania 24, 3, 81–92. Veth, P. 1993. Islands in the interior: the dynamics of prehistoric adaptations within the arid zone of Australia. Ann Arbor: University of Michigan Press. Veth, P. 1995. Aridity and settlement in northwest Australia. Antiquity 69, Special Number 265, 733–746. Veth, P. 2006. Social dynamism in the archaeology of the Western Desert. In: B. David et al. (eds.), The social archaeology of Australian Indigenous societies. Canberra: Aboriginal Studies Press, 242–253. Veth, P. et al. 2009. Excavations at Parnkupirti, Lake Gregory, Great Sandy Desert: OSL ages for occupation before the Last Glacial Maximum. Australian Archaeology 69, 1–10. Veth, P. et al. 2011. The role of information exchange in the colonisation of Sahul. In: R. Whallon et al. (eds.), Information and its role in hunter-gatherer bands. Los Angeles: Cotsen Institute of Archaeology Press, 203–220. Villa, P. et al. 2009. The Still Bay points of Blombos Cave (South Africa). Journal of Archaeological Science 36, 441–460. Voormolen, B. 2008. Ancient hunters, modern butchers. Schöningen 13II-4, a kill-butchery site dating from the Northwest European Lower Palaeolithic. Journal of Taphonomy 6, 2, 71–247. Voris, H. K. 2000. Maps of Pleistocene sea levels in Southeast Asia: shorelines, river systems and time durations. Journal of Biogeography 27, 1153–1167. Vos, J. de 1985. Faunal stratigraphy and correlation of the Indonesian hominid sites. In: E. Delson (ed.), Ancestors: the hard evidence. New York: Alan R. Liss, 215–220. Vos, J. de, A. Bouteaux & A. Bautista 2007a. The mammalian faunas chronology in island Southeast Asia. In: A.-M. Sémah et al. (eds.), First islanders: human origins patrimony in Southeast Asia. Paris: Semenanjung – HOPsea, 81–84. Vos, J. de, L. W. van den Hoek Ostende & G. D. van den Bergh 2007b. Patterns in insular evolution of mammals: a key to island palaeogeography. In:W. Renema (ed.), Biogeography, time, and place: distributions, barriers, and islands. New York: Springer, 315–345. Vu The Long, J. de Vos & R. L. Ciochon 1996. The fossil mammal fauna of the Lang Trang caves,Vietnam, compared with Southeast Asian fossil and recent mammal faunas; the geographical implications. Bulletin of the Indo-Pacific Prehistory Association 14, 101–109. Wadley, L. 2001. What is cultural modernity? A general view and a South African perspective from Rose Cottage Cave. Cambridge Archaeological Journal 11, 201–221. Wadley, R. L. 2004. Punitive expeditions and divine revenge: oral and colonial histories of rebellion and pacification in Western Borneo, 1886–1902. Ethnohistory 51, 3, 609–636. Wadley, L. 2010. The effect of organic preservation on behavioural interpretations at the South African Middle Stone Age sites of Rose Cottage and Sibudu. In: K. V. Boyle et al. (eds.), The Upper Palaeolithic revolution in global perspective. Cambridge: McDonald Institute Monographs, 53–61. Wadley, L.T. Hodgskiss & M. Grant 2009. From the cover: implications for complex cognition from the hafting of tools with compound adhesives in the Middle Stone Age, South Africa. PNAS 106, 24, 9590–9594. Waguespack, N. M. et  al. 2009. Making a point: wood-versus stone-tipped projectiles. Antiquity 83, 321, 786–800.

314

References Wahida, G. et al. 2009. A Middle Paleolithic assemblage from Jebel Barakah, coastal Abu Dhabi Emirate. In: M. D. Petraglia & J. I. Rose (eds.), The evolution of human populations in Arabia: paleoenvironments, prehistory and genetics. Dordrecht: Springer, 117–124. Wallace, A. R. 1876. The geographical distribution of animals. London: Harper and Brothers. Wallace, A. R. 1889. Darwinism. An exposition of the theory of natural selection with some of its applications. London: Macmillan. Wallace, A. R. 1913. The Malay Archipelago. 10th ed. London: Macmillan. Wallis, L. 2001. Environmental history of northwest Australia based on phytolith analysis at Carpenter’s Gap. Quaternary International 83–85, 103–117. Walsh, G. L. 1988. Australia’s greatest rock art. Bathurst: E. J. Brill – Robert Brown & Associates. Walters, I. 1988. Fish hooks: evidence for dual social systems in Southeastern Australia. Australian Archaeology 27, 98–114. Wang, W. et al. 2006. Sequence of mammalian fossils, including hominoid teeth, from the Bubing Basin, South China. Journal of Human Evolution 52, 4, 370–379. Wang, Y. et  al. 2008. Millennial- and orbital-scale changes in the East Asian monsoon over the past 224,000 years. Nature 451, 1090–1093. Wang, X. M. et al. 2009.Vegetation on the Sunda Shelf, South China Sea, during the Last Glacial Maximum. Palaeogeography, Palaeoclimatology, Palaeoecology 278, 88–97. Watanabe, H. 1985. The chopper-chopping tool complex of eastern Asia: an ethnoarchaeological approach. Journal of Anthropological Archaeology 4, 1–18. Watchman, A. & E. Hatte 1996. A nano approach to the study of rock art: ‘The Walkunders’, Chillagoe, north Queensland, Australia. Rock Art Research 13, 85–92. Watchman, A. L., S. O’Connor & R. Jones 2005. Dating oxalate minerals 20–45 ka. Journal of Archaeological Science 32, 369–374. Watchman, A. L. et al. 2001. Spatial and compositional variations within finely laminated mineral crusts at Carpenter’s Gap, an archaeological site in tropical Australia. Geoarchaeology 16, 803–824. Wayland, J. 1934. Rifts, rivers, rains and early man in Uganda. Journal of the Royal Anthropological Institute of London 64, 333–352. Weaver, T. D. 2009. The meaning of Neandertal skeletal morphology. PNAS 106, 16028–16033. Webb, S. G. 1989. The Willandra Lakes hominids. Canberra: ANU. Webb, S. 1990. Cranial thickening in an Australian hominid as a possible palaeoepidemiological indicator. American Journal of Physical Anthropology 82, 403–411. Webb, C. & J. Allen 1990. A functional analysis of Pleistocene bone tools from two sites in south west Tasmania. Archaeology in Oceania 25, 75–78. Webb, S., M. L. Cupper & R. Robins 2006. Pleistocene human footprints from the Willandra Lakes, southeast Australia. Journal of Human Evolution 50, 405–413. Weidenreich, F. 1943. The skull of Sinanthropus pekinensis; a comparative study on a primitive hominid skull. Pehpei: Geological Survey of China. Weidenreich, F. 1945. The Keilor skull: a Wadjak type from Southeast Australia. American Journal of Physical Anthropology 3, 1, 21–32. Weidenreich, F. 1951. Morphology of Solo Man. Anthropological Papers of the American Museum of Natural History 43, 222–290. Welch, D. M. 1996. Material culture in Kimberley rock art, Australia. Rock Art Research 13, 2, 104–123. Wendorf, F., R. Schild & A. Close 1993. Egypt during the last Interglacial: the Middle Paleolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press. Westaway, M. 2002. Preliminary observations on the taphonomic processes at Ngandong and some implications for a late Homo erectus survivor model. Tempus 7, 189–193. Westaway, M. & C. P. Groves 2009. The mark of ancient Java is on none of them. Archaeology in Oceania 44, 40–51. Westaway, K. E. et al. 2007a. Age and biostratigraphic significance of the Punung rainforest fauna, East Java, Indonesia, and implications for Pongo and Homo. Journal of Human Evolution 53, 709–717. Westaway, K. E. et al. 2007b. Establishing the time of initial human occupation of Liang Bua, Western Flores, Indonesia. Quaternary Geochronology 2, 1–4, 337–343. Westaway, K. E. et al. 2009. Reconstructing the geomorphic history of Liang Bua, Flores, Indonesia: a stratigraphic interpretation of the occupational environment. Journal of Human Evolution 57, 5, 465–483.

315

References Whalen, N. et al. 1981. Saudi Arabian archaeological reconnaissance 1980: preliminary report on the western province survey. Atlal 5, 43–58. White, J. P. 1972. Ol Tumbuna. Canberra: ANU. White, J. P. 1977. Crude, colourless and unenterprising? Prehistorians and their views on the stone age of Sunda and Sahul. In: J. Allen, J. Golson & R. Jones (eds.), Sunda and Sahul: prehistoric studies in Southeast Asia, Melanesia and Australia. London: Academic Press, 13–30. White, R. 1985. Upper Paleolithic land use in the Perigord. A topographic approach to subsistence and settlement. Oxford: Archaeopress. White, M. J. 2000. The Clactonian question: on the interpretation of core and flake assemblages in the British Lower Palaeolithic. Journal of World Prehistory 14, 1–63. White, R. 2007. Systems of personal ornamentation in the early Upper Palaeolithic: Methodological challenges and new observations. In: P. Mellars et  al. (eds.), Rethinking the human revolution. Cambridge: McDonald Institute for Archaeological Research, 287–302. White, J. P. 2011. Revisiting the ‘Neolithic problem’ in Australia. Records of the Western Australian Museum, Supplement 79, 86–92. White, T. D. et al. 2003. Pleistocene Homo sapiens from Middle Awash, Ethiopia. Nature 423, 742–747. Wickler, S. 1990. Prehistoric Melanesian exchange and interaction: recent evidence from the Northern Solomon Islands. Asian Perspectives 29, 135–154. Wickler, S. 2001. The prehistory of Buka: a stepping stone island in the Northern Solomons. Canberra: Centre of Archaeological Research, Australian National University. Willoughby, P. R. 2007. The evolution of modern humans in Africa: a comprehensive guide. Plymouth: Altamira Press. Winney, B. J. et al. 2004. Crossing the Red Sea: phylogeography of the hamadryas baboon, Papio hamadryas hamadryas. Molecular Ecology 13, 2819–2827. Wiessner, P. 1983. Style and social information in Kalahari San projectile points. American Antiquity 48, 253–276. Wiessner, P. & A.Tumu 1998. Historical vines. Enga networks of exchange, ritual, and warfare in Papua New Guinea. Washington, D.C.: Smithsonian Institution Press. Winterhalder, B. & E. A. Smith (eds.) 1992. Evolutionary ecology and human behavior. New York: Aldine de Gruyter. Witter, D. 2007. Aboriginal Dreaming and aridity. In: C. Dickman et al. (eds.), Animals of arid Australia: out on their own? Mosman, Sydney: Royal Zoological Society of New South Wales, 14–29. Wolfe, P. 1991. On being woken up: the dreamtime in anthropology and in Australian settler culture. Comparative Studies in Society and History 33, 2, 197–224. Wolpoff, M. H. 1999. Paleoanthropology. 2nd ed. Boston: McGraw Hill. Wolpoff, M. H. & R. Caspari 1997a. Race and human evolution: a fatal attraction. New York: Simon and Schuster. Wolpoff, M. H. & R. Caspari 1997b. What does it mean to be modern? In G. A. Clark and C. M. Willermet (eds.), Conceptual issues in modern human origins research. New York: Aldine de Gruyter, 28–44. Wolpoff, M. H.,Wu Xinzhi & A. G.Thorne 1984. Modern Homo sapiens origins: A general theory of hominid evolution involving the fossil evidence from East Asia. In: F. Smith & F. Spencer (eds.), The origins of modern humans: a world survey of the fossil evidence. New York: Alan R. Liss, 411–483. Wolpoff, M. H., J. Hawks & R. Caspari 2000. Multiregional, not multiple origins. American Journal of Physical Anthropology 112, 129–136. Wolpoff, M. H. et al. 1994. The case for sinking Homo erectus. 100 years of Pithecanthropus is enough! Courier Forschungsinstitut Senckenberg 171, 341–361. Wolpoff, M. H. et al. 2001. Modern human ancestry at the peripheries: a test of the replacement theory. Science 291, 293–297. Woodroffe, C. D. 1990. The impact of sea-level rise on mangrove shorelines. Progress in Physical Geography 14, 4, 483–520. Wright, R. V. S. (ed.) 1971. Archaeology of the Gallus Site, Koonalda Cave. Canberra: Australian Institute of Aboriginal Studies. Wroe, S. et al. 2003. An alternative method for predicting body mass: the case of the Pleistocene marsupial lion. Paleobiology 29, 3, 403–411. Wu, X. 2004. On the origins of modern humans in China. Quaternary International 117, 131–140.

316

References Wu Xinxhi & F. E. Poirier 1995. Human evolution in China: a metric description of the fossils and a review of the sites. New York/Oxford: Oxford University Press. Wunderly, J. 1939.The cranial and other skeletal remains ofTasmanians in the collections of the Commonwealth of Australia. Biometrika 30, 305–337. Wunderly, J. 1943. The Keilor fossil skull: anatomical description. Memoirs of the National Museum, Melbourne 13, 57–69. Wurster, C. M. et al. 2010. Forest contraction in north equatorial Southeast Asia during the last glacial period. PNAS 107, 35, 15508–15511. Wurz, S. G. D. 1999. The Howiesons Poort at Klasies River: an argument for symbolic behaviour. South African Archaeological Bulletin 54, 38–50. Wurz, S. G. D. 2002.Variability in the Middle Stone Age lithic sequence, 115,000–60,000 years ago at Klasies River, South Africa. Journal of Archaeological Science 29, 1001–1015. Xhauflair, H. 2009. Analyse tracéologique et recherche de résidus: contribution à l’étude de l’industrie lithique de Tabon Cave, Palawan, Philippines. Unpublished master’s thesis, Muséum national d’Histoire naturelle, Paris. Xiang Anqiang 1990. Palaeolith discovered in the Lishui Basin, the Dong-ting Lake region. Southern Ethnology and Archaeology 3, 249–270. Xie Guangmao 1990. Relation between the stone tools from Baise and the Lower Palaeolithic cultures from South and Southeast Asia. Southern Ethnology and Archaeology 3, 237–247. Yamada, S. 1993. The formation process of ‘use-wear polishes’. In: P. Anderson et al. (eds.), Traces et fonction: les gestes retrouvés. Liège: Université de Liège, 433–445. Yellen, J. et  al. 2005. The archaeology of Aduma Middle Stone Age sites in the Awash Valley, Ethiopia. PaleoAnthropology 10, 25–100. Yengoyan, A. P. 1968. Demographic and ecological influences on Aboriginal marriage sections. In: R. B. Lee & I. DeVore (eds.), Man the hunter. Chicago: Aldine, 185–199. Yi, S. & G. Clark 1983. Observations of the Lower Palaeolithic of Northeast Asia. Current Anthropology 24, 181–202. Yi, S. et al. 2008. New data on the Hoabinhian: investigations at Hang Cho cave, northern Vietnam. Bulletin of the Indo-Pacific Prehistory Association 28, 73–79. Yokoyama Y. et al. 2008. Gamma-ray spectrometric dating of late Homo erectus skulls from Ngandong and Sambungmacan, Central Java, Indonesia. Journal of Human Evolution 55, 274–277. Zarins, J., A. J. Murad & K. S. Al-Yish 1981. The comprehensive archaeological survey program. The second preliminary report on the southwestern province. Atlal 5, 9–42. Zilhão, J. 2007. The emergence of ornaments and art: an archaeological perspective on the origins of behavioral modernity. Journal of Archaeological Research 15, 1–54. Zilhão, J. et al. 2010. Symbolic use of marine shells and mineral pigments by Iberian Neandertals. PNAS 107, 3, 1023–1028. Zuraina Majid, D. P. 1982. The West Mouth, Niah, in the prehistory of Southeast Asia. Sarawak Museum Journal 31 (n.s. 52), 1–200. Zuraina Majid, D. P. 1994. The excavation of Perak man, an epi-palaeolithic burial at Gua Gunung Runtuh. In: D. P. Zuraina Majid (ed.), The excavation of Gua Gunung Runtuh and the discovery of the Perak Man in Malaysia. Kuala Lumpur: Department of Museums and Antiquity, Malaysia, 23–47. Zuraina Majid, D. P. 1997. The discovery of Bukit Jawa, Gelok, a middle-late palaeolithic site in Perak, Malaysia. Journal of the Malaysian Branch of the Royal Asiatic Society 70, 2, 49–52. Zuraina Majid, D. P. 2001. Man and technology in Late Pleistocene Malaysia. In: T. Simanjuntak et al. (eds.), Sangiran: man, culture, and environment in Pleistocene times. Jakarta:Yayasan Obor Indonesia, 185–192. Zuraina Majid, D. P. 2003. Archaeology in Malaysia. Pulau Pinang: Pusat Penyelidikan Arkeologi Malaysia. Zuraina Majid, D. P. & H. D. Tjia 1998. Kota Tampan, Perak. The geological and archaeological evidence for a Late Pleistocene site. Journal of the Malaysian Branch of the Royal Asiatic Society 61, 123–134. Zuraina Majid, D. P. et al. 1998. Some interesting Late Pleistocene-Early Holocene finds from excavations in Balambangan Island, Sabah, Malaysia. Sabah Society Journal 15, 29–40.

317

Index

abalone, 158 Abbie, Andrew Arthur, 27 ABOX, 96, 99, 104 abstract thought, 6, 203, 204, 247, 255 Acheulean, 8, 14, 15, 16, 19, 72, 73, 74, 75, 137, 249, 250 adaptation, 20, 55, 65, 69–70, 72, 75, 76, 83, 88, 118, 121, 126, 146, 147, 149, 152, 164, 167, 173–174, 177, 180, 188, 190, 192, 196, 198, 202, 206, 211, 212, 213, 217, 222–224, 226, 253, 257, 262–264 Admiralty Islands, 217 admixture, 45, 70, 73, 76, 85, 231, 232, 233, 234, 235, 236, 237, 238, 242 adornment. See personal ornamentation Aduma, 78 Africa, 63, 115, 193, 203, 209, 210, 211, 246, 260 clothing, 197 environment, 62, 65 fossil record, 2, 10, 26, 54, 106, 236 Late Stone Age, 166, 197 lithic technology, 5, 8, 10, 54, 60, 62, 63, 65, 73, 77–78, 80, 81, 82, 83, 87, 88, 166, 249, 261, 262 Middle Stone Age, 5, 6, 54, 65, 73, 77, 78, 87, 161, 163, 186, 197, 201, 202, 209, 210, 240, 261, 262 modern human origins, 2, 5, 6, 9, 10, 30, 48, 51, 55, 56, 70, 77, 85, 257, 261 out of, 4, 5, 31, 39, 46, 48, 49, 51, 54, 55, 56, 57, 58, 59, 62, 63, 64, 65, 67, 70, 71, 73, 74, 76, 77, 78, 119, 174, 213, 228, 229, 236, 238, 239, 242, 248, 260, 261 subsistence strategies, 229 Ailuripoda melanoleuca, 40 Airport Mound, 217 Aitape, 29 Allen, Harry, 29 Allen’s Cave, 158, 169 Altamira, 211 AMH. See anatomically modern humans Amud Cave, 34

anatomically modern humans, 2, 25, 30, 80, 89, 118, 126, 133, 135, 139, 140, 142, 146, 148, 166, 189, 228, 229, 231, 232, 233, 234, 235, 236, 237, 238, 239, 240, 242, 262 Andaman Islands, 23, 57 Andra Pradesh, 82 Anir Islands, 218 anthropoid apes, 25 Anyathian culture, 13 Arabian Peninsula, 2, 33, 49, 51, 54, 56, 57, 76, 238 colonisation, 4, 51, 56–63, 65, 76, 80, 237, 239 environment, 51, 61, 62 fossil record, 48, 54, 55, 56, 63, 238 lithic technology, 54, 57–61, 62, 80–81, 87, 240 Middle Palaeolithic, 59, 60 subsistence strategies, 61, 62 Upper Palaeolithic, 57, 58, 59 Arabian Sea, 69 archaic hominins, 6, 10, 46, 51, 70, 112–113, 136, 140, 231, 235, 244, 246 Arcy-sur-Cure, 161 aridification, 61, 150, 151, 152, 153, 158 Arnhem Land, 46, 86, 87, 152, 153, 155, 159, 170 art, 106, 149, 152, 153, 159, 160, 163, 175, 177, 184, 192–193, 197, 200, 201, 261 mobiliary, 160, 193, 197, 209, 210, 211 Aru Islands, 173 Arubo, 137, 138 Arudy, 211 ASTT. See Australian Small Tool Tradition Aterian industry, 59, 60, 77, 78, 81 Aurignacian, 80, 146, 240 Australia, 3, 6, 33, 42, 43, 150, 243 art, 154–155, 157, 159–160, 184, 205–206, 251 clothing, 193–194 colonisation, 3, 6, 31, 46, 55, 56, 76, 82, 83, 87, 108, 115, 140, 146, 174, 186, 189, 229, 233, 241, 242, 244, 245–247 environment, 150, 151–152, 158, 162, 167, 177–179, 180, 194 fauna, 112, 114, 116, 167, 168, 169, 170

319

Index Australia (cont.) fossil record, 26, 28 lithic technology, 16, 26, 28, 73, 85–87, 88, 164–171, 177, 180, 185, 194, 206, 240, 248–250, 255, 263 skeletal evidence, 28–31, 46–47, 55, 67, 179–180, 234 skeletal record, 248 skull collecting, 22 subsistence strategies, 150, 152, 167–168, 169, 170–171, 177, 182–184, 229, 252–253 Western perceptions, 3, 165–167 Australian Aborigines, 150, 151, 152, 154, 155, 190, 193 as relics, 27–28 genetics, 54, 70, 233, 234, 235, 241 morphology, 29–31, 46–47, 55 Western perceptions, 3, 23, 25, 26, 27–28, 164, 166, 263 Australian Core Tool and Scraper Tradition, 164, 166, 169 Australian Small Tool Tradition, 166 Australoid, 30, 93 Australopithecines, 26 Australopithecus afarensis, 247 Australopithecus africanus, 10 Australopithecus tranvaalensis, 10 Austronesians, 114–115, 116, 143 awl, 168, 191, 196, 197 Aybut Al Auwal, 60, 80 Bab el Mandeb, 56, 57, 61, 69, 78 baboons, 62, 63 Badger Island, 180 Bailiangdong Cave, 45 Baki, 220 Baler shell, 154 Bali, 229, 241, 248 Ballawinne, 159 Ballawinne Cave, 184 Balof 2, 206 bamboo tool hypothesis, 8, 142, 147, 171, 174 Bamburi, 72 Bandung, 119, 120 Bangladesh, 48 Banks, Joseph, 22 Bass Strait islands, 27, 180 Batadomba Lena, 39, 65, 67, 72 Batari, 206 Bau Caves, 90, 92 beads, 39, 73, 82, 106, 144, 154, 157, 158, 160, 161, 162, 192, 193, 197, 200, 204, 206, 207, 208, 209, 220 Beaton Shelter, 180 behavioural flexibility, 6, 126, 129, 132–134, 139, 146, 147, 149, 173, 174, 185, 188, 210, 211, 212, 222, 224, 226, 244, 262

320

behavioural modernity, 6, 51, 65, 77, 106, 142, 144, 164, 193, 198, 200, 202, 209, 210, 243, 244, 255, 257, 258, 260, 261, 262, 263 markers of, 6, 77, 87, 106, 146, 175, 177, 185, 186–187, 188, 189–190, 192, 194, 195, 196, 197, 198, 201, 203, 204, 207, 208, 209, 210, 211, 248, 249, 260, 264 origins of, 260–262 package of traits, 3, 65, 73, 106, 118, 134, 137, 143, 146, 148–150, 189, 190, 197, 198, 200, 201, 202, 204, 209, 211, 244, 247, 260–261 Belilena Kitugala, 39 Bennett’s wallaby, 152, 183, 194, 208 Berry, Richard J. A., 24, 25 Bhan-Kao, 14 biface, 8, 59, 60, 61, 78, 79, 80, 85, 87, 137, 138, 161, 162, 166, 170, 250 biogeography, 51, 62, 108–116, 135, 151, 161 Bir Tirfarwi, 78 Birdsell, Joseph, 26, 27, 214 Bismarck Archipelago, 114, 115, 154, 213, 214, 218, 219, 225, 226 Black Current, 140 blade, 2, 3, 8, 13, 15, 58, 59, 60, 65, 66, 67, 72, 73, 74, 75, 77, 78, 79, 80, 82, 83, 85, 87, 142, 143, 166, 167, 171, 174, 177, 182, 188, 192, 196, 197, 209, 249, 260 Blombos Cave, 2, 82, 162 Blumenbach, Johann Friedrich, 22 Bobongara, 85, 86, 170, 214, 216, 221, 223 body painting, 193, 194 Boker Tachtit, 80 Bolomor Cave, 188 bonding behaviour, 152, 153–155, 160, 161, 162, 163 Bone Cave, 180, 181, 182, 183, 195 bone tools, 65, 73, 77, 82, 106, 146, 162, 164, 168, 173, 182, 188, 192, 194, 195, 196, 198, 200, 201, 204, 206, 208 boomerangs, 210 Borneo, 3, 5, 33, 43, 48, 55, 67, 85, 109, 112, 120, 121, 122, 129, 135, 138, 140, 146 bounding behaviour, 152, 153, 155–160, 161, 162, 163 Bowler, Jim, 29 Braholo Cave, 126 Breuil, Henri, 11, 12 British Museum, 93, 99 Broca, Paul, 22, 24 Broken Hill, 10 Broom, Robert, 10 Broome, 116 Brothwell, Don, 93, 99 Brown, Peter, 30, 31 Buang Merabak, 158, 206, 207, 214, 216, 218, 219, 220, 225

Index Buka Island, 217, 218 Bukit Jawa, 15 burial, 29, 33, 39, 40, 45, 46, 55, 67, 76, 78, 92, 95, 105, 106, 139, 149, 152, 153, 154, 157, 158, 160, 169, 175, 177, 179, 180, 192, 200, 203, 204, 205, 206, 209, 260, 261 Burma, 10, 11, 12, 13, 14, 19, 48, 49 Busk, George, 90 butchery, 96, 104, 172, 182, 183, 185, 187 C99, 169 Cabalwanian industry, 137 Cagayan Valley, 137, 142 Callao Cave, 5, 44, 55, 118, 119, 126, 132, 136, 137, 139–142, 143, 146, 171, 173 Cape Range, 169 Cape York Peninsula, 155 Carnarvon Gorge, 154 Carpenter’s Gap, 85, 151, 154, 158, 170, 173, 205 Caucasoid, 30 cave art. See rock art Cave Bay Cave, 180, 206 Cave Period, 26 cave taphonomy, 40, 98 Cebu Island, 141 Celebes Sea, 135 cemeteries, 149, 152, 153, 157, 158, 160, 163, 206 Central Asia, 10, 17, 34, 69, 232, 233 Chao Phraya River, 48 Chardin, Teilhard de, 10, 11, 12, 13, 15, 18 Châtelperronian, 160, 161 Cheetup, 206 Chillagoe, 154, 159 China, 8, 10, 12, 15, 18, 19, 30, 33, 48, 49, 119, 232, 235–236 colonisation, 5, 19, 77, 239 fossil record, 6, 9, 19, 26, 27, 44–46, 235, 236, 238, 240, 242, 248 lithic technology, 8, 138 chopper-chopping tool industries, 14 Choukoutien. See Zhoukoudian cicatrices. See scarification climate change, 54, 104, 112, 132, 150, 151, 160, 162, 168, 180, 183, 196, 197, 213, 221, 253 clo unit, 191 Cloggs Cave, 194 clothing, 4, 106, 168, 182, 189–198, 206, 210, 249 complex, 191, 192, 193, 194, 196, 197, 198 simple, 191, 193, 194, 197, 198 Coastal Colonisation Model, 115 cognition, 2, 6, 143, 166, 167, 174, 175, 186, 187, 196, 200, 201, 208, 210, 243, 244, 245–247, 250, 255, 257, 258, 259, 260, 261, 263 Cohuna, 28, 29 cold exposure, 190

cold tolerance, 190, 191 complex behaviour, 119. See also behavioural modernity composite tool technology, 146, 147 Congo River, 161 Coobool Creek, 31, 160, 206, 210, 234 Cook, James, 22, 190 Coon, Carleton, 28, 30 core, 8, 13, 14, 15, 58, 60, 72, 77, 78, 80, 82, 83, 85, 86, 87, 88, 137, 139, 143, 166, 169, 171, 172, 173, 180, 220, 250 discoidal, 61, 72, 77, 78, 87, 171 horsehoof, 137, 166, 249 multiplatform, 59, 80, 83, 85, 86 prepared, 13, 72, 73, 83, 137 radial, 78, 83, 85, 86, 87 reduction, 60, 77, 80, 85, 137, 166, 171 Cosquer, 184 cranial deformation, 205, 206, 209, 210 craniometry, 22–25, 27, 28 cremation, 29, 46, 169, 205 Crocuta crocuta ultima, 40 Cro-Magnon, 25 CTST. See Australian Core Tool and Scraper Tradition Cuddie Springs, 154, 170 cultural complexity, 201, 202 cultural modernity. See behavioural modernity culture contact, 152, 159, 161 curation, 144 Curnoe, Darren, 31 cuscus, 110, 206, 225, 246 Darling River lakes, 169 Dart, Raymond, 10, 26 Darwin, 116 Darwin, Charles, 6, 9, 18, 21, 22, 24, 90, 92, 190, 258, 259 Darwin Glass, 154, 182, 207 Davis, Joseph Barnard, 24 Denisova Cave, 49, 133, 232, 233, 248 Denisovans, 48, 49, 232, 248 distribution, 133 interbreeding, 49, 70, 114, 186, 232–236, 242, 248 descent of man,The, 258, 259 Devil’s Lair, 85, 106, 158, 167, 168, 173, 194, 206 Devon Downs, 26, 29, 166 Dhofar, 60, 80, 238 di-hybrid theory, 248 dingo, 116 Discontinuous Dreaming Network Model, 154 Dobzhansky, T., 259, 261 Doi Ta Ka Cave, 40 Dreaming tracks, 152, 153, 154, 155 Drift Period, 26

321

Index Dubois, Eugène, 26, 43 Duke of York Islands, 218 Duoi Oi, 41 Duruthy, 211 Duyong Cave, 138, 143 Early Man, 155 East Asia, 10, 14, 17, 19, 44–46, 49, 116, 248 biogeography, 10 colonialism, 16–17 colonisation, 48, 238, 240 Early Palaeolithic, 8, 11 fossil record, 48, 55, 235 human origins, 9, 10, 49 lithic technology, 8, 10, 16 Western perceptions, 18 East Timor, 85, 114, 146, 174, 219 fauna, 174 lithic technology, 171–172 subsistence strategies, 171–172, 229 Egypt, 78, 80, 87 El Castillo, 211 Elephas maximus, 41 emblemic behaviour, 155, 157, 158, 159, 160, 161, 162 Enkapune Ya Muto, 78 Enlène, 211 environmental barriers, 49, 69, 83, 109, 113, 151, 225, 237, 247 environmental change, 51, 60, 61–62, 63, 66, 69, 83, 119–121, 126, 128, 129, 150, 151, 160, 161, 162, 179, 180, 192, 226 Eoanthropus dawsoni, 259 estuaries, 115 Etheridge, R., 26 Ethiopia, 78, 106 ethnography, 106, 153, 155, 158, 181, 182, 190, 193, 218, 250, 252 Europe, 2, 10, 11, 16, 17, 22, 26, 50, 71, 149, 166, 175, 177, 196, 244, 249 art, 160, 193 clothing, 197 colonisation, 3, 76, 166, 228, 231, 237 environment, 178, 196 fossil record, 27, 240 lithic technology, 10, 11, 15, 82, 143, 197 Middle Palaeolithic, 160, 163, 196, 202, 208, 209, 210, 211, 244 Neolithic, 198 subsistence strategies, 188, 194 Upper Palaeolithic, 4, 160, 161, 163, 186, 196, 197, 198, 201, 202, 210, 239, 244, 260, 263 Upper Palaeolithic industry, 77 Everett, A. Hart, 90, 92 evolution, 27, 51, 54, 108, 110, 175, 185, 186, 243, 255, 256, 257, 258, 259, 260, 262, 263

322

adaptive radiation, 110, 114 missing link, 90, 92 theory of, 21, 22 evolutionary superiority, 22 expedient tool technologies, 137, 143, 144, 146, 167, 171, 172, 220, 222 expression of emotions in man and animals,The, 259 extinction, 49, 54, 56, 71, 108, 112, 187, 232 Fa Hien Cave, 39, 67 Fasad technology, 60 faunal dispersal, 109–110, 112 faunal translocation, 110, 175, 206, 225, 246 fibre, 168, 169, 172, 173, 174 Fiji, 114, 115, 235 fire, 190, 192, 197 fish hook, 122, 132, 167, 172, 174, 219 fisher-gatherer-hunter economies, 116 fisher-gatherer-hunter-farmer economies, 116 fishing, 132, 171, 172, 204, 219 pelagic, 114, 132, 146, 171, 175, 206, 207, 219 Flinders, Matthew, 26 Florentine River, 179 Flores, 15, 33, 109, 110, 112, 113, 136, 140, 244, 245, 247, 248, 255 colonisation, 112, 113 fauna, 174 lithic technology, 112, 248 Flower, William, 24 foliate points, 78 forward planning, 132, 174, 183, 186, 188, 203, 204, 206, 221, 244 founder effects, 66, 82, 87, 249 Fox, Robert, 136, 137, 138 France, 11, 17, 161, 178, 184, 244 Gabarnmang, 170 Gaitou Cave, 44 Galapagos Island, 111 Ganges River, 49 Ganges Valley, 72 Ganges-Brahmaputra Delta, 69 Ganjian Cave, 44, 48 Gargas, 184 Gebe Island, 172 genetics, 5, 6, 19, 31, 48, 49, 50, 51, 54, 56–57, 62, 63, 64, 67, 70–72, 73, 74, 75, 82, 84, 88, 114, 132, 133, 186, 187, 192, 232, 233, 243, 248, 257, 261, 262, 264 aDNA, 49 dating, 2, 4, 54, 57, 65, 70, 71, 76, 87, 228, 230, 231, 236, 238, 239, 240, 241, 242 divergence, 70 diversity, 56, 64, 65, 67, 70, 229, 236 DNA, 30, 49, 67, 70, 76, 228, 230, 236, 248

Index gene flow, 30, 45, 48, 49, 70, 109, 231, 232, 235 gene trees, 228, 230 genetic drift, 47 haplogroup, 65, 71, 228, 236 haplogroup L2, 65 haplogroup L3, 65, 70, 71, 236, 237, 238, 239, 242 haplogroup L4, 57 haplogroup M, 65, 71, 73, 229, 236, 239, 240, 241, 242 haplogroup N, 65, 71, 73, 229, 236, 239, 240, 241, 242 haplogroup P, 225 haplogroup Q, 225 haplogroup R, 71, 229, 240, 241, 242 haplogroup U, 71 isolation, 225, 235 L3 lineage, 232 M168 lineage, 232 mitochondrial DNA, 30, 46, 54, 57, 65, 70, 114, 225, 228, 229, 230, 231, 232, 233, 236, 237, 238, 239, 241, 242, 248, 255 mutation rates, 56, 57, 62, 70 phylogeography, 230–231, 236, 241, 242 Y chromosome, 70, 230, 231, 232, 236, 237, 241, 242 Gigantopithecus, 40 glabello-cerebral index, 25 Gobi Desert, 49 Gogora rock shelter, 78 Golo Cave, 172, 173, 174 Gould, Stephen J., 24 grassland, 70, 120, 133, 178, 221, 222, 224 grave robbing, 24 GRE8, 170 Green Gully, 29, 30 Green islands, 218 grindstones. See seed grinding Grotte du Renne, 161 ground-edge axe, 87, 166, 170, 174, 209 ground-edge technology, 167 ground stone technology, 167, 170 Gua Balambagan, 129 Gua Lawa Cave, 126 guanaco, 190 Gulf of Carpentaria, 154, 253 Gulu, 220 Gunung Sewa, 126 Gunung Subis massif, 105 Gunung Tjantelan, 126 Guri Cave, 138, 143 Gwion Gwion, 159 Had Pu Dai, 14 hafting, 137, 144, 146, 147, 166, 168, 170, 174, 175, 177, 186, 188, 209, 221 Hale, Herbert, 166

Halmahera, 172 handaxe, 8, 11, 13, 14, 15–16, 72, 80, 83, 137, 138, 166, 173, 177, 249, 250 Hang Boi, 127, 128 Hang Hum Cave, 41 Hang Troˆ´ng, 127 harpoon, 211 Harrisson Excavation Archive, 96, 97, 99, 104 Harrisson, Barbara, 92, 93, 95, 104, 105 Harrisson, Tom, 43, 85, 92–95, 96, 99, 104, 105 hatchets, 174 Hathnora, 34, 67 headbinding, 31 Herto, 2, 106 Himalayan Mountains, 69 Hindu Kush Mountains, 69 Hoabinhian industry, 15, 142, 166 Holocene, 43, 48, 55, 56, 59, 60, 61, 67, 69, 73, 95, 112, 114, 119, 126, 127, 129, 138, 139, 143, 150, 151, 161, 164, 166, 173, 189, 190, 195, 196, 201, 210, 214, 220, 235, 237, 246, 254 hominin fossil record, 9, 26, 27, 54, 88, 108, 118, 132–133, 171, 235, 236, 237, 238, 247–248, 255 lack of, 2, 5, 39, 40, 48, 49, 55, 56, 62, 63, 76, 77, 80, 238, 239, 240, 242 morphological variability, 54 Homo antecessor, 247 Homo erectus, 14, 19, 26, 27, 28, 29, 30, 40, 41, 42, 46, 47, 49, 88, 112, 132, 140, 232, 233, 235, 244, 247, 248, 259 extinction, 42, 43, 47 interaction with Homo sapiens, 42, 43, 48, 132 interbreeding, 19 Homo ergaster, 30 Homo floresiensis, 5, 15, 33, 44, 108, 112, 133, 140, 247 Homo fossilis, 25 Homo georgicus, 247 Homo habilis, 5, 44, 247 Homo heidelbergensis, 54, 166, 232, 233 Homo helmei, 261 Homo neanderthalensis. See Neanderthals Homo primigenius, 25 Homo rhodesiensis, 54 Homo rudolfensis, 247 Homo sapiens, 4, 259, 261 archaic, 6, 14, 233 morphological variability, 2, 25, 54, 55, 180, 258 morphological variation, 139, 179, 192, 234 skeletal evidence, 4, 5 Homo soloensis, 27 Honshu Island, 87 Hormuz Strait, 114 Horn of Africa, 77, 78, 80, 87 Howieson’s Poort industry, 5, 54, 65, 73, 77, 78, 161, 162, 197

323

Index Huanglong Cave, 45, 48 human revolution,The, 260 Hunter, John, 22 Hunterian Museum, 24 Huon Peninsula, 86, 170, 216 Huon Terraces, 85 Huxley’s Line, 139, 234 hybridisation, 48, 49 hypothermia, 190 ICPMS analysis, 99 Idaltu, 78 Ille Cave, 126, 137, 138, 139, 144, 146 Inamgaon, 72 India, 2, 11, 14, 34, 39, 48, 49, 63, 64, 67, 70, 71, 74, 76, 81, 82, 83, 120, 239 colonisation, 4, 76, 82, 174, 228, 239–240, 242 lithic technology, 2, 4, 5, 8, 39, 87, 239, 240 Indian Ocean, 21, 55, 62, 229, 242 Indo-Malay faunal species, 109 Indonesia, 172, 213, 233, 244, 247, 248 Indonesian Throughflow, 110, 113 Indus River, 69, 83 innovation, 19, 66, 83, 134, 162, 173, 200, 204, 206, 209, 210, 222, 224, 225, 255, 307 insolation, 69 interglacial, 54, 55, 61, 68, 69, 119, 120, 178, 187, 228, 229, 237, 239, 243, 244 Iran, 48, 49, 71, 80, 85 lithic technology, 87 Irian Jaya, 216 Irrawaddy Valley, 10, 11, 12, 13, 14 ISEA. See Island Southeast Asia Island Southeast Asia, 106, 108, 118, 119, 132, 167 colonisation, 43, 114, 115, 132, 136, 137, 139, 146, 164, 171 fauna, 112, 126, 135 Japanese in, 115 lithic technology, 137, 171–173, 174 skeletal record, 42–44, 135 subsistence strategies, 121, 126, 174 islands accessibility, 109 colonisation difficulty, 109 endemism, 108, 109, 135, 174 faunal vulnerability to humans, 112 impoverished faunal community, 109, 135, 219, 225 insular dwarfing, 247 isolation, 112 retention of early faunal lineages, 112 isolation, 225, 255 Israel, 4, 33, 34, 188 Isturitz, 211 Ivane Valley, 217, 218, 222, 223, 224, 226, 240

324

Jansz, 169 Japan, 8, 87 Japanese defeats, 115 Java, 9, 10, 11, 12, 13, 14, 15, 19, 26, 27, 30, 42, 43, 47, 109, 112, 115, 119, 120, 126, 132, 133, 136, 137, 140, 233, 238, 248 Java man, 26 Jebel Faya, 58, 59, 60, 62, 63, 80, 237, 238, 242 Jebel Qattar, 80 Jebel Qattar 1, 60 Jerimalai Cave, 85, 87, 114, 129, 171–172, 173, 174, 219 Jones, Rhys, 28, 29 Jordan, 61 Jo’s Creek, 216 Jubbah, 60, 61, 62 Judds Cavern, 184 Jurreru Valley, 83, 239 Jwalapuram, 39, 65, 67, 72, 73, 74, 82 Kafiavana, 206 Kafuan industry, 10 Kalimantan, 106 Kalumburu, 116 Kampong Ngebung, 14 Kanam, 10 Kangaroo Island, 26, 166 Kanjera, 10 Kapthurin, 249 Kara Kum Desert, 49 Karachi, 83 Karakorum Mountains, 69 Kartan culture, 26 Kebara Cave, 34, 80 Keilor, 28, 29 Keith, Sir Arthur, 10 Keyhole Cavern, 159, 184 Kilu Cave, 206, 207, 217, 219, 220 Kimberley, 116, 155, 159, 170, 173 King Island, 179, 180 Klasies River Mouth, 77 Klein Kliphuis, 82 Koenigswald, G. H. R. von, 10, 12, 13–14, 15, 43, 93 Korea, 8 lithic technology, 8 Kosipe, 85, 87, 170 Kosipe Mission, 217 Kota Tampan, 14, 15, 132 Kow Swamp, 29, 30, 31, 158, 160, 179, 206, 210, 234 Kromdraii, 10 Ksar Akil, 34 Kuching, 90, 92, 93 Kupona na Dari, 216, 217, 220, 226 Kurnool District, 82 Kutikina Cave, 180, 182, 183, 195

Index Kybra, 154 Kyzl Kum Desert, 49 La Madeleine, 210 La Riera, 211 Lachitu rock shelter, 173, 214, 216, 218, 219, 226 Lake Eyre, 153 Lake Mungo, 29, 30, 31, 46, 55, 67, 74, 76, 85, 86, 106, 154, 169, 179, 205, 206, 207, 249 WLH 50, 42, 47 landscape management, 106, 170, 207, 221, 223, 224, 226 Lang Rongrien, 85, 126, 133 Lang Trang Cave, 40, 41 Langsonia liquidens, 41 language, 6, 186, 187, 188, 244, 245, 246, 251, 255, 258, 260, 262 Lanne, William, 22 Laos, 40, 55, 67 fossil record, 248 skeletal record, 40–41 Lapita, 115 pottery, 115 Last Glacial Maximum, 56, 67, 83, 106, 120, 127, 150, 151, 152, 157, 158, 159, 160, 162, 169, 172, 180, 192, 193, 194, 196, 197, 198, 201, 206, 207, 208, 210, 211, 213, 216, 222, 224, 235, 236 Late Stone Age industry, 8, 57 Laugerie-Basse, 210 Laurente Cave, 143 Leakey, Louis, 10, 11 Leang Burung 2, 172 Leang Sarru rock shelter, 132, 172, 173 Lebanon, 34 Lene Hara, 129, 171–172, 174 Lenggong, 136, 140 Les Harpons, 211 Lespugue, 211 Lesser Sundas, 248 Levallois, 59, 60, 72, 77, 78, 79, 80, 83, 85, 86, 87, 180, 249 Levant, the, 2, 4, 6, 33, 54, 55, 56, 58, 59, 60, 62, 65, 67, 68, 78, 85, 197, 237, 239 colonisation, 4, 5, 34 “Egbert”, 34 fossil record, 237 lithic technology, 59, 80, 87 Middle Palaeolithic, 5, 34, 79 Neanderthals, 2, 4 skeletal evidence, 33–34, 79, 237 Upper Palaeolithic, 58, 87, 239 LGM. See Last Glacial Maximum Liang Bua, 247, 248 Liang Lemdubu, 173 Libya, 78

lice, 192 Lida Ayer, 41 Limeuil, 210 Linnaeus, Carolus, 258 Lipuun Point, 138 Liujiang, 48, 235, 238, 242 Longlin 1, 235 Longlindong, 46 Lourdes, 211 lowland forest, 103, 104, 120 Lupemban industry, 161 Luzon, 5, 44, 118, 119, 135, 136, 137, 138, 139, 140, 141, 142, 143, 146, 171, 248, 285, 290, 295, 301, 302, 305, 308, 310 Luzon Strait, 135 Lydekker Line, 109 Lyell, Charles, 90, 259 Ma U’Oi Cave, 14, 41 Macintosh, Norman, 28, 29, 30 Madagascar, 111, 112, 114 Mae Hong Son Province, 127 Mae Tha, 14 Magdalenian, 210, 211 Maghreb, 78 Maharashtra, 82 Mainland Southeast Asia. See Southeast Asia Malakunanja, 46, 77, 85, 170, 205, 206 Malangangerr, 170 Malaysia, 15, 115, 120, 133, 136, 140, 166, 213 Maludongdong, 46 Mandu Mandu Creek, 150, 154, 158, 169, 170, 206 mangrove, 103, 104, 120, 132 Mannalargenna Cave, 180, 183 Manus, 225 Manus Island, 114, 217, 225 Marine Isotope Stage 1, 69 Marine Isotope Stage 2, 69, 120, 178, 182, 197, 222 Marine Isotope Stage 3, 54, 55, 57, 58, 59, 60, 61, 62, 63, 65, 69, 71, 72, 120, 121, 133, 140, 146, 161, 178, 182, 197, 213, 218, 221, 222, 224, 240 Marine Isotope Stage 4, 4, 34, 46, 54, 55, 56, 57, 59, 61, 62, 63, 65, 69, 71, 120, 132, 140, 161, 182, 197, 238, 240 Marine Isotope Stage 4–3 Boundary Model, 65, 67, 68, 69, 71, 73, 75 Marine Isotope Stage 5, 4, 34, 46, 54, 55, 57, 59, 60, 61, 62, 63, 65, 69, 70, 71, 74, 80, 119, 120, 132, 140, 182, 197, 236, 237, 238, 239, 242 Marine Isotope Stage 5 Model, 65, 67, 68, 70, 71, 73, 74, 75 Marine Isotope Stage 6, 4, 34, 60, 61, 69, 119, 192 Marine Isotope Stage 7, 60, 62 marine resource exploitation, 129, 133, 229, 249

325

Index maritime technology, 116, 214, 219, 226 Martin, J., 3 Mas d’Azil, Bédeilhac, 211 Matenbek, 219, 220, 225 Matenkupkum, 207, 214, 216, 219, 220 Mauer, 9 Mayr, Ernst, 257, 259, 261 megafauna, 112, 119, 179, 183, 184, 224 Mekong River, 48 Melanesia, 86, 114, 115, 207, 225, 233, 234, 235, 241, 248 colonisation, 229, 240 lithic technology, 85–87 microlith, 2, 5, 26, 39, 59, 63, 65, 67, 72, 73, 74, 75, 77, 78, 80, 82, 83, 84, 85, 87, 88, 166, 177, 188 microlithic first hypothesis, 82 Microlithic Model, 54, 55, 57, 59, 62, 63, 73 Micronesia, 114 microwear analysis. See use-wear analysis Middle Palaeolithic first hypothesis, 82, 83 Middle Palaeolithic industry, 2, 8, 65, 80, 82, 83, 85, 185, 187, 202, 260 Middle Palaeolithic Model, 54, 55, 56, 57, 59, 62, 63, 73 Middle Son River Valley, 83 Middle Stone Age industry, 8, 88, 238, 242, 261 Middle Stone Age toolkit, 83 Minatogawa, 235 Mindanao, 135, 140 Mindoro, 140 Minori Cave, 143 modern human dispersal, 51–63, 88, 108, 112–114, 119, 239–242, 263 coastal route, 54, 59, 66, 67, 69, 73, 82, 114, 115, 214, 219, 229–230, 242 corridors, 62, 69, 70, 74, 77, 121, 151, 160, 237 difficulty, 140 failed, 4, 55, 67, 68 models, 51 processes, 51–63, 109, 213 rate, 56, 114, 115, 229, 242, 254 routes, 61, 67, 68–69, 70, 76, 78, 85, 140, 174, 213, 214, 218, 228, 230, 237–239, 240, 241, 242, 245, 247, 248, 250, 255 modesty, 192 Moh Khiew Cave, 40, 126 Mojokerto, 9, 13 Moluccas, 115 Money Cowry, 158 Mongolanthropus, 233, 235 Mongolia, 233 fossil record, 248 Mongoloid, 30, 46, 235 Montagu, Ashley, 29 montane forest, 120, 132, 142, 222, 224

326

Mopir, 220 mortuary ritual, 106 Mossgiel, 29 Mount Kinabalu, 105 Mousterian, 2, 26, 80, 160, 161, 166, 188, 201, 249, 250 Movius, Hallam L., 3, 8, 10, 11, 12–13, 14, 15, 19 Movius Line, 8, 9, 12, 19, 20 Mt Riveaux, 184 Mu O’oi, 40 Mudukian culture, 26 multiple exits theory, 236–237 multi-regional evolution model, 5, 6, 28, 30, 33, 48, 49, 231–236, 248, 260 Mulu Cave, 90 Mulvaney, John, 28 Mumba rock shelter, 78 Murray River, 26, 28, 160, 253 Murray Valley, 30, 31 Murundian culture, 26 Myanmar. See Burma Nanwoon Cave, 179 Narmada Valley, 11, 34, 67 Nauwalabila, 46, 77, 85, 170, 206 Nawarla Gabarnmung, 85, 170 Nazlet Khatar 4, 78 NCP. See Niah Caves Project N’Dhala Gorge, 154 Neanderthals, 2, 4, 19, 25, 27, 30, 34, 49, 67, 71, 79, 106, 161, 166, 177, 180, 185, 209, 210, 229, 232, 233, 238, 242, 244, 259 behaviour, 185–188, 209 distribution, 232, 237, 244 interaction with Homo sapiens, 48, 161, 186 interbreeding, 48, 71, 186, 231–232, 233, 236, 238, 242 needles, 4, 182, 191, 195, 196, 197 Nefud Desert, 80 Negrito, 27, 234 Negroid, 30 Nejd Leptolithic, 59 Nelson Bay Cave, 77 Neolithic farmers, 109 Neolithic farming, 112, 114 nets, 167, 168, 169, 171, 172, 174, 206 neural mutation, 54 New Britain, 115, 154, 213, 214, 217, 218, 219, 220, 221, 225 New Caledonia, 115 New Guinea, 2, 3, 29, 44, 85, 86, 115, 158, 170, 173, 174, 213, 217, 218, 235, 243, 251, 252 colonisation, 83, 164, 214–226, 240 environment, 222 highlands, 3, 85, 87, 217, 235 lithic technology, 85–87, 216, 220–222

Index subsistence strategies, 170, 218–224, 225–226, 252, 253 New Ireland, 158, 213, 214, 216, 218, 219, 220, 221, 222, 225, 226, 241, 246 New South Wales, 29, 46, 169, 174 New Zealand, 111 Ngandong, 6, 13, 19, 27, 42, 43, 47, 132, 233 Ngandong Fauna, 42 Ngebung, 15 Niah Cave, 3, 6, 33, 41, 43, 44, 48, 55, 67, 74, 85, 118, 120, 121, 126, 133, 144, 146 burials, 92, 95, 105 dating, 96 Deep Skull, 5, 43, 92–106, 118, 121 fauna, 121 geochemistry, 96, 98, 99, 100–101, 105 Hell Trench, 43, 92, 95, 96, 101, 104 Metal Age, 95 Neolithic, 95 palynology, 96, 98, 99, 101–104, 105 sediment, 96–101, 105 skeletal evidence, 93, 105–106 subsistence strategies, 106, 121, 133 West Mouth, 90, 92, 95, 96, 97, 104, 105, 120, 144 x-ray fluorescence, 101 Niah Caves Project, 96, 99, 104 Nile Corridor, 78 Nile Valley, 78 Nissan, 218 noble savage, 21, 22 Nombe rock shelter, 217, 224, 226 non-synbolic behaviour, 204, 206, 207, 209, 211 Northern Territory, 249 notational pieces, 205. See also art Nubian industry, 60, 63, 77, 78, 80, 238 Nueva Ecija Province, 137 Nullarbor Plain, 158 Nunamira Cave, 180, 181, 182, 183 obsidian, 143, 154, 207, 217, 220, 225, 226 Oceania, 114 ochre, 6, 52, 82, 106, 144, 154, 161, 162, 184, 185, 193, 194, 205, 207, 271, 285, 293, 299, 308 Okladnikov Cave, 233 Oldowan industry, 10, 249 Olduvai Gorge, 11 Oman, 60, 63, 80, 238, 239 Omo Kibish, 2 On the origin of species, 21, 22 open woodland, 42, 138, 139, 146, 167 optimal foraging theory, 184, 220, 224, 253, 263 organic technology, 106, 142, 164, 167, 168, 169, 170, 171, 173, 174, 177, 180, 204, 206, 210, 211 origins of species,The, 258 ORS 7 rock shelter, 180, 183

osseous technologies. See bone tools ostrich eggshell, 73, 82, 162 Owen, Richard, 23, 24 Pacific Ocean, 21 Pacitan, 15 Pacitan River, 14 Pajitan, 14, 15 Pakistan, 19, 48, 49, 81, 83 palaeoenvironments, 74, 75, 137, 192, 213, 220 reconstruction of, 61–62, 69, 103–104, 121, 138, 139, 142, 151–152, 177–179, 183–184, 221, 222 palaeolake, 61, 81, 152 palaeontological record, 110 palaeontology, 51, 135 Palawan, 43, 121, 125, 126, 133, 135, 136, 137, 138, 139, 140, 143, 144, 145, 146, 269, 278, 280, 290, 298, 301, 302, 311, 313, 317 Pallawa Trounta Cave, 180, 182, 194 palynology, 96, 98, 99, 101–104, 105, 120, 142, 167, 179, 223 Pamwak Cave, 217, 218 Panaramitee tradition, 154, 155, 159 Panxian Dadong, 15 Papua New Guinea. See New Guinea Paranthropus robustus, 10 Pardoe, Colin, 31 Parmerpar Meethaner rock shelter, 180, 182 Patne, 65, 72, 73, 75, 82 Patpara, 72 pearl, 154, 158 Pearl River, 48 Peche Merle, 184 Peking Man, 26 Peñablanca, 136, 139, 143, 144 Pengelly, William, 90 Périgord, 210 Persian Gulf, 61, 80 personal ornamentation, 4, 149, 152, 153, 157–158, 160, 161, 162, 163, 192–193, 194, 196, 197, 201, 203, 205, 206, 209, 219, 251 Philippine Sea, 135 Philippines, the, 5, 33, 43–44, 55, 114, 118, 119, 121, 126, 132, 133, 135–147, 171, 234, 235 art, 143 colonisation, 77, 135, 137–140, 171 Early Palaeolithic, 137, 138 environment, 138, 140–142 fauna, 126, 135, 138–139, 140–142, 174 fossil record, 138, 139 lithic technology, 137–140, 142–146, 147, 171, 174 personal ornamentation, 143 skeletal evidence, 126, 135–137, 139, 143, 171 subsistence strategies, 138–139, 146 Upper Palaeolithic, 143

327

Index phytolith, 59, 151 pigment use, 39, 105, 106, 121, 144, 146, 193, 200, 205, 209, 211 Pilanduk Rockshelter, 143 Pilgonaman, 169 Piltdown Man, 10, 12, 28, 92 Pinnacle Point, 186 Pirrian culture, 26 Pithecanthropus erectus, 14, 25, 26, 27, 28, 259 plant processing, 97, 106, 121, 133, 206, 226 Plesianthropus transvaalensis, 259 polished adzes, 167 pollen. See palynology polygenism, 24 Polynesia, 114, 115, 234, 235 Pongo pygmaeus, 40, 41, 120 population growth, 82, 83, 149, 150, 253–254 population pressure, 66, 72, 87, 150, 151, 152, 153, 157, 158, 159, 160, 161, 162, 163, 175, 206, 208, 209, 225, 251 Porc Epic, 78 possum-fur cloaks, 193 Prime Seal Island, 180 projectile technology, 54, 143, 144, 146, 173, 177, 206 Punung, 42, 43, 120, 132, 136, 140, 240 Punung Cave, 238 Punung Fauna, 42, 43, 119, 120 Puritjarra rock shelter, 154, 169, 249 pursuit hunting, 126 Qafzeh Cave, 4, 34, 46, 47, 65, 67, 80, 106, 236, 237, 238, 239, 260 Queensland, 27, 28, 153, 154, 170, 249 Quinkan, 159 Rabaul, 115 Rabel Cave, 143 race, 22, 24, 28, 259 Americans, 22 Caucasians, 22 Ethiopians, 22 Malays, 22 Mongolians, 22 racial prejudice, 10 racial purity, 25, 27 racial superiority, 28 racism, 259 rainforest, 20, 42, 70, 112, 119, 120, 121, 132, 138, 146, 173, 180, 217, 221 Red River, 48 Red Sea, 2, 33, 61, 62, 78, 229, 237 reduction sequence, 59, 60 refugia, 47, 58, 59, 63, 69, 120, 151, 152, 157, 158, 159, 160, 161, 162, 235, 236, 237, 253 replacement model, 5, 19, 33, 48, 49, 231–236, 260

328

residue analysis, 144–146, 170, 220, 222 retouch, 59, 60, 78, 83, 85, 169, 170, 171, 172, 173, 180, 194 Rhinoceros cf. sondaicus, 41 Rhinoceros cf. unicornis, 41 risk minimisation strategies, 152, 162, 221 Riwi Cave, 150, 154, 158, 170, 206, 207 Robertson, A. W. D., 24, 25 Rochereil, 210 rock art, 153, 154, 157, 159, 160, 161, 184, 188, 193, 200, 204, 205, 206, 208, 209, 210, 211, 243, 251, 252. See also art Rose Cottage Cave, 78 Royal Society of Victoria, 25 Rub’ al Khali, 81 Russia, 192, 248 Sa’gung, 138 Sabah, 105 sagaies, 211 Sahul, 3, 4, 7, 77, 85, 108, 109, 114, 115, 140, 148, 149, 150, 154, 158, 160, 162, 163, 167, 168, 171, 173, 174, 175, 177, 188, 190, 195, 200–212, 213, 234, 235, 241, 243, 244, 251, 255, 256 colonisation, 3, 46, 67, 114, 115, 139, 140, 148, 150, 164, 169, 171, 173, 174, 201, 204, 213–226, 235, 241, 243, 245–247, 250, 255 environment, 150, 162 fauna, 109, 112 lithic technology, 4, 86, 210, 220–222 subsistence strategies, 218–224, 252–253, 255 Salkhit, 233 Salween River, 48 Salzgitter-Lebenstedt, 188 Sambungmacan, 15, 42, 233 Sandy Creek, 155, 170 Sangiran, 9, 13, 14, 15 Sarawak, 90, 92, 105, 118 Sarawak Museum, 92, 96 Saudi Arabia, 59, 60, 61, 81 savannah, 120, 132, 138, 139, 146 savannah corridor, 120 scarification, 193 Schöningen, 16 Schwaner Mountains, 105 scraper, 4, 60, 72, 78, 79, 80, 83, 85, 86, 87, 166, 168, 169, 171, 177, 180, 188, 191, 194, 197, 221, 222 steep-edged, 173 thumbnail, 168, 180, 182, 194, 195 sea crossings. See seafaring seafaring, 3, 4, 5, 44, 61, 87, 112, 114, 115, 116, 135, 138, 140, 146, 172, 175, 203, 214, 218, 219, 225, 226, 230, 241, 244, 245, 246 Second World War, 259, 261 sedimentology, 96, 99

Index seed grinding, 170, 175, 206, 209 Senckenberg Institute, 43 sewing, 4, 191, 195, 196 shell midden, 138, 219, 225, 229 shell tools, 173, 174, 206 shellfishing, 204, 209, 219 Shi’bat Dihya 1, 7, 59, 60 shivering, 190 Shuidonggou, 12 Siberia, 49, 133, 232 Simple Figurative, 159 Sinai, 61 Sinanthropus pekinensis, 9, 13, 26, 27, 28, 259 Singapore, 115 Singhbum, 72 single exit hypothesis, 232, 234, 236–237, 238, 239, 241, 242 single origin theory, 248 Site 55, 72, 83 16R Dune, 72, 74 Skhu¯l Cave, 4, 12, 34, 46, 65, 67, 78, 80, 106, 236, 237, 238, 239, 260 skull measuring. See craniometry Smith, Sir Eliot Grafton, 10 Snake Cave. See Thum Wikan Nakin Cave Soa Basin, 112 Soan Valley, 11, 13, 14 social Darwinism, 22, 188, 259 social evolutionism. See social Darwinism Sodmein Cave, 78 Solo Man, 27, 29, 233 Solo River, 42 Solomon Islands, 114, 115, 213, 214, 217, 218, 220, 225 Son Valley, 72, 83 Song Gupuh Cave, 120, 126, 132, 140 Song Hong River delta, 127 Song Keplek Cave, 126 Song Terus Cave, 120, 126, 132, 140 South Africa, 2, 10, 26, 77, 82 South Asia, 54, 232, 237 colonisation, 40, 213, 236, 242 fossil record, 67–68 lithic technology, 65, 68, 72–73, 81–85, 87 Middle Palaeolithic, 65, 67, 68, 72, 73, 74, 75, 83 skeletal record, 34–39 Upper Palaeolithic, 65, 72, 75 South Australia, 166 South Australian Museum, 26 South China Sea, 119, 135 South Kov, 217 Southeast Asia, 2, 3, 14, 15, 42, 48, 49, 50, 76, 87, 118, 239, 244 colonisation, 19, 20, 114, 121, 140, 213, 239 Early Palaeolithic, 15

environment, 119–121, 132 fauna, 40, 41, 129 fossil record, 132–133, 240, 242 lithic technology, 19, 81–85, 142, 240 skeletal record, 39–42 subsistence strategies, 121 Western perceptions, 3, 19 Spain, 188 spearthrowers, 211 speciation, 48, 231 spiral decorated rods, 211 Spirit Cave, 127 Spring Creek, 206 Sri Lanka, 7, 39, 48, 73, 81, 82, 83 lithic technology, 39, 87 Stegodon florensis, 112 Stegodon-Ailuripoda Fauna, 41, 45 Steinheim, 9 Sterkfontein, 10 Still Bay industry, 77, 78, 161, 162 Stone Cave, 181 Strickland River, 253 sub-alpine forest, 222 sub-montane forest, 120, 132 Sulawesi, 109, 112, 113, 132, 133, 135, 140, 172 Sulu Sea, 135, 139 Sumatra, 4, 41, 65, 109, 112, 115, 120, 140 Sunda, 3, 5, 6, 7, 47, 67, 108, 109, 119, 120, 121, 135, 140, 213, 225, 244, 255 Sungir, 192 swamp forest, 103, 104, 119, 120, 222 Swanscombe, 10 symbolic, 160, 162, 163 symbolic behaviour, 6, 65, 73, 75, 82, 149, 150, 151, 152, 153, 160–162, 163, 175, 188, 193, 197, 198, 201, 204, 205, 206, 208, 209, 211, 245, 247, 251–252, 253, 261, 262, 263, 264, 277, 317 symbolism, 2, 192, 194, 203, 204, 205, 210, 243 Tabon Cave, 43, 121, 126, 133, 136, 138, 139, 143, 144, 146 Tabonian industry, 142, 143 Tabun Cave, 34, 80 Taiwan, 114, 135, 248 Taklamakan Desert, 49 Talaud Islands, 132, 133, 146, 164, 172 fauna, 174 lithic technology, 172, 174 Talaud-Sangihe Archipelago. See Talaud Islands Talgai, 28, 29 Tam Pa Ling Cave, 40, 41, 55, 67 Tamil Nadu coast, 82 Tanzania, 78 taphonomy, 148, 200–202, 204–208, 209, 210–211 Taramsa Hill, 78

329

Index Tartanga culture, 26, 29 Tasmania, 2, 3, 28, 152, 154, 159, 164, 166, 167, 168, 173, 177, 207, 243, 253, 254 art, 106, 184, 193 clothing, 4, 106, 182, 190, 194, 195–196, 198 environment, 177–179, 183–184, 194, 196 isolation, 190, 254, 255 lithic technology, 28, 180, 185, 194–195 skeletal record, 179–180 subsistence strategies, 179, 182–184 Tasmanian Aborigines, 22, 23, 166, 177, 185, 186, 188, 190 comparisons with mainland population, 24–25 craniometry, 23–25 lithic technology, 4, 26 Lower Palaeolithic “folk”, 26 non-skeletal data, 24 Western perceptions, 23, 177, 188, 190 taurodontism, 41 Terra, Helmut de, 10–12, 13, 14, 19 Thailand, 14, 19, 40, 41, 48, 85, 120, 126, 127, 133 art, 106 lithic technology, 40 skeletal record, 40 Tham Hai Cave, 41 Tham Khuyen Cave, 14, 41 Tham Lod rock shelter, 127 Tham Om Cave, 41 Thar Desert, 69, 70, 72, 83 thermal adaptations, 192 thermal insulation, 190 Thorne, Alan, 29, 30 Thum Wikan Nakin Cave, 14, 40, 41 Tianyuan. See Tianyuandong Tianyuandong, 44, 48 Tibetan Plateau, 10 Tierra del Fuego, 23, 190 Timor, 109, 110, 113, 129, 133, 241 Tindale, Norman B., 26, 28, 29, 166, 284, 312 Tito Bustillo, 211 Toba eruption, 2, 4, 65, 70, 80, 83, 228, 236, 239–240, 242 Toe Cave, 218, 226 Tongtianyan Cave, 44–45 Topinard, Paul, 22, 24, 25 trade, 153, 154, 155, 161, 175, 182, 205, 207, 208, 209, 226, 251 Tràng An park, 127 trapping, 121, 126, 133, 172, 181 très-dolichocéphalique, 24 très-prognathe, 24 Tridacna costata, 62 tri-hybrid theory, 27, 248 Trinil, 9, 14 tropical environments, 118

330

Tunnel Cave, 168 Turkey, 34 Tweedie, Michael, 92 Two-Phase Model, 54, 55 Two-Stage Model, 54, 57, 59, 62 Tylor, Edward Burnet, 26 Tyson, Edward, 258 Uçagizli Cave, 34 Uganda, 10 Umboi, 218 United Arab Emirates, 61, 80 Upper Palaeolithic industry, 2, 8, 57, 78, 249, 260, 261 Upper Palaeolithic Model, 54–55, 57, 58, 59, 62, 63 Upper Palaeolithic Revolution, 2, 4, 55, 77, 82, 167, 175–177, 260, 261 Upper Swan, 173 Ursus thibetanus, 40 use-wear analysis, 137, 142–146, 192, 195 Vietnam, 14, 40, 48, 127, 166 skeletal record, 41–42 Vilakuav, 217 Visadi, 72 Visayas, the, 135, 140 Wadi Surdud, 59, 62 Wadjak, 43 waisted axe, 86, 170, 173, 174, 206, 208, 221, 223 Walkunder Arch Cave, 154 Wallace Line, 6, 109, 139, 174 Wallace, Alfred Russel, 6, 90, 108, 258, 262 Wallacea, 109, 112, 119, 129, 133, 134, 135, 136, 146, 173, 174, 214, 234, 235, 246, 299 fauna, 109 Wargata Mina, 159 Warreen Cave, 180, 182, 183, 195 watercraft, 3, 5, 6, 114, 116, 132, 138, 146, 174, 187, 201, 207, 214, 245, 246, 247, 276 Watru Abri, 72 Webb, Stephen, 31 Weidenreich, Franz, 27, 28, 29, 30, 43 West Indies, 111 Western Australia, 154, 169 Western Ghats, 69, 72 Western perceptions, 8, 15, 16, 17–18 Widgingarri 1 Rockshelter, 170, 173 Willandra Lakes, 31, 46, 47, 106, 169, 206 Willaumez Peninsula, 217, 220 Williamson, George, 23, 24 wind chill, 193, 194 wind-chill index, 190 Wolo Sege, 112 Wolpoff, Milford, 30

Index woodland, 70, 132, 133 woodworking, 173, 195 World War II, 114 Wyrie Swamp, 210 Xinzhe, Wu, 30 Yangtzi River, 236 Yemen, 7, 56, 59, 60, 228

Yombon, 217, 221, 222, 226 Younger Toba Event. See Toba eruption Zagros Mountains, 34 Zhirendong, 19, 45–46, 48, 140, 238, 242 Zhoukoudian, 910, 12, 13, 14, 26, 27, 44, 235 Zirendong, 119 zooarchaeology, 126 Zuraina Majid, D. P., 95

331