Origins of Agriculture in Western Central Asia : An Environmental-Archaeological Study [1 ed.] 9781934536513, 9781934536162

In Origins of Agriculture in Western Central Asia, archaeologist David R. Harris addresses questions of when, how, and w

221 91 7MB

English Pages 321 Year 2010

Report DMCA / Copyright

DOWNLOAD PDF FILE

Recommend Papers

Origins of Agriculture in Western Central Asia : An Environmental-Archaeological Study [1 ed.]
 9781934536513, 9781934536162

  • 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

Origins of Agriculture in

Western Central Asia

Professor V. M. Masson introducing school children from Ashgabat to the excavations at the Neolithic site of Jeitun, Turkmenistan, April 1990.

David R. Harris

Origins of Agriculture

in Western Central Asia

An Environmental-Archaeological Study

with contributions from: Eleni Asouti, Amy Bogaard, Michael Charles, James Conolly, Jennifer Coolidge, Keith Dobney, Chris Gosden, Jen Heathcote, Deborah Jaques, Mary Larkum, Susan Limbrey, John Meadows, Nathan Schlanger, and Keith Wilkinson

University of Pennsylvania Museum of Archaeology and Anthropology Philadelphia

© 2010 University of Pennsylvania Museum of Archaeology and Anthropology Philadelphia, PA 19104-6324 Published for the University of Pennsylvania Museum of Archaeology and Anthropology by the University of Pennsylvania Press. All rights reserved. Published 2010. Production of this book was supported by a publication grant from the Academic Committee of the Iran Heritage Foundation (London) and an award from the Stein-Arnold Expedition Fund of the British Academy. The drawing on p. 304 of the head of a wild bezoar goat is from Harris 1962, Fig. 3a.



library of congress cataloging-in-publication data

Harris, David R. Origins of agriculture in western central Asia : an environmental-archaeological study / David R. Harris. p. cm. Includes bibliographical references and index. ISBN-13: 978-1-934536-16-2 (hardcover : alk. paper) ISBN-10: 1-934536-16-4 (hardcover : alk. paper) 1. Agriculture—Turkmenistan—Origin. 2. Agriculture—Asia, Central—Origin. 3. Agriculture, Prehistoric—Turkmenistan. 4. Agriculture, Prehistoric—Asia, Central. 5. Excavations (Archaeology)— Turkmenistan. 6. Excavations (Archaeology)—Asia, Central. 7. Turkmenistan—Antiquities. 8. Asia, Central—Antiquities. I. Title. GN855.T85H37 2010 306.3’4909585--dc22 2010009780

This book was printed in the United States of America on acid-free paper.

Contents Illustrations Tables Contributors Preface and Acknowledgments

vii x xi xiii

Introduction

1

PART I: Physical Environment and Ecology

3

1. The Present Environment

5

2. Environmental Changes in the Pleistocene and Holocene

19

3. The Local Environment of Jeitun, with Susan Limbrey

27

4. The Local Environment of the Bolshoi Balkhan Sites, with Jen Heathcote

35

PART II: Prehistoric Archaeology

41

5. History of Archaeological Research, with Jennifer Coolidge

43

6. The Mesolithic and Neolithic Periods: Sites, Sequences, and Subsistence, with Jennifer Coolidge

53

PART III: Neolithic Crop Plants and Domestic Animals 7. Areas of Origin of the Crops and Domestic Animals

71 73

PART IV: Archaeological-Environmental Investigations in Turkmenistan 1989–98

93

8. Jeitun, the Sumbar and Chandyr Valleys, and the Bolshoi Balkhan Region: Excavation and Survey, with Chris Gosden

95

9. Jeitun: Dating and Analysis of Excavated Materials

119

9.1 Dating the Site: Radiocarbon Chronology, with Chris Gosden and John Meadows

119

9.2 Yard Deposits and Building Materials at Jeitun, Susan Limbrey

125

9.3 Buried Soils at Jeitun, Susan Limbrey

131

9.4 Investigation of a Palaeosol Sequence at Jeitun: Excavation of a Ditch-like Feature and Measurement of Magnetic Susceptibility, Keith Wilkinson

137

9.5 Phytolith Analysis of Samples from On- and Off-Site Deposits at Jeitun, Mary Larkum

142

9.6 Charred Plant Macro-Remains from Jeitun: Implications for Early Cultivation and Herding Practices in Western Central Asia, Michael Charles and Amy Bogaard

150

9.7 Remains of Wood Charcoal from Jeitun: Identification and Analysis, Eleni Asouti

166

contents

9.8 Pollen and Charcoal-particle Analysis: Sampling Off-Site Deposits at Jeitun, David Harris

171

9.9 The Vertebrate Assemblage from Excavations at Jeitun, 1993 and 1994, Keith Dobney and Deborah Jaques

174

9.10 The 1994 Knapped-Stone Assemblage from Jeitun, James Conolly

180

9.11 The Pottery from Jeitun, Jennifer Coolidge

186

9.12 Summary Discussion of the Excavation Evidence from Jeitun, with Chris Gosden

190

10. The Bolshoi Balkhan Sites: Analysis of Excavated Materials

197

10.1 Charred Seeds from the Dam Dam Cheshme Rockshelters, Michael Charles

197

10.2 Wood Charcoal from the Dam Dam Cheshme Rockshelters, Eleni Asouti

199

10.3 Animal Remains from the Dam Dam Cheshme Rockshelters, Keith Dobney and Deborah Jaques

201

10.4 Stone Tools from the Dam Dam Cheshme Rockshelters, James Conolly and Nathan Schlanger

203

10.5 Pottery from the Dam Dam Cheshme Rockshelters and Other Sites in the Bolshoi Balkhan Region, Jennifer Coolidge

206

PART V: Synthesis and Conclusions

209

11. Neolithic Settlement and Subsistence

211

12. The Beginnings of Agriculture in Western Central Asia

225

Appendices

237

Appendix 3.1, Susan Limbrey

237

Appendices 8.1–-8.5, Chris Gosden

239

Appendices 9.1–9.2, Susan Limbrey

244

Appendices 9.3–9.5, Mary Larkum

250

Appendix 9.6, Michael Charles and Amy Bogaard

256

Appendices 9.7–9.10, Keith Dobney and Deborah Jaques

260

Appendix 9.11, James Conolly

266

Bibliography

269

Author Note

297

Index

299

vi

Illustrations Professor V. M. Masson introducing school children to the excavations at Jeitun, 1990

Frontispiece

figures

1.1

The Central Asian arid zone

4

1.2

Turkmenistan and adjacent areas

4

1.3

The Kopetdag mountains and piedmont

1.4

Takyrs on the Kopetdag piedmont

6

1.5

A takyr in the Central Karakum

6

1.6

Pentagonal cracks on a takyr surface

6

1.7

Turkmenistan: mean annual precipitation

8

1.8

Riparian tugai forest

Color figure I

1.9

Groves of pistachio trees, Badghyz Reserve

Color figure I

1.10

Pistachio tree with wild barley, Badghyz Reserve

Color figure I

1.11

Steppe grassland on the Kopetdag piedmont

Color figure I

1.12

White saksaul, Central Karakum

Color figure II

1.13

Black saksauls, Trans-Unguz Karakum

Color figure II

1.14

Ancient black saksaul, Repetek Reserve

Color figure II

1.15

Distribution of wild Asiatic mouflon, urial, and argali sheep

12

1.16

Distribution of wild bezoar, ibex, and markhor goats

13

3.1

Location of Jeitun and the Kara Su

28

3.2

Kara Su gorge near Jeitun

Color figure II

3.3

Solonetz soil in the Kara Su gorge

Color figure II

3.4

Overgrazed semi-shrub vegetation at Jeitun

Color figure II

3.5

Tamarisk and reeds beside the Kara Su

Color figure III

3.6

Reed-tamarisk swamp near Jeitun

Color figure III

4.1

Bolshoi and Maly Balkhan massifs and the Uzboi lowland

36

4.2

Escarpment and piedmont of the Bolshoi Balkhan massif

Color figure III

4.3

View from the DDC 1 rockshelter

Color figure III

4.4

Southern Bolshoi Balkhan, piedmont, and lower Uzboi

37

4.5

Bedded sands and gravels below the DDC rockshelters

38

4.6

Downstream view of the Uzboi channel

4.7

Scattered juniper trees above the DDC 4 rockshelter

4.8

Grove of fig trees at the head of the DDC 4 canyon

6.1

Distribution of Neolithic and Mesolithic sites

Color figure I

Color figure III 39 Color figure III 56

illustrations

6.2

Distribution of Keltiminar and other Neolithic sites

65

8.1

Jeitun: plan of Masson’s second level

96

8.2

Jeitun: locations of 1989–94 excavations

97

8.3

Jeitun: plan of the central area excavation

98

8.4

Jeitun: part of the central area excavated 1989–90

99

8.5

Jeitun: section in trench 1991a

8.6

Jeitun: northern face of section in trench 1991a

8.7

Jeitun: plan of 1993 House A excavation

8.8

Jeitun: excavated House A oven

Color figure IV

8.9

Jeitun: dog-like figurine and truncated cone from House A

Color figure IV

8.10

Jeitun: plan of 1994 House B excavation

8.11

Jeitun: House B from the west

Color figure IV

8.12

Jeitun: incomplete tortoise carapace from House B

Color figure IV

8.13

Southwestern Turkmenistan

108

8.14

Sumbar and Chandyr valleys: location of archaeological sites

109

8.15

Approach to the Jebel rockshelter

Color figure V

8.16

DDC 1 from the Bolshoi Balkhan piedmont

Color figure V

8.17

Entrance of DDC 1

Color figure V

8.18

Interior of DDC 1

Color figure V

8.19

Entrance of DDC 2

Color figure V

8.20

Plan of DDC 2

114

8.21

Section of Area 1 in DDC 2

115

8.22

Entrance of DDC 3 during excavation

9.1

Jeitun: calibration of radiocarbon results

120

9.2

Jeitun: Bayesian model of on-site radiocarbon dates

121

9.3

Jeitun: duration of occupation

122

9.4a,b

Micromorphology: silica skeletons

Color figure VI

9.5

Micromorphology: residual plant structure in yellow organic matter

Color figure VI

9.6

Micromorphology: spherulitic fabric

Color figure VI

9.7

Jeitun: excavation of buried-soil section

9.8

Jeitun: ditch-like feature in buried-soil section

Color figure VI

9.9

Micromorphology: framboids in limpid yellow coatings

Color figure VI

9.10

Jeitun: excavated ditch-like feature

137

9.11

Jeitun: buried-soil section and magnetic susceptibility

138

9.12

Jeitun: section through buried-soil sequence

139

9.13

SEM micrographs of Jeitun phytoliths and spherulites

143

101 Color figure IV 103

105

Color figure V

131

viii

illustrations

9.14

Jeitun: glume wheat spikelet forks

152

9.15

Jeitun: glume wheat grain

152

9.16

Jeitun: barley material

153

9.17

Triangular plot of Jeitun barley and wheat samples

154

9.18a,b Scatter plots of samples from Jeitun and Amorgos

156

9.19a,b Correspondence analysis of samples and taxa

159

9.20

Correspondence analysis of taxa by seasonality

161

9.21

Correspondence analysis of samples by seasonality, crop material, and taxa

161

9.22

Correspondence analysis of samples by context type

161

9.23

SEM micrographs of Jeitun charcoal specimens

169

9.24

Jeitun: evidence of bone bead production

9.25

Jeitun: representative stone artifacts

9.26

Stone tools from Jeitun, 1994

9.27

Jeitun blade widths

182

9.28

Jeitun flake lengths

182

9.29

SEM micrograph of edge wear on a Jeitun blade

184

9.30

Jeitun pottery: vessel forms and types of ware

187

10.1

SEM micrographs of DDC charcoal specimens

200

10.2

Stone tools from DDC 2, 1997

203

Color figure VII 180 Color figure VIII

Color figures are located after p. 208

ix

Tables 2.1

Late Pleistocene/Holocene fluctuations of the Caspian Sea and glacial-postglacial sequence

21

2.2

Lateglacial fluctuations of the Caspian Sea

21

6.1

Mesolithic rockshelter and cave sites

55

6.2

Periods and principal sites of the Jeitun Culture in Turkmenistan

60

9.1

Jeitun: on-site AMS radiocarbon results

120

9.2

Jeitun: off-site AMS radiocarbon dates from three palaeosols

124

9.3

Jeitun: phytolith samples

128

9.4

Jeitun: particle-size analysis of mudbrick and mortar

129

9.5

Jeitun: pH determinations

129

9.6

Jeitun: samples from 1993 buried-soils section

132

9.7

Jeitun: chemical and particle-size analysis of 1993 samples

134

9.8

Jeitun: AMS radiocarbon dates from three off-site palaeosols

139

9.9

Jeitun: phytolith samples

146

9.10

Relative frequency of glume-wheat grain and glume bases

155

911

Classification of wild plant taxa

155

9.12a,b Amorgos and Jeitun: discriminant function of samples

156, 157

9.13a,b Taxa used in correspondence analysis

158

9.14

Seasonality of Jeitun wild plants

160

9.15

Ecological data for Jeitun wild plants

163

9.16

Jeitun: charcoal samples

168

9.17

Jeitun: pollen types and charcoal particles

172

9.18

Jeitun: vertebrate taxa identified

176

9.19

Jeitun: skeletal elements with traces of bone working

178

9.20

Jeitun: stone-tool color variability

181

9.21

Jeitun: lithic categories

181

9.22

Jeitun: dimensions of stone-tool types

183

10.1

Charred seeds from DDC 2 and 3

198

10.2

Charcoal samples from DDC 1, 2 and 3

199

10.3

Animal remains from DDC 2 and 3

202

10.4

Stone artifacts from DDC 2

205

Contributors Eleni Asouti, School of Archaeology, Classics and Egyptology, University of Liverpool, UK Amy Bogaard, School of Archaeology, University of Oxford, UK Michael Charles, Department of Archaeology, University of Sheffield, UK James Conolly, Department of Anthropology, Trent University, CANADA Jennifer Coolidge, Oxford Institute for Energy Studies, University of Oxford, UK Keith Dobney, Department of Archaeology, University of Aberdeen, UK Chris Gosden, School of Archaeology, University of Oxford, UK David Harris, Institute of Archaeology, University College London, UK Jen Heathcote, Head of Research Policy (Freshwater and Wetlands), English Heritage, London, UK Deborah Jaques, Palaeoecology Research Services, Hull, County Durham, UK Mary Larkum, Department of Anthropology, University of Massachusetts, Amherst, USA Susan Limbrey, Institute of Archaeology and Antiquity, University of Birmingham, UK John Meadows, Institute of Archaeology, University College London, UK Nathan Schlanger, INRAP (Institut national de recherches archéologiques préventives), Paris, FRANCE Keith Wilkinson, Department of Archaeology, University of Winchester, Winchester, UK

Preface and Acknowledgments

T

oday the products of domesticated plants and animals sustain—however precariously—over six billion people, predicted to rise to over nine billion by 2050, but dependence on agriculture is a recent phenomenon in the 100,000-year evolution of anatomically modern humans. The question of how this innovation came about has puzzled scholars in many disciplines since the early 19th century, and although there have been great advances in our understanding of agricultural origins in a few regions of the world, particularly the Southwest Asian “Fertile Crescent,” there are many gaps in our knowledge of how agriculture began in other regions. One such region is the vast arid zone of Central Asia, where, with the notable exception of southern Turkmenistan and parts of northeastern Iran and northwestern China, there has been little archaeological investigation of prehistoric agricultural settlement. The purpose of this book is to help fill this gap by adding new data to the earlier research undertaken in Turkmenistan and adjacent areas and by synthesizing what is currently known about when, how, and why agricultural settlements originated in western Central Asia. Much of the new data presented in this book was generated under the auspices of an international project carried out between 1989 and 1998 in Turkmenistan by a team of British archaeologists in collaboration with Russian and Turkmen colleagues. The seeds of the project were sown at the first meeting of the World Archaeological Congress that took place in England, at the University of Southampton, in 1986. During the Congress its National Secretary, the late Professor Peter J. Ucko, had a series of meetings with one of the Russian participants, Professor V. M. Masson, Director of the Institute of the History of Material Culture in Leningrad and a Corresponding Member of the Academy of Sciences of Turkmenistan.

They discussed the idea of developing collaborative field projects in the Soviet Union, and the following year Peter Ucko, accompanied by his Southampton colleague Tim Champion, visited Russia at the invitation of Professor Masson. He showed them a series of archaeological sites, including Jeitun (Djeitun, Dzheitun) at the southern edge of the Karakum desert, which he had partially excavated in the late 1950s and early 1960s. Although not radiocarbon dated, Jeitun was interpreted as the earliest site of the Neolithic “Jeitun Culture” of southern Turkmenistan and—on the basis of finds of charred wheat and barley grains and their impressions in mudbrick, and of the bones of domestic goats and sheep—it came to be regarded as the oldest agricultural settlement in Central Asia. Masson recognized that this interpretation of Jeitun’s significance should be tested by undertaking new excavations at the site using modern methods of retrieving, analyzing, and directly dating plant and animal remains. He therefore invited me to coorganize a collaboration in which a group of British environmental archaeologists would work at Jeitun with him, several of his colleagues from Leningrad, and a Turkmen archaeologist from Ashgabat, Dr. K. K. Kurbansakhatov, who had been one of his doctoral students. In April 1989 I visited Jeitun with my colleague from the Institute of Archaeology in London, Gordon Hillman, to assess the potential of the site for archaeobotanical work. The deposits we sampled proved to contain abundant charred cereal grains and chaff and numerous animal bones. Gordon and I therefore returned the following year, accompanied by Michael Charles, then one of our doctoral students in archaeobotany at the London Institute, Susan Limbrey, a geoarchaeologist from the University of Birmingham, andTony Legge, a zooarchaeologist from the extra-

preface and acknowledgments

Kurbansakhatov, and we recall with gratitude the generous assistance we received from the directors and other members of the Desert Research Institute, the Central Botanic Garden, and the Institutes of Botany, Geology, History, and Zoology in Ashgabat. Not all the individuals who helped can be listed here, but particular thanks for their major contributions during the excavations are due to Yuri Berezkin, Natasha Solovyova, Sasha Maretin, Vladimir Timofeyev, Nikolai Savvanidi, Vladimir Zavyalov, Patrick Blackman, Jane Kaye, Colin Merrony, Nathan Schlanger, and Jane Sidell. In addition to Professor Masson himself, the initial team at Jeitun included Russian colleagues Galina Korobkova, Ogulsona Lollekova, and Tamara Sharovskaya who worked on the lithic and ceramic finds, and Alexei Kasparov who analyzed the animal bones recovered in 1989 and 1990. The environmental team from Britain was joined in the later years of the project by zooarchaeologist Keith Dobney, and geoarchaeologists Keith Wilkinson, Jen Heathcote, and Sarah O’Hara, while Michael Charles took over as principal archaeobotanist from Gordon Hillman after the 1992 season. For the final excavation season in 1997, in the Bolshoi Balkhan, we were joined by Saïd Khamrakuliev from Ashgabat, Amy Bogaard, then a doctoral student in archaeobotany at Sheffield, Jennifer Coolidge, then a doctoral student at Oxford who specialized in ceramic analysis and who in 1997 and 1998 visited many of the Jeitun-Culture sites on the Kopetdag piedmont with Dr. Kurbansakhatov, Simeon Mellalieu, a doctoral student from London who assisted Jen Heathcote with geoarchaeological survey, and Greger Larson, an American student who was then visiting Turkmenistan. Several long-distance field excursions, undertaken to enhance our understanding of the ecology of other regions of Turkmenistan, were made possible by Russian and Turkmen colleagues, particularly the Director and Deputy Director of the Desert Research Institute in Ashgabat, Dr. A. G. Babaev and Dr. N. S. Orlovsky. Their generous assistance enabled some of us to visit four areas of particular ecological interest: the Natural Reserve of Badghyz southeast of the Kopetdag, established in 1941 and inhabited by the largest surviving population of the goitered gazelle, as well as onagers and wild sheep, where we were expertly guided by V. I. Kuznetsov and K. P. Popov; Repetek Sand Desert Reserve in the Southeast Karakum, established in 1912, where we gained valuable

mural department of the University of London. Our initial results encouraged us to try to continue the project after Turkmenistan became an independent nation in 1991. This proved possible, but participation by our Russian colleagues from (the now re-named) St. Petersburg was perforce gradually reduced. In subsequent years the project became almost entirely a British venture, dependent on the help of Dr. Kurbansakhatov who continued to provide essential logistic support from Ashgabat and participate in our later surveys and excavations at Jeitun and elsewhere. In 1991 the team was joined by Chris Gosden (then at La Trobe University in Australia, now Professor of European Archaeology at the University of Oxford). He became co-director of the project with me after Masson’s withdrawal, directed our excavations at Jeitun using modern methods of fine-grain excavation, led a reconnaissance survey in the Sumbar and Chandyr valleys in 1996, and oversaw our excavations at sites in the Bolshoi Balkhan massif in western Turkmenistan in 1997 (see Chapter 8). After ten years of co-operative effort in Turkmenistan, and despite many logistical, administrative, and financial challenges along the way, the fieldwork phase of the environmental-archaeological project we started in 1989 was completed. Thereafter, analytical work on the organic and inorganic materials from our excavations continued—necessarily intermittently— in several laboratories in England, and other circumstances delayed full publication of all the results. But this unintended prolongation of the post-excavation phase of the project was, in the event, beneficial because it enabled me, in this book, to relate its results more closely to relevant archaeological, environmental, and genetic research published since 1998, and to complete a more comprehensive account of the origins of agriculture in western Central Asia than originally envisaged.

Acknowledgments The project would not have been possible without the permission and support of various government departments in Ashgabat, which we gratefully acknowledge and which enabled many Russian and British archaeologists to participate in the excavations and field surveys. We are especially indebted to the late Professor Masson and the Turkmenistan Academy of Sciences, to the late Dr. G. F. Korobkova and to Dr.

xiv

preface and acknowledgments

photographs and some of the drawings. All the photographs were taken by David Harris, with the exception of those individually attributed in their captions. Without financial support from many sources the project would have been impossible. Research and travel grants were received from the British Academy, the British Institute for Persian Studies, the British-Soviet Archaeological Committee, the Gordon Childe Fund of the University of London Institute of Archaeology, the Dean’s Fund of University College London, the University of Birmingham, the University of Oxford, and La Trobe University (Melbourne), all of which are gratefully acknowledged, as are publication grants received from the Iran Heritage Foundation (London), and the British Academy’s Stein-Arnold Expedition Fund. Financial and logistic support was also generously provided by the Institute of the History of Material Culture of the Russian Academy of Sciences (St. Petersburg), the South Turkmenistan Multi-disciplinary Archaeological Expedition of the Turkmenistan Academy of Sciences (Ashgabat), and the Desert Research Institute (Ashgabat). I am especially grateful to the Leverhulme Trust (London) for awarding me an Emeritus Fellowship in 2001 to help finance the final analyses and dating of organic samples from our excavations and the preparation of this volume. Without this support I would have been unable to complete the project. I also wish to thank the Iranian archaeologists (mentioned in Chapters 5 and 6) who kindly provided information about their recent surveys and excavations in northern Iran; the Polish and Uzbek archaeologists (also mentioned in Chapters 5 and 6) who generously allowed me to refer to the preliminary results of their excavations at Ayakagytma in the Kyzylkum desert; Dr. Michael Gregg of the University of Pennsylvania Museum for information about his research on unpublished materials from two rockshelters in northern Iran excavated by C. S. Coon in 1949 and 1951; Dr. Richard Meadow of Harvard University for commenting on a draft of the section on domestic animals in Chapter 7; and my colleague Dr. Dorian Fuller of the UCL Institute of Archaeology for commenting likewise on the section on crops in Chapter 7 and for helping me to keep up to date with the burgeoning genetic literature on plant and animal domestication. Lastly, and most warmly, I thank my daughter Zoë, who participated in four of our field seasons (and learned a good deal of archaeology in the process!), and my wife Helen, who accompanied me on our first

insights from the director and his colleagues into the geobotany and ecology of this area of desert long protected from intensive grazing and wood cutting; the Southwest Kopetdag Reserve at Kara-Kala in the Sumbar valley where, courtesy of its director, we visited the local experimental station of the N. I Vavilov Institute of Plant Industry in the wooded environment of the Sumbar valley; and a south-north transect across the Karakum from Ashgabat to Tashauz, Urgench, and Khiva organized for us and led by Nicolai Orlovsky. These excursions added greatly to our understanding of the diverse landscapes of Turkmenistan and have informed the interpretation of environmental and ecological changes offered in this book. So too has the work of the many 20th century Russian naturalists who made detailed studies of the flora and fauna of Turkmenistan, to which V. Fet’s and K. I. Atamuradov’s 1994 volume on the biogeography and ecology of the country provides an invaluable introduction. In addition to all the practical and intellectual support we received, we are especially indebted to those who made the project possible by acting as skillful and patient interpreters, principally Irina Annisimova, Yuri Berezkin, Isabella Moskalyeva, Liya Orlovskaya, and Natasha Solovyova. I also want particularly to thank Katharine Judelson who has always been willing to translate crucial passages in Russian academic papers, who has ensured accuracy and consistency in the transliteration of the titles of Russian publications cited in the Bibliography, and who, in the course of her frequent visits to Turkmenistan and helped by Dr. Kurbansakhatov, tracked down several obscure references. Jennifer Coolidge, too, made a major contribution by translating many of the earlier publications on Jeitun and other aspects of the prehistory of Turkmenistan (reflected particularly in Chapter 5). I am most grateful to the following for drawing the maps, plans, and sections: Catherine D’Alton and Miles Irving, UCL Department of Geography Drawing Office (Figs. 1.1, 1.2, 1.7, 1.15, 1.16, 6.1, 6.2); Colin Merrony, Department of Archaeology, University of Sheffield (Figs. 3.1, 8.2); Alison Wilkins, Institute of Archaeology, University of Oxford (Figs. 8.1, 8.3, 8.5, 8.7, 8.10, 8.20); and Keith Wilkinson, Department of Archaeology, University of Winchester (Figs. 4.1, 4.4, 8.13, 8.14). Other figures were prepared by the authors of the chapters and sections in which they appear. Special thanks are also due to Stuart Laidlaw and Ash Rennie of the UCL Institute of Archaeology for their help with the preparation of digital files of most of the

xv

preface and acknowledgments

But the majestic River floated on, Out of the mist and hum of that low land Into the frosty starlight, and there moved, Rejoicing, through the hushed Chorasmian waste, Under the solitary moon: he flowed Right for the Polar Star, past Orgunjè, Brimming, and bright, and large: then sands begin To stem his watery march, and dam his streams, And split his currents; that for many a league The shorn and parcelled Oxus strains along Through beds of sand and matted rushy isles— Oxus forgetting the bright speed he had In his high mountain cradle in Pamere, A foiled circuitous wanderer:—till at last The longed-for dash of waves is heard, and wide His luminous home of waters opens, bright And tranquil, from whose floor the newbathed stars Emerge, and shine upon the Aral Sea.

visit to Jeitun in 1989 and came again in 1994, and who has for many years kept the home fires burning during my frequent periods of absence on fieldwork in Turkmenistan and elsewhere in the world. It has been a great privilege to undertake field research with Russian and Turkmenian colleagues, and I hope this book will add to knowledge and international awareness of the role the region has played in the prehistory of Eurasia. The desert lands of Central Asia, with their dramatic landscapes and fabled history, have long excited the imagination of Western scholars and travelers, and I cherish personal memories of poring over a world atlas as a schoolboy and being fascinated by the mysteriously named Karakum and Kyzylkum deserts, separated by the valley of the Amudarya—the ‘Oxus’ of Alexander the Great. At school, too, my imagination was fuelled by the lyrical description of the Oxus in the final lines of Matthew Arnold’s epic poem Sohrab and Rustum (1853). It has stayed with me ever since, and though the Aral Sea is no longer the “luminous home of waters” of Arnold’s imagination, those lines may still serve as a fitting preface to this book:

xvi

Introduction

T

he term “western Central Asia” denotes the geographical scope of this book, but how much of the vast Central Asian arid zone it encompasses is not self-evident. In English usage Central Asia comprises the entire area of deserts, plateaus, and mountains from the Caspian Sea through the Karakum, Kyzylkum, Takla Makan, and Gobi deserts to Mongolia and northwest China, whereas Russian authors have traditionally divided the area into a western sector, from the Caspian to the western Tien Shan mountains, known as Middle Asia, and an eastern sector from the eastern Tien Shan to Mongolia and northwest China, referred to as Central Asia. Some Russian authors also include southern Kazakhstan, northeastern Iran, and northern Afghanistan in Middle Asia, and that slightly broader connotation equates closely with what is meant here by western Central Asia. Thus, as defined in this book, the region is bounded in the west by the Caspian Sea and in the south by the Iranian plateau, the mountain systems of Hindu Kush and Pamir, and the southwestern ranges of the Tien Shan, while in the north it extends beyond the valley of the Syrdarya river into the southern steppes of Kazakhstan. Most of the region consists of lowlands less than 200 m above mean sea level that include the Karakum and Kyzylkum deserts, the valleys of the Amudarya and Syrdarya rivers, and the Aral Sea basin (Fig. 1.1). Although western Central Asia encompasses the whole of Turkmenistan and Uzbekistan and parts of five other nation states, this study is concerned mainly with the environments and prehistory of Turkmenistan and adjacent areas of Iran, Afghanistan, and Uzbekistan, with a particular focus on Turkmenistan. There are several reasons for this focus: first, because a great deal of archaeological research over many

decades has been undertaken in Turkmenistan, including the environmental-archaeological project described in detail in Chapters 8, 9, and 10; second, because the climate, vegetation, and animal life of its deserts, plateaus, and mountains is replicated in much of the rest of western Central Asia; and third, because its physical geography and ecology have been very thoroughly studied, particularly under the auspices of the Desert Research Institute in Ashgabat. As explained in the Preface, a large part of this book is devoted to an account of the fieldwork and analytical results of the surveys and excavations undertaken in Turkmenistan between 1989 and 1998, but its broader aims are to set the results of this new research in the chronological and geographical context of the Neolithic period in western Central Asia as a whole, and to provide an introduction to the region for readers unfamiliar with its environments, ecology, and early prehistory. These aims determine the sequence of chapters and the overall division of the volume into five parts. Present and past environmental and ecological conditions are described in Part I. The physiography, climate, vegetation, and animal life of the region are outlined in Chapter 1, and what is currently known about changes in the region’s hydrology, climate, and vegetation during the Pleistocene and Holocene epochs is summarized in Chapter 2. In Chapters 3 and 4 attention shifts from the regional to the local scale to examine the environmental settings of the two main areas in Turkmenistan where we undertook surveys and excavations: near Ashgabat where the early Neolithic site of Jeitun is located at the junction of the Kopetdag piedmont and the Karakum desert, and in the Bolshoi Balkhan massif close to the eastern coast of the Caspian.

2

origins of agriculture in western central asia

Part II is concerned with the prehistoric archaeology of the region with, in Chapter 5, an outline of the history of archaeological research from the late 19th century to the present, and in Chapter 6 an account of the main sites, archaeological sequences, and subsistence economies of the Mesolithic and Neolithic periods. Part III (Chapter 7) focuses more specifically on the crop plants and domestic animals known to have been present in the region during the Neolithic. In it, the question of whether any of them might have been domesticated locally from their wild progenitors is discussed in the light of current biogeographical, genetic, and archaeological evidence from within and beyond western Central Asia. Having described the physical environments, ecology, and prehistoric archaeology of the region, a comprehensive account is given in Part IV of the investigations we undertook in Turkmenistan. Chapter 8 provides detailed descriptions of our excavations at the site of Jeitun, which was occupied by 6000 (calibrated) BCE (Before the Common Era), and which gave its name to the other Neolithic Jeitun-Culture settlements of the Kopetdag piedmont zone. This section of the chapter is followed by short accounts of the surveys and test excavations we carried out at several sites in the Sumbar and Chandyr valleys in the Kopetdag range, and in the Bolshoi Balkhan massif, where we hoped (but failed) to find Mesolithic or early Neolithic antecedents of the Jeitun-Culture settlements. In Chapters 9 and 10 the results of analyses of the materials we excavated at Jeitun and at three rockshelters in the Bolshoi Balkhan are described by the members of the team who undertook them. These chapters include sections on AMS (accelerator mass-

spectrometric) radiocarbon chronology, sediments and soils, phytoliths and pollen, macro plant and animal remains, stone tools, and pottery. Part V brings the book to a close with a synthesis, and general conclusions, presented in two chapters. Chapter 11 offers a model of Neolithic settlement and subsistence. The occupation and economy of Jeitun is considered first, then the Jeitun Culture as a whole, and, third, connections between it and other Neolithic sites in adjacent parts of Uzbekistan, Iran, and Afghanistan—a topic that raises interesting questions about possible interactions between agro-pastoral settlements and hunter-fisher-gatherer groups who may also have herded domestic livestock. Finally, in Chapter 12, the fundamental questions of how and why agriculture began in western Central Asia are debated, in relation to four themes: the biogeography of the Neolithic crops and domestic animals, the overall Neolithic settlement pattern and economy, and the roles of environmental changes and cultural processes in the establishment of settlements mainly dependent on cereal cultivation and livestock herding. The book brings together and interprets a wide array of evidence from several disciplines to investigate the origins of agriculture in part of the “dry heart” of Asia, between the more thoroughly studied areas of early agriculture in Southwest and East Asia. It does so in the hope that it will make a valuable contribution to the comparative worldwide study of agricultural origins, as well as reporting fully the results of the environmental-archaeological project undertaken in Turkmenistan by the British-RussianTurkmen team.

part i

Physical Environment and Ecology

1.1 The Central Asian arid zone, showing the location of the principal deserts, modern nation states, and (enclosed by a dashed line) the approximate extent of western Central Asia as defined in this book.

1.2 Turkmenistan and adjacent parts of Afghanistan, Iran, Kazakhstan, and Uzbekistan, showing the principal physical features and towns mentioned in the text.

1

The Present Environment

T

he purpose of this chapter is to set the scene for those that follow by describing the main features of the physiography, climate, vegetation, and animal life of western Central Asia.

Physiography The principal features of the physical geography of Turkmenistan and the neighboring parts of Iran, Afghanistan, and Uzbekistan are shown in Figures 1.1 and 1.2. The dominant structural contrast is between the high mountains and plateaus that enclose the region in the south, and the extensive lowlands of the Karakum and Kyzylkum deserts that occupy the central area. The southern highlands are part of the great chain of mountains formed in the late Mesozoic to early Tertiary that stretches from west to east across Eurasia. They consist of series of folded parallel ranges and valleys composed of varying combinations of igneous, metamorphic, and sedimentary rocks. Four major mountain systems form the southern boundary zone of western Central Asia: the Elburz (Alborz) south of the Caspian, the Kopetdag south of the Karakum, and the northern ranges of the Hindu Kush-Pamir and southwestern ranges of the Tien Shan south and east of the Kyzylkum. The Elburz, northern Hindu Kush-Pamir, and western Tien Shan reach elevations of over 5,000 m and the highest peaks of the Kopetdag exceed 2,500 m. Areas of lower plateaus and foothills separate the four major mountain systems and form transitional zones between the high ranges and the deserts to the north. The lower valleys of the Gorgan and Atrek rivers occupy the gap in the mountain barrier between the eastern end of the Elburz and the western end of the Kopetdag; the valleys

of the Tedzhen and Murghab rivers cross the more extensive break in the mountain barrier between the eastern Kopetdag and the western Hindu Kush; and the upper valley of the Amudarya separates the Hindu Kush-Pamir from the southwestern ranges of the Tien Shan (the Turkestan, Alay, and Hissar mountains). In southern Turkmenistan the northernmost range of the Kopetdag forms the abrupt down-faulted edge of the Iranian plateau and overlooks a submontane zone that varies in width from 10 to 40 km. The zone consists of an upper piedmont made up of coalesced alluvial fans, and a more gently sloping lower piedmont plain that merges northward into the sands of the Karakum (Fig. 1.3, color). The north-facing slopes of the Kopetdag are drained by some 30 rivers that cross the piedmont, most of which formerly dissipated in the sands of the southern Karakum. Today, however, the flow of many of them is interrupted by the Karakum Canal and diverted for irrigation and to supply water to the capital, Ashgabat, and other piedmont towns. The rivers are fed by groundwater, as well as rainfall, and are less prone to extreme flooding than longer rivers such as the Tedzhen and Murghab that rise in the mountains much farther south in Iran and Afghanistan. The piedmont rivers flood after spring rains, and they have deeply dissected the coarse sediments of the upper piedmont, but because they are partly fed by groundwater their flow remains more stable through the year, with maximum discharges between March and May and minimum from June to October (Dolukhanov 1981:366, 375). The channel of one of the rivers, the Kara Su, passes close to the site of Jeitun (this volume, pp. 28–29). To the northwest, closer to the eastern shore of the Caspian, the Bolshoi (“great”) and Maly (“little”) Balkhan massifs are isolated outliers of the Kopetdag

6

origins of agriculture in western central asia

1.4 Aerial view of takyrs on the lower Kopetdag piedmont west of Ashgabat, April 1994.

1.6 Pentagonal cracks formed following evaporation of standing water on the clay surface of the takyr shown in Figure 1.5. 1.5 A takyr north of Erbent in the Central Karakum, April 1993.



the present environment

mountain system (this volume, pp. 35–36). They are separated from the Caspian Sea by the Caspian or West Turkmen lowland and the Krasnovodsk plateau, and from each other by the valley of the former Uzboi river which in the past flowed from the Sarykamysh depression to the Caspian (Fig. 1.2). North of the Krasnovodsk plateau and the Kara-Bogaz Gol bay, the hilly, dissected Ustyurt plateau extends into southwestern Kazakhstan and northwestern Uzbekistan. The Karakum desert stretches from the Uzboi channel in the northwest to the valley of the Amudarya in the east, and from the Kopetdag mountains in the south to the oasis of Kwarazam (or Khiva) in the north. It has a total area of c. 500,000 km2 and is divided by a chain of large depressions known as the Unguz into a northern area of dissected terrain (the Trans-Unguz Karakum) from the more extensive sand desert of the Lowland Karakum to the south, which is itself divided into the Central Karakum between the Unguz and the Kopetdag piedmont, and the Southeast Karakum (Fig. 1.2; Walter and Box 1983, and see Babaev 1994 for more detailed descriptions of the Karakum and other landscapes of Turkmenistan). The Central Karakum increases in height from 20 m above mean sea level (msl) in the west to 200 m in the east, and its topography consists mainly of sand ridges, often interspersed with takyrs: level or very gently sloping areas of saline clayey soils on which water accumulates and gradually evaporates. Algae and lichens often grow on takyrs, and they are most extensively developed in the southern half of the desert and on the adjoining lower Kopetdag piedmont (Figs. 1.4 1.5, 1.6, and this volume, p. 29). Takyrs are valued as sources of surface and subsurface water and cultivable land. During rains they can become temporary lakes, often used for watering livestock, and crops are grown on them, irrigated from wells and pits dug into their smooth surfaces. Where superficial deposits are saturated, seasonally or for longer periods, solonchak soils occur (cf. Bridges 1997:120–22). The Southeast Karakum, which reaches elevations of 350 m in the south, consists mainly of sand ridges, interrupted only by the lower valleys and cultivated oases of the Tedzhen and Murghab rivers and the clay plains that stretch between them. Farther south, approaching the border with Afghanistan, the sands of the Karakum give way to the plateaus of Badghyz east of the Tedzhen and Karabil east of the Murghab (Fig. 1.2). The Badghyz plateau rises to 1,255 m and its surface is broken by hills often

7

separated by enclosed depressions containing takyrs, solonchaks, and even salt lakes, two of which occur in the largest depression, Yeroyulanduz. The Karabil plateau is lower, its hilly relief is more uniform, and it contains fewer depressions than Badghyz. In the southeastern corner of Turkmenistan, east of the Amudarya valley, the Kugitang mountains rise to c. 3,000 m. They are part of the higher Hissar range, one of the westernmost spurs of the Tien Shan in western Tajikistan and southern Uzbekistan east of the Kyzylkum lowland. The Kyzylkum desert is much smaller (c. 200,000 2 km ) than the Karakum. No rivers debouch into it today, but in the past the Zeravshan, which rises in the ranges of the western Tien Shan, flows though Samarkand, and forms a deltaic zone at Bokhara, was a tributary of the Amudarya, which it joined close to the present city of Chardzhou (Fig. 1.2). In the southern Kyzylkum there are many now-dry lake beds and watercourses, such as the ancient Lake Lyavlyakan and right-bank tributaries of the Zeravshan. Neolithic sites exist in the vicinity of these former water bodies, as they do also around lake beds and wind-deflated depressions among a group of isolated low mountain ridges in the central Kyzylkum (Dolukhanov 1986: Fig. 2, and this volume, pp. 64–65). The Amudarya and Syrdarya—the Oxus and Jaxartes of Classical authors—are the longest rivers of western Central Asia. They rise in the Pamir and western Tien Shan respectively and flow northwest, parallel with each other, to their deltas in the Aral Sea basin. Under natural conditions, before their flow regimes were altered for irrigation, they flooded in spring and early summer after winter-spring rainfall and the melting of snow and glacier ice in the mountains. They discharged large quantities of sediment into their lower valleys, forming shifting islands and actively eroding their banks. Southwest of the Amudarya, the Murghab and the Tedzhen likewise rise in the mountains to the south and are subject to variable annual floods, but they are shorter, less powerful, and dissipate in the sands of the Southeast Karakum after feeding irrigation systems around the cities of Mary (ancient Merv) and Tedzhen. The only other large river system in western Central Asia is formed by the Atrek and its main tributaries, the Sumbar and the Chandyr, which drain the western Kopetdag. The Atrek rises east of Quchan in northeastern Iran, flows through a broad intermontane valley, and after its junction with the Sumbar crosses the Cas-

8

origins of agriculture in western central asia

pian lowland where, together with the Gorgan, it has formed a broad deltaic plain (Fig. 1.2). Today the Atrek seldom reaches the Caspian because most of its flow is diverted for irrigation, and, for the same reason, the Sumbar does not now connect with the Atrek, except when exceptionally high spring floods occur.

Climate The climate of western Central Asia is characterized by the extreme aridity, with very hot summers and cold winters, that is typical of mid-latitude continental interiors distant from the moderating influence of the ocean. The dryness of the climate is accentuated by the “rain-shadow” effect of the massive mountain ranges and plateaus to the south that act as barriers to the influx of humid air generated by the Indian monsoon, and even the effect of the Caspian Sea is limited to slight increases in summer humidity and breeziness along its

eastern shore. It is only in the deep valleys and at lower elevations in the southern mountains and plateaus that more moderate conditions prevail. There, summer and winter temperatures are less extreme and annual precipitation is higher. For example, in Turkmenistan average annual precipitation exceeds 300 mm in parts of the central Kopetdag, while the piedmont and southeastern plateaus receive between 200 mm and 300 mm, and precipitation diminishes north across the Karakum to less than 100 mm (Fig. 1.7). This contrast between aridity in the desert and somewhat higher precipitation and less extreme temperatures in the mountains to the south is duplicated in Uzbekistan and southwestern Tajikistan by the contrast between very low rainfall in the Kyzylkum and more humid conditions and more moderate temperatures in the southern plateaus and mid-altitude mountain ranges. Throughout western Central Asia most of the annual precipitation occurs in the form of rain and snow between October and April, with maxima in

1.7 Turkmenistan: mean annual precipitation (based on data in the Atlas of the Turkmenian SSR, Moscow: Main Agency for Geodesy and Cartography affiliated to the Council of Ministers of the USSR, 1982, p. 11).



the present environment

late winter/early spring. There is very little or no rain in the summer months, and annual and monthly totals of precipitation vary greatly from year to year. In most of the region, particularly in the lowlands, severe drought prevails during the summer (June–August). It is accentuated by very high air temperatures that, for example, reach average maxima of 35–40o C in June and July in the Central and Southeast Karakum. In autumn (September–November) average monthly air temperatures decrease rapidly, associated with some precipitation, and in the winter (December–February) average monthly maxima fall to between 0o C and 10o C, with minima generally below 0o C. In the spring (March–May) average monthly temperature increases very rapidly, reaching 25o C in May in the Karakum, and in March and April the monthly maxima of precipitation are also reached. Spring is thus the optimal season for plant growth. The high temperatures and drought of summer are associated with low relative humidity of the air (20–30% in the Karakum) and high annual rates of evaporation from water surfaces (e.g., 2,000–2,300 mm in the Karakum, falling to c. 1,500 mm on the Kopetdag piedmont and 1,000 mm along the Caspian shore). Soil surfaces also reach very high temperatures during the summer, with correspondingly high rates of evaporation of soil moisture that severely curtail the growth of most plants other than very deep-rooting desert shrubs (Orlovsky 1994:30–37).

Vegetation The vegetation of western Central Asia consists mainly of xerophytic communities of woody and herbaceous plants adapted to the strongly seasonal climatic regime, with prolonged summer drought, cold winters, and unevenly distributed rainfall. The principal types of forest, woodland, grassland, and desert vegetation are described in this section, much of which is based on more detailed descriptions of the vegetation of Turkmenistan by Fet (1994), Popov (1994), and Rustamov (1994). Latin binomials given for the plants referred to follow those published in Nikitin and Geldykhanov (1988).

Forest and Woodland Forests of broadleaf deciduous and evergreen coniferous trees are largely restricted to the mountain sys-

9

tems of the Elburz, Kopetdag, northern Hindu Kush, and southwestern Tien Shan, which acted as refugia for mesophytic taxa as the climate became more arid and continental from the late Tertiary through the Quaternary period. There is palaeobotanical evidence that some broadleaf trees, such as oak (Quercus spp.), became extinct in the Kopetdag (this volume, p. 33), and today fragments of forest containing Turkmen maple (Acer turcomanicum), Syrian ash (Fraxinus syriaca), cornel (Thelycrania meyeri), elm (Ulmus carpinifolia), and, less commonly, oriental plane (Platanus orientalis) persist in deep valleys on north-facing slopes of the western and central Kopetdag. They are associated with smaller trees, shrubs, climbers, and herbs that include many wild species with edible nuts and fruits such as walnut (Juglans regia), almond (Amygdalus spp.), apple (Malus spp.), apricot (Armeniaca vulgaris), cherry (Cerasus spp.), fig (Ficus carica), pear (Pyrus spp.), plum (Prunus spp.), hawthorns (Crataegus spp.), blackberries (Rubus spp.), and grape (Vitis sylvestris). The survival of many nut and fruit trees is now threatened by over-exploitation and habitat destruction, and several are included in the Red List of Trees of Central Asia (Eastwood, Lazkov, and Newton 2009) where their status is recorded, according to the IUCN Red List categories, as Endangered (e.g., almond, apricot, pear), Vulnerable (e.g., apple), and Near Threatened (e.g., walnut). Broadleaf deciduous trees also occur at lower altitudes along some of the rivers and streams of the Kopetdag, in the Atrek, Sumbar and Chandyr valleys, and farther east in the valleys of the Tedzhen, Murghab, Amudarya, and Syrdarya. They form narrow riparian forests known as tugai (Fig. 1.8, color) that are dominated by tamarisks (Tamarix spp.), poplars (Populus spp.), willows (Salix spp.), and, more rarely, oriental plane (Platanus orientalis) and Russian olive (Elaeagnus orientalis), associated with tall grasses and reeds (Erianthus ravennae, Arundo donax, and Phragmites australis, syn. P. communis). Xerophytic woodlands comprise another type of arboreal vegetation. They are characteristically dominated by one or more drought-tolerant species, particularly juniper and pistachio, and they occur mainly at lower elevations in the mountains, foothills, and desert plateaus. For example, stands of Turkmen juniper (Juniperus turcomanica) occur in the western and central Kopetdag above 200 m and also in the Bolshoi Balkhan massif above 400 m (this volume, p. 39). Juniper wood is valued for construction as well as firewood and this, combined with damage inflicted

10

origins of agriculture in western central asia

by browsing livestock, has greatly reduced the extent of juniper woodlands in recent centuries (Kamakhina 1994:145). Today pistachio (Pistacia vera) woodlands occur in Turkmenistan mainly on the arid Badghyz plateau, with only scattered groups of trees remaining elsewhere in the foothills of the Kopetdag and Kugitang mountains. Before they were reduced in extent and their regeneration prevented by browsing livestock, logging, and the harvesting of their edible nuts, they grew extensively on the central and eastern Kopetdag piedmont, although not in the western Kopetdag or the Balkhan mountains. In Badghyz the pistachio trees grow sparsely in open-canopy groves associated with wild barley (Hordeum spontaneum subsp. spontaneum) (Figs. 1.9, 1.10, color) and other grasses and sedges (e.g., Poa bulbosa, Carex pachystylis), in a formation sometimes referred to as pistachio semi­ savanna (Popov 1994:175). Another category of woody vegetation that occurs in the foothills, plateaus, and valleys of southern Turkmenistan and northeastern Iran is known as shiblyak. It consists of a diverse range of small trees and shrubs, many palaeobotanically of Mediterranean origin, and it is most highly developed in the southwestern Kopetdag. Taxa typical of shiblyak include wild apple, almond, cherry, fig, pear, and plum, and species of Berberis, Celtis, Colutea, Cotoneaster, Crataegus, Jasminum, Lonicera, and Paliurus. In parts of the Sumbar, Chandyr, and Atrek valleys, which experience a summer-dry Mediterranean climate, other species with edible fruit occur, such as mulberry (Morus alba), jujube (Zizyphus jujuba), pomegranate (Punica granatum), medlar (Mespiles germanica), and quince (Cydonia oblonga, syn. C. vulgaris). Turkmen maple is also present as a small tree or shrub in shiblyak in the western Kopetdag, but it has been reduced by intensive cutting and its growth is suppressed by thickets of the invasive shrub Christ’s thorn (Paliurus spina-christi). Shiblyak communities also exist in the central and eastern Kopetdag, Bolshoi Balkhan, and Kugitang mountains, but in these areas they are less floristically diverse and contain fewer species that provide edible fruits.

Steppe Steppe dominated by perennial grasses, associated with a variety of other herbaceous plants (forbs), occurs on plateaus and some of the more level, less dissected surfaces in the foothills of the major mountain

systems. Undisturbed steppe grasslands are dominated by species of Stipa and Festuca, associated with other grasses such as Elytrigia trichophora and Poa bulbosa and a variety of forbs including Galium verum, Helianthemum salicifolium, and Thymus transcaspicus, whereas over-grazing by domestic livestock leads to the replacement of Stipa and Festuca species by E. trichophora and other indicators of disturbance such as Hordeum bulbosum, Convolvulus subhirsutus, Ferula oopoda, and Perovskia abrotanoides. Much of the steppe that formerly occupied extensive areas in piedmont zones has been severely degraded by grazing or plowed up for cereal cultivation, but some relatively undisturbed patches still exist, for example in more remote areas of the upper Kopetdag piedmont (Fig. 1.11, color).

Desert Vegetation Desert vegetation, adapted to the extreme aridity, high rates of evaporation, and uneven seasonal distribution of precipitation, occupies well over half the land surface of western Central Asia. Shrubs that range in size from small, low-growing semi-shrubs to large tree-like forms dominate extensive areas of sand, gravel, and rocky desert. They include both sand-tolerant psammophytes and salt-tolerant halophytes, as does the associated herbaceous flora. The clay surfaces of takyrs lack vascular plants although they can support colonies of algae and lichens. The only areas devoid of vegetation are the most mobile sand-dune surfaces. The most extensive type of desert vegetation is that dominated by semi-shrubs, particularly sagebrush (Artemisia spp.) growing on rocky and sandy surfaces in the lowlands and foothills, and halophytic species of Salsola, Anabasis, and other genera on more saline surfaces among sand dunes and on alluvial soils along stream channels. Other plants commonly associated with the sagebrush communities include small shrubs such as species of Ephedra, Astragalus, and Halothammus, perennial herbs such as species of Stipagrostis, Iris, Carex, Allium, Tulipa, and Ferula, and a variety of annuals. Species of Salsola and Anabasis are characteristic of the halophytic communities, as well as the desert sedge Carex pachystylis and low-growing forms of tamarisk and black saksaul (Haloxylon aphyllum). A second major type of desert shrub vegetation is that dominated by larger shrubs of the genera Haloxylon, Ammodendron, Calligonum, Salsola, and Ephedra.



the present environment

They grow extensively in sand deserts, clay lowlands, and on the alluvial soils of modern and ancient deltas, and they are divisible into saksaul and psammophyte deserts. The former vary according to whether the white or black saksaul (Haloxylon persicum and H. aphyllum) is the dominant species. White saksaul typically grows on sand dunes, less frequently in intradune depressions and other lowland areas. It is well developed on the dune systems of the Karakum where it supports a structurally diverse, layered community consisting of large shrubs, smaller shrubs, and a herbaceous layer of tall grasses such as Stipa pennata and various perennial and annual forbs. White saksaul can reach 5 m in height but does not form dense thickets (Fig. 1.12, color). Although its leaves are bitter, it is browsed by sheep, goats, and camels, especially in autumn and winter, and this, together with its exploitation for firewood, has led to its replacement in some areas by species of Calligonum. Unlike H. persicum, black saksaul seldom grows on sandy soils. It occurs principally on alluvial lowlands, where it sometimes forms a transitional belt between desert and riparian woodland. It is found in the valleys and deltas of the Amudarya, Syrdarya, Murghab, and Tedzhen rivers, along the Uzboi channel and around the Sarykamysh depression, as well as in smaller depressions in the deserts (Fig. 1.13, color). Its root system enables it to tap ground water to great depths and it can reach 9 m in height (Fig. 1.14, color), but it supports less complex layered communities of large and small shrubs and herbs than white saksaul. Black saksaul is not an important source of fodder for domestic livestock although camels do browse on it, but it provides high-quality firewood and in the past was extensively burned for charcoal. The second category of large-shrub desert vegetation—the psammophyte deserts—is less extensive and less floristically diverse than the saksaul deserts. They are restricted to stable and mobile sand dunes and they too are dominated by species of Ammodendron, Calligonum, Salsola, and Ephedra. Some species, such as Ammodendron connollyi, are adapted to growth on mobile sands and can reach a height of 10 m, whereas the lower-growing species of Calligonum and Salsola are effective sand stabilizers able to grow on both stable and mobile sands. Many communities dominated by Calligonum are thought to have replaced white saksaul communities destroyed by human activities. Throughout the deserts, except in areas where natural reserves have been established (for example, in

11

Turkmenistan in Badghyz and the Southeast Karakum at Repetek), the vegetation is grazed and browsed by sheep, goats, and camels. Grazing continues through the year, and over the past century the number of livestock has greatly increased and overgrazing has intensified, as has the cutting of wood for fuel by the increased human population. These pressures have been particularly intense in the south-central Karakum, including the zone immediately north of the Kopetdag piedmont where the site of Jeitun is located. Today the desert vegetation in this zone is severely degraded, and it is therefore very difficult to infer at all precisely, from its present condition, the likely state of the vegetation near the site in Neolithic times (see this volume, pp. 31–32).

Animal Life Wild Mammals The varied desert, plateau, and mountain environments of western Central Asia provide habitats for a great diversity of vertebrates and invertebrates. The wild mammals include large herbivorous herd animals such as sheep, goat, gazelle, onager, saiga, and deer, and various smaller species such as hare and a great variety of rodents. Predators and scavengers including wild pig, brown bear, hyaena, jackal, wolf, fox, leopard, tiger, cheetah, lynx, wild cats, badger, otter, marten, and polecat are also native to the region. However, the populations of most of these animals have been much reduced in recent centuries by hunting and habitat destruction, and the numbers of many wild species have continued to decline in recent years as human pressures on their populations have intensified. In this section, brief descriptions are given of the distribution of the principal species, especially those that have been exploited by humans for food and other purposes. Scientific nomenclature and information on geographical distribution is from Clutton-Brock (1999), Heptner and Naumov (1989, 1992, 1998, 2002), and Wilson and Reeder (2005). Information is also given on the recent population status and distribution of threatened species based on the 2008 IUCN Red List of Threatened Species (http//www.iucnredlist.org), with additional data for Turkmenistan from Rustamov and Sopyev (1994). Red List categories are referred to as follows: CR (Critically Endangered), EN (Endangered), VU (Vulnerable), LR (Lower Risk), and NT (Near Threatened).

12

origins of agriculture in western central asia

1.15 Generalized recent distribution of wild Asiatic mouflon, urial, and argali sheep in Southwest and western Central Asia (based on data in Clutton-Brock 1999:69–72 and Nadler et al. 1973:117–21).

Several of the wild mammals described below were identified in the bone assemblages from the Neolithic site of Jeitun (see p. 17).

Urial Sheep (Ovis vignei Blyth) The range of the urial stretches eastward from the Elburz mountains in northern Iran, where it overlaps with the Asiatic mouflon (Ovis orientalis) at the eastern limit of the latter’s range and where a hybrid race of the two species exists (Clutton-Brock 1999:72), through the mountains and plateaus of Turkmenistan, southwestern Kazakhstan, Uzbekistan, and Tajikistan, and south across Afghanistan to western Pakistan and Kashmir (Fig. 1.15). In Turkmenistan the urial occurs in the Bolshoi Balkhan massif and through the Kopetdag to Badghyz, Karabil, and Kugitang, but its total population is now estimated to be below about 3,000 (Red List category VU). According to Nadler et al.

(1973:118–20) the Balkhan urials belong to a steppe subspecies, O. v. arkal, which formerly inhabited the Krasnovodsk plateau and the northern Karakum as far east as the delta of the Amudarya, a view previously advanced by Zeuner (1963:159–60) who suggested that the present restriction of sheep to mountainous areas was partly due to “either persecution or to the absorption of lowland races into domesticated breeds” (ibid., p. 160).

Bezoar Goat (Capra aegagrus Erxleben) The bezoar is native to the mountains of Southwest and western Central Asia from Turkey to Iran, Turkmenistan, Afghanistan, and southwestern Pakistan (Fig. 1.16); and in the Late Pleistocene/Early Holocene its range reached into the southern Levant (Uerpmann 1987:114). It was formerly widespread in Turkmenistan from the Krasnovodsk plateau and



the present environment

13

Am ud ary a

Ti g

ris

Ind us

Nile

E

up hra tes

Bezoar Capra aegagrus Markhor Capra falconeri Ibex Capra ibex

0

km

1000

1.16 Generalized recent distribution of wild bezoar, ibex, and markhor goats in Southwest and western Central Asia (redrawn from Harris 1962: Fig. 1).

Balkhan massifs east to Badghyz, but during the 20th century its population was greatly reduced (Red List VU). Rustamov and Sopyev (1994:210) suggested that the total population of wild bezoars in Turkmenistan probably did not exceed 2,000, almost all in the central Kopetdag with small isolated groups elsewhere in the Kopetdag and the Bolshoi Balkhan, but Korshunov (1994:231) estimated the total population as about 8,000. Although often regarded as strictly mountain animals, bezoars are not limited to higher elevations. They also favor rocky slopes, cliffs, and deep valleys where they can feed all year round on juniper and a wide variety of shrubs. In Turkmenistan they formerly ranged from the highest peaks of the Kopetdag to semidesert foothills as low as 100–200 m, for example around the base of (as well as within) the Bolshoi Balkhan massif, which in the past sustained particularly large numbers of wild goats.

Markhor (Capra falconeri Wagner) This goat is native to the mountains of western Central Asia including southeastern Turkmenistan, southern Uzbekistan and Tajikistan, northern Afghanistan, and into northern and central Pakistan (Fig. 1.16). It may formerly have been even more widely distributed in the Central Asian mountains, but fewer than 2,500 mature animals now exist in small sub-populations (Red List EN). In Turkmenistan it is now restricted to the western slopes of the Kugitang mountains where a population of fewer than 200 of the subspecies C. f. heptneri recently survived.

Goitered Gazelle (Gazella subgutturosa Güldenstädt) This gazelle is the largest of the three western Asiatic species. It has a more easterly distribution than the smaller dorcas and Arabian species (G. dorcas and G. gazella) that includes Iran, Turkmenistan, where it

14

origins of agriculture in western central asia

is known as the geran (dzheiran), Uzbekistan, southern Kazakhstan, Afghanistan, and east to northwestern China and Mongolia. In the 19th century large herds ranged widely over the lower mountain slopes, plateaus, and desert lowlands of western Central Asia, including the Kopetdag piedmont, feeding mainly on steppe and desert herbs and shrubs, but their population has since been drastically reduced (Red List VU). In the early 1940s the total population in Turkmenistan was estimated as still over 100,000, but since then, despite a ban on hunting gazelles introduced in 1950, their population and range has continued to decrease. Now probably fewer than 3,000 animals survive, most of them in the Badghyz Natural Reserve where they are formally protected, although still vulnerable to illicit hunting. Before the railway between Krasnovodsk (Turkmenbashi) on the Caspian and Chardzhou (Turkmenabat) on the Amudarya was built—a barrier later reinforced by construction of the Karakum Canal (Fig. 1.2) which began in 1954—gazelles migrated seasonally from summer grazing grounds in the foothills of the Kopetdag to winter grazing in the Karakum. In southeastern Turkmenistan these seasonal migrations continued until the 1950s, with herds moving from the Paropamiz mountains in northwestern Afghanistan across Badghyz to the Karakum, but this pattern has been disrupted and the surviving gazelle population is now restricted to very limited seasonal movements in the Badghyz Reserve (Valerii Kuznetsov, pers. comm. 1992).

Onager (Equus hemionus Pallas) The onagers or hemiones are a group of Asiatic wild asses that formerly ranged over extensive areas of steppe and desert between the Levant and Mongolia. Five subspecies are generally recognized: the Syrian onager (E. h. hemippus), which is now extinct, the Iranian onager (E. h. onager), the Turkmen kulan (E. h. kulan), the Indian khur (E. h. khur) of the Thar desert, and the Mongolian kulan (E. h. hemionus) of northern Mongolia and the Gobi desert (the Gobi population being sometimes classified as a separate subspecies, E. h. lutens). In historical times onagers inhabited most of western Central Asia but in recent centuries their range declined substantially, and now only small isolated sub-populations exist in northern Iran, southeastern Turkmenistan, southern Uzbekistan, and (re-introduced) southern Kazakhstan (Red List EN). The Turkmen kulan formerly occurred throughout Turkmenistan and in the 19th century their popula-

tion still numbered many thousand, but by 1941, when the Badghyz Natural Reserve was established to save them from extinction, only 250 remained. By the early 1990s there was said to be a breeding population of 2,000–3,000 in Badghyz, and a few small herds had been re-established on drier parts of the eastern Kopetdag piedmont where there is little competition from pastoralists for the sparse grazing available (Valerii Kuznetsov, pers. comm. 1992).

Saiga (Saiga tatarica Linnaeus) The saiga is a small bovid often referred to as an antelope, although normally classified with goats and sheep in the subfamily Caprinae (Clutton-Brock 1999:209). Recent phylogenetic analysis of mitochondrial DNA sequences in the family Bovinae has however demonstrated that the saiga is most closely related to gazelles (Kuznetsova, Kholodova, and Luschekina 2002). Saigas are native to Eurasian steppes and semideserts and they ranged historically from southeastern Europe to Mongolia and western China. Today small populations remain only in Kazakhstan, Turkmenistan, Uzbekistan, and Mongolia. The species is divided into western and eastern (Mongolian) subspecies: S. t. tatarica and S. t. mongolica. Large populations of the western subspecies formerly inhabited extensive areas in Central Asia, migrating seasonally to feed on the steppes to the north in spring and summer and the deserts to the south in autumn and winter. Between the Caspian and Aral Seas large herds moved south annually across the Ustyurt plateau to the Karakum for winter grazing. Their winter migrations reached the middle Amudarya valley, and in the severe winter of 1946–1947 they penetrated into the southwestern Karakum. Single animals were observed even in the foothills of the western and central Kopetdag (Sapamuradov 2005), an observation reinforced by Valerii Kuznetsov’s report (pers. comm. 1992) that some saiga herds used to migrate south in the summer from the Karakum to the southern margin of the desert and the Kopetdag piedmont. These observations are particularly interesting in relation to reported finds of saiga bones at the Neolithic site of Jeitun on the southern edge of the Karakum and at the Chalcolithic/ Bronze Age sites of Ilgynly-depe and Altyn-depe on the eastern Kopetdag piedmont (Kasparov:1992:51, 61; 1994:148). During the 19th and 20th centuries, overhunting for meat, hides, and horns greatly reduced saiga populations. Although they recovered after 1919 when



the present environment

a Soviet law was passed prohibiting their hunting (Linnard 1963), their numbers have more recently been drastically reduced, especially by killing them for their horns which are highly prized in Chinese traditional medicine. In 2002 the saiga was added to the Red List Critically Endangered category (Milner-Gulland 2004; Milner-Gulland et al. 2001), the total population having fallen from 1,240,000 in the mid 1970s to about 50,000, mostly in Kazakhstan. Between 2,000 and 3,000 are believed still to winter in Turkmenistan, and there are plans to establish a saiga sanctuary in the north of the country from the Kara-Bogaz Gol bay to the Sarykamysh lake (Sapamuradov 2005).

Bokhara Deer or Khangul (Cervus elephus bactrianus Lydekker) This subspecies of red deer formerly inhabited areas of tugai (riparian forest) in the valleys of the Syrdarya, Amudarya, Murghab, Tedzhen, and Atrek, but now they survive (if at all) only in very small numbers in a reserve in the Amudarya valley (no Red List category available).

Eurasian Wild Pig (Sus scrofa Linnaeus) Wild pigs inhabit broadleaf forests and woodlands across Eurasia and in North Africa. Over 15 locally adapted subspecies have been described, two of which are native to western Central Asia: S. s. attila and S. s. nigripes, and a third, S. S. davidi, which is marginally so, ranging from eastern Iran to Pakistan and western India and perhaps north into Tajikistan (Clutton-Brock 1999:91–93; Groves 1981:29–24). Wild pigs are still present throughout the region, where they live mainly in tugai and other forested, wooded, and shrubby areas where there is ground cover and access to water.

Brown Bear (Ursus arctos Linnaeus) Brown bears formerly occupied mountain, steppe, and desert habitats throughout western Central Asia but their numbers were substantially reduced during historical times, and more drastically in the recent past. Isolated populations remain in northern Iran and Afghanistan, southern Kazakhstan and Uzbekistan, Tajikistan, and Kirgizstan. They no longer live and breed in Turkmenistan, but they occasionally visit the western Kopetdag from northern Iran. Records of such visits were common in the 1920s, but between 1961 and 1972 fewer than 10 sightings of bears or their footprints were reported, from the Sumbar and Chandyr valleys, and in 1980 a female

15

bear with cubs was encountered in the eastern Kopetdag close to the Iranian frontier (Rustamov and Sopyev 1994:206). In prehistoric and historical times breeding populations probably existed throughout the Kopetdag.

Striped Hyaena (Hyaena hyaena Linnaeus) This carnivorous scavenger was once widespread from northern and eastern Africa through Southwest Asia to Central and South Asia, and in the early 20th century it was still well established in northern Iran, southern Turkmenistan and Uzbekistan, and Tajikistan. In Turkmenistan hyaenas inhabited the Balkhan and Kopetdag mountains and the Badghyz and Karabil plateaus until recently, but now probably fewer than 100 remain in reserves in the Kopetdag and Badghyz (Red List NT).

Golden Jackal (Canis aureus Linnaeus) Smaller than the hyaena and closely related to the wolf, the common jackal is, like them, a hunter and scavenger. It is widely distributed across mid- and low-latitude Asia from Turkey to Indonesia, inhabiting mainly valleys, plains, and foothills. It is still well established in western Central Asia, especially in the river valleys. It feeds on a wide variety of rodents and other small mammals, and preys on domestic sheep, goats, and cattle.

Gray Wolf (Canis lupus Linnaeus) Before they were intensively hunted in recent centuries, wolves ranged right across Eurasia from Ireland to Japan (and into North America). Many local races evolved, but in general, northern wolves tend to be larger and more gregarious than the wolves of Southwest, South, and Central Asia. Wolves still exist throughout western Central Asia where they used to prey on the formerly abundant herds of gazelle and other wild ungulates, as well as many smaller animals, but their numbers have been greatly reduced by pastoralists to protect their herds of sheep and goats from predation.

Red Fox (Vulpes vulpes Linnaeus) and corsac fox (V. corsac Linnaeus) The red fox occupies a wide range of habitats throughout western Central Asia, whereas the corsac fox is restricted to desert habitats in the region. Both species prey on a variety of rodents and other small mammals, birds, reptiles, and insects, and they have traditionally been hunted for their pelts.

16

origins of agriculture in western central asia

Persian Leopard (Panthera pardus saxicolor Pocock)

Eurasian Lynx (Lynx lynx Linnaeus)

As recently as the late 19th and early 20th centuries this subspecies occurred widely across western Central Asia, in northern Iran, including the Elburz mountains and the southern fringe of the Caspian Sea, Turkmenistan, and parts of Uzbekistan and Tajikistan. In Turkmenistan leopards ranged from the Balkhan mountains through the Kopetdag to Badghyz and Karabil, but now probably fewer than 50 animals survive in reserves in the Kopetdag and Badghyz (Red List EN). In the past leopards preyed extensively on gazelles and were also the principal predators of bezoar goats.

The lynx was formerly widespread in western Central Asia, but now it only survives in isolated populations in northern Iran and in some mountain habitats farther east. In Turkmenistan lynx are still occasionally encountered in the western and central Kopetdag, but (like the brown bears) these animals are probably visitors from northern Iran.

Caspian or Turanian Tiger (Panthera tigris virgata Illiger) This subspecies formerly inhabited tugai and other forest habitats in much of western Central Asia. It is now extinct, having been almost completely exterminated early in the 20th century. In Turkmenistan tigers formerly lived in the Atrek, Sumbar, Chandyr, Tedzhen, Murghab, Amudarya, and Syrdarya valleys, where they preyed particularly on Bohkara deer and wild pigs. It is interesting to note that a tiger bone was identified in the assemblage of wild and domestic animal bones excavated at the Chalcolithic site of Ilgynlydepe on the eastern Kopetdag piedmont, not far from the Tedzhen river (Kasparov 1994:148).

Asiatic Lion (Panthera leo persica Meyer) This subspecies formerly ranged from Southwest into western Central Asia, but it is now extinct in the region. There are a few historical references to lions being killed in southern Turkmenistan, where their main prey would have been gazelles and onagers (Valerii Kuznetsov, pers. comm. 1992).

Cheetah (Acinonyx jubatus venaticus Griffith, syn. A. j. raddei Hilzheimer) This Asian subspecies formerly ranged widely in desert, steppe, and woodland habitats in Southwest and Central Asia, including Iran, Turkmenistan, Uzbekistan, and Afghanistan, but its numbers were drastically reduced in recent centuries by depletion of their wild prey, especially gazelles, loss of habitat, and overhunting. Cheetahs remained fairly abundant in Badghyz and Karabil until the late 1950s and some survived in northwestern Turkmenistan until the early 1970s, but this swiftest of all land mammals is now extinct throughout western Central Asia, except in north-central Iran where 60–100 animals are estimated still to survive (Red List CR).

Caracal or Sand Lynx (Caracal caracal Schreber) Smaller than the Eurasian lynx, this carnivore is widely distributed in Africa and Southwest Asia. It reaches the eastern limit of its range in the deserts and other lowland habitats of western Central Asia, where it preys mainly on small and medium-size mammals, including the desert hare (Lepus tolai). Its populations have been severely reduced by hunting and irrigation schemes, and probably fewer than 200–300 animals survive in Turkmenistan.

Manul or Steppe Cat (Otocolobus manul Pallas, syn. Felis manul Pallas) This wild cat once ranged widely in western Central Asia, where it lived mainly in montane and riverine habitats. It is now rare, partly as a result of being hunted for its fur, but small isolated populations remain (Red List NT). In Turkmenistan it was formerly present in the Kopetdag, the Bolshoi Balkhan, Badghyz, Karabil, and parts of northern Turkmenistan, but it now survives only in very small numbers in the Kopetdag, the Bolshoi Balkhan, and possibly Badghyz.

Sand Cat (Felis margarita Loche) This cat, which is thought to be closely related to the manul, has a wide distribution from North Africa to Southwest and Central Asia. It inhabits desert areas east of the Caspian in northern Iran, Turkmenistan, Uzbekistan, and southeastern Kazakhstan as far east as the Syrdarya valley. Kasparov (1992:51, 56) identified it in the bone assemblage from Neolithic Jeitun, but records of the present status of the sand cat population in western Central Asia are not available (Red List NT).

Wild or Yellow Cat (Felis silvestris ornata Gray) The distribution of this subspecies of the African and Eurasian wild cat extends from east and south of the Caspian through Iran, Turkmenistan, Uzbekistan, southern Kazakhstan, and northern Af-



the present environment

ghanistan, east to northwestern China and Mongolia, and south into western India. It inhabits deserts, semi-deserts, woodlands, and riparian forests, and it preys mainly on small mammals, chiefly rodents. It is sometimes trapped for its fur, and its populations are decreasing.

Jungle or Swamp Cat (Felis chaus Schreber) The jungle cat is larger than the wild cat and ranges from Egypt through Southwest and Central Asia into India, Southeast Asia, and southern China. In western Central Asia it inhabits mainly densely vegetated wetlands in river valleys and coastal lowlands close to the Caspian and Aral Seas, but desertification and agricultural development, especially reclamation of wetlands, have caused its populations to decline and it is now largely restricted to the southern Caspian lowland and the valleys of the Atrek, Amudarya, and Syrdarya.

17

Red List VU), also occur across Eurasia, but in western Central Asia their numbers are declining as a result of hunting and habitat loss.

Other Small Mammals Many other small mammals occur in western Central Asia, especially in the Karakum, Kyzylkum, and other arid lowlands. They include the long-eared hedgehog (Hemiechinus auritus Gmelin) and a wide variety of rodents, among which are the desert hare (Lepus tolai Pallas); the long-clawed ground squirrel (Spermophilopsis leptodactylus Lichtenstein); jerboas (e.g., Dipus sagitta Pallas and Jaculus blanfordi Murray); gerbils (e.g., Meriones libycus Lichtenstein, M. meridianus Pallas, and the larger Rhombomys opimus Lichtenstein); voles (e.g., Blanfordimys afghanus Thomas and Microtus transcaspicus Satunin); a porcupine (Hystrix indica Kerr); a hamster (Calomyscus mystax Kashkarov); and a rare dormouse (Myomimus personatus Ognev, Red List 2006 VU, 2008 data deficient).

Mustelideae Many members of this family of small carnivores, which includes badgers, otters, martens, and polecats, are native to western Central Asia. The distribution of the Asian badger (Meles leucurus Hodgson), which is now classified as a species distinct from the European badger (Meles meles Linnaeus), extends from east of the lower Volga river to China, Mongolia, and Korea and includes most of western Central Asia, except for Iran, Turkmenistan, and northern Afghanistan which are within the range of the European badger. Both species remain abundant. The honey badger (Mellivora capensis Schreber) is distributed from Africa through Southwest and Central Asia to the Indian peninsula, but it is now rare across most of its range. It is legally protected in Turkmenistan, Uzbekistan, and Kazakhstan. All three badger species are opportunistic foragers that live in a wide variety of habitats from deserts to dense woodlands. Eurasian otters (Lutra lutra Linnaeus) inhabit river systems from western Europe to eastern Asia, but their numbers are declining in most areas (Red List NT), and in Central Asia they are endangered. They still live in the catchments of the Atrek, Sumbar, Tedzhen, Murghab, Amudarya, and Syrdarya, but hunting for their pelts has greatly reduced the total population, which, in Turkmenistan, probably now numbers fewer than 200. The rock or stone marten (Martes foina Erxleben), which is hunted for its fur, and the marbled polecat (Vormela peregusna Güldenstädt,

Wild Mammals Identified at the Neolithic Site of Jeitun The following were identified in the bone assemblages excavated at Jeitun by Shevchenko (1960), Kasparov (1992), and Dobney and Jaques (this volume, pp. 174–79). The initials S, K, and D/J denote these authors and are preceded by a question mark if the author regarded the identification as uncertain. Urial sheep S, K, D/J; bezoar goat S, K; goitered gazelle S, K, D/J; saiga K; wild boar S, K; wolf S; red fox S, K; corsac fox K, ?D/J; wild (yellow) cat S, K, ?D/J; manul (steppe) cat ?D/J; sand cat K; rock marten K; tolai (desert) hare S, K, D/J; long-eared hedgehog K, D/J; long-clawed ground squirrel ?D/J.

Domestic Animals Sheep and goats are the most numerous and widely raised domestic animals, providing milk, meat, skins, hair, and wool. They are usually managed in mixed herds that combine the propensity of goats to browse woody vegetation and of sheep to graze herbaceous plants. In some parts of western Central Asia, such as the Bolshoi Balkhan, they still sustain systems of transhumance in which people move with their flocks and herds between summer pastures in the mountains and winter pastures in the lowlands. Local breeds are well adapted to aridity and seasonal

18

origins of agriculture in western central asia

extremes of heat and cold, being protected by thick, hairy coats, and, in the sheep, also by the ability to lay down fatty deposits in the rump and the tail. Most domestic sheep are of the fat-rumped type, although the Karakul, which is kept to produce lambs’ skins, is a fat-tailed breed. The goats are mostly of the screwhorned, lop-eared type characteristic of Southwest Asia, but the Central Asian cashmere or pashmina goat is also raised for its fine woolly undercoat or down. Large sheep dogs are employed both to control herd movements and to ward off predators, and some still wear spiked collars for protection against wolves. Smaller domestic dogs are also commonly kept in villages and towns. Horses, mules, and asses are used for a wide variety of agricultural tasks, and there is a long history of breeding slender-limbed racing horses of Arabian type, such as the famous Akhal-Teke breed in Turkmenistan. The only other domestic herd animals that have important roles in the agro-pastoral economy are cattle and both the one-humped and the twohumped camel. Cattle are raised in less arid areas, mainly in cultivated zones in the river valleys and better-watered piedmont areas, where they are used as draft animals as well as for milk and meat production. Both species of camel are highly adapted to life in the deserts as a result of being able to store fat in their humps and go without drinking water for several days. The two-humped or Bactrian camel (Camelus bactrianus Linnaeus) is native to Central Asia and was present in Turkmenistan as a domestic animal in prehistoric times (this volume, pp. 81–83), but from the 2nd century BCE it began to be replaced by a long-haired, cold-resistant breed of the one-humped Arabian camel or dromedary (Camelus dromedarius Linnaeus). Dromedaries, which are valued principally for their milk and their capacity to subsist in very dry, sparsely vegetated areas, now greatly outnumber two-humped camels and play an important part in the pastoral economy of the Karakum desert. In addition to dogs and the herd animals, cats, chicken, ducks, and geese are kept, mainly in the larger settlements.

Other Animals The varied environments of western Central Asia provide habitats for a diverse non-mammalian fauna of birds, amphibians, reptiles, fish, and invertebrates, most of which are not described here because they contribute little or nothing to human subsistence. Several hundred species of resident and migratory birds have been recorded in the region. Avian diversity is higher in the mountains and foothills and along the major river valleys than in the desert lowlands, but only a few species, such as the pheasant (Phasianus colchicus Linnaeus) and various ducks, geese, and other waterfowl, have been regularly hunted. Many species of reptiles also inhabit the region. For example, some 80 species are known in Turkmenistan. They include 50 species of snakes, 27 species of geckos and other lizards, 2 water turtles, and 1 tortoise (Ataev, Rustamov, and Shammakov 1994:330–31; Shcherbak 1994:307). Many are adapted to desert conditions, while others occupy mountain and riverine habitats. Few reptiles are hunted or gathered, but both the steppe tortoise (Agrionemys horsfieldii Gray [syn. Testudo horsfieldii Gray], Red List VU), which hibernates from late May/early June to late March, and the largest lizard, the gray monitor (Varanus griseus Daudin), which occurs mainly in the desert lowlands and foothills and can attain 1.5 m in length, are sometimes exploited for food. Remains of both species have been found at Jeitun (Kasparov 1992:51; this volume, Table 9.18). Historically, fish populations large enough to contribute significantly to human food supplies were restricted to the Caspian and Aral Seas and the river systems of the Syrdarya, Amudarya, Murghab, Tedzhen, Atrek, Gorgan, and their tributaries. However, since the 1950s large-scale irrigation schemes, such as those in Turkmenistan prompted by the construction from 1954 of the Karakum Canal and the filling in 1963 of the Sarykamysh depression to form Lake Sarykamysh, coupled with the introduction and naturalization of many new species of fish from eastern Russia and China, have transformed the composition and ecology of the ichthyofauna (Salnikov 1994).

2

Environmental Changes in the Pleistocene and Holocene

H

aving outlined the present physical environment of western Central Asia, and in order to make sound inferences about prehistoric settlement and subsistence in the region, we need next to consider environmental changes that occurred during the Pleistocene and Holocene, the two epochs of the Quaternary period.1 At present there is little local evidence available of such changes, with the exception of data derived from two main sources: studies of the palaeohydrology of the Caspian and Aral Sea basins, and of palaeosol (buried soil) sequences in the Tajik-Afghan basin. These studies provide the basis for much of the following description of major Pleistocene-Holocene environmental changes in the region. Studies of Tertiary geography, climate, and biota (e.g., Atamuradov 1994:52–61; Kurbanov 1994:124–27) indicate that the climate became progressively more continental and the flora more xerophytic through the Pliocene (the final epoch of the Tertiary). There is fossil evidence that broadleaf trees such as oak and liquidambar were components of mesophytic forests in the Kopetdag mountains, but they became extinct as the trend toward a more continental climate, coupled with tectonic uplift, continued through the Quaternary. This trend caused the vegetation to become differentiated into lowland desert forms and more mesic communities in the mountains and along the main river valleys. By the end of the Pliocene tectonic activity had created the major mountain ranges and lowlands, including the Aral and Caspian Sea basins and the Sarykamysh depression; the ancestral (proto-) Syrdarya, Zeravshan, Amudarya, Uzboi, Murghab, Tedzhen, Atrek, and Gorgan river systems were in existence; and the lowlands had become generally

desertic (Atamuradov 1994:56 –60; Boomer et al. 2000:1260–66). Then, during the Pleistocene and Holocene, the climate, drainage pattern, and plant and animal communities underwent a series of further changes that led to the present-day configuration and ecology of the region.

Pleistocene–Holocene Changes in Hydrology The alternation of glacial and interglacial climates that occurred across the mid and high latitudes of the northern hemisphere during the Pleistocene was manifested in Central Asia in alternating phases of ice accumulation and wastage in the higher mountains, associated respectively with cold-arid and warmer semi-arid conditions in the lowlands. The glacial/ interglacial succession was also linked to changes in river flow and the levels of the Caspian and Aral Seas. During the Early Pleistocene the proto Amudarya flowed west to the Caspian and built a large alluvial plain in the Lowland Karakum, before it shifted northward, later in the Pleistocene, and began flowing via the Sarykamysh depression into the Aral Sea basin, which had previously been largely dry. There is uncertainty about when in the Late Pleistocene this major change of direction occurred (Aladin et al. 1996:34), but it caused the Lowland Karakum, in which the Murghab and Tedzhen rivers built deep series of deltaic deposits, to be gradually transformed by wind action into a sand desert (Atamuradov 1994:62).

20

origins of agriculture in western central asia

In the Pleistocene and Holocene the extent and levels of the Caspian and Aral Seas fluctuated repeatedly. Their palaeogeography has been extensively studied, but the question of how their fluctuations related to the succession of glacial, interglacial, and postglacial periods has been difficult to resolve, partly for lack of radiometric dating of diagnostic deposits. In recent years increased application of radiocarbon, uraniumionium (ionium = thorium-230; Aitken 1990:132), and, to a lesser extent, thermoluminescence dating has enabled such correlations to be more firmly established, although much uncertainty remains. Mamedov (1997) incorporated dating evidence then available in a review of the Late Pleistocene-Holocene history of the Caspian. He recognized a sequence of seven largescale transgressions and regressions and inferred that the final four—the Early Khvalynian, Yenotavian, Late Khvalynian, and Neocaspian—occurred between c. 32,000 years ago and the present, when the level of the Caspian stands at c. 27 m below mean sea level (msl). Mamedov estimated the level and areal extent of each of these four transgression/regression events and correlated them very approximately with the Bryansk interval (interstadial) in the Valdai (Würm) glacial, the Last Glacial Maximum (LGM), the Lateglacial, and the Postglacial (Table 2.1). Archaeologically the sequence extended from the Upper Palaeolithic through the Mesolithic, Neolithic, and subsequent prehistoric and historical periods to the present. Since 1997 more radiometric, geomorphological, sedimentary, and molluskan fossil data relating to changes in Caspian Sea levels have been published (e.g., Chepalyga 2007; Dolukhanov et al. 2009). As a result, Mamedov’s Early–Late Khvalynian periodization has been superseded by a more detailed Late Pleistocene sequence from c. 20,000 to c. 11,500 calibrated years BP (before present)2 which recognizes three progressively less extensive Khvalynian transgressions separated by two brief regressions, the Eltonian and Yenotavian (Table 2.2). The main differences between the two sequences are the recognition of a Middle Khvalynian transgression and, more significantly, the shorter chronology of the new sequence. By confining the sequence to the Lateglacial it contrasts with Mamedov’s correlation (1997:165) of the Early Khvalynian transgression with the Briansk interstadial, the Yenotavian regression with the LGM, and the Late Khvalynian transgression with the Lateglacial. The shorter sequence is itself subject to modification as more dating evidence becomes available

and should not be regarded as definitive, but it provides an appropriate framework for the summary that follows of environmental changes associated with Late Pleistocene–Holocene fluctuations of the Caspian. The Early Khvalynian transgression was preceded in the Late Pleistocene by the Atelian regressive and Khazarian transgressive fluctuations of the Caspian. What is known of their palaeogeography is not discussed here because it is not relevant to the scope of this book, but the changes that occurred from the beginning of the Early Khvalynian, from c. 20.0 ka onward, are relevant and are briefly described. At each of the Khvalynian transgressions the Caspian rose sufficiently to flood parts of the lowland adjacent to its present eastern coast in Turkmenistan, northern Iran, and southwestern Kazakhstan (as well as extensive areas around its northern shores in northwestern Kazakhstan and southern Russia). It reached its highest level, approximately 50 m above present msl, during the Early Khvalynian transgression. At that time it extended east into the lowland between the Bolshoi Balkhan massif and the western Kopetdag, probably as far as the modern town of Kizyl Arvat (Fig. 1.2). During his travels east of the Caspian in 1903, W. M. Davis (1905:42–43) observed former “shoreline-terraces” at Krasnovodsk and “on the mountain flank near Jebel” that may correlate with this extensive transgression. Hecontrasted the evidence of ancient shorelines near Jebel with a lack of similar evidence near Kizyl Arvat and concluded (ibid.) that “no shorelines occur along the base of the Kopet Dagh [there or] farther eastward”—an observation which suggests that Kizyl Arvat marks the approximate eastern limit of the transgression. The Early Khvalynian transgression was followed by the Eltonian regression during which the level of the Caspian fell to c. 50 m below present msl, to be succeeded by a rise to c. 20 m above msl during the Middle Khvalynian transgression, a fall to c. 100 m below msl in the Yenotavian (Enotaevka) regression, and a rise to the Caspian’s present level at the peak of the Late Khvalynian transgression (Chepalyga 2007). The Caspian did not spread east in the Middle and Late Khvalynian as far as it had during the Early Khvalynian, but the sea again invaded the lowland between the Bolshoi Balkhan and the Kopetdag during these later transgressions, having withdrawn during the Eltonian and Yenotavian regressions (Table 2.2). The Khvalynian transgressions, principally the first, are likely to have been caused mainly by large influxes of glacial meltwater into the Caspian. These



environmental changes in the pleistocene and holocene

21

Table 2.1 Late Pleistocene fluctuations of the Caspian Sea according to Mamedov (1997) tentatively correlated by him with the Briansk interstadial–Lateglacial sequence (Velichko and Kurenkova 1990) in uncalibrated radiocarbon years BP; equivalent cal. BP years are indicated by [ka] for dates within calibration range, including the (widely accepted) date of 11.5 ka for the beginning of the Holocene/Postglacial (see Roberts 1998:22–23). Late Pleistocene Transgressions/ Regressions

Approximate Years BP

Late Pleistocene/ Holocene Phases

Approximate Years BP

Early Khvalynian trangression

32,000–24,000

Briansk interstadial

30,000–24,000

Yenotavian regression

24,000–17,000 [20 ka]

Last Glacial Maximum

23,000–16,000 [19 ka]

Late Khvalynian transgression

16,000 [19 ka]–8,000 [9 ka]

Lateglacial

16,000 [19 ka]–12,000 [14 ka]

Neocaspian regression

8,000 [9 ka]–present

Postglacial (Holocene)

11.5 ka–present

Table 2.2 Lateglacial fluctuations of the Caspian Sea (after Chepalyga 2007 and Dolukhanov et al. 2009). Lateglacial Transgressions/ Regressions

Approximate Years BP

Approximate Years cal. BP

Early Khvalynian transgression

17,000–14,500

20.0–17.5 ka

Eltonian regression

14,500–14,000

17.5–17.0 ka

Middle Khvalynian trangression

14,000–13,000

17.0–15.5 ka

Yenotavian (Enotaevka) regression

13,000–12,500

15.5–14.6 ka

Late Khvalynian transgression

12,500–10,500

14.6–12.3 ka

Mangyshlak regression

10,500–10,000

12.3–11.5 ka

occurred when the northerly flow of Siberian rivers, which had been blocked by the continental polar ice sheet of the Valdai glaciation, formed huge proglacial lakes from which water spilled southward into the Caspian, partly via the Volga drainage and partly by way of the Turgai channel into the Aral Sea and thence via the Uzboi channel (Fig. 1.2) into the Caspian (Grosswald 1980:15–19; Mangerud et al. 2004:1,322–28). Two clusters of Mesolithic rockshelter sites close to the southeastern coast of the Caspian, one in western Turkmenistan and the other in northern Iran, were occupied, probably intermittently, during the Lateglacial (this volume, pp. 55–58). Their stratified deposits reveal changes in the abundance of terrestrial and marine animals that reflect shifts in the availability of food resources arising from the Lateglacial (and early Postglacial) changes in position of the Caspian coast and the estuaries and deltas of the larger rivers such

as the Uzboi, Atrek, and Gorgan. The beginning of the Postglacial is signaled in the history of the Caspian by a rapid fall in level at the end of the Late Khvalynian transgression. This phase, known as the Mangyshlak regression (Table 2.2), represents the earliest part of the prolonged Neocaspian regression that continued through the Holocene, when the Caspian became more stable but continued to fluctuate below the present level of c. 27 m below msl. The number and dates of individual fluctations are not well established, but—on the basis of a palaeogeographical analysis of Neocaspian terraces and shorelines, and age estimates from radiocarbon dating of marine shells—Rychagov (1997:167) recognized a series of oscillations in the Holocene between c. 20 and c. 32 m below msl. He identified six transgressions separated by regressive episodes of varying length. Although the number, duration, and causes of

22

origins of agriculture in western central asia

these oscillations remain uncertain, there appears to be quite a close correlation between the regressive episodes recognized by Rychagov and several short-term phases of drier and colder climate that occurred in the Holocene, such as the 8.2 and 4.2 ka events (discussed in the next section of this chapter). The Pleistocene–Holocene evolution of the Syrdarya and its connections with the Aral Sea have been less intensively studied than changes in the Amudarya-Caspian-Aral hydrological system. During the Pleistocene the Syrdarya shifted northeast across the Kyzylkum from the southeastern part of the desert, where it deposited thick accumulations of sand and clay as it emerged from its upper course in the western Tien Shan; created a large delta on the eastern side of the Aral Sea that connected in its early stages with the Amudarya’s Akchadarya delta; and only reached its modern position in the Late Holocene (Boomer et al. 2000:1266–67). Through the Pleistocene and Holocene the level and extent of the Aral Sea was largely determined by changes in the discharge of the Syrdarya and Amudarya, and the intermittent connections of the latter with the Caspian. It is difficult to correlate closely the hydrology of the Aral Sea basin with the fluctuations of the Caspian, but Boomer et al. (2000) infer that the basin probably remained dry through much of the Pleistocene, when evaporite deposits formed in it; and that during the Khvalynian transgression of the Caspian (not subdivided by Boomer at al. 2000:1,276) several of the Amudarya’s major channels reached the basin, where they formed the Sarykamysh, Horezma (Khoresmia), and Akchadarya deltas southwest and south of the Aral Sea. Boomer et al. also summarize (ibid., p. 1,265), in a series of sketch maps, their reconstruction of changes in the palaeogeography of the Aral Sea through the Holocene from c. 10.0 ka to the present. A different reconstruction of part of the sequence of changes in the Aral Sea basin has been proposed by the leaders of an Uzbek-Polish archaeological-palaeoenvironmental research team working in the Kyzylkum desert (Szymczak and Khudzhanazarov 2006a). Prompted by the striking absence of Upper Palaeolithic and Mesolithic settlements in the Turanian Lowland (the area approximately occupied today by the Kyzylkum and Karakum deserts), they have suggested that the apparent lack of sites may be due to the lowland having been flooded for several millennia from the beginning of the Holocene until c. 5.0 ka. They infer

that this transgression, which they name the “Io Sea” and attribute to the spilling of proglacial lakes through the Turgai channel (referred to above), reached a maximum altitude of c. 210 m above msl at about 7.5 ka (c. 5500 cal. BCE). This corresponds to the altitude of many Neolithic sites which, they suggest, were located at the margins of the Io Sea, and they point out that the two sites in the Kyzylkum that have been reliably radiocarbon dated—Ayakagytma and Uchashchi (Fig. 6.2)—were occupied at that time. They observe that the Neolithic settlements at the southern margins of the Kyzylkum (Keltiminar sites) and Karakum (JeitunCulture sites) cluster close to the 200 m contour, and they infer that at least some of them, including Jeitun itself, would have been at or near the coast of the Io Sea (see this volume, p. 191, for a comment on this idea in relation to Jeitun). This speculative hypothesis deserves consideration, but it is not supported by Rychagov’s (1997) reconstruction of the Holocene oscillations of the Caspian Sea in that the proposed timing of the Io flood precedes rather than coincides with the maximum level of the second transgression dated by Rychagov to 7.0 ka. Also, Boomer et al. affirm (2000:1,275) that “during much of the Neolithic the Aral Sea was probably at about 72–73 m a.s.l.” However, the inferred timing of the flood does place it in the mid Holocene, early in the Climatic Optimum (see below), during the wet “Lavlakansky phase” (or “Lyavlyakan pluvial”) (Boomer et al. 2000:1,267) when increased flow of the Syrdarya and Amudarya rivers would have discharged much larger amounts of water into the Turanian lowland. It is to be hoped that future field research and radiometric dating will help to resolve these contradictory reconstructions, as the Io Sea hypothesis merits further investigation.

Pleistocene–Holocene Changes in Climate and Vegetation In addition to its impact on the hydrology of western Central Asia, the Late Pleistocene–Holocene glacial/interglacial succession profoundly affected the climate and vegetation of the region. Evidence for this comes from windblown deposits of loess that mantle extensive areas in the northern ranges of the Central Asian mountains, especially in the Tajik-Afghan basin in southwestern Tajikistan between the Pamirs and the Hindu Kush. There, at mid altitudes (2000–3000



environmental changes in the pleistocene and holocene

m), numerous palaeosols have been found buried in loess deposits over 200 m deep. Some of the palaeosols contain Lower Palaeolithic stone tools; and pollen analysis of the deposits has demonstrated alternating periods of xerophytic herbaceous (steppe) and mesophytic woody (broadleaf forest) vegetation representing, respectively, cold/dry and warmer, more humid conditions (Bronger et al. 1995; Dodonov et al. 2006; Dodonov and Baiguzina 1995; Forster and Heller 1994; Vishnyatsky 1999:87–90). Ding et al. (2002:388) infer that the sands of the Karakum and Kyzylkum were the main source of the Tajik loess deposits. They regard the rate of loess deposition as a good proxy indicator of aridity and desertification, and cite data from the site of Chashmanigar in southern Tajikistan which show that the rate increased during glacial periods through the Pleistocene as aridity increased, vegetation cover diminished, the deserts expanded, and dust storms became more frequent (Ding et al. 2002:394–95). Such conditions would have been inimical to human occupation, and it is not surprising that such evidence as there is of human presence, in the form of chipped stone tools, is found in the interglacial and interstadial palaeosols. Dating the palaeosol sequences has proved difficult. For example, thermoluminescence dating failed to produce reliable absolute chronologies for sequences at Urkutsay in Uzbekistan (Zhou, Dodonov, and Shackleton 1995) and Remisowka in Kazakhstan (Machalett et al. 2006), but palaeomagnetic dating has been more successful. In one of the sequences, at Kuldara in southern Tajikistan, chipped stone tools were excavated from palaeosols dated palaeomagnetically to approximately 900,000 years ago in the Lower Pleistocene (Dennell 2009:326; Ranov and Dodonov 2003; Vishnyatsky 1999:89). Stone tools have also been found in Middle Pleistocene pedocomplexes in the same area at Darai Kalon, Karatau, Lakhuti, and several other sites (Davis and Ranov 1999; Dennell 2009:326–29; Dodonov et al. 2006; Ranov 2001). They were first dated, by correlation with the marine oxygen-isotope and north Chinese loess sequences, to between c. 500,000 and c. 600,000 years ago, thus demonstrating that Middle Pleistocene populations were present in western Central Asia in a semi-arid interglacial environment at that time (Davis and Ranov 1999:187; Shackleton et al. 1995). More recently, Ding et al. (2002) correlated the loess-palaeosol sequence at Chashmanigar closely with the Chinese loess deposits and the deep-sea oxygen-isotope record and inferred a

23

basal age of 1.77 million years (Dennell 2009:217–19). These results confirmed that throughout the Pleistocene the loess horizons and the palaeosols correspond respectively to cold, dry, windy glacials and warmer, more humid and calmer interglacial periods. Although uncertainty remains about the detailed chronology of much of the Tajik palaeoenvironmental evidence, it is a uniquely valuable record of oscillations between glacial/stadial and interglacial/interstadial climates and vegetation and provides what is probably a reliable proxy for changing environmental conditions in western Central Asia as a whole. It also provides a framework for understanding the episodic nature of human occupation during the Pleistocene (see below). Despite the many climatic oscillations of the Pleistocene, xerophytic grass and shrub communities progressively, if episodically, increased in the lowlands, plateaus, and foothills at the expense of mesophytic woodlands and forests. For example, on the piedmont of southern Turkmenistan dry steppe communities dominated by grasses of the genera Stipa, Festuca, and Poa expanded as, in the desert lowlands, did sagebrush communities dominated by species of Artemisia; and riparian tugai forests dominated by tamarisks and poplars became established in lowland valleys (Kurbanov 1994:127).

The Last Glacial Maximum The trend in the Pleistocene toward increased aridity reached a climax during the LGM, between c. 23.0 ka and c. 19.0 ka. Models of worldwide climatic changes suggest that sea-surface temperatures in the North Atlantic and North Pacific were 5–10°C lower than at present and temperatures on land were up to 20°C lower (COHMAP Members 1988:1,045; Kutzbach, Behling, and Selin 1993). Although very little palaeoenvironmental research relating directly to the LGM has been undertaken in western Central Asia, there is no doubt that the climate was exceptionally cold and dry, with ice sheets over the higher mountains and a broad zone of periglacial steppe and tundra stretching north of the Aral and Caspian Seas to the continental ice sheets far to the north (Velichko and Kurenkova 1990: Plate 1). In the deserts and plateaus east of the Caspian extreme cold and aridity accentuated the trend toward more xerophytic vegetation by reducing the stands of steppe grasses and favoring the further spread

24

origins of agriculture in western central asia

of sagebrush, a process that would have lowered the capacity of the lowlands to support large herbivores such as gazelle, onager, and saiga. These changes would have made it increasingly difficult for huntergatherers to sustain life, except where they had access to perennial water sources and the plant and animal resources around them. Very few Upper Palaeolithic sites are known in western Central Asia, in contrast to numerous Middle Palaeolithic and Mesolithic ones (Movius 1953; Ranov and Davis 1979:257–58; Vishnyatsky 1999:112); and Davis (1990:272–73; also in Davis and Ranov 1999:191–92) has suggested that Central Asian populations were substantially reduced and may have disappeared altogether in some areas during the LGM. It is quite possible that the lowlands were too dry to sustain long-term human occupation and that hunter-gatherer groups were largely or wholly restricted to uplands and to riparian habitats along river channels and in the deltaic margins of the Caspian and Aral Seas (where traces of former occupation would now be hard to detect).

The Lateglacial Following the extreme aridity and low temperatures of the LGM, a fluctuating trend toward warmer and wetter conditions set in across high- and midlatitude Eurasia at the beginning of the Lateglacial, c. 19.0 ka. Forest, woodland, and steppe vegetation began to re-expand during the Lateglacial Bølling interstadial, but this process was punctuated by reversals to colder and drier conditions during the Older and Younger Dryas stadials (separated by the warm/ wet Allerød interstadial). The alternation of cold/dry and warm/wetter conditions, with associated changes in vegetation, affected Upper Palaeolithic settlement and subsistence across northern Eurasia, including western Central Asia (Dolukhanov 1997:182–84). The Younger Dryas stadial, between c. 13.0 ka and c. 11.5 ka, was the most abrupt and pronounced of these Lateglacial episodes. Its occurrence across the northern hemisphere is well established and it was probably a global phenomenon (Peteet 1993; Roberts 1998:70–76). It disrupted the Lateglacial warming trend, temporarily halted the process of vegetation recolonization, and in Southwest Asia and possibly China appears to be implicated in the beginnings of cereal cultivation (Bar-Yosef and Belfer-Cohen 2002; Harris 2003; Hillman 2000:375–99; Hillman et al. 2001). Palynological

or other local palaeoenvironmental evidence of the Younger Dryas in western Central Asia is lacking at present, although the rapid fall in the level of the Caspian at the end of the Late Khvalynian transgression and during the Mangyshlak regression (Table 2.2) may be, in part at least, a result of its impact. There is little doubt that the Younger Dryas had severe ecological effects and (like the LGM but less drastically) probably reduced or even eliminated Upper Palaeolithic huntergatherer populations in the most arid areas, including much of the Karakum and Kyzylkum deserts.

The Postglacial The end of the Younger Dryas marks the beginning of the Holocene and the rapid resumption of the fluctuating trend toward warmer and wetter conditions. It used to be thought that this trend was only interrupted by minor climatic shifts identified as the succession of cool/dry (Pre-Boreal), warm/ dry (Boreal), warm/wet (Atlantic), warm/dry (SubBoreal), and cool/wet (Sub-Atlantic) stages originally proposed for northern Europe by Blytt and Sernander early in the 20th century on the basis of peat stratigraphy. But more recent research elsewhere in the world has revealed greater climatic variability in the Holocene (Lamb et al. 1995; Roberts 1998:117–19). This re-evaluation of Holocene climate has been based on varied sources of palaeoenvironmental data, but it is the Greenland ice cores that have yielded the most detailed chronology of northern hemisphere climatic variability through the Late Pleistocene and Holocene. They have not only provided evidence of major changes in the Lateglacial, notably the Younger Dryas stadial, but have also revealed a series of abrupt, short (century-scale) oscillations in the Holocene between cool/dry and warm/wet phases. The first and largest of these, with about half the amplitude of the Younger Dryas, occurred at c. 8.2 ka (Alley et al. 1997; Alley & Ágústsdóttir 2005; Kobashi et al. 2007). This 8.2 ka event is recorded in the Greenland ice cores and many other northern-hemisphere palaeoclimatic records, including data from tree rings, pollen, diatoms, foraminifera, and speleothems. These records show that the change to cooler/drier conditions began very abruptly, within a decade, and that warmer/wetter conditions returned within a century or two. Despite the brevity of the 8.2 ka event, pollen records indicate that it caused shifts in vegetation toward more cold-



environmental changes in the pleistocene and holocene

and drought-tolerant plant communities. No local records of it are available from western Central Asia, but it is recorded in pollen cores and deposits of windblown silts at sites farther east in the Central Asian arid zone in northwestern China (X. Q. Liu et al. 2002; C. L. Liu et al. 2003). The possible significance for the beginnings of agriculture in western Central Asia of this sudden change to colder and drier conditions toward the end of the early Holocene is considered in Chapter 12. There is also evidence that other global phases of drier (and cooler?) climate occurred later in the Holocene, at approximately 7.0–6.4 ka in the mid Holocene, and 4.8–4.4, 3.5–2.9, and 2.5–2.2 ka in the late Holocene (Wünnemann et al. 2007:22). It was pointed out in the previous section of this chapter that some of the short-term regressions of the Caspian Sea during the Holocene correlate quite closely with these dry climatic phases, but what other environmental (and cultural) effects they may have had elsewhere in western Central Asia remains to be established. Although very little evidence of climatic variability in the Holocene is available from sites in western central Asia, palaeoclimatic records derived from such sources as loess and dune deposits, lacustrine and fluvial sediments, and shoreline and river terraces are available from farther east in the Central Asian arid zone. Evidence from lake sediments is particularly valuable because it tends to have better temporal resolution and such sediments are less subject to discontinuities than aeolian and fluvial deposits. In a comparative analysis of data from nine lakes in eastern arid Central Asia (together with records from the Aral Sea and Lake Van in eastern Turkey), Chen et al. (2008) tracked changes in effective moisture through the Holocene. They inferred that in the early Holocene, c. 11.0–8.0 ka, the climate was drier than today; that in the mid Holocene, between c. 8.0 ka and c. 5.0 ka, moisture reached a maximum, with lake levels at their highest and vegetation at its most dense; and that in the late Holocene, from c. 5.0 ka to the present, effective moisture decreased, although (despite brief dry phases) the climate was wetter than in the early Holocene. Chen et al. (2008) also compared Holocene moisture trends in Central Asia with lake-sediment, peat-bog, and cave-speleothem records from nine sites in monsoonal Asia and found that there, in contrast to arid Central Asia, maximum moisture occurred during the early to mid Holocene. They attributed the difference to the dominant effects respectively of the

25

Indian and East Asian summer monsoons south and east of Central Asia, and of the mid-latitude westerlies coupled with the orographic (“rain-shadow”) effect of the Tibetan plateau. They also emphasized that the Holocene climate of arid Central Asia was strongly influenced, via the westerlies, by changes in sea-surface and air temperatures around the North Atlantic (evidenced by marine, ice-core, and pollen records) that reached maxima during the mid Holocene and were the main cause of the increased moisture and higher temperatures experienced at that time in Central Asia. It had long been recognized that warmer and wetter conditions than prevail today had occurred in mid-latitude Eurasia in the mid Holocene. The phenomenon is often referred to as the Climatic Optimum (or Altithermal) and it has been widely regarded as the temporal equivalent of the Atlantic stage of the Blytt-Sernander sequence, but estimates of its duration and of how widely it occurred across Eurasia have varied. Chen et al.’s (2008) comparative analysis confirms and clarifies the occurrence of the Climatic Optimum in arid Central Asia and is broadly consistent with other less comprehensive studies of Holocene climatic change in the region (e.g., Feng, An, and Wang 2006 and Herzschuh et al. 2004). Apart from the evidence (discussed above) of postglacial changes in the hydrology of the Caspian and Aral Seas, there are no lake-level or other long records from sites in western Central Asia with which to build a local chronology of environmental changes through the Holocene. But there are extensive traces of former lakes and now-dry watercourses in the region, mainly in the Kyzylkum and Karakum. In the southern Kyzylkum in central Uzbekistan, Vinogradov and Mamedov (1975) and Vinogradov (1981) demonstrated a close correlation between clusters of Mesolithic and Neolithic sites, around the ancient Lyavlyakan lake, where palaeosols indicative of denser vegetation have been found (Dolukhanov 1994:206), and also along former channels of the Zeravshan river (Fig. 6.2). In the Karakum, former drainage networks are particularly evident along the eastern margins of the desert close to the present channel of the Amudarya, and in the north between the Sarykamysh depression and the Caspian where the channel of the former Uzboi river is the dominant feature. Geomorphological and geoarchaeological investigations in southern Turkmenistan, which combined satellite imagery, aerial photographs, and ground survey, have revealed traces of a palaeochannel of the Amudarya trending

26

origins of agriculture in western central asia

northwest from its present valley just south of Chardzhou, as well as evidence of a large former lake (Marcolongo and Mozzi 1998:3). The channel and the lake may have been part of the drainage system, already referred to, that connected the Amudarya to the Caspian via the Sarykamysh depression in the Pleistocene, or it may represent a later change in the drainage network that occurred in the Holocene. Uncertainty about the age of former lakes and extensive now-dry fluvial networks in the deserts demonstrates the need for well-dated evidence from stratified Quaternary deposits with which to disentangle the complex hydrological history of the desert lowlands, but the probability remains that much of the former drainage pattern constitutes evidence of the more humid conditions of the Climatic Optimum. It is very difficult to determine how far precipitation and air temperature increased in western Central Asia during the Climatic Optimum, but according to Dolukhanov (1994:203), reporting estimates by Zabakov (1986), annual precipitation was 100–150 mm higher, and mean annual temperature about 1°C and winter temperature about 1.2°C higher than present values. Higher temperatures and rainfall in winter and spring would have increased the length of the growing season and favored the development of Neolithic crop cultivation and pastoralism. The milder, more humid conditions would also have led to expansion of shrub, woodland, and forest vegetation in the deserts, along valleys, and in the mountains. The Postglacial warming trend that culminated in the raised temperatures and precipitation of the Climatic Optimum began to give way in the late Holocene, from c. 5.0 ka, to cooler conditions with short-term oscillations between wetter and drier phases. Attempts have been made to relate such short-term climatic and hydrological changes to changing settlement patterns, for example by Tolstov and Kes (1960), who tried to correlate natural and artificial alterations in the Amudarya drainage system to population movements from Neolithic to medieval times. But particularly from the Bronze Age onward when large-scale irrigation systems were developed—and often proved difficult to maintain—such correlations tend to be problematic. Nevertheless, it is probable that short-term phases of greater aridity, such as those that are thought to have occurred at 4.8–4.4, 3.5–2.9, and c. 2.5–2.2 ka (see above), did cause changes in river flow and sedimentation that led to the breakdown of irrigation works and caused some population displacements.

In addition to such short-term climatic events, people were active agents of environmental change through the late Holocene. Even in the mid Holocene, the establishment of agro-pastoralism on the Kopetdag piedmont in the Neolithic, and the subsequent development in the Bronze Age of urban settlements and irrigation agriculture associated with continuing pastoralism, had had some impact on soils, stream flow, vegetation, and animal life. This process intensified with the increases of population, land clearance, transport, trade, and warfare associated with the ancient and medieval Achaemenian, Parthian, and Islamic civilizations. They exerted cumulative environmental pressures through the grazing and browsing of domestic livestock, the collection of wood for fuel, the felling of timber for the construction of buildings and irrigation works, and the modification of soils, sediments, and drainage systems for agricultural, industrial, and military purposes. Many of these processes intensified further in more recent times, especially during the 20th century when the environment became subject to new pressures, such as the exploitation of natural gas and mineral resources and the development of large-scale irrigation schemes following the construction of the Karakum Canal. As a result, vegetation and animal life has been extensively degraded, resulting in major losses of biodiversity, and soil erosion and salinization have rendered large tracts of land uncultivable (Kharin 1994). The description of Pleistocene/Holocene environmental changes in western Central Asia given in this chapter provides a broad temporal framework for much of the more detailed archaeological-environmental evidence presented in the rest of the book. Much remains to be learned about the extent and intensity of such changes and about past human responses to them, but the evidence available now does allow some tentative exploration, in later chapters, of possible correlations between environmental change and the beginnings of agriculture and sedentary settlement in the Neolithic.

notes 1. The Quaternary period is conventionally defined as the past 1.8 million years, but many specialists regard 2.6 million years ago as a more appropriate geochronological definition of the Tertiary/Quaternary boundary because it encompasses the time during which the Earth’s climate has been influenced by periodic glaciations (Pillans and Naish 2004). 2. In the rest of this chapter, calibrated radiocarbon dates before present are indicated by the abbreviation ka, thus 1 ka=1,000 cal. BP or AD 1000, and 3 ka=1000 BCE.

3

The Local Environment of Jeitun with Susan Limbrey

T

he Jeitun mound lies at the junction of the Kopetdag piedmont and the Karakum desert in an area of small sand hills just north of a major dune ridge which, in this area, forms the southern edge of the desert. Looking south from the site, the steep mountain front of the Kopetdag range is clearly visible and to the east the desert extends past the terminal fans of the Tedzhen and Murghab rivers to the valley of the Amudarya, some 600 km away. By the 1990s, extensive cultivation of grapevines, irrigated with water brought from the Amudarya by the Karakum Canal, had encroached on Jeitun, and there was a wide drainage ditch about 100 m east of the site (Fig. 3.1). In this chapter, the present climate, soils, vegetation, and native terrestrial fauna in the vicinity of Jeitun are described to place the site in its landscape, but changes in the topography, drainage, and ecology of the area, especially the recent development of large-scale irrigation agriculture, severely limit the extent to which the Neolithic environment of Jeitun can be inferred from present-day conditions. However, there is some local evidence (see below) of different environmental conditions in the Neolithic, such as the existence of buried-soil horizons close to the site and charcoal and animal bones excavated on site, that testify respectively to the former presence of riparian trees no longer represented in the local vegetation and of wild animals now very rare or extinct in the area.

Climate No climatic record is available for Jeitun itself, but data for Ashgabat, at 227 m altitude on the middle

piedmont 28 km south-southeast of Jeitun, provide a guide to precipitation and temperature at the site, although allowance must be made for the lower altitude (c. 120 m) of Jeitun and its position at the southern edge of the Karakum. Ashgabat receives mean annual precipitation of 230 mm, but at Jeitun the mean falls to c. 200 mm, most of which occurs between October and May, with a maximum in March and April. At Ashgabat the January mean air temperature is just above freezing, with extremes down to -26 o C, and the July mean is c. 31o C, with extremes up to 47o C. Winter and summer temperatures at Jeitun probably exceed somewhat these means and extremes. Potential evaporation at Ashgabat reaches a total of 1,629 mm (unpublished data from the Desert Research Institute, Ashgabat, 1990) and is at a maximum from June to August. Precipitation exceeds potential evapotranspiration only in January, and by March–April soil moisture falls under rising temperature, even though this is the period of maximum rainfall. Rainfed wheat is said to have been grown in recent times in localized areas of the Kopetdag, but rainfall decreases sharply away from the mountain front (Fig. 1.7), and at Jeitun cereal crops would probably have been grown on areas of high groundwater, or with the aid of irrigation (for further discussion of conditions for crop growth at Jeitun, see this volume, p. 136 and pp. 163–64).

Topography The topography of Jeitun’s local environment is formed of alluvial and aeolian deposits of silts, clays, and sands. The distributary fan of the Tedzhen river

28

origins of agriculture in western central asia

3.1 Sketch map showing the relationship of Jeitun to the present-day channel of the Kara Su and to the nearby modern irrigation drainage ditch.

extends westward to beyond the location of Jeitun, and its deposits, which overlie the Amudarya sediments that form the substrate of the Lowland Karakum, interdigitate with those of the rivers that flow across the piedmont from the nearby mountain front of the Kopetdag. Flooding by the Kopetdag streams in

May and early June was observed by the geographer W. M. Davis, who traveled in the region in 1903. He described “the variable floods that sweep suddenly, unguided by channels, down the piedmont slope; now eroding, now depositing; here sweeping along coarse blocks, there depositing fine silts”; and he also observed the Tedzhen in flood, having “overflowed its channel and spread a thin sheet for miles over the [piedmont] plain” (Davis 1905:44, 54). The Tedzhen and the piedmont streams have been utilized for irrigation since prehistoric times, with the result that the natural supply of sediment has been interrupted, and deep cultural deposits have been formed in the irrigated areas. For example, at the Chalcolithic (Eneolithic)–Bronze Age site of Anau on the middle piedmont east of Ashgabat, where Pumpelly’s expedition of 1904 dug a series of deep shafts through the stratigraphy of the north and south mounds and the surrounding alluvial deposits of the piedmont, it was shown that “22 feet or more of irrigation silts” had accumulated since the earliest occupation (Pumpelly 1908:34). Water flow from the piedmont has also been exploited by vertical wells and by gently sloping tunnels (qanats or karez) since prehistoric times, so the natural extent and volume of water and deposits reaching the desert margin is difficult to estimate. Also, ground water stored in the deep gravels of the upper piedmont contributes to stream flow, with the result that seasonal variation would have been somewhat smoothed. Today the belt of cultivated land irrigated from the Karakum Canal occupies a much more extensive area of the lower piedmont than did oases watered by streams, qanats, and wells, and this has further interrupted the natural expression of alluvial morphology. The alluvial surface of the lower piedmont has a very low slope and is formed of fine sand, silt, and clay, the silt content being increased by desert loess. It is fed by the rivers that drain the Kopetdag, cross the piedmont, and emerge into the desert, where the water is lost by infiltration and evaporation. The boundary between the lower piedmont and the desert has little topographic expression, and windblown sand overlies alluvium in the form of small sand hills and major dune systems, through which the rivers formerly maintained their courses. The watercourse close to Jeitun, the Kara Su, is one of these rivers. It penetrated the southernmost dune ridge and petered out in the alluvial fan which provided the water and the soils exploited by Jeitun’s Neolithic farmers. Recently it



the local environment of jeitun

was impounded in a series of lakes that probably utilized one of the takyrs that extend in a belt along the lower piedmont, and from which it is now provided with a piped overflow into its original course through the dune ridge. The gorge (Fig. 3.1) occupied today by this artificial version of the Kara Su exposes a series of dune-sand and alluvial deposits, with a major buried-soil horizon that emerges from beneath sands in the lower part of the gorge, about 1 km south of Jeitun (Fig. 3.2, color). The buried soil is a truncated solonetz (Fig. 3.3, color) that shows development, including clay translocation and rubification. Where the stream emerges from the gorge it forms a terrace, within which a lower terrace of stratified sands, silts, and clays is itself dissected by the stream. The deposits of the lower terrace are similar to those of the recent alluvial fan and represent an earlier phase of fan construction. Less than 100 m west of Jeitun, and extending about a kilometer to the north, a former takyr has been used to form a series of shallow lakes, the precise location and size of which varies. People grazing camels, donkeys, goats, and sheep make small hoe dams to impound and direct the floods, so that the position and extent of the water varies according to the people’s needs and to the volume of the outlet feeding the Kara Su.

Soils and Sediments The area around Jeitun is dominated by clayey, silty, and sandy soils in the form of saline takyrs, solonchak soils, sandy alluvial sediments, and dune sands.

Takyrs Takyrs are areas of saline hydromorphic soils on level ground or very low slopes (Figs. 1.5, 1.6). Because they are important sources of cultivable land and water for livestock, they attracted the attention of Russian soil scientists from the 1920s onward. Lobova (1960 [1967]) summarized how understanding of their formation developed, the nature of their pedological regime, and their variants. They form where clay-rich materials at shallow depth impede water penetration, water evaporates, and saline conditions develop, resulting in distinctive forms of takyr solo-

29

netz soils that may develop further into solonchaks. The takyr surface layer is a thin porous saline crust, and at shallow depth the clay-rich horizon has a characteristic platy structure and cracks into polygonal blocks in the summer. Gypsum is a significant component of the salt content, with a maximum in the crust and a second maximum at the depth reached by sub-crustal moistening. Takyrs support few or no higher plants, surface growth being limited to algae and lichens that form a mat which cracks, rolls up, and is blown about as it dries. Takyrs gather water directly from rainfall and from seasonal stream flow, or by percolation through windblown sand that may overlie surrounding areas of the fine alluvial layers. Takyrs are exploited by digging pits and wells to act as reservoirs for watering livestock and for irrigation of crops. Rain is often concentrated into periods of two or three days during which run-off to these reservoirs can be substantial. Takyrs with particularly high impedence to infiltration, especially those in the piedmont zone where both precipitation and runoff are relatively high, have provided very valuable areas of cultivable land; for example, Lobova (1960 [1967:274]) calculated that in the piedmont zone one square kilometer can collect 9,000 to 25,000 l of water per annum. Under irrigation, takyrs are exploited by plowing up the layer of low permeability and digging ditches to provide through-drainage and remove salts, with addition of sand also being beneficial. Tension between pastoralists’ use of takyrs for their livestock and policies for the extension of arable cultivation, evident today, must have a long history in southern Turkmenistan. In the Lowland Karakum, takyrs are formed on the clay-rich components of alluvial deposits of both Pleistocene and Holocene age. It is thought (Lobova 1960 [1967:246]) that where stream flow contributes to takyr wetting on the fringes of the distributary fans, as it does at Jeitun, humus content as well as clay is increased by fine sediment derived from the soils of the stream catchment (in this case the catchment of the Kara Su in the Kopetdag). The name Kara Su, meaning “black water,” may reflect a high humus content in its water. A thin layer of fine humus-rich material can be the primary cause of takyr development, rather than more substantial clay-rich layers. This suggests that in the evolution of a distributary fan, takyrs of small extent can form rapidly and may be of limited persistence if buried or disrupted as the fan develops.

30

origins of agriculture in western central asia

Sandy Soils and Sediments The materials of the dunes and sandy alluvial sediments around Jeitun are dominated by micaceous fine sands. Derived from areas of granite, limestones, sandstones, and gneiss, the light fraction is rich in unweathered feldspars, the micas are dominated by biotite rather than muscovite, and the high heavy mineral content is predominantly horneblende. Calcite of pedological origin and as detrital grains of limestone is common, and gypsum, of local pedological origin as well as in the form of detrital grains transported by the wind from soils elsewhere, also occurs. The dune ridges and smaller dunes are partially stabilized by the root systems of woody shrubs and the soil development they promote. Tap roots and woody lateral roots many meters in length are exposed by sand-blow throughout the areas between the shrubs. The shrubs trap sand and finer dust that, together with their own organic detritus, improves water retention. They also trap snow, increasing moistening from snow-melt around them. The shrubs therefore occupy mounds of soil of finer texture and higher humus content than the spaces between them, and thus provide suitable conditions for herbaceous vegetation, contributing further to stability. On the lower slopes of the mounds and between them, where grasses, sedges, and ephemerals are more sparsely distributed, the sand is more mobile and the surface is constantly changing. The mounds also offer shelter and nesting sites for burrowing animals that range in size from insects and insect larvae through small lizards and rodents to monitor lizards, foxes, and jackals. Bioturbation and the introduction of organic materials into burrows affects the sands to considerable depth, and in the case of the larger animals brings about the eventual collapse of the stabilized area. Stability of the sands is also highly sensitive to grazing. The approaches to pastoralists’ settlements, such as the one at Artykhodja 1.5 km north of Jeitun, whose livestock are grazed around the site, are completely denuded of vegetation, and there the sands are mobile. At Jeitun, stability is sufficient for the deep-rooted woody shrubs to persist, but soil development is limited to a primitive aerosol without a thick sod horizon. Lobova (1960 [1967:334]) refers to evidence of luvic processes in sandy desert soils that result in films of hydrated iron oxides forming on sand grains. At Jeitun, such films were observed in the lower buried soils (this volume, Section 9.3), but not in the upper buried soil or the blown sand within

the site deposits, suggesting that the “ferruginization” process described by Lobova has not been active in the soils since Neolithic times. In addition to the sand hills and the former takyr or lake bed, the terrain close to Jeitun includes areas of varying extent in which blown sand lies only thinly over the alluvial sands and clays. In some of these areas small takyrs are developed, while in others solonchak soils carry thickets of tamarisk and other shrubs.

Soil/Sediment Profiles Profiles were recorded in the vicinity of Jeitun by hand auguring and digging pits (see Appendix 3.1). The variability of the alluvial sediment and blownsand layers in the upper 1–2 m exemplified the nature of fan sedimentation. Wherever takyr-like characteristics prevailed, whether in the larger area of the herders’ lakes or in small hollows in the honeycomb pattern of sand hills, a thin layer of clay or sandy clay occurred immediately below a surface puffed by salt efflorescence. Below that, sand, or interstratified sand and clay or sandy clay, extended down until groundwater limited further access. In two of the pits, fine gray sand with very high mica content lay beneath the initial takyr clay, whereas this sand was encountered at greater depth elsewhere. The proximity of the irrigation drainage ditch and the now ponded and managed waters of the Kara Su have created artificial groundwater levels that provide no guidance to the levels that affected soils, and were accessible to vegetation, in Neolithic times. Lisitsina (1965:25; this volume, Appendix 3.1) reported a humus-stained horizon at 15–35 cm depth beneath the takyr surface and suggested that it might be the soil cultivated during Jeitun’s occupation. No equivalent horizon at that depth was encountered during our investigations, where local variability was in any case evident; but the darker upper part of the sand immediately below the takyr clay in profile 1, with its inclusion of charcoal flecks, may be comparable. There is no evidence, however, that the evolution of the alluvial complex at the time the site was occupied produced a takyr in the same location as today. Buried-soil horizons with potsherds and charcoal in them were investigated in the side of the irrigation drainage ditch east of the site (this volume, pp. 131–41), and they provide secure evidence both of contemporaneity with the site’s occupation and cul-



the local environment of jeitun

tivation by Jeitun’s farmers. Levelling data show that these soils lie 1–2 m higher than the takyr surface.

Vegetation Today the vegetation in the vicinity of Jeitun consists mainly of xerophytic communities of shrubs, semi-shrubs, and herbs, the structure and floristic diversity of which have been severely degraded by browsing and grazing livestock and by people cutting wood for fuel. In the past, and especially during the Neolithic period when, during the mid-Holocene Climatic Optimum, the climate was warmer and less arid than it is today (this volume, pp. 25–26), the local vegetation cover would have been more continuous and floristic diversity greater. However, in the absence of palynological evidence of long-term vegetation change (this volume, pp. 171–73), we cannot determine with any precision how the Neolithic plant communities differed from their present-day equivalents. Nevertheless, we can be confident that the main types of vegetation now present in the area would (with the exception of the irrigated crops of modern agriculture) have been present in the past. Also, by examining the vegetation that exists today in protected areas such as the reserves at Repetek in the Southeast Karakaum and on the Badghyz plateau east of the Tedzhen river, we can gain some impression of the greater structural and floristic complexity that would have characterized the vegetation of southern Turkmenistan during the Neolithic. As was emphasized at the beginning of this chapter, Jeitun is located where the gentle slope of the lower piedmont merges into the southern edge of the Karakum. It is thus situated on the ecotone between two major ecological zones: the desert, and the piedmont and mountain front. Desert vegetation surrounds the site itself, and would have done so in the Neolithic, but the plant (and animal) communities of the piedmont and of the nearer valleys and uplands of the Kopetdag would also have been accessible to Jeitun’s Neolithic inhabitants. In this section, both the desert vegetation close to Jeitun and the more distant plant communities of the piedmont and Kopetdag are described, the former in relation to the topography of the local area and the latter more briefly, focusing on plants that may have been exploited by the Neolithic population.

31

The Vegetation Close to Jeitun The vegetation growing around Jeitun on sand hills, dunes, and alluvial surfaces covered with fine sands is dominated by xerophytic shrubs and smaller semi-shrubs with a scattering of grasses, sedges, and other herbs on open patches between the shrubs (Fig. 3.4, color). The tallest stands consist mainly of species of Ammodendron, Calligonum, Ephedra, Haloxylon, and Salsola, although not all these genera are always present. For example, both white saksaul (Haloxylon persicum), which typically grows on sand dunes, and black saksaul (H. aphyllum), which grows mainly on alluvial soils and is more salt-tolerant and deep rooting, now occur infrequently in the vicinity of Jeitun. Both species are cut for firewood and provide fodder for domestic livestock; white saksaul is browsed by sheep, goats, and camels, whereas only camels browse black saksaul because they can tolerate the bitterness of its leaves. The wood of black saksaul is very dense and makes good charcoal (unlike white saksaul which burns to ash) and it was extensively cut for this purpose in the past. In areas where it is protected today, as in the Repetek Sand Desert Reserve, black saksaul forms tall stands (Fig. 1.14, color), and it is likely that it did so near Jeitun in Neolithic times. Haloxylon was not identified in our archaeological samples of charcoal from the site (this volume, p. 167), but its absence may be due to charcoal burning having customarily taken place close to black saksaul trees because their wood is so dense and heavy. This practice has been documented at Repetek where layers of charcoal have often been encountered beneath the surface when soil pits are dug near the trees (pers. comm., Dr. Suhane, Director of the Repetek Reserve, April 1994, who also reported that, before they were over-exploited for fuel, mature stands of black saksaul were much more extensive in the Karakum and on the Kopetdag piedmont). On the dune ridge south of Jeitun, and on other large dunes in the area, a less floristically diverse psammophyte shrub community occurs. It is dominated by Calligonum spp. (which often replace over-exploited stands of white saksaul), Ammodendron conollyi, Ephedra strobilacea, and Salsola richteri which provides valuable browse for sheep and goats. Semi-shrub vegetation consisting principally of species of sagebrush (Artemisia), with other low-growing shrubs such as Alhagi sp., Ephedra distachya, Astragalus turcomanicus, and Halothamnus subaphyllus, is more extensive in the Jeitun area than are stands of the taller shrubs, and pe-

32

origins of agriculture in western central asia

rennial and annual herbs grow between and beneath clumps of sagebrush (Fig. 3.4, color). The perennials include grasses (Stipagrostis spp.), sedges (Carex spp.) and species of Iris, Tulipa, Ferula, and Allium. Some species, such as Sophora alopecuroides, are too toxic to be browsed by sheep or camels. Around Jeitun the semi-shrub community as a whole shows signs of heavy overgrazing, accentuated by wind erosion following disintegration of the soil surface. As a result, soil has been removed around many shrubs, leaving them isolated on soil mounds or pedestals which function as micro habitats that sustain herbaceous plants and provide shelter for a variety of animals. Between the mounds there are deflated sandy areas where the root systems of shrubs are exposed and where perennial and annual herbs grow sparsely. As well as the shrub and semi-shrub communities that dominate the vegetation in the vicinity of Jeitun, salt-tolerant (halophytic) taxa form distinctive communities on the more saline alluvial surfaces: on areas of solonchak soils, in areas where the groundwater level is high, such as around the lakes just west of Jeitun that occupy former takyrs, and along the now artificially fed stream channel of the Kara Su. Today its channel is intermittently bordered by stands of tamarisk (Tamarix spp.) and reed (Phragmites australis) (Fig. 3.5, color), and these plants also grow around the lakes and swamps near Jeitun (Fig. 3.6, color) that are fed by overflow from the extensive area of grapevine cultivation south of the site. In prehistoric times more complex riparian vegetation of tugai type that also included poplar (Populus), willow (Salix), and probably alder (Alnus) may have grown along the channel of the Kara Su and in local marshy areas liable to flooding. This inference is supported by the presence of wood charcoal of these trees in our samples from Jeitun, in addition to much more abundant tamarisk charcoal (a discrepancy that may be partly due to differential preservation of the denser wood of tamarisk; this volume, p. 169). The local tugai vegetation would have been an important source of fuel and timber for the inhabitants of Jeitun, and continued exploitation of the larger trees in later prehistoric and historical times is likely to have progressively reduced their abundance. It is also possible that the frequently changing topography of the Kara Su’s terminal fan, subjected to sometimes violent seasonal floods, reduced the chances of poplars and willows re-establishing themselves after periods of over-exploitation because they withstand flooding less well than

tamarisks. Among the sand hills and dunes near Jeitun, there are small areas of saline soil that support such halophytes as Salsola gemmescens, S. orientalis, Anabasis salsa, the sedge Carex pachystylis, and the stem succulent Halocnemum strobilaceum, whereas the takyr surfaces tend only to be colonized by algae and lichens.

Plant Communities of the Near Piedmont and Kopetdag Much of the piedmont, and several of the gorges through the front range of the Kopetdag eroded by the headwaters of piedmont rivers such as the Kara Su, are within 40 km of Jeitun and would have been physically accessible to the site’s occupants (as well as to people living at other early Neolithic sites on the piedmont). We cannot determine from the archaeological evidence whether Jeitun’s inhabitants routinely gathered and hunted wild plants and animals on the piedmont and in the near intermontane valleys of the Kopetdag, or, if so, at what season(s) of the year they did, but it is probable that some wild food resources were regularly procured there by people who spent at least part of the year living at Jeitun (see Chapter 11 for discussion of early Neolithic patterns of settlement and subsistence). The diverse plant communities of the piedmont and Kopetdag would have offered a wide variety of foods that could supplement the products of crop cultivation and livestock herding at Jeitun, and many of the plants that may have been exploited in the early Neolithic are noted in this section. The main types of forest, woodland, shrub, and steppe vegetation of the piedmont and Kopetdag have already been described in Chapter 1. Although today forests of broadleaf trees are restricted to the western and central Kopetdag, where stands of maple, ash, walnut, elm, and other broadleaf trees persist in some deep north-facing valleys, these mesophytic communities would have been more extensive in the Neolithic. The canopy-forming larger trees are associated with a diverse flora of smaller trees, shrubs, and climbers, many of which produce edible fruits. Those that are likely to have been gathered, casually or more systematically, in summer and autumn include wild apple, pear, plum, cherry, fig, hawthorn, blackberry, and grape. Some of the herbaceous plants may also have been valued as sources of food, for example, wild garlic (Allium sativum) and, for their edible roots, Leontice ewersmannii and Bongardia chrysogonum.



the local environment of jeitun

On the lower and drier slopes of the mountains and on the piedmont, xerophytic woodlands dominated by such drought-tolerant trees as wild almond (Amygdalus communis) and pistachio (Pistacia vera), both of which yield edible nuts, were formerly extensive, although they are now restricted to remote areas least vulnerable to over-exploitation. For example, stands of wild almond now only survive in the southwestern Kopetdag (where the nuts were until recently systematically harvested), and pistachio woodlands survive mainly on the Badghyz plateau, where they form open-canopy groves with wild barley and other grasses such as Poa bulbosa in what Popov (1994:175) has described as pistachio “semisavanna” (Figs. 1.9, 1.10, color). Before it was extensively destroyed by overgrazing and logging in recent times, pistachio woodland occupied large areas of the central and eastern piedmont, and the nuts were regularly harvested, as they are likely to have been in the early Neolithic. Shiblyak constitutes a third type of vegetation that is a source of wild plant foods. It is a floristically diverse formation of small trees and shrubs that is most highly developed in intermontane valleys of the southwestern Kopetdag, such as the Sumbar and Chandyr, which experience a summer-dry Mediterranean climate. Shiblyak vegetation contains many species with edible nuts and fruits, notably wild almond, cherry, quince, medlar, fig, pomegranate, and jujube (Zizyphus jujuba). Steppe consisting of perennial grasses of the genera Stipa, Festuca, and Elytrigia together with species of many other herbaceous genera used to be extensive on the more level surfaces of the piedmont, and it is possible (although we have no direct archaeobotanical evidence of this) that seeds of some of the grasses may have been harvested for food in the remote past before cereal cultivation began in the Neolithic. Today most of the former areas of steppe are under cultivation or degraded by persistent grazing, but some undisturbed areas still exist in remote parts of the upper piedmont (Fig. 1.11, color). The main importance of the steppe to Neolithic populations was probably as a habitat of wild herbivores that were hunted, particularly gazelle (see below). It is evident that the main types of vegetation characteristic of those parts of the piedmont and Kopetdag within relatively easy reach of Jeitun contained a wide variety of wild plant foods, principally fruits but also edible roots, tubers, and leaves, that could have been gathered by its inhabitants, despite the fact that no traces of any of these foods were detected in the

33

archaeobotanical samples from Jeitun (see Chapter 9, Sections 9.5–9.8). This negative evidence does not exclude the possibility that such plant foods were gathered casually, perhaps during hunting trips, or more systematically in summer and autumn when fruits and nuts matured, but it does suggest that they were not regularly transported to, and consumed or stored at, Jeitun. To test the hypothesis that wild plants made a more than casual contribution to Neolithic diet in this part of southern Turkmenistan would require further excavation at other sites on the piedmont, using modern archaeobotanical techniques, but in the absence of such new data we should not dismiss the possibility that wild plant products did supplement the staple foods of Neolithic diet derived from cereal cultivation, livestock raising, and hunting.

Native Terrestrial Animals The distribution and present status of terrestrial animals native to western Central Asia has been described in Chapter 1, with emphasis on the wild mammals that have been exploited for food and other purposes and on the extent to which the populations of most of the larger mammals have been reduced in recent centuries by hunting and by destruction of their habitats. This has resulted in the extinction or near-extinction in Turkmenistan of many of the carnivorous species that would have been present in the Neolithic (e.g., Persian leopard, Caspian tiger, Asiatic lion, cheetah, striped hyaena, brown bear, gray wolf, Eurasian lynx, sand lynx or caracal, steppe cat or manul, jungle or swamp cat, rock marten, marbled polecat). Some of these predators would have been hunted for their pelts, but the main food prey of Neolithic (and Mesolithic) hunters, and of most of the larger animal predators, were the native herbivores: bezoar goat, urial sheep, goitered gazelle, onager, saiga, and Bokhara deer or kangul. All these herd animals exist now only in small numbers in remote, mainly mountainous areas, and some, such as the Bokhara deer that inhabited riparian forest (tugai), are either extinct or very close to extinction. Wild pigs survive today, mainly in tugai and other wooded habitats, but were more widespread in the past. Other mammals, such as red fox, corsac fox, desert or tolai hare, and many small rodents maintain viable populations in Turkmenistan, as do many species of reptiles, principally snakes and lizards and also the steppe tortoise.

34

origins of agriculture in western central asia

It is clear that in prehistoric times the terrestrial fauna of southern Turkmenistan was much more diverse, and the populations of individual species much larger, than they are today. The inhabitants of Jeitun were well placed to exploit for food, and for furs and pelts, many different animals living in the desert environment near the site, on the piedmont, and in the near valleys and mountains of the Kopetdag. The shrub-dominated plant communities around Jeitun would have provided habitats for many medium-size and small mammals (red fox, corsac fox, sand lynx, desert hare, porcupine, long-eared hedgehog, long-clawed ground squirrel, jerboa, dormouse, hamster, gerbils, and voles), as well as many reptiles (steppe tortoise and numerous snakes and lizards, including the large gray monitor lizard). The desert near Jeitun would also have been visited by herds of onager, gazelle, and probably saiga, with their attendant predators. The gazelles migrated north in the winter from their summer grazing grounds in the foothills of the Kopetdag to graze in the Karakum, and the saiga moved south into the desert in autumn and winter from their spring and summer pastures in the Russian steppe. The Neolithic occupants of Jeitun and other Neolithic sites on the piedmont were well positioned to intercept migrating herds of gazelle and saiga—the latter at least when it migrated that far south in hard winters (this volume, p. 14). However, although these species are represented in the animalbone assemblages from Jeitun (this volume, p. 176; Kasparov 1992:51), there is insufficient evidence from the site, or from other excavated Neolithic piedmont sites, to estimate how intensively and extensively they may have been hunted. Tugai vegetation in the channels of piedmont streams, and more fully developed along the Sumbar and Chandyr rivers and in other intermontane valleys in the Kopetdag, would have been the principal habitat of wild pigs and Bokhara deer and their predators, particularly the Caspian tiger, but there is little zooarchaeological evidence to suggest that people regularly hunted boar or deer in the Neolithic; nor, at Jeitun at least, is there any evidence that fish or waterfowl were important sources of food (this volume, p. 179; Kasparov 1992).

The forests, woodlands, shiblyak, and steppe of the Kopetdag and piedmont sustained a greater variety and larger populations of herbivores and carnivores than the areas of tugai and desert vegetation. Bezoar goats are likely to have been restricted to the mountains, whereas in the Neolithic urial sheep probably grazed on the piedmont steppe as well as in the mountains, and both would have been preyed upon by the “big cats” and also by hyaenas, lynx, wolves, foxes, and steppe cats. The piedmont and the mountains and valleys of the Kopetdag were also home to brown bears and small carnivores such as polecats, rock martens, badgers, and otters. Bezoar and urial bones have been reported from Jeitun (Kasparov 1992:61–63) but none were identified in the samples from our excavations. The extent to which wild goats and sheep may have been hunted in the Neolithic remains unknown, but their remains have been reported from both Mesolithic and Chalcolithic sites in Turkmenistan and neighboring regions (this volume, pp. 84 and 87), which suggests that they continued to be hunted, probably on a diminishing scale, through the Neolithic. This illustrates the difficulty of assessing at all precisely the role of hunting in the predominantly agro-pastoral economy of Jeitun and other Neolithic piedmont sites, but it is clear, even from the limited zooarchaeological evidence from Jeitun, that quite a wide range of wild animals was exploited for food, furs, and pelts and that they were procured not only from desert areas near the site but also from the piedmont and Kopetdag. The wild animals identified in the bone assemblages excavated at Jeitun include bezoar goat, urial sheep, goitered gazelle, saiga, wild boar, wolf, red fox, corsac fox, manul (steppe) cat, wild (yellow) cat, sand cat, rock marten, tolai (desert) hare, long-eared hedgehog, long-clawed ground squirrel, steppe tortoise, and (gray monitor?) lizard. They include most of the herbivores that are likely to have been hunted (with the interesting exception of the onager which may have eluded hunters more easily than other slower animals), and the list suggests that wild animals made a significant and varied contribution to the diet and to other aspects of life at Neolithic Jeitun.

4

The Local Environment of the Bolshoi Balkhan Sites with Jen Heathcote

C

lose to the eastern shore of the Caspian Sea in western Turkmenistan, the Bolshoi Balkhan massif rises to an altitude of just over 1,880 m. It is an isolated outlier of the Kopetdag mountain system and consists predominantly of a core of Jurassic limestones and an outer belt of Cretaceous limestones, surrounded by a piedmont zone of varying width formed by the coalescence of alluvial fans. Its southern flank consists of a steep escarpment penetrated by canyons and ravines cut by intermittently flowing streams that have built, and dissected, a series of fans at the foot of the escarpment, thus creating the piedmont zone (Fig. 4.1 and Fig. 4.2, color). The fans have created a semi-continuous sedimentary apron that slopes down to the narrow floodplain of the former Uzboi river, which here forms a corridor typically less than 5 km wide that bisects the lowland between the Bolshoi and Maly Balkhan massifs. The lowland varies in width from c. 25 km to c. 30 km and much of it lies close to sea level. Saline areas of solonchak soils, developed on superficial deposits of Quaternary age, occur in the lowland and along the eastern flank of the Bolshoi Balkhan massif (Fig. 4.1). Farther southwest, toward the present shore of the Caspian, the lowland becomes wider where it encompasses the former delta of the Uzboi and there are extensive areas of sand ridges, mobile dunes, and solonchaks. In this chapter we give a brief account of the landforms, sediments, climate, vegetation, and large mammals of the Bolshoi Balkhan massif and its

southern environs, based on our field observations and on published sources. The rockshelters investigated by the British team in 1997—Jebel and the Dam Dam Cheshme (DDC) sites (this volume, pp. 113–16)—are located at the southwestern extremity of the massif near the town of Nebit Dag (Balkhanabat), east of the present shore of the Caspian (Fig. 4–1). Jebel, which is the type site for the Caspian Mesolithic in Turkmenistan, is a shallow rockshelter at c. 300 m above msl (Fig. 8.15, color). Some 9 m above it there is evidence, in the form of a wave-cut platform, of a marine terrace indicative of a previous high stand of the Caspian Sea during the Pleistocene. We could find no evidence of raised beaches or deltaic deposits higher on the mountain side, and because it is at such a high altitude, the wave-cut platform may record a very ancient (Aspheronian?) transgression of the Caspian, although it may have been created later and tectonically uplifted since its formation. The DDC rockshelters (Figs. 8.16–8.19, color, and Fig. 8.22, color) are located approximately 20 km southeast of Jebel, and at about the same altitude (c. 300 m above msl), at the base of the southern escarpment of the massif, where they overlook the piedmont and the channel of the Uzboi (Fig. 4.3, color). We did not observe any local evidence of Caspian transgressions close to the DDC sites, but at times of higher sea level they would have been in closer proximity to the lower channel and delta of the proto-Uzboi and/or to the sea itself.

36

origins of agriculture in western central asia

km

4.1 The Bolshoi and Maly Balkhan massifs and the Uzboi lowland. The locations of the Dam Dam Cheshme sites and of Oyukli, Adjikuli, and Bashkovdan are based on GPS readings taken at them, whereas the positions of Jebel, Charla’uk, and Joyruk are approximations based on field notes in the absence of GPS data.

Landforms and Sediments During our 1997 field season in the area it was only possible to undertake preliminary geomorphological observations. We decided to focus attention on the piedmont slope at the foot of the southern escarpment and

on the lower Uzboi channel where it runs parallel to the escarpment (Fig. 4.4), as these zones would have been directly accessible to the inhabitants of the rockshelters. Methods employed in the field included geomorphological mapping, limited hand-augering of sub-surface sediments, and description of stratigraphic sections.



the local environment of the bolshoi balkhan sites

37

km

4.4 The southern Bolshoi Balkhan, piedmont zone, and lower Uzboi channel; points A and B mark the extent of the transect walked along the channel.

The Southern Piedmont Slope A large quarry close to the approach road to the DDC rockshelters (Fig. 4.4) provided an opportunity to examine sections through an area of alluvial-fan deposits. They comprised sequences of unconsolidated sands and gravels that in places were exposed at the surface and elsewhere covered with a thin (< 0.50 m) veneer of poorly sorted sediment. Exposures within the quarry showed bedded sequences of sands and gravels up to 4 m thick (Fig. 4.5), although they may well be thicker because this was simply the depth of quarrying represented by the height of unquarried sections of sediment. At the base of these sections there was no evidence of the nature of the underlying geology. The gravel units consist predominantly of pebble- to boulder-sized stones in a sand matrix. They alternate with coarse to medium, moderately sorted sands which only occasionally show fining upward of particle-size distributions within the units. A 12-mlong exposure in the quarry provided a longitudinaloblique section through coarse alluvial-fan deposits, the depth of which suggests the previous existence of a sizeable channel. The extent of the quarry (c. 1.5 km long) attests to substantial accretion of coarse-fan ma-

terial that is indicative of relatively high-energy fluvial conditions and may represent deposits laid down by melt water during the final deglaciation of the region, which began c. 19,000 cal. BP. Unfortunately, deposits suitable for radiometric dating were not identified in any of the exposures examined.

The Lower Uzboi Channel Despite the fact that topographic maps of the area often show the Uzboi as a river, today there is no continuous watercourse in the valley, and although ephemeral flow can occur in the channel, it is evidently sporadic and likely to be seasonal. In order to trace and study the character of the relict Uzboi channel we examined it along an 80-km transect walked parallel to the southern escarpment of the massif from approximately 10 km east of its easternmost limit nearly to Nebit Dag. (Fig. 4.4, points A to B). The channel is incised locally to a depth of up to 10 m, is strongly sinuous, and contains well-established vegetation which suggests that flooding is infrequent (Fig. 4.6, color). Although maps often show the course of the Uzboi as

38

origins of agriculture in western central asia

4.5 Natural exposure of bedded sands and gravels in the quarry on the approach road to the Dam Dam Cheshme rockshelters where alluvial-fan deposits were examined, March 1997. (Photo by Jen Heathcote).

a single, strongly meandering channel, on the ground the morphology of the channel is far more complex. For example, a significant degree of lateral channel migration is indicated downstream from location C–D (Fig. 4.4), where the channel frequently divides, and there is evidence of cut-off meander loops across the whole valley floor. This basic survey was undertaken to identify the nature and morphology of any channel features, the number of terraces present, and the nature of the terrace deposits. Another aim was to identify locations and deposits that showed potential for further work, to establish their inherent sedimentary properties, and/ or to determine their palaeoenvironmental associations. Although several such deposits were identified, it was not possible to undertake analytical work on either the environmental associations or the relative or absolute chronology of the deposits. Nevertheless, the data acquired allow some basic characteristics of the

fluvial regime and relative timings to be suggested. At the eastern limit of the survey (A) the Uzboi channel has vertical sides. It is constrained by lithified deposits of probable Tertiary age, and its floor lies at c. 10 m below the present desert surface. The basal 2.5 m of these lithified deposits are well bedded and frequently show localized contortion. However, this morphology exists only for a limited portion of the easternmost part of the valley, and in the rest of the transect the channel is typified by vertical and lateral accretion of fine-grained, unconsolidated sediments indicative of a low-gradient river channel. These accretionary events may have taken place during one of the regressive phases of the Caspian, perhaps when its level fell rapidly at the end of the Late Khvalynian transgression (this volume, p. 21). Traveling southwest from point A, two terraces were identified (first observed at location E–F, Fig. 4.4): an upper terrace c. 6 m above the present valley floor and a lower one at c. 2.5 m. They show clear differentiation in terms of their composite deposits. The upper terrace consists of a pale mica-rich clayey coarse silt, the color of which varies slightly from bluegray to pale gray. This unit lacks clear sedimentary structure, although some localized horizontal weak iron staining may be associated with subtle textural changes and/or weak bedding. It contains pockets of abundant remains of small vertebrates (mammals and amphibians rather than fish). The deposits of the lower terrace are very different. They comprise brown salty clays that lack a significant mica content. This may indicate that they are derived from sediments which have undergone a significant amount of aeolian reworking prior to being deposited by fluvial activity. They are predominantly massive but display bedding toward the top of the unit, which shows a sharp (erosion) contact with the overlying deposits of the upper terrace. No animal or plant remains were identified in the deposits of the upper terrace, the full depth of which is unknown as no basal contact was identified along the whole of the walked transect.

Gravel Mounds In addition to the present channel of the Uzboi, a further type of (probably alluvial) geomorphological feature was identified closer to the piedmont zone. To the west of the GAI road check-point, mounds of gravel occur that may be relict coarse channel or bar



the local environment of the bolshoi balkhan sites

deposits of an early proto-Uzboi river system (Fig. 4.4). The mounds are c. 500 m long and c. 200 m wide, have flat tops, and are surrounded by flat desert. A section through the edge of one of them showed poorly sorted gravels with occasional fine, discontinuous bedding toward the top. The origin of the mounds is problematic as they are too distant from the piedmont to be explained as coarse alluvial fans generated from the Bolshoi Balkhan massif. Although they might be explained as such, produced from a previous cold climatic episode dating to the Early or Middle Pleistocene, considerable erosion would need to have taken place to remove the sedimentary apron that would have existed between them and the present piedmont along the southern escarpment. It is difficult to envisage an erosional mechanism that could remove such a large amount of sediment from such a wide zone while leaving the gravel mounds upstanding. An alternative hypothesis, which seems more probable, is that they represent remnants of a terrace of a large, possibly braided, Lateglacial proto-Uzboi fluvial system associated with one of the Khvalynian transgressions (Table 2.2).

Climate and Vegetation Climate As a result of its northwestern location in Turkmenistan, and its isolation from the Kopetdag mountains, the Bolshoi Balkhan massif experiences slightly lower average annual precipitation (c. 130 mm) and temperature (c. 15° C) than areas of equivalent altitude in the Kopetdag. Nevertheless, evaporation exceeds precipitation through most of the year, and the moderating inf luence of the Caspian, which causes summer humidity to be locally higher along its eastern shore than it is inland, has little effect in the Bolshoi Balkhan. Aridity is the prevailing characteristic of the climate, although the higher slopes and summits receive somewhat more precipitation, especially in the form of winter snowfall, than the rest of the massif.

to the surrounding piedmont, surfaces consist mainly of sands, gravels, and boulders, with no or very little soil development. Vegetation is generally sparse and woodland is today confined to the higher altitudes, where it is dominated by stands of the drought- and cold-tolerant Turkmen juniper (Juniperus turcomanica). Junipers occur above about 800 m on the southern side and above about 400 m on the northern side of the massif, although isolated trees survive on steep slopes and ledges lower down (Fig. 4.7), and old tree stumps indicate that juniper woodland used to extend to the base of the mountains (Popov 1994:177). According to Popov (ibid., p. 178), as recently as the 1930s juniper woodlands covered 40,000 ha in the massif, but they have since decreased drastically as a result of unrestricted wood cutting and grazing, despite the trees’ ability to regenerate from stumps if they are not cut too frequently. Associated with the juniper woodlands, a distinctive type of xerophytic vegetation occurs which consists of low shrubs known as cushion plants that grow extremely slowly and can withstand both drought and frost. The cushion plants of the Bolshoi Balkhan include Onobrychis cornuta and Tragantha marshalli, as well as four other species, three of which are endemic to the massif (Popov 1994:181). At lower altitudes in the mountains, several types of small trees and shrubs grow sporadically on rocky slopes and along stream channels in canyons, often as single plants or forming small degraded patches of shiblyak vegetation. They consist mainly of wild cherry (Cerasus microcarpa), associated with species of Berberis, Calligonum, Cotoneaster, Ephedra, Euphorbia, Rhamnus, and Zygophyllum. Fig trees (Ficus carica) also grow here

Vegetation in the Mountains The mountain landscape is dominated by steep, bare rock faces, and even on more gentle slopes and along the seasonally dry watercourses that lead down

39

4.7 Scattered juniper trees surviving on steep slopes and ledges above the Dam Dam Cheshme 4 rockshelter, March 1997.

40

origins of agriculture in western central asia

and there, especially at the heads of canyons where water seeps out of the limestone and is tapped by the figs’ deeply penetrating roots. These sites are often under rock overhangs that provide shelter for shepherds (Fig. 4.8, color), and the dispersal of fig trees to such sites may well have been assisted not only by birds but also, inadvertently or deliberately, by pastoralists. Although the lower slopes of the mountains are today largely bare of vegetation, they would have been more heavily vegetated in the past, with a greater diversity of taxa represented, when browsing by livestock and the exploitation of trees and shrubs for fuel and construction was less intensive.

Vegetation of the Piedmont and the Channel and Former Delta of the Uzboi At the base of the massif, where the rocky terrain of the mountains gives way to the gravelly slopes of the piedmont, scattered trees and shrubs are replaced by a sparse cover of low-growing desert shrubs, semishrubs, grasses, and forbs. The shrubs include several species of Artemisia, Calligonum, and Salsola, and the herbaceous component includes perennial species of Stipagrostis, Carex, and Iris as well as many annuals. Floristic diversity is low as a result of aridity, thin soil cover, and seasonal grazing on the piedmont by sheep, goats, and camels. In the channel of the Uzboi, at the lower edge of the piedmont, the vegetation is somewhat more varied. The most obvious contrast with the piedmont is the presence of individual trees and clumps of tamarisk (Tamarix sp.), which is well adapted to growth on the saline channel deposits (Fig. 4.6, color). Scattered halophytic semi-shrubs, principally Salsola gemmascens, occupy much of the channel floor, and the sand sedge Carex physodes and the grass Stipagrostis pennata also occur, sometimes associated with bushes of white saksaul (Haloxylon persicum) (Fig. 1.12, color). The deeper rooting black saksaul (H. aphyllum) (Figs. 1.13, 1.14, color), which grows mainly on alluvial soils and is well adapted to areas of high water table and temporary standing water, is also present. It may have been more abundant when the climate was less arid during the

mid-Holocene Climatic Optimum (this volume, p. 25). At that time, when the Uzboi flowed through its delta into the Caspian, riparian tugai forests were probably established along its channels. That tugai forests wereformerly more extensive can also be inferred from the fact that even under present-day aridity stands of poplar, tamarisk, and reed (referred to by Walter and Box [1983:115] as “remains of an earlier floodplain forest”) grow around the edges of marshes and lakes fed by rainfall in the Bolshoi Balkhan massif that drains above and below ground toward the Uzboi (ibid., p. 113). Today, in the area of the former delta of the Uzboi southwest of the massif, there are extensive sand accumulations and saline depressions with solonchak soils and very sparse vegetation. Tamarisk and white saksaul are established on some of the dunes, and the leafless stem succulent Halocnemum strobilaceum is a distinctive feature of halophytic vegetation on solonchaks.

Large Mammals The Bolshoi Balkhan massif formerly provided habitats for a wide variety of wild mammals, but their populations have been drastically reduced in recent centuries. This is particularly true of the larger herbivores and carnivores. Both the urial sheep and the bezoar goat still exist in the mountains in very small numbers, but their predators—principally (apart from humans) leopards but also lynx, striped hyaena, red fox, wild cat, and manul or steppe cat (Korshunov 1994:243)—are now either locally extinct or rare (Rustamov and Sopyev 1994:207–13). Wolf populations have been much reduced by pastoralists seeking to protect their livestock from predation, but jackals and wild pigs are still relatively common in the area, where they hunt and forage in the mountains, on the piedmont, and along the channel of the Uzboi. Populations of the larger mammals would have been greater in the past, especially during the Climatic Optimum, when tugai and juniper forests were more extensive along the Uzboi channel and in the mountains where they would have been hunted by the Mesolithic occupants of the Jebel and Dam Dam Cheshme rockshelters.

part ii

Prehistoric Archaeology

5

History of Archaeological Research with Jennifer Coolidge

T

he prehistory of western Central Asia remained almost entirely unknown until after the Russian conquest of most of the region in the second half of the 19th century. Archaeological research soon followed, and by the 1880s investigations of prehistoric sites were underway. During the Soviet era systematic research on the prehistoric archaeology of Soviet Central Asia began, and, by means of a series of field campaigns of survey and excavation, a spatial framework and a chronological sequence for the prehistoric past of Turkmenistan and the other Central Asian republics were established. Great advances in knowledge were made, particularly after the Second World War, when the investigation of tell (depe) sites on the Kopetdag piedmont revealed the former existence there of numerous Neolithic agro-pastoral villages and larger, more urban settlements of the Chalcolithic (Eneolithic in Russian terminology) period and the Bronze Age. Most Western archaeologists remained unaware of these new discoveries, and the way in which research results were interpreted and reported was constrained by Marxist ideology, as is very evident when the voluminous Russian archaeological literature of the period is studied. In this chapter the history of research on prehistoric archaeology in Turkmenistan and adjacent areas in Uzbekistan, northeastern Iran, and northern Afghanistan is summarized, as a prelude to describing in Chapter 6 the principal Mesolithic and Neolithic sites, sequences, and subsistence economies. Our summaries (which draw on Chapters 2 and 3 in Coolidge 2005) are based on a comprehensive reading of the relevant Russian literature published in the former Soviet Union, as well as of the more limited body of

work published by archaeologists from elsewhere who have worked in the region. Most of the detailed Russian publications have not been synthesized by Soviet scholars, and very few Western archaeologists have summarized the Russian literature at all comprehensively, with the notable exceptions of P. L. Kohl (1981, 1984), F. T. Hiebert (1994a, 2003), and F. Brunet (1999, 2002, 2003, 2005). There are some basic overviews in Russian, but these tend to omit much of the significant detail in the original research publications. This situation appears to be due, in part at least, to rivalry and lack of collaboration during the Soviet era between the Institute of Archaeology of the Academy of Sciences of the USSR in Moscow and the branch of the Institute in Leningrad, as well as the strong although unofficial direction exerted by Moscow and Leningrad over the institutes of archaeology of the Academies of Science in the Central Asian republics. A secondary reason for the lack of Soviet attempts to synthesize archaeological data (although this situation has changed following the dissolution of the Soviet Union) is the formulaic system of conducting archaeological research and interpreting archaeological data that was prescribed by Soviet ideology. It required investigators to fit an archaeological site into a three-tier hierarchy of local variant, archaeological sub-culture, and ultimately a designated group of archaeological cultures, and it encouraged comparison of like with like within the hierarchy across geographical regions while discouraging comparison of like with unlike. This favored the perception and study of archaeological sites and prehistoric cultures as separate entities and militated against attempts at regional synthesis.

44

origins of agriculture in western central asia

Turkmenistan and Uzbekistan Pre-Soviet Investigations Archaeological investigations in the present territory of Turkmenistan began during the 1880s. The initial stage of activity was marked by the plundering of sites for the international art market as well as the use of ad hoc or haphazard techniques of excavation. More systematic investigations began when General A. V. Komarov, the Imperial Russian governor of Transcaspia, organized exploratory excavations at several sites. They included the great urban center of ancient Merv in the Murghab delta northeast of the present-day city of Mary, which occupied a pivotal position on the Silk Route between China and the West. Then in 1886, he turned his attention to the large mounds (referred to as kurgans) at Anau on the Kopetdag piedmont east of Ashgabat (Fig. 6.1), where he bisected the north mound with a massive trench, expecting it to contain a royal burial, possibly that of Alexander the Great. Although no such burial was found, he realized that he had unearthed evidence of Stone and Bronze Age occupation, and his trench was the earliest large-scale excavation of prehistoric deposits undertaken in Turkmenistan (Hiebert 2003:24–25; Kohl 1984:17). In the first decade of the20th century, the earliest multidisciplinary scientific investigation into Turkmenistan’s prehistory took place when in 1904 the American geologist Raphael Pumpelly, assisted by the German archaeologist Hubert Schmidt, carried out excavations at the Anau mounds. They excavated at both the north (Chalcolithic/Eneolithic) and south (Bronze and Iron Age) mounds, as well as at the historical site of Anau city, and thereby established the first prehistoric cultural sequence for Turkmenistan (Kohl 1984:18). Pumpelly and Schmidt approached the task of excavation and the retrieval of finds systematically, and many of the finds were then sent to specialists for detailed analysis: the animal bones to J. U. Duerst in Bern, pottery and mudbricks with imprints of plant remains to H. C. Schellenberg, also in Bern, the human bones to G. Sergi in Rome and T. Mollison in Zurich, and the metal objects to F. A. Gooch in New Haven, Connecticut, for chemical analysis (Hiebert 2003:25–28). Pumpelly published the results of his investigations in two thoroughly documented volumes to which the scientific specialists also contributed (Pumpelly

1908). As a geologist, he had already traveled extensively in Central Asia, including a reconnaissance expedition in 1903 when he chose Anau for detailed investigation (Pumpelly 1905), and he had been impressed by the widespread evidence, in the form of dry river channels and lake basins, that the climate had formerly been less arid. In his studies of the physiography of deserts and oases he was assisted by the geographer Ellsworth Huntington, who was later to become a champion of environmental determinism (Huntington 1915, 1919), and Pumpelly became convinced that climatic changes had profoundly affected patterns of human settlement and migration in Central Asia. In particular, he suggested that progressive desiccation had stimulated cereal cultivation and the beginnings of (irrigation) agriculture in Central Asian oases (Pumpelly 1906:668; 1908:65–66). In proposing this hypothesis he foreshadowed Gordon Childe’s famous “oasis theory” of the origins of agriculture in the Near East, which Childe first referred to in 1928 and stated more explicitly in 1934 and in later publications (Childe 1928:42; 1934:24–25). The prehistoric cultural sequence that Pumpelly established at Anau strongly influenced later Soviet archaeological research and his two-volume 1908 publication brought Central Asia, as a center of early civilization, to the attention of Western archaeologists. In retrospect, he deserves to be recognized as an innovator well ahead of his time, whose methods anticipated the multidisciplinary and systemic approaches that were widely adopted in the second half of the 20th century by American and British archaeologists investigating the origins of agriculture (e.g., Binford 1968; Braidwood and Howe 1960; Flannery 1968; Harris 1969; Higgs and Jarman 1969; MacNeish 1967).

Soviet Investigations between the Revolution and the Second World War No further significant investigation into the pre­historic archaeology of Turkmenistan took place until after the Russian Revolution of 1918, but changes in the aims and structure of archaeological research soon followed the Revolution and the ensuing Civil War. For example, unauthorized excavations were forbidden, an archaeological map of Central Asia was compiled in 1920–21, and, after the political boundaries of the Central Asian republics were de-



history of archaeological research

marcated in 1924–25, separate institutions were set up in each to organize archaeological research (Kohl 1984:18). With the establishment of Soviet administration, archaeology became a highly structured, well-funded, state-controlled enterprise, oriented toward revealing the origins of communism. The Soviet approach provided a very clear if rigid formula for archaeological excavation, the cataloguing and analysis of artifact assemblages, and the dissemination of results. Prehistoric archaeology in the Turkmen Soviet Socialist Republic can be said to have begun when D. D. Bukinich, who had discovered the easternpiedmont site of Namazga-depe in 1916, returned in 1924 to start investigating it (Bukinich 1924; Kohl 1984:18). Two years later he found cultural layers buried 7 m below the level of the piedmont surface near Anau, which suggested that active sedimentation could have masked archaeological sites on the alluvial fans that decline northward across the piedmont from the base of the Kopetdag mountains. In 1925, an ethnographic-archaeological association was created in Ashgabat, but it was soon superseded by the creation of an archaeological section of Turkmen’kult, the Institute of Turkmen Culture. In the 1930s and during the Second World War Turkmen’kult together with Turkomstaris, the Turkestan Committee for the Preservation of Monuments of Antiquity and Art, established in 1921, organized expeditions to investigate archaeological sites on the Kopetdag piedmont, in the Balkhan-Atrek region in western Turkmenistan, between the Tedzhen and Murghab rivers, and along the Amudarya (Kohl 1984:19). During this period A. A. Marushchenko and A. F. Ganyalin worked at the piedmont sites of Old and New Nisa and Ak-depe west of Ashgabat, as well as at Altyn-depe at the eastern end of the piedmont. The most significant development in western Central Asian archaeology at this time was the formation by the Soviet Academy of Sciences in 1937 of the Khoresmian Archaeological Ethnographic Expedition, directed by S. P. Tolstov. This project investigated many sites south of the Aral Sea in northern Turkmenistan and central Uzbekistan, including the stratified site of Janbas (Dzhanbas) 4 in the ancient delta of the Akchadarya (part of the former drainage of the lower Amudarya). In 1939, following his excavations at Janbas 4, Tolstov identified the Keltiminar Culture of hunter-fisher-gatherers (Tolstov 1946, 1958; this volume, pp. 64–68). The project continued after

45

1945 under the direction of M. A. Itina, and in 1968 A. V. Vinogradov published a comprehensive account of the Neolithic Keltiminar sites found up to 1965. By the early 1980s over 60 Keltiminar sites had been identified by aerial reconnaissance, ground survey, and excavation (Kohl 1984:59). As Kohl remarked (1984:19–20), “This project…pioneered numerous field techniques in Central Asian archaeology, including settlement pattern studies, aerial reconnaissance and photography, and the use of mechanized digging equipment”; it also generated important publications and made a major contribution to the training of local archaeologists.

Soviet Investigations after the Second World War In 1946 a major new archaeological initiative was taken when the Southern Turkmenistan Complex Archaeological Expedition (YuTAKE) was set up under the direction of the renowned orientalist M. E. Masson. The work of YuTAKE was divided into brigades that focused on particular sites and periods. One of its earliest missions was undertaken by the IXth Brigade, led by A. P. Okladnikov, who worked extensively on Palaeolithic and Mesolithic sites on the Krasnovodsk plateau and in the Bolshoi Balkhan region between 1947 and 1952 (Okladnikov 1951, 1953). His excavation of the rockshelter site of Jebel (Djebel) close to the Caspian Sea at the southwestern end of the Bolshoi Balkhan massif yielded a long stratigraphic sequence of Mesolithic and Neolithic deposits (Okladnikov 1956), and although the chronology was insecure, it came to be regarded as the type site for the Turkmenistan Caspian Mesolithic. Okladnikov also examined two other rockshelters at the foot of the southern escarpment of the Bolshoi Balkhan—Dam Dam Cheshme 1 and 2—but it was not until the 1960s that these two important Mesolithic– Neolithic sites were comprehensively excavated, by G. E. Markov of Moscow State University (Markov 1966a, 1966b, 1981). (These three sites are described more fully in this volume, pp. 57–58 and 113–15). YuTAKE sponsored many other field projects between 1952 and 1965. One of the most significant of these was B. A. Kuftin’s work at Namazga-depe in 1952, published posthumously in 1956. It established a chronological sequence divided into six phases termed Namazga I–VI, which became the accepted di-

46

origins of agriculture in western central asia

vision of the period from the Chalcolithic (Eneolithic) to the Late Bronze Age. The important Bronze Age site of Altyn-depe, farther east on the piedmont, which had been discovered in 1929 by A. A. Semenov, was explored in 1947 by the VIIth Brigade of YuTAKE, directed by M. E. Masson and G. A. Pugachenkova, and excavated intermittently between 1951 and 1961 by S. A. Ershov and A. F. Ganyalin (Ganyalin 1967). The latter also worked at the nearby site of Ilgynly-depe (Ganyalin 1959). Between 1954 and 1956 M. E. Masson’s son, V. M. Masson, discovered and excavated the Iron Age Yaz I culture complex in Margiana (the part of the ancient delta of the Murghab river that extends north of the classical and medieval cities of Merv). Continuing in his father’s footsteps, the younger Masson went on to direct the XIVth Brigade with V. I. Sarianidi and I. N. Khlopin, excavating 11 sites between 1955 and 1962, 9 in the Geoksyur oasis (the ancient delta of the Tedzhen river), Kara-depe on the piedmont near the town of Artykh, and the Early Neolithic site of Jeitun at the edge of the Karakum desert northwest of Ashgabat. Further work was undertaken by the XIVth Brigade at Altyn-depe from 1965 to 1967, after which V. M. Masson became the sole director of the excavations there. He continued to work actively at Altyn-depe until 1978, and returned to the site intermittently thereafter (Masson 1988). V. I. Sarianidi went on to excavate at the Bronze Age piedmont sites of Ulug-depe and Khapuz-depe in the late 1960s, and I. N. Khlopin, together with his wife, L. I. Khlopina, turned his attention to the Sumbar valley where in 1977 he discovered the Early Bronze Age cemetery of Parkhai II (Khlopin 1981 and Kohl 1984:20 for much of the information in this paragraph). A second phase of research in Margiana began in the early 1970s when the Institute of Archaeology of the Academy of Sciences of the USSR in Moscow and the Institute of History of the Turkmen Academy of Sciences formed the Margiana Archaeological Expedition (MAE), with V. I. Sarianidi as director. Further work was undertaken at the previously excavated sites of Auchin-depe and Takhirbai-depe, and an extensive topographic survey located over 100 Bronze Age sites, including the large settlement of Gonur 1. In 1978 it was decided to divide the MAE into two teams, one, led by Sarianidi, began excavations at the southern and eastern sites around Togolok and Gonur, and the other, led by I. S. Masimov of the Turkmen Academy of Sciences, began excavations farther north in the Kelleli group of sites (Hiebert 1994a:16–17).

It was during the decade from 1963 to 1973 that the Neolithic sites of what came to be called the Jeitun Culture of the Kopetdag piedmont were systematically studied. V. M. Masson laid the foundations for this research with his excavations at the type site of Jeitun in the late 1950s and early 1960s (Masson 1957, 1971; this volume, pp. 95–97), but it was O. K. Berdiev, who died in a car accident at the age of 38 in 1973, who extended research to the Jeitun-Culture sites of the piedmont as a whole. He deserves special recognition for carrying out, with great dedication, a comprehensive survey of the Neolithic piedmont sites, from Bami in the west to the Meana-Chaacha district in the east (Fig. 6.1). His initial fieldwork was reported in his doctoral dissertation (Berdiev 1963a). Then, between 1963 and 1973, he excavated and published the results of work at the sites of Bami (1963b), Chagylly (1964a, 1966), Togolok (1964b), New Nisa (1965), Chakmakli (1968a, 1968b), Chopan (1972a), Pessedjik (1968b, 1973), and Monjukli (1972b), and also published various more general papers (Berdiev 1964c, 1968c, 1970, 1971a, 1971b, 1971c, 1976; Atagarryev and Berdiev 1967, 1970). Thus, in a decade of extraordinarily productive fieldwork and publication, tragically terminated by his early death, Berdiev made an unmatched contribution to the definition and interpretation of the Jeitun Culture, much of which he synthesized in the volume that he published in 1969 entitled (in English translation) The Most Ancient Agriculturalists of Southern Turkmenistan. Also on the piedmont, investigation of the north mound at Anau was resumed in 1977 when the Turkmen archaeologist K. K. Kurbansakhatov, who was trained by V. M. Masson and became a leading figure in Turkmenian archaeology, began work there as part of his doctoral research. Between 1977 and 1982 he excavated areas north of Komarov’s trench, exposed successive layers of building, and obtained a long stratigraphic sequence by means of a deep sounding that reached the lowest deposits (Kurbansakhatov 1987; Hiebert 2003:30–31).

Collaboration between Foreign Archaeologists and Russian, Turkmen, and Uzbek Colleagues In the early 1970s foreign scholars began again, some seven decades after Pumpelly’s work at Anau, to become involved in the archaeology of western Cen-



history of archaeological research

tral Asia. C. Lamberg-Karlovsky of Harvard University and M. Tosi of the University of Bologna were both interested in “spheres of interaction” of resource exploitation and trade, and were eager to learn more about possible links between the Iranian sites at which they had worked—respectively Tepe Yahya in southeastern Iran and Shar-i Sokhta in east-central Iran—and sites to the north of the Iranian plateau in Turkmenistan. They had examined parallels in the material inventories of sites that circumscribed the plateau (LambergKarlovsky and Tosi 1973; Lamberg-Karlovsky 1975), which strengthened Masson and Sarianidi’s prior suppositions (1972) of the possible existence of prehistoric trade networks connecting Iran and Turkmenistan. This initial research provided the background and stimulus for both Lamberg-Karlovsky and Tosi to develop, later in the 1970s and the 1980s, academic collaboration with Soviet archaeologists. This process was sparked in 1977 by a meeting between Lamberg-Karlovsky and two archaeologists from the Institute of Archaeology in Moscow, R. Munchaev and N. Merpert, at a conference in Denmark on the origins of agriculture at which Tosi was also present. This initial contact led, after a further meeting at Harvard University in 1979 to plan collaboration, to P. L. Kohl of Wellesley College, Massachusetts, being invited to travel to Soviet Central Asia in 1979 and 1980 to visit archaeological sites, study museum collections, and meet Soviet colleagues (Lamberg-Karlovsky 1994:xviii–xx). Although collaborative excavations were not possible at that time, given the political climate of the Cold War, Kohl’s study tours resulted in the publication of two of the most significant volumes on Central Asia produced by a Western archaeologist, one of which focused on the Bronze Age while the other provided a comprehensive overview from the Palaeolithic to the beginnings of the Iron Age (Kohl 1981, 1984). Following Kohl’s visits, the first joint USA-USSR archaeological symposium was organized by Kohl and Lamberg-Karlovsky in 1981 at Harvard’s Peabody Museum, and three more joint symposia then took place, in Samarkand in 1983, Washington, DC, in 1986, and Tbilisi in 1988. The second of these resulted, in 1985, in the first participation of American archaeologists (Lamberg-Karlovsky and Kohl) in a Soviet excavation, at Sarazm in Tajikistan, where a French team was already working (LambergKarlovsky 1994:xxi–xxii). Another outcome of the developing collaboration with Soviet archaeologists that followed Lamberg-

47

Karlovsky’s original initiative was the opportunity taken by one of his research students, F. T. Hiebert, to become involved in Central Asian archaeology. Hiebert was able to undertake a study tour of archaeological sites in Soviet Central Asia, and, after meeting V. I. Sarianidi, was invited to join his 1989 field season at the Bronze Age site of Gonur 1 in Margiana. Hiebert went on to complete in 1992 a doctoral dissertation on the Bronze Age settlements of Central Asia, which gave rise to a substantial monograph (1994a). He then initiated a second collaborative field project in Turkmenistan: a re-study of the Chalcolithic (Eneolithic)Iron Age north mound at Anau. Here he was following in the footsteps of Pumpelly and later investigators and this led to the publication, with Kurbansakhatov and others, of a second major monograph (2003). Tosi also built on his early contacts with Soviet colleagues and in 1989 a protocol of cooperation was signed in Rome between the Academy of Sciences of the USSR and the Italian Institute of Middle and Far East Studies (IsMEO). G. Koshelenko, head of the Classical Archaeology section of the Academy of Sciences, had already conceived the idea of creating an archaeological map of the Merv oasis, and after many years of excavation within the walls of ancient Merv, he, in association with A. Gubaev of the Turkmenistan State University, had begun exploring fortresses and other settlements on the northern borders of the oasis. The area had been extensively occupied during the Iron Age, as V. M. Masson’s discovery (already referred to) of the Yaz I culture complex in the 1950s had shown. The proposal was now to explore the whole of the vast northern area and to produce an “Archaeological Map of the Northern Limits of the Merv Oasis.” Tosi visited the area with several Italian colleagues in October 1989, and the new project was set up under the joint direction of Gubaev, Koshelenko, and Tosi. Archaeological and geological fieldwork was carried out from 1990 to 1995 by the Italian-Russian-Turkmen team, which included B. Marcolongo, director of the Institute for Applied Geology at Padova. Innovative use was made of satellite and other remotely sensed images to detect ancient irrigation features and to locate new archaeological sites, and in 1998 the project produced its first major publication, a volume of text and figures and another of maps, edited by Gubaev, Koshelenko, and Tosi (1998). Tosi and Marcolongo also conducted smaller-scale geoarchaeological studies in southern Turkmenistan, in the Sumbar valley, and the Meana– Chaacha district (Marcolongo, pers. comm. 2000).

48

origins of agriculture in western central asia

Independently of the American and Italian collaborative projects that stemmed from the initial meetings with Soviet archaeologists in the late 1970s and early 1980s, British archaeologists also started two joint field projects with Russian and Turkmen colleagues in the late 1980s and early 1990s, the first at the Neolithic site of Jeitun and the second at the multiperiod urban center of Merv. Both were, initially, collaborative ventures between the Institute of Archaeology, University College London (UCL), the Institute of the History of Material Culture in St. Petersburg (formerly Leningrad) and the Turkmen Academy of Sciences in Ashgabat. Participation in the two projects from the London Institute was organized and led by D. R. Harris for Jeitun and G. Herrmann for Merv, in close collaboration with Masson from St. Petersburg and Kurbansakhatov in Ashgabat. The Jeitun Archaeological-Environmental Project was initiated in 1989 and the International Merv Project (IMP) in 1992, and in subsequent years British participation was broadened to include colleagues from several other university departments and from the British Museum. Whereas the Jeitun project was concerned with prehistoric, principally Neolithic, archaeology, the IMP focused on the development in historical times of the ancient and medieval cities of the Merv oasis. The genesis and development of the Jeitun project is described in the Preface to this volume, and a brief account of the recent investigations at Merv is given here. The spectacular site of Merv, with its many standing buildings and extensive walls enclosing a complex of ancient cities (Erk Kala and Gyaur Kala) and their medieval successors (Sultan Khan Kala and Abdullah Khan Kala), has been a focus for archaeological investigations since 1890 when V. A. Zhukovsky began excavating there, following earlier reconnaissance by General Komarov. Work continued at the site intermittently, intensifying after the Second World War when YuTAKE conducted excavations in the area of the ancient cities. The size and complexity of the site presents a great archaeological challenge, and one of the main aims of the IMP was to carry out detailed surface surveys of the distribution of ceramics and other materials and to document all the standing buildings and other visible structures such as canals and water cisterns. Members of the project also undertook selective excavations of house foundations, as well as of metal-working areas where early evidence of steel manufacture was unearthed.

The initial Turkmenian-British agreement for collaboration at Merv was for the years 1992–94, but the project was later extended for a second, and then a third, three-year period, under the co-direction of Herrmann and Kurbansakhatov, with St. John Simpson of the British Museum in charge of the excavations. Reports of each season’s work appeared in the journal Iran from 1993 to 2001, and specialized papers and substantive volumes were also published (Brun 2005; Herrmann 1999; Puschnigg 2006). In 1999 the international importance of Merv was recognized by Unesco when it was formally declared a World Heritage Site. It was then agreed that greater priority should henceforth be given to conservation and interpretation of the whole area of the walled cities and outlying monuments that now constitute an archaeological park. Accordingly a new five-year Turkmenian-British collaboration between the Institute of Archaeology, UCL, and the Ministry of Culture of Turkmenistan was started in 2001, co-directed by Kurbansakhatov and T. D. Williams of UCL (Williams et al. 2002, 2003). This was extended for a further five years to 2010, co-directed by Williams and Dr. Mukhammed Mamdeov of the Ministry of Culture. Research and selective excavation continue to be part of the project, together with conservation work on the standing buildings, the training of local staff, and the development of a management plan for the archaeological park (Williams 2007, 2008). East of Merv, across the Amudarya in Uzbekistan, another international project is underway. Between 1995 and 1999 the Polish-Uzbek Archaeological Expedition, under the direction of Karol Szymczac of Warsaw University’s Institute of Archaeology and Mukhiddin Khudzhanazarov of the Uzbek Academy of Sciences’ Institute of Archaeology in Samarkand, conducted archaeological and palaeoenvironmental field research in the Kyzylkum. The aim of the expedition was to establish an accurate chronology and better understanding of the prehistoric occupation of the desert in the early and mid Holocene. The project continued, under the same direction, from 2002 onward and expanded to include French colleagues. Research on the prehistoric archaeology of the Kyzylkum was initiated in 1937 by S. P. Tolstov, who first identified sites of the Keltiminar Culture (this chapter, p. 45), and it was continued after the Second World War by several scholars, notably A. V. Vinogradov (1968, 1981; Vinogradov and Mamedov



history of archaeological research

1975). Most of the many Keltiminar sites discovered in the desert lack stratified cultural deposits, but Vinogradov (1981) found several where stratigraphy was preserved, including the site at the northeastern edge of the Ayakagytma depression in the southern Kyzylkum that was chosen by the Polish-Uzbek team for intensive investigation. A monograph reporting initial results of the project, including 13 radiocarbon dates and preliminary accounts of the excavation and of the structures, artifacts, and animal bones found, has been published (Szymczac and Khudzhanazarov 2006b, and see this volume, pp. 67–68). This account of the history of archaeological investigations in Turkmenistan and part of Uzbekistan demonstrates the great advances in knowledge that have been achieved since Komarov carried out the first excavations at Anau in 1886. Pumpelly’s pioneering investigations at Anau in the first decade of the 20th century inaugurated a new scientific approach to the study of Central Asian prehistory, and the extensive surveys and excavations carried out by Russian, Turkmen, and Uzbek archaeologists from the 1920s onward established the framework, and filled in much of the detail, of the prehistoric and historical past of the region. With the exception of Pumpelly, foreign archaeologists played little part in the early stages of this exciting process of discovery, but in more recent decades that situation changed and an increasing number of Western archaeologists (ourselves included) have had the privilege of participating, with local colleagues, in the continuing archaeological investigation of the region’s past.

Northeastern Iran Here and in the final section of this chapter a brief account is given of the history of research on the prehistoric archaeology of those areas of northeastern Iran and northern Afghanistan where investigations have been conducted that relate closely to the prehistory of Turkmenistan. The most relevant sites in Iran are located in the province of Mazandaran south of the Caspian Sea and in the northernmost part of the province of Khorassan close to the Iran-Turkmenistan frontier. Archaeological explorations of northern Iran began in the 19th century when Western and Iranian soldiers and travelers examined some of the more conspicuous sites, but systematic survey and excava-

49

tion did not begin until the 1930s. Among the first sites to be test-excavated were Tureng Tepe and Shah Tepe on the plain of the Gorgan river close to the southeastern coast of the Caspian (Fig. 6.1). They were investigated respectively by the American archaeologist F. R. Wulsin and the Swedish archaeologist T. J. Arne, and Arne also mapped the location of 310 sites on the western Gorgan plain (Arne 1945; Daher 1969; Kohl 1984:22; Wulsin 1932). More comprehensive excavations were carried out at Tureng in the 1960s and 1970s by a French Archaeological Mission under the direction of J. Deshayes (1963 and reports in Iranica Antiqua 1965, 1966; Iran 1967, 1968, 1970, 1972, 1973, 1974, 1976; and Paléorient 1974), and in 1960 and 1962 the British archaeologist D. Stronach excavated at another site, Yarim Tepe, on the eastern Gorgan plain where he found sherds resembling Jeitun pottery in the deepest levels (Stronach 1972; this volume, p. 62). South of the Elburz mountains on the Damghan plain, the mainly Bronze Age site of Tepe Hissar was excavated in 1931 and 1933 by an American team led by E. F. Schmidt (1937), and reinvestigated in 1976 by an American-Italian-Iranian team directed by R. H. Dyson, Jr., and M. Tosi (Dyson and Howard 1989). The Damghan plain was also surveyed at that time (Trinkaus 1981). Investigations of three rockshelters close to the southeastern shore of the Caspian (Fig. 6.1), which led to the recognition of the Iranian Caspian Mesolithic, began in the years 1949–51 when the American archaeologist C. S. Coon excavated two of the sites— Ghar-i Kamarband (Belt Cave) and Hotu—a short distance west of the town of Behshahr (Coon 1951, 1952, 1957:129–204; Dupree 1952). The third site, Ali Tappeh, located just east of Behshahr, was found in 1962 and excavated by the British archaeologist C. B. M. McBurney (1964, 1968). The three sites, particularly Ali Tappeh, provided valuable evidence of changes in Caspian sea levels, climate, and fauna during Lateglacial and Postglacial times. During the early 1970s, under the auspices of a Japanese-Iranian Joint Archaeological Mission, S. Masuda excavated two small mounds at Sang-i Čakmaq (Sang-i Chakmak) on the north Iranian plateau near Bastam north of Shahroud (Fig. 6.1), both of which proved to have been occupied during the Neolithic (Masuda 1972, 1974a, 1974b, 1976, 1977, 1984). The pottery, stone, and bone tools and animal figurines found in the lower levels of the eastern mound resemble equivalent assemblages found

50

origins of agriculture in western central asia

at Jeitun-Culture sites on the Kopetdag piedmont (this volume, p. 63). In the 1970s too, the Iranian archaeologist E. O. Negahban excavated a site much farther west, Zaghe on the Qazvin plain south of the Elburz range. Its lower levels contained the remains of small rectangular mudbrick houses with plastered platforms resembling the houses at Jeitun, and, together with the nearby sites of Sagzabad and Ghabristan (Qabristan), it provided a stratigraphic sequence that appeared to extend from the Neolithic to the Achaemenian period (Negahban 1977, 1979; Shahmirizade 1977). In the late 1970s reconnaissance surveys and test excavations were carried out close to the frontier with the USSR in northeastern Iran. An Italian team from the University of Turin led by R. V. Ricciardi, assisted by R. Biscione, investigated sites in the upper Atrek valley near the town of Quchan (Fig. 6.1). During three seasons of fieldwork in 1976, 1977, and 1978 they identified some 180 mounds, made extensive surface collections, and carried out test excavations at two of the larger mounds, Yam and Khorramabad, but found no evidence of pre-Chalcolithic occupation (Ricciardi 1980). Also in 1978 P. L. Kohl and D. L. Heskel (1980) surveyed the Dargaz (Darreh Gaz) plain northeast of Quchan, which had previously been explored by Negahban and which affords easy access to the Kopetdag piedmont. They located 40 sites, only 4of which were found to contain prehistoric material. One of these (DG19) yielded a few sherds resembling ceramics from Jeitun and also from Yarim, far to the west (Fig. 6.1). The Iranian Revolution of 1979 brought to a halt investigations in Iran by foreign archaeologists, although some Iranian projects continued in the 1980s. From the mid 1990s onward fieldwork was resumed, and since 2000 a dozen joint Iranian-foreign missions have been initiated under the auspices of the Iranian Center for Archaeological Research (ICAR) in Tehran (Azarnoush and Helwing 2005:189). Examples of this renewed activity are excavations and surveys by H. Fazeli of the Institute of Archaeology of the University of Tehran and colleagues from English and Australian universities that have established a new chronology for the northern Central Plateau. This new phase of research included a resumption of excavation, coupled with a program of AMS radiocarbon dating, at the site of Zaghe on the Qazvin plain which demonstrated that it was first occupied around 5300 cal. BCE in the Transitional Chalcolithic period

and that, contrary to earlier interpretations (Negahban 1977, 1979; Shahmirizade 1977), there was no evidence of Neolithic occupation at the site (Fazeli, Wong, and Potts 2005). Also, extensive settlement surveys on the Tehran plain have discovered many prehistoric (mainly Chalcolithic) sites, while new excavations at Cheshmeh-Ali just south of Tehran and at Tepe Pardis farther southeast close to the city of Garchak, have yielded an absolute chronology for the Late Neolithic–Transitional Chalcolithic–Early Chalcolithic (c. 6200–5500–4700 cal. BCE), with evidence of (Late) Neolithic settlement at Cheshmeh-Ali and Sadeghabadi, and of the Late Neolithic/Transitional Chalcolithic interface at Tepe Pardis (Coningham et al. 2004, 2006; Fazeli, Coningham, and Batt 2004, Fazeli et al. 2007). Iranian prehistorians have also carried out new surveys and excavations closer to southern Turkmenistan, in Mazandaran and Khorassan provinces. Near Behshahr in eastern Mazandaran, close to the rockshelters excavated by Coon, Mesolithic (Epipalaeolithic) deposits and Neolithic potsherds and sickle blades were found in a cave (Qar-i Komishan) near the Hotu rockshelter (Fig. 6.1) during a landscape survey directed by A. Mahfroozi of the Iranian Cultural Heritage and Tourism Organization (Azarnoush and Helwing 2005:193, 201; Mahfroozi 2004; Shidrang 2004). Also at Tugh Tepe, 25 km west of Behshahr, Mahfroozi found sherds of Jeitun type above the deepest levels, which lack pottery, and below levels containing late Neolithic pottery of Cheshmeh-Ali type, while at the large site of Gohar Tepe, 5 km west of Behshahr, he found late Neolithic Cheshmeh-Ali sherds below Iron Age, Bronze Age, and Chalcolithic levels (A. Mahfroozi, pers. comm. 2007). Northeast of these sites, on the Gorgan plain, a rescue excavation carried out in 2000 at another large mound, Aq Tepe, by S. Malek Shahmirizade from the University of Tehran uncovered three Neolithic occupation phases; and on the northern side of the Elburz mountains above the Gorgan plain in the Golestan National Park five sites with evidence of Neolithic occupation have been reported by O. Rekavandi, one of which, Armadlu Tepe, is the first Neolithic site to be found in a forested environment (Azarnoush and Helwing 2005:199). In northern Khorassan several new projects are underway. Following earlier surveys in the Shahroud area, during which a site, Deh Keir Tepe, was found with a ceramic assemblage very similar to the one at



history of archaeological research

Sang-i Čakmaq (Rezvani 1999), K. Roustaei of ICAR carried out a further landscape survey in the Shahroud–Bastam area in 2005. He found pottery with painted designs that resemble Jeitun-Culture ceramics at six Neolithic sites, including Deh Keir and Čakmaq. Then, early in 2006, soundings at Deh Keir by Rezvani and by Roustaei at one of the other sites, Kalate Khan about 15 km southwest of Čakmaq, revealed a Neolithic sequence closely comparable to the one at the eastern mound at Čakmaq (Roustaei In press and pers. comm. 2007; Azarnoush and Helwing 2005:199). Farther east O. Garazhian has carried out extensive surveys in the vicinity of Bojnurd, Neyshabour, and Dargaz (Fig. 6.1). Two sites have been sounded: Qaleh Khan near Ashkhaneh west of Bojnurd where ceramics resembling Middle/Late Jeitun pottery were found in Neolithic levels, and Borj Tepe 25 km southeast of Neyshabour where 2–3 m of what are probably Neolithic deposits underlie the lowest (Chalcolithic) level so far excavated, and dated to 4530 cal. BCE. Also, in his survey of the Dargaz plain, Garazhian identified 20 prehistoric sites, 3 of which, Yarim Tepe, Novroz Tepe, and Novkhandan Tepe, produced pottery that resembles Jeitun-Culture ceramics (Garazhian 2006, In press, and pers. comm. 2007). Although no conclusive, radiocarbon-dated evidence of links with Neolithic sites in southern Turkmenistan has so far been produced by these many new surveys and excavations, together they suggest that close connections existed between northeastern Iran and southern Turkmenistan during the Neolithic (see Chapter 11).

Northern Afghanistan Systematic search for prehistoric sites in northern Afghanistan did not get underway until after the Second World War. In 1954 C. S. Coon prospected for Palaeolithic cave sites in the northern foothills of the Hindu Kush. He found and excavated a rockshelter, Kara Kamar, in the valley of the Samangan river near Haibak southeast of Mazar-i Sharif, where he recovered stone tools and animal bones of Upper Palaeolithic and Mesolithic age (Coon 1957:205–37; Coon and Ralph 1955; Davis 1978:50–54, 63–67). In 1959 L. B. Dupree, who had worked with Coon at Ghar-i Kamarband and Hotu in northern Iran and who had already excavated at the Chalcolithic site of Deh Morasi Ghundai near Kandahar in southern

51

Afghanistan (Dupree 1963), carried out an extensive survey in the northern foothill zone that Coon had partially explored. Dupree found several sites with deposits of Palaeolithic, Mesolithic, and Neolithic age (Dupree 1964; Dupree and Howe 1963), and between 1962 and 1965 he investigated and partly excavated four of them near the village of Aq Kupruk in the Balkh valley south of Mazar-i Sharif. He interpreted the sites as spanning collectively a long sequence from the Upper Palaeolithic (divided into Kuprukian A and B) at the base, through Neolithic and Chalcolithic levels, to the Iron Age and Early Islamic period. He claimed that sheep and goats had been domesticated locally during the Neolithic (Dupree 1967, 1972; Perkins 1972), but there were inconsistencies between the radiocarbon dates and the stratigraphy, and the domestic status of the sheep and goats in the Neolithic levels was later questioned (Meadow 1989a:34). In 1966 Dupree excavated a rockshelter, Dari-i Kur (Darra-i Kur), farther east in the foothill zone beyond the town of Kunduz, where he found evidence of Middle Palaeolithic occupation and supposedly also of a Neolithic “goat cult” assemblage, which was later shown to date to c. 2000 BC uncalibrated (Dupree 1967:24; 1972:11−13, 32, 79, 82; Shaffer 1978:81–83; and see this volume, pp. 58–59, for further discussion of the Aq Kupruk sites). In 1965 a member of an Italian archaeological mission to Afghanistan, S. M. Puglisi, test-excavated a rockshelter, Dara-i Kalon, in the vicinity of Haibak, southwest of the site of Kara Kamar, which contained Mesolithic assemblages (Alessio, Bachechi, and Cortesi 1967), and in 1969 Dupree and R. S. Davis visited the same area and found a Mesolithic surface site, Kok Jar, between Dara-i Kalon and Kara Kamar (Davis 1978:38–39, 63–66). Surveys of the plain around Kunduz, and farther east, were conducted in the 1970s by a French team led by J.-C. Gardin (Gardin and Lyonnet 1978–79), and several Bronze Age sites were found in the Shortugai and Talugab areas (Kohl 1984:23). Soviet archaeologists were also active in northern Afghanistan in the 1970s. The Soviet-Afghan Archaeological Expedition, led by A. V. Vinogradov, found numerous lithic surface sites on the arid plain between the foothill zone and the Amudarya, most of which were dated typologically to the Mesolithic (Vinogradov 1979). With the Soviet invasion of Afghanistan in 1979, survey and excavation ceased and continuing political

52

origins of agriculture in western central asia

upheavals and military action have since prevented any significant resumption of archaeological activity in northern Afghanistan. Many questions about the prehistory of the region and early connections with Turkmenistan, Uzbekistan, and northern Iran remain

unresolved, including when and how an agro-pastoral economy became established in the valleys and plains north of the Hindu Kush. Resolution of such questions must await political conditions that will allow archaeological research to be resumed in the region.

6

The Mesolithic and Neolithic Periods: Sites, Sequences, and Subsistence with Jennifer Coolidge

T

he main purpose of this chapter is to provide a summary account of the principal Mesolithic and Neolithic sites, sequences, and subsistence economies of Turkmenistan and adjacent parts of Uzbekistan, Iran, and Afghanistan, but first we review briefly the Palaeolithic foundations of the prehistory of western Central Asia as a whole.

Palaeolithic Prelude At present, most of the evidence of human populations in western Central Asia during the Palaeolithic comes from the more mountainous areas in northern Afghanistan, southern Uzbekistan, Tajikistan, and Kirgizstan, although there is also some evidence from lowland areas in Kazakhstan and Turkmenistan (Dennell 2009; Moloney, Olsen, and Voloshin 2001; Vishnyatsky 1996, 1999).

Lower Palaeolithic In the Tajik depression at the site of Kuldara in southern Tajikistan, chipped stone tools stratified in palaeosols buried in deep deposits of loess have been dated palaeomagnetically to c. 900,000 years ago in the Lower Palaeolithic; and such tools have also been found in buried soils estimated to be 600,000–400,000 years old within loess at other sites in the region, such as Karatau, Lakhuti, and Chashmanigar (Davis and

Ranov 1999:187–88; Dennell 2009:217–20, 327–29; Ranov 2001; Ranov and Dodonov 2003; Shackleton et al. 1995; Vishnyatsky 1999:87–90; this volume, pp. 22–23). Variations in the thickness of the loess deposits and the presence in them of the palaeosols provide evidence of oscillating cold dry, and warmer, more humid conditions associated with the fluctuations of Pleistocene ice sheets, and the stone tools indicate that hunters occupied the area during semi-arid interglacial and interstadial intervals.

Middle Palaeolithic Numerous Middle Palaeolithic sites have been discovered in western Central Asia, some with hominin remains such as the famous Teshik Tash cave south of Samarkand in Uzbekistan, where in 1938 and 1939 A. P. Okladnikov excavated a Mousterian burial of a young Neanderthal child surrounded by wild-goat horn cores (Vishnyatsky 1999:80–82), and in northern Afghanistan where in 1966 L. B. Dupree excavated Middle Palaeolithic deposits in a rockshelter, Dari-i Kur (Darra-i Kur), and found a fragment of a Neanderthal skull (Angel 1972; Dupree 1972:11−13, 79). More recently, 10 hominin teeth and a humerus fragment were excavated at a large cave site, Sel-Ungur, in the Fergana depression in southern Kirgizstan, but it is uncertain whether they derive from a Neanderthal or a pre-Neanderthal population. Large numbers of stone tools resembling the pebble and flake indus-

54

origins of agriculture in western central asia

tries of the Central Asian loess sites were found at Sel-Ungur and more than 4000 fragments of poorly preserved mammal bones, including wild sheep, goat, deer, cattle (aurochs), horse, wolf, fox, and several now extinct taxa such as cave bear, cave lion, and cave hyaena. The fauna, together with a uranium-thorium date of 126,000 ± 5000 BP, points to occupation of the site by hunters in the late Middle Pleistocene (Davis and Ranov 1999:189–90; Dennell 2009:329–30, 453; Islamov 1990; Vishnyatsky 1999:94–95). In addition to the few cave sites that have yielded hominin remains, many Middle Palaeolithic sites with pebble tools and flakes of Mousterian type have been discovered in the region. Hand axes and other bifacial tools are generally absent, although undated surface finds of Acheulean bifaces, some of which may be of Middle Palaeolithic age, have been reported from central Kazakhstan, Uzbekistan, and western Turkmenistan (e.g., on the Krasnovodsk plateau; Vishnyatsky 1989; 1999:71–72). Vishnyatsky (1999:110–11) pointed out that most of the faunas represented in the Middle Palaeolithic Mousterian sites of the region are dominated by wild goat and sheep, which may reflect dominance of steppe vegetation under semiarid conditions during the Middle Pleistocene, an inference supported by palynological data from the Sel-Ungur site (Vishnyatsky 1999:94). Davis and Ranov (1999:192–93) concluded that in the Middle Palaeolithic Central Asia was “part of the eastern extension [from Southwest Asia] of the Neandertals along the mid-altitude foothills of the Zagros, Kopet Dagh, Tien Shan, and Altai mountain ranges.” However, the relatively small number of adequately dated sites prevents more detailed interpretation of the changing relationship between the distribution of hunter-gatherer populations and the environmental oscillations of the Pleistocene, beyond the generalization that continuous human occupation would have been possible only during interglacial periods (Dennell 2009:334, 461).

Upper Palaeolithic The Upper Palaeolithic presents a striking contrast to the preceding Middle Palaeolithic and the following Mesolithic (Epipalaeolithic) period because it is largely invisible in the Central Asian archaeological record (Derevyanko and Zun 1992). Very few stratified sites with deposits of demonstrably Upper Paleolithic age are known, and only two have been extensively

excavated and reported: the open-air Samarkand site within the city of Samarkand in Uzbekistan, which has been intermittently excavated since 1939, and Shugnou in eastern Tajikistan, an open-air site excavated by Ranov from 1968 to 1970 (Davis and Ranov 1999:191; Vishnyatsky 1999:86–87, 93–94). The stratigraphy of the Samarkand site is complex and poorly understood and its large assemblages of stone artifacts and animal bones probably represent a mixture of Upper Palaeolithic occupational episodes. Shugnou is located at about 2000 m above msl in the narrow upper valley of the Yaksu river and the dominance of horse remains in the small bone assemblage, coupled with palynological evidence of steppe vegetation, suggests that it was a site from which hunters intercepted horses and other migrating herbivores. A few other possibly Upper Palaeolithic sites have been reported, for example, Kharkush in Tajikistan and Siabcha in Uzbekistan (Davis and Ranov 1999:191), and Kara Kamar in northern Afghanistan, a rockshelter dug by C. S. Coon in 1954 (this volume, p. 51), but their rarity reinforces the general impression that during the Upper Palaeolithic the human population was very sparse. The few sites with evidence of Upper Palaeolithic occupation are poorly dated, but Davis (1990:272–73) maintained that none can confidently be placed within the Last Glacial Maximum (c. 23,000–c. 19,000 BP), and he suggested that the very cold and dry conditions that prevailed then led to a depopulation of Central Asia (as probably also occurred in previous glacial maxima during the Pleistocene).

The Mesolithic The period that follows the Upper Palaeolithic is customarily referred to in the archaeological literature on Southwest Asia and, less frequently, Central Asia as the Epipalaeolithic, and is usually regarded as synonymous with the Mesolithic. However, some archaeologists working in Central Asia, for example, V. A. Ranov (Ranov 2003; Ranov and Davis 1979:252; and see Brunet 1999, 2002, 2003), have defined the Epipalaeolithic and Mesolithic as separate, successive periods, which they distinguish by the absence of geometric microliths in the former and their presence in the latter—for example the change from non-geometric to geometric forms in levels 7 and 6 at the Dam Dam Cheshme 2 rockshelter in the Bolshoi Balkhan massif (see below). In this volume, which is



the mesolithic and neolithic periods: sites, sequences, and subsistence

not primarily concerned with lithic technology, we follow the more widespread practice (in Eurasian archaeology) of regarding the two terms as synonymous, while acknowledging that their usage varies by region; for example, Epipalaeolithic is used in the Levant and Mesolithic in Europe. In this volume, Mesolithic subsumes Epipalaeolithic. In their introduction to the Mesolithic of western Central Asia, Masson and Sarianidi (1972:26) described the period as a link between “the two great epochs in the history of man: the period of a foodgathering economy and the period of a production economy.” They defined four regional variants of the Mesolithic, which are, from west to east: the southeast Caspian region in western Turkmenistan and northern Iran, western Tajikistan, the Fergana valley in southwestern Kirgizstan, and the Pamir mountains in southeastern Tajikistan. Kohl (1984:39–43) and more recently Brunet (1999, 2002) also summarized the Mesolithic of these regions, as well as a fifth group of sites in the foothills and on the plains north of the Hindu Kush in northern Afghanistan. Only the first and last of these five regions are directly relevant to the theme of this book. The Mesolithic in western Central Asia is defined chronologically here as the period between approximately 10,000 and 7000 years ago (c. 11,500–8000 cal. BP), and Table 6.1 lists the rockshelter and cave sites with Mesolithic deposits in northern Iran, western Turkmenistan, and northern Afghanistan that are mentioned in the following sections of this chapter.

The Southeast Caspian Region In the coastal lowlands and adjacent uplands that fringe the Caspian Sea in northern Iran and western Turkmenistan there are two groups of rockshelters that have yielded evidence of what is referred to as the Caspian Mesolithic. The Iranian sites are located in Mazandaran province on the narrow coastal plain between the northern foothills of the Elburz mountains and the Caspian coast near the town of Behshahr (Fig. 6.1). Two rockshelters west of Behshahr, Ghar-i Kamarband (Belt Cave) and Hotu, were excavated by C. S. Coon in the fall of 1949 and the spring of 1951, and a third, Ali Tappeh, just east of the town, was found by C. B. M. McBurney in 1962 and excavated in 1963 and 1964 (this volume, p. 49). All three sites contained deep stratified deposits overlain by disturbed,

55

Table 6.1 Mesolithic rockshelter and cave sites referred to in the text. Northern Iran

Western Turkmenistan

Northern Afghanistan

Qar-i Komishan

Jebel

Kara Kamar

Ali Tappeh

Dam Dam Cheshme 1

Aq-Kupruk

Ghar-i Kamarband

Dam Dam Cheshme 2

Dar-i Kalon

Hotu

Kailyu, Hodja-Su

later prehistoric and historical materials. Recently, as part of a national renewal of archaeological research in Iran, another rockshelter, Qar-i Komishan, which contains Mesolithic deposits as well as some Neolithic potsherds, has been found near Hotu (this volume, p. 50). Preliminary investigation suggests that the Mesolithic assemblage represents a sequence similar to those recorded at Hotu and Ghar-i Kamarband, described below. At Ali Tappeh McBurney recognized 23 stratigraphic layers of Mesolithic age which he correlated with a series of radiocarbon dates that spanned the period c. 12,400–10,800 BP, later corrected by dating new charcoal samples (Hedges et al. 1994:348–49) to c. 11,300–10,200 BP (c. 13,000–11,800 cal. BP). Remains of large and small mammals and mollusks were recovered throughout the sequence (McBurney 1968:396 –99; Uerpmann and Frey 1981:146 –51), and changes in their relative abundance suggested a division of the occupation into five main stages that McBurney tentatively correlated with part of the north-European Lateglacial/early Postglacial climatic sequence (this volume, p. 24). The most conspicuous faunal contrasts were between abundance of gazelle in stages I, III, and IVb (correlated respectively with the Bølling interstadial, and the Older and Younger Dryas stadials) and a sharp increase in seal at the beginning of stage IVa (correlated with the first part of the Allerød interstadial). Although the correlation suggested by McBurney now appears too direct and oversimplified, the changes in abundance of gazelles and seals recorded at Ali Tappeh probably do reflect contemporaneous Lateglacial/ early Postglacial changes in local forest/steppe vegetation and Caspian Sea levels. Other large mammals represented in the sequence with varying frequencies include wild sheep, aurochs, onager, wild pig,

56

origins of agriculture in western central asia

6.1 Distribution of Mesolithic sites in the Bolshoi Balkhan massif and the southeast Caspian region, Neolithic Jeitun-Culture sites on the Kopetdag piedmont, related Neolithic sites in northeastern Iran, and the Chalcolithic–Bronze Age site of Anau near Ashgabat. The locations of sites recently reported by Iranian archaeologists working in northeastern Iran in the provinces of Mazandaran and Khorassan (and mentioned in the text) are not shown.

and deer (Uerpmann and Frey 1981:151), but there is no evidence of domestic animals at this early Mesolithic site. This situation contrasts with the nearby site of Ghar-i Kamarband where the stratigraphic sequence (divided by Coon into 28 levels) extends from the early Mesolithic into the Neolithic. Coon’s excavations coincided with the earliest attempts to date organic remains by the radiocarbon method, when large samples (often aggregated from more than one level) were required and when dates carried large probability errors. After an initial trial following his 1949 excavation at Ghar-i Kamarband, which produced four stratigraphically inconsistent dates from samples of charred bone, he obtained another five dates after his 1951 season which enabled him to establish a more consistent chronology. This provided approximate dates for the earliest (Mesolithic) occupation at c. 11,500 years ago (c. 13,400 cal. BP), and (at level 10) the beginning of the Neolithic at c. 8500 years ago (c. 9500 cal. BP), with gaps in occupation within and at the end of the Mesolithic (Coon 1951:30–32; 1957:165–67;

Ralph 1955). Domestic goat and sheep appear early in the Neolithic, followed later by pig (Uerpmann and Frey 1981:148). Uerpmann and Frey attempted (1981:159–62) to correlate the faunal sequences at Ali Tappeh and Ghar-i Kamarband and on that basis they tentatively dated the Mesolithic occupation of Ghar-i Kamarband as from c. 9500 to c. 8600 years ago (c. 10,900–c. 9700 cal. BP), thus reducing Coon’s estimate for the beginning of the Mesolithic at the site by some two millennia. (This large discrepancy is probably due to Coon’s radiocarbon dates having been produced by the solid carbon method which proved to be unreliable.) The sequence at the Hotu rockshelter was better dated (22 radiocarbon dates from 17 levels; Ralph 1955) than that at Ghar-i Kamarband, on the basis of which Coon (1952; 1957:200) established a chronology for two periods of Mesolithic occupation between c. 11,000 and c. 8000 years ago (c. 12,800 cal.–c. 8900 cal. BP), succeeded by the so-called Sub-Neolithic and Software Neolithic periods. He suggested (1957:202) that “ox, sheep, pig, and dog” found in the Sub-Neolithic



the mesolithic and neolithic periods: sites, sequences, and subsistence

horizon were probably domesticated, but because no analysis of the animal remains from Hotu was ever published this remains an unverified speculation. The uncertainties inherent in Coon’s radiocarbon chronologies for Ghar-i Kamarband and Hotu, and the fact that he did not publish full reports of his excavations, has inhibited interpretation of the stratigraphic sequences at these important transitional Mesolithic–Neolithic sites. However, materials from the excavations, together with records of the contexts in which finds were made, have been preserved in the collections of the University of Pennsylvania Museum. This archive offers an opportunity to establish sound chronologies for the two sites by dating samples of organic materials from known contexts, and generally to re-assess the significance of Coon’s discoveries. This task is being undertaken by a postdoctoral research fellow of the Museum, Michael Gregg, who is examining pottery, lithics, and faunal remains from the excavations. By 2009 he had already identified 54 samples of ash, bone, and wood still labeled with information about the contexts from which they were recovered by Coon. Gregg aims to date sufficient samples to establish a high-resolution radiocarbon chronology; to determine whether the “Software” pottery from the earliest ceramic levels resembles pottery found at other Mesolithic rockshelters east of the Caspian (such as the DDC sites, see below) or Neolithic Jeitun-type ceramics; and, when he has assessed the shape, size, and technological attributes of the earliest ceramics from Ghar-i Kamarband and Hotu, to conduct molecular and isotopic analyses of organic residues that may be preserved in them. This ambitious research program can be expected to add much valuable new evidence relevant to the transition from hunting, fishing, and gathering to food production south and east of the Caspian. The second group of Caspian-Mesolithic sites consists of three rockshelters at the foot of the southwestern escarpment of the Bolshoi Balkhan massif east of the present shore of the Caspian near Nebit Dag (Fig. 6.1). The westernmost, Jebel (Djebel), is located in a rock outcrop at the junction of the mountains and the piedmont (Fig. 8.15, color), and there is evidence above and below it of marine terraces formed during previous high stands of the Caspian (this volume, pp. 20 and 35). It was excavated by A. P. Okladnikov between 1947 and 1950 and is regarded in Turkmenistan as the type site for the Caspian Mesolithic. Okladnikov (1956) recognized l0 main layers containing cultural deposits and he divided

57

the sequence, on the basis of a shift in lithic assemblages from a predominantly pebble-tool tradition to the incorporation of geometric microliths, into early and late Mesolithic periods, succeeded in the upper layers (5–1) by Neolithic–early Bronze Age occupation. In the early 1950s the radiocarbon dating technique was in its infancy and was not yet widely applied, but Okladnikov nevertheless submitted a sample of (saksaul) charcoal for trial dating by the liquid-scintillation method. It came from his early Neolithic layer 4 and yielded a radiocarbon age of 6030 ± 240 BP, calibrated to 5235–4715 BCE (Protopopov and Butomo 1959:32; Kohl 1992:157), and re-calibrated for this volume to 5500–4400 BCE (at 95% probability). In 1976 P. M. Dolukhanov collected four more charcoal samples from the rockshelter at depths in the deposits between 0.20 and 0.85 m (the lowest sample consisted of “charred twigs”), which, when calibrated, gave a date range from AD 20 to 5255 BCE (Hurst and Lawn 1984:224–25). The dates were recalibrated for this volume with OxCal 3.10 (Bronk Ramsey 2005), using the atmospheric calibration curve of Reimer et al. (2004). This produced dates (at 95% probability) between AD 250–1050 at 0.20–0.25 m to 4890–5300 BCE at 0.80–0.85 m. The lowest sample was associated with flint blades, an ornamented bone, and a few potsherds that, according to G. F. Korobkova who examined them, were similar to those from Okladnikov’s layer 4 (Pavel Dolukhanov, pers. comm. May 2009). This, together with Okladnikov’s date, indicates that the transition from Mesolithic to Neolithic occupation at Jebel occurred at least 500 years after Jeitun was founded c. 6100 cal. BCE. The other two rockshelters, Dam Dam Cheshme (DDC) 1 and 2 overlook the piedmont zone that fringes the southern escarpment of the Bolshoi Balkhan massif and slopes down to the ancient channel of the Uzboi river which formerly flowed into the Caspian (Fig. 4.4, and see also Fig. 4.3, color). Both sites (Figs. 8.16–8.19, color) were excavated in the 1960s by Markov (1966a, 1966b, 1981), following preliminary investigations by Okladnikov in the 1950s. Markov identified five cultural layers at DDC 1 and nine at DDC 2, and, in the absence of radiocarbon dates, he tried to establish a relative chronology that encompassed both sites. This proved difficult, especially because the stone-tool assemblages were poorly preserved, but he was able to infer, mainly from the stone tools and the pottery, that the sequence at DDC 1 extended from the Upper

58

origins of agriculture in western central asia

Palaeolithic to the early Bronze Age, whereas the somewhat shorter sequence at DDC 2 extended from the early Mesolithic (or possibly the Upper Palaeolithic in layer 9) at the base of the fill to the Bronze Age in the uppermost horizon (layer 2), after which the rockshelter ceased to be occupied. According to Okladnikov (1956:200–202) the earliest (Mesolithic) inhabitants of the Jebel rockshelter depended on hunting and gathering, supplemented in the early Neolithic by fishing and in the late Neolithic and Bronze Age by the herding of domestic sheep and goats and the collection of wild grain (inferred from the presence of grindstones). The main hunted prey, judging from the animal remains found, were goitered gazelle and sheep/goat, although because the bones were very poorly preserved the zoologist who studied them was often unable to differentiate between remains of gazelle and sheep/goat (Tsalkin 1956). A few bones of wild cattle, onager, and a species of wild cat (manul, caracal, or sand cat) were also recovered. Markov provided very little direct evidence of the subsistence economy of the inhabitants of DDC 1 and 2, but he did report (1966a:123) that in the DDC 2 sequence, Tsalkin identified sheep/goat bones in increasing numbers from Mesolithic layers 6 and 5 through the Mesolithic–Neolithic transition at the end of layer 5 and the beginning of layer 4, to the Bronze Age in layer 2. He stated that Tsalkin regarded some of the bones in layer 4 as from “indisputably” domesticated goats and some of those in layer 3 as from “indisputably” domesticated sheep. Tsalkin also identified onager bones in layers 6 and 2, and dog and bird bones in layer 2. Markov concluded (1966b:91; 1981:42), on the very inadequate osteoarchaeological evidence, that by the beginning of the Neolithic the inhabitants of the DDC rockshelters hunted, fished, and probably domesticated first goats and later sheep. He also inferred (1966b:91), from the presence of a grindstone in layer 2, that by the Bronze Age cereals may have been cultivated or their seeds collected, although by then the economy was, he suggested, based mainly on steppe pastoralism (i.e., the seasonally mobile herding of domestic sheep and goats). No archaeobotanical evidence of cereals or other crops having been cultivated was found at any of the three rockshelters, but it must be remembered that techniques for the recovery of charred plant remains were either not known or not applied in the Soviet Union when the sites were excavated.

North of the Bolshoi Balkhan massif there are two more sites with stratified Mesolithic deposits: Kailyu cave, which also contained Neolithic deposits, located close to the Caspian shore on the Krasnovodsk plateau, and Hodja (Khodzha)-su on the eastern coast of the bay of Kara-Bogaz Gol (Masson and Sarianidi 1972:26–29; Okladnikov 1953:31–32). At Kailyu Okladnikov dug through 4 m of mainly Neolithic deposits to a Mesolithic layer underlain by marine pebbles that contained deposits similar to the Mesolithic levels at DDC 2. The Mesolithic inhabitants of both sites are thought to have subsisted mainly by hunting and fishing. At Kailyu large quantities of fish bones and scales, particularly of sturgeon, were recovered, and stone blades and many shell beads were found in human burials on the remains of a nearby marine terrace. The two sites represent the northernmost manifestation of the Caspian Mesolithic. Viewed as a whole, the evidence from the Mesolithic levels at the rockshelter sites of the southeast Caspian region suggests that they functioned as hunting and fishing camps probably occupied seasonally—an interpretation supported by Dolukhanov (1986:124)—and that they provided convenient shelter and easy access to terrestrial, marine, and riverine resources. Furthermore, there is no conclusive evidence that Jebel, DDC 1, and DDC 2 were inhabited year-round and long-term from the Neolithic to the Bronze Age (see Chapters 11 and 12 for further discussion of their role in the Neolithic pattern of settlement and subsistence).

Northern Afghanistan Between the Hindu Kush and the upper valley of the Amudarya there is an extensive area of foothills and plains with evidence of Mesolithic (and Neolithic) occupation. The rockshelter of Kara Kamar southeast of Mazar-i Sharif, discovered by Coon in 1954, contained Mesolithic stone tools and bones of gazelle and wild sheep stratified above Upper Palaeolithic deposits (Coon 1957:225–37; Davis 1978:64–67), and more comprehensive evidence was obtained by Dupree in the early 1960s when he partially excavated four prehistoric sites north of the village of Aq Kupruk in the Balkhab valley south of Mazar-i Sharif: three rockshelters and one open-air site, known respectively as Aq Kupruk I, Ghar-i Mar or Snake Cave; Aq Kupruk II, Ghar-i Asp or Horse Cave; Aq Kupruk IV or Skull



the mesolithic and neolithic periods: sites, sequences, and subsistence

Cave; and Aq Kupruk III on a terrace of the Balkh river (Dupree 1964,1967,1972). Dupree interpreted the deposits he excavated at Aq Kupruk—principally at Ghar-i Mar (AK I) and Ghar-i Asp (AK II)—as representing the Upper Palaeolithic at the base of the sequence, succeeded by what he designated “non-ceramic Neolithic” and “ceramic Neolithic” levels, overlain by Chalcolithic, Iron Age, and Early Islamic deposits. He believed that by c. 10,200 and c. 8600 years ago at AK II and AK I respectively, in the “non-ceramic Neolithic,” the local economy included domesticated sheep and goats (identified as such in a preliminary report by Perkins 1972), and possibly the cultivation of wheat and barley (Dupree 1972:76–77, 80–81). This interpretation was challenged by Vinogradov and Ranov (1985:71), who maintained that the so-called Neolithic economy at Aq Kupruk was based on hunting and gathering without animal husbandry or crop cultivation; and Meadow (1989a) pointed out that the osteoarchaeological evidence was insufficient to determine whether the remains of sheep and goats included any domesticated animals. Unfortunately, too, although plant remains were recovered during the excavations at Aq Kupruk (Dupree 1972:80), no archaeobotanical report was ever published and the question of whether cereals or other crops were cultivated there during the Mesolithic or Neolithic remains unresolved. Farther north, where the Amudarya forms the northern frontier of Afghanistan, there is an east-west zone of takyrs and sand dunes north of Mazar-i Sharif that receives the terminal runoff of streams flowing north out of the Hindu Kush. During three field seasons, in 1969, 1975, and 1976, A. V. Vinogradov (1979) undertook systematic surface surveys across this zone and found over 20,000 stone tools widely distributed in clusters of (larger) sites and (smaller) stations, as well as in many isolated scatters. He dated the sites typologically according to morphological similarities of the tools with stratified lithic assemblages found elsewhere, particularly those at Tutkaul in Tajikistan and at the Caspian rockshelters of Jebel and DDC 1 and 2, and he concluded that although most were Mesolithic in age, they ranged from the Middle Palaeolithic to, possibly, the Late Neolithic (Vinogradov 1979; Kohl 1984:40-41). Vinogradov (1979:60), quoted by Kohl (1984:41), asserted that “At least for the Early Neolithic (sixth millennium B.C.) it is completely impossible to exclude the partial presence of a food-producing economy,

59

particularly one based on stockbreeding.” He also saw similarities between the Early Neolithic stone-tool industries of northwestern Afghanistan and those of the Keltiminar Culture in Khoresmia (the area of the lower Amudarya and the former Akchadarya delta south of the Aral Sea) and the southern Kyzylkum desert. This led him to suggest that in the Early Neolithic the Keltiminar may have been stockbreeders as well as hunters, fishers, and gatherers—a view that Kohl regarded as “reasonable” (1984:41) although there was “no direct evidence to support it” (see below and Chapters 11 and 12 for further discussion of the Keltiminar Culture and Keltiminar-related sites).

The Neolithic The Neolithic period is central to this study of the beginnings of agriculture in western Central Asia. Although it has long been recognized that there is an evolutionary continuum from subsistence systems focused primarily on wild plant and animal resources to those in which crop cultivation and animal management predominate (Harris 1989, 1996), in southern Turkmenistan there is a clear contrast between the preceding Mesolithic hunter-fisher-gatherer economy and the mainly agro-pastoral economy of the Neolithic. The latter is manifested in the sedentary Jeitun Culture of the Kopetdag piedmont. But farther north and east, in the northern Karakum and the area of the former Uzboi river and Sarykamysh depression, as well as eastward across the Amudarya in Uzbekistan as far as the western Kyzylkum desert, there is evidence of a different, more mobile type of Neolithic occupation— the Keltiminar Culture—that lacked agriculture and was dependent on hunting, fishing, and gathering. Dolukhanov (1986) contrasted a northern zone of foragers with a southern zone of farmers in the Neolithic of “west-Central Asia,” and Masson (1996:92) divided the region similarly into two large “super-zones…the agriculturalists and stock-breeders of the south (Djeitun), and the hunters, fishermen and food gatherers of the north (Keltiminar, Hissar, Ferghana).” The Jeitun and Keltiminar cultures overlapped chronologically, but whereas the former was the precursor of the Bronze Age urban civilization of southern Turkmenistan (Hiebert 1994a, 1994b, 2003; Kohl 1981), the latter persisted little changed from the early Neolithic into the Bronze Age long after agriculture and urban life had become established in the south. The Keltimi-

60

origins of agriculture in western central asia

nar Neolithic is described in more detail below, after the settlement pattern and principal sites of the Jeitun Culture have been summarized.

The Jeitun Culture It is in the context of the Jeitun Culture that we encounter the earliest evidence in western Central Asia of settled village life, when people lived in houses built of mudbrick and supported themselves mainly by cereal cultivation and the herding of sheep and goats (caprine pastoralism). Pottery also makes its appearance, and in contrast to the preceding Mesolithic period, the stone-tool assemblages now include numerous microlithic sickle blades and grindstones. This cultural complex has been identified at a series of tell sites and lithic and ceramic surface scatters across the Kopetdag piedmont for a distance from west to east of some 450 km, and there are also several sites with Jeitun-Culture traits in northeastern Iran (see below). The Jeitun Culture was divided by Berdiev (1969), on the basis of changes in stone tools, architecture, and ceramic styles, into three phases referred to as Early, Middle, and Late Jeitun. With the benefit of recently obtained radiocarbon dates, Hiebert was able to demonstrate that the Jeitun Culture spanned some 1400 years, from c. 6100 to c. 4500 cal. BCE, and he introduced the terms KD (Kopetdag) 1, 2, and 3 for three successive periods that are equivalent to Berdiev’s Early, Middle, and Late Jeitun (Hiebert 2002a:28–29). Berdiev associated his three temporal phases with a division of the piedmont into three geographical zones: a central zone (the Geok-tepe region) which contained sites of the Early and Middle Jeitun phases, a western zone with Late Jeitun sites, and an eastern (Meana-Chaacha) zone with sites of the Late Jeitun phase and the following Chalcolithic (Eneolithic) Anau IA phase. Kohl (1984:46) adopted a threefold division of the distribution of Jeitun-Culture sites broadly similar to Berdiev’s, except that he grouped together the western and Geok-tepe zones, defined a central Darreh Gaz (Dargaz) piedmont zone which lacked known Jeitun-Culture sites, and, like Berdiev, recognized an eastern Meana-Chaacha zone. Hiebert (2002a:26) adopted essentially the same regional division as Berdiev, but he refers to the three regions by their Turkmen names: Arkash for the western, Akhal for the central, and Etek for the eastern region—a practice that is followed here. The regional

location and chronology of the sites are summarized in the next three subsections, which are based on numerous Russian publications supplemented by observations in 1997 and 1998 when a member of the British team (Coolidge) and our Turkmen colleague Kurbansakhatov visited most of the Jeitun-Culture sites on the piedmont (Table 6.2 and see Coolidge 2005:30–40 for brief descriptions of the principal sites other than Jeitun).

The Akhal Region of the Central Piedmont This region extends for some 50 km northwestsoutheast where the piedmont forms a narrower strip between the Kopetdag escarpment and the southernmost sands of the Karakum than it does in the western and eastern regions. Several of the rivers that cross the central piedmont issue from gorges in the mountains which give access to valleys within the Kopetdag and farther south to the broad valley of the upper Atrek river and the Iranian plateau. The sites of the Akhal

Table 6.2 Periods and principal sites of the Jeitun Culture in the central, western, and eastern regions of the Kopetdag piedmont and the southern margin of the Karakum desert. Periods

Jeitun-Culture Sites

Regions

KD4 Anau IA 4500 cal. BCE: Neolithic/Chalcolithic boundary KD 3 Late Jeitun

Bami Pessedjik Chagylly Chakmakli Monjukli Gadymi

Western Central Eastern Eastern Eastern Eastern

Bami Chopan Togolok Pessedjik New Nissa Gievdzhik Chagylly Monjukli

Western Central Central Central Central Central Eastern Eastern

Chopan Togolok Jeitun

Central Central Central

5100 cal. BCE KD2 Middle Jeitun

5700 cal. BCE KD1 Early Jeitun

6100 cal. BCE: Mesolithic/Neolithic boundary



the mesolithic and neolithic periods: sites, sequences, and subsistence

61

region include Jeitun itself (the type site of the Jeitun Culture, 28 km north-northwest of Ashgabat), and, in a cluster west and south of Jeitun (Fig. 6.1), Chopan (Berdiev 1972a), Togolok (Berdiev 1964b), Pessedjik (Berdiev 1968b, 1973), Gievdzhik (Korobkova 1975a; not visited during our reconnaissance survey in 1998), Kelyata, Kantar and Kepele (both surface scatters briefly documented by Berdiev [1971b] but which could not be located in 1998 and are not marked on Fig. 6.1), and New Nisa (Berdiev 1965). Another site, Yarti–Gumbez (which we also failed to find in 1998 and is undocumented in the published literature), is located on Kohl’s (1984:47) and Hiebert’s (2002a:26) maps a few kilometers southwest or northwest of Ashgabat respectively, but is omitted from Figure 6.1. Jeitun is attributed entirely to the Early Jeitun phase and Chopan and Togolok are interpreted as having Early Jeitun levels beneath Middle Jeitun levels. Pessedjik, Gievdzhik, Kelyata, Kantar, Kepele, New Nisa, and Yarti-Gumbez are considered representative of the Middle Jeitun phase and also possibly of the Late Jeitun phase.

Fig. 6.1) to the Tedzhen river. Across most of this zone the piedmont is wider than in the central region and the Kopetdag forms a steep, continuous escarpment uninterrupted by through valleys. This part of the piedmont lacks Jeitun-Culture sites today, although it is possible that Neolithic settlements have been buried beneath alluvial-fan deposits. However, at the eastern extremity of the region there is a cluster of four JeitunCulture tell sites between the Meana and Chaacha rivers: Gadymi (Gademi) (Lollekova 1982; not visited in 1998 because of its proximity to the Iranian border and the need for special permits and armed guards to enter the area), Monjukli, Chagylly, and Chakmakli (Fig. 6.1). The stratigraphy of the four sites is representative of the Late Jeitun phase and the following Early Chalcolithic (Early Eneolithic) period. Berdiev suggested (1964b:276) that the eastern (Meana-Chaacha) zone was settled by people from the central (Geokdepe) region, and that the most striking difference in artifacts between the two was a significant decrease in the percentage of decorated ceramics in the eastern compared with the central region.

The Arkash Region of the Western Piedmont

Jeitun-Culture Phases: Changes in Artifact Assemblages

Only three Jeitun-Culture sites are known in this region: Naiza, Bami, and a surface scatter at Bacha Well northwest of Kizyl Arvat (Fig. 6.1). Naiza (which we were unable to find in 1998) and the surface scatter at the Bacha Well are reported by Berdiev and Kohl to be quite small, but Bami (near the present town of Bami) is a large circular mound that rises some 6–7 m above the surrounding plain and is closer to the Kopetdag mountains than any other Jeitun-Culture site on the piedmont. In 1998 we descended into the remains of a 5-m-deep trench that had been excavated in 1960 by A. A. Marushchenko assisted by Berdiev (1963b). It appeared that approximately 2 m of sediment at the bottom of the trench had caved in, obscuring the lowest Neolithic levels, but we did not have sufficient time to reopen and shore up the caved-in trench. Because Bami is the westernmost known Jeitun-Culture site on the piedmont and has quite deep stratigraphy, it is a promising site for future excavation that could throw light on possible links during the Neolithic with related sites farther west in northern Iran and the southeast Caspian region.

The Etek Region of the Eastern Piedmont This region extends from just east of Ashgabat (and the Chalcolithic–Bronze Age site of Anau North;

The Early Jeitun phase represented at Jeitun, lower Chopan, and lower Togolok is marked by a lithic assemblage with numerous sickle blades and occasional geometric microliths, a predominance of painted and unpainted handmade ceramic bowls, terracotta counters and discs (the latter possibly used as spindle whorls), bone scrapers and piercers, and turquoise and shell beads (Kohl 1984:48). The Middle Jeitun phase is identified in the upper levels of Chopan and Togolok, at Bami levels 1, 2, and possibly 3, New Nisa, Pessedjik, and the lower levels of Monjukli and Chagylly. It is considered comparable to the Sialk I complex of the north-central Iranian plateau and is characterized by fewer painted vessels than are found at sites of the Early Jeitun phase, a greater variety of types of ceramic vessels, a decrease in scraping tools, and an increase in denticulate blades (Kohl 1984:49). The Late Jeitun phase is represented at Bami levels 4 and 5, the upper levels of Chagylly, and possibly the middle levels of Monjukli. It is characterized by fewer geometric microliths, an increase in the number of denticulate blades, an absence of bone scrapers, increasingly naturalistic motifs on the pottery, mudbricks (rather than the long cylindrical proto-bricks

62

origins of agriculture in western central asia

used previously, e.g., at Jeitun), stone door sockets, and small pieces of copper (Kohl 1984:48–49).

Evidence of Domestic Animals and Crop Plants There is insufficient archaeobotanical and zooarchaeological evidence of crops and domesticated animals at Jeitun-Culture sites to make comparable statements about changes through the Early, Middle, and Late phases, principally because when the sites were first excavated plant and animal remains were not systematically retrieved. Also, identifications of the remains that were found and reported are often unreliable. Remains of domestic goats, sheep, and sometimes dogs were recovered at Early, Middle, and Late Jeitun sites, including Jeitun itself, but bones of domestic cattle were only reported (by V. I. Tsalkin) at two sites, Chagylly and Chopan, from Middle and Late Jeitun levels (Berdiev 1966:26–27; 1972a:78–79). No remains of domestic pigs, horses, or camels have been reported from Jeitun-Culture sites, but Berdiev (ibid.) did record the presence of onagers (kulan) in the bone assemblage from Chagylly. Very few systematic attempts were made to recover and identify plant remains at Neolithic (or later prehistoric) sites in Turkmenistan prior to our use of the flotation technique at Jeitun from 1989 onward and Miller’s (2003) at Anau North in 1997. Previously, the main source of information on Neolithic crops was a study of the development of irrigation agriculture by Lisitsina (1978). She studied cereal remains and cereal impressions in mudbrick at Jeitun and other Jeitun-Culture sites and reported (1978:92) the presence of wheat and barley at Jeitun, Chopan, and Bami (identified just as Triticum sp. and Hordeum sp.), and of bread and club wheat (identified as T. aestivum L. and T. compactum Host.) and two-row barley (identified as H. distichum L.) at Chagylly. Masson summarized these findings in his monograph on Jeitun (1971:97) and emphasized the continuity of wheat and barley cultivation from the Early Neolithic phase at Jeitun and Chopan, the Middle Jeitun phase at Bami, the Late Jeitun phase at Chagylly, and on through the Chalcolithic and Bronze Age in southern Turkmenistan. Lisitsina’s identifications should not be regarded as reliable, and our investigations have since shown that at Jeitun glume wheat—both einkorn (T. monococcum) and another newly recognized type—was the principal crop, together with naked and hulled (probably six-row) barley (Hordeum vulgare [syn. sativum]); also a few chaff fragments of what appears to be free-threshing wheat (present in three samples) have

been tentatively identified as of T. aestivum/durum type (this volume, pp. 151–53).

Northeastern Iran West of the Kopetdag piedmont in northeastern Iran, pottery and other artifacts resembling JeitunCulture materials have been found at several tell sites. Two of them, Yarim Tepe (Stronach 1972; Crawford 1963) and Tureng Tepe (Wulsin 1932; Deshayes 1963; and many other publications, see this volume, p. 49), are situated on the plain traversed by the Gorgan river between the eastern Elburz mountains and the southeastern coast of the Caspian (Fig. 6.1). Fragments of pottery resembling Jeitun ceramics were recovered from the basal levels at both sites, but at neither site were these levels radiocarbon dated. At Yarim the lowest levels, associated with “the remains of a Neolithic village,” were reached in a small excavation (Trench X) that contained sherds of straw-tempered pottery of Jeitun type painted with curvilinear patterns (Stronach 1972:22 and pers. comm. 2007). At Tureng, groundwater prevented excavation of the basal levels but small fragments of straw-tempered pottery that closely resembled, in their temper and decoration, pottery from Jeitun were found incorporated into mudbricks used to construct later buildings at the site. In a brief comment on Tureng, Sarianidi (1992:113) stated that the pottery fragments from the site correspond to the Late Jeitun phase, but Deshayes (1967:123–25) only compared them to Jeitun as a whole and did not refer to any of the Jeitun-Culture phases. Despite the lack of a secure chronology at both Yarim and Tureng, Sarianidi stated (1992:114) that the two sites “provide indirect evidence that in the sixth millennium BC the Gorgan valley contained scattered early agricultural settlements that were based fully on production,” and Kohl (1984:46) suggested that “More sites must exist [on the Gorgan plain], but they may prove difficult to locate because of the overburden of later remains and burial due to alluviation.” Near the southern margin of the plain there is a third site, Shir-i Shayn (Fig. 6.1), which has produced ceramics that Sarianidi said (1992:115) “closely resemble and possibly predate those in the lower strata of Hissar” (a mainly Bronze Age site on the Iranian plateau near Damghan) and to be “to some extent reminiscent of the pottery of Jeitun,” but only “exploratory digs” were carried out and further excavation would be required



the mesolithic and neolithic periods: sites, sequences, and subsistence

before the site could add significantly to knowledge of the Neolithic in the area. More recently, Iranian archaeologists carrying out surveys on the Gorgan plain and Caspian lowland south of Behshahr have located several sites with evidence of Neolithic occupation, at one of which, Tugh Tepe near Behshahr, sherds resembling Jeitun pottery have been found (this volume, p. 50). Excavation of such sites can be expected to add substantially to knowledge of Neolithic settlement in these lowland areas and throw more light on any connections they may have had with the JeitunCulture sites of the Kopetdag piedmont (discussed in Chapter 11). South of the Gorgan plain on the north Iranian plateau two tell sites provide more, but still inconclusive, evidence of links with the Jeitun Culture. They are part of a cluster of small mounds at the edge of a terminal fan at Sang-i Čakmaq (Sang-i Chakmak) in the village of Bastam 8 km northeast of Shahroud (Fig. 6.1). Two of the mounds, about 200 m apart, were excavated in 1971, 1973, and 1975 by S. Masuda (1972, 1974a, 1974b, 1976, 1977, 1984). The eastern mound is about 6 m high and Masuda’s excavations revealed six occupation strata between the surface and bedrock, all but the uppermost of which contained the remains of mudbrick structures and pottery. In a description of the ceramics of each stratum, H. Kamuro stated (in Masuda 1977) that there were striking similarities between the material from the third stratum and Middle-Jeitun pottery at Bami and Chopan, and between cruder pottery from the sixth (lowest) stratum and sherds from Jeitun itself. Masuda (1976:63) mentioned that sherds from the lower levels were straw-tempered like Jeitun pottery, and that many of the artifacts made of stone, bone, and clay resembled artifacts from Jeitun-Culture sites (1974b:223). He also pointed out (ibid.) that similar materials had been found by Stronach at Yarim Tepe, but (as mentioned above) only sherds resembling Jeitun-type pottery were found in the lowest levels there. The finds from the eastern mound at Sang-i Čakmaq included clay figurines of animals and wooden sickle handles decorated with animal designs. Most of the bone implements from the upper strata were made from cattle bone, whereas many of those found in the lower levels were of deer bone and antler (Masuda 1976:64). The western mound is 3 m high with five strata of mudbrick structures but, unlike the eastern one, was almost devoid of pottery. After his first field season Masuda (1972:2) was inclined to attribute the

63

occupation to an aceramic Neolithic phase, but following further excavation he concluded (1974b:223) that more research was needed to establish how the site related to the regional chronology of Neolithic settlement. In his publications Masuda only refers to three radiocarbon dates and his statements about them are not always consistent. In his first brief report he mentioned (1972:2) two dates—7240 ± 150 and 7270 ± 125 BP—as from the “West Tappeh,” but it is clear from the context that in the text West should have been East. Nor did he state from which level(s) the charcoal samples that were dated came. In his later reports he mentioned a third date—c. 7800 BP (lacking a ± margin of error)—from the western mound, but attributed it in different reports to the “second layer” and the “third level” (1974a:25; 1976:65). When calibrated, the two dates from the eastern mound give almost identical dates (at 95% probability) within the range 6450–5850 cal. BCE, whereas the single date from western mound gives a calibrated range (at 95% probability) of 7100–6350 cal. BCE. Despite the inadequacy of this chronology for the two mounds, the dates do indicate, when compared with the 11 on-site dates we obtained at Jeitun (this volume, pp. 120–23) which range from c. 6300 to c. 5600 cal. BCE, that the western mound at Sang-i Čakmaq was probably occupied at least several centuries earlier than Jeitun. This comparison also demonstrates that Sang-i Čakmaq is at present—before new data from the current investigations in the area by Iranian archaeologists become available1—the earliest dated and westernmost known Neolithic settlement with an artifact assemblage very similar to that of the Jeitun Culture. This conclusion is reinforced by Aurenche’s observation (1985:236) that the rectangular shape, internal platforms, and raised hearths of the buildings at Sang-i Čakmaq resemble those at Jeitun. Masuda regarded the north Iranian plateau as the area of origin of the culture represented at Sang-i Čakmaq (1976:65). However, the almost complete absence of pottery in the western mound led Gupta (1979, II:49–52) to suggest that the site might represent a transitional stage between the southeast Caspian Mesolithic—as revealed in the rockshelters of Ghar-i Kamarband, Hotu, and Ali Tappeh—and the Jeitun Neolithic. Although there are some suggestive similarities in the lithic assemblages and in the presence of small cones of baked clay, this hypothesis appears very tenuous in the absence of more support-

64

origins of agriculture in western central asia

ing evidence from further investigations of Mesolithic and Neolithic sites in both areas. Fortunately, such investigations have recently been initiated by Iranian archaeologists who have conducted regional surveys and exploratory excavations on the Shahroud plain close to Sang-i Čakmaq and farther east in northern Khorassan near Bojnurd, Neyshabour, and Dargaz (Darreh Gaz) (Fig. 6.1). They have so far identified some 30 Neolithic sites, at least 10 of which have yielded sherds that resemble JeitunCulture pottery (this volume, pp. 50–51). It is to be hoped that this new phase of research will lead to the excavation of some of these new sites and the retrieval, identification, and direct (AMS) dating of animal and plant remains from them, in order to gain a better understanding of Neolithic settlement and subsistence in the region. It is clear that the Neolithic sites of the Gorgan plain, at Sang-i Čakmaq and the newly discovered ones near Shahroud and farther east in northern Khorassan, provide evidence of similarities in artifacts and architecture with the Jeitun-Culture settlements of southern Turkmenistan. It therefore seems surprising that no such Neolithic settlements have been discovered in the intermontane valleys and plains of the middle Atrek catchment and especially its northern tributaries, the Sumbar and Chandyr, which offer accessible routes across the mountains to the western and central Kopetdag piedmont (Fig. 6.1). These long, eastwest trending valleys receive more precipitation and have deeper, more fertile soils than the piedmont, and are well suited to rainfed agriculture. The apparent absence of Neolithic sites in the Sumbar and Chandyr valleys cannot be ascribed only to a lack of archaeological prospecting. In the past, several Soviet archaeologists explored the middle Sumbar valley, where they made surface finds of Palaeolithic stone tools (Lyubin 1984:28–31) and also discovered a few settlement sites, notably the Chalcolithic (Eneolithic)–Bronze Age cemetery of Parkhai II (Khlopin 1981, 1989). Then, in 1996, members of our team prospected for early sites in the Sumbar and the Chandyr valleys north of the Iranian frontier, but despite carrying out trial excavations at eight rockshelters and one open site, we found no evidence of Neolithic (or earlier) occupation (this volume, pp. 107–13). Also, until Iranian archaeologists recently began to undertake extensive surveys in northern Khorassan, very few Neolithic sites had been found in the upper Atrek valley or elsewhere in the region, despite the investigations in the late 1970s by

Italian and American archaeologists near Quchan and on the Dargaz plain (this volume, p. 50). The apparent scarcity of Neolithic sites in the intermontane valleys may partly be due to many of them having been buried by more recent alluviation, because the Kopetdag mountains are subject to intense tectonic activity and severe erosion; and other sites may have been destroyed by post-Neolithic agricultural settlement and cultivation. However, more comprehensive geoarchaeological prospecting in the intermontane valleys both north and south of the Iran-Turkmenistan frontier is needed and may reveal the existence of many as yet undiscovered Neolithic, and perhaps Mesolithic, sites in the region, as the recent surveys by Iranian archaeologists suggest.

The Keltiminar Culture The Keltiminar (Kel’teminar) Culture is a less well defined and more heterogeneous entity than the agro-pastoral Jeitun Culture (Kohl 1984:57–64; Brunet 2005; 1999:40–42). The term Keltiminar has been applied to a large number of sites thought to have been occupied by seasonally mobile hunterfisher-gatherers across a broad swathe of northern Turkmenistan and Uzbekistan from the Caspian Sea to the Kyzylkum desert. Sites attributed to the Keltiminar Culture were first identified in 1939 in the ancient Akchadarya delta south of the Aral Sea where the type site, Janbas (Dzhanbas) 4, was excavated by Tolstov, and subsequent surveys led to the discovery of over 60 Keltiminar sites in the area of the ancient delta (this volume, p. 45). Keltiminar or Keltiminar-related sites have also been found in four other main areas: around the ancient Lyavlyakan lake and other depressions in the southern Kyzylkum, along the former northwestern extensions of the lower Zeravshan river, in the area of the Sarykamysh depression and the former channels of the upper Uzboi in the northern Karakum, and near now-dry depressions and watercourses on the Ustyurt plateau west of the Aral Sea (Fig. 6.2) (Brunet 2005; Dolukhanov 1986:125–26; Korobkova 1975b; Masson 1996:96–101; Vinogradov 1968). The Neolithic levels in the rockshelters of the Bolshoi Balkhan massif and the Krasnovodsk plateau (Jebel, Dam Dam Cheshme, and Kailyu), and open-air sites north (e.g., Oyukli) and south (e.g., Joyruk) of the massif (Fig. 4.1), share some features of their material culture



the mesolithic and neolithic periods: sites, sequences, and subsistence

65

6.2 Distribution of Keltiminar, possibly Keltiminar, and other Neolithic sites in the desert and steppe environments of western Central Asia; only sites mentioned in the text are named (site locations from Brunet 2005: Fig. 2, and Dolukhanov 1986: Fig. 2).

with Keltiminar-Culture sites, but Korobkova (1975b) showed that these Caspian sites have a distinctive stone-tool tradition. A few sites in western Kazakhstan between the Aral Sea and the Caspian have also been attributed by some authors to the Keltiminar Culture (Brunet 2005:89–90). Investigation of the Keltiminar Culture and of its relationship with the Jeitun Culture of southern Turkmenistan has been hampered by severe difficulties in establishing a secure chronology. Most of the Keltiminar and Keltiminar-related sites have been deflated by wind erosion and are unstratified. Indeed, Kohl (1984:59) pointed out that cultural levels

and architecture had (by then) only been recorded at nine sites: Oyukli, north of the Bolshoi Balkhan, and, in the Kyzylkum, Janbas 4 and 31, Kavat 7, Lyavlyakan 26, Beshbulak 2 and 3, and Darbazakir 1 and 2 (Fig. 6.2). In his discussion of the chronological problems associated with the dating of Keltiminar sites Kohl also emphasized (1984:61) that in addition to a general lack of well-dated stratigraphy there was a dearth of convincing parallels with better-dated sites in southern Turkmenistan, and that attempts to date the beginning of the Keltiminar Culture had depended on whether ceramic or lithic comparisons were accepted as primary evidence. Soviet scholars dif-

66

origins of agriculture in western central asia

fered in their interpretations of when the Keltiminar Culture began and how long it lasted, as well as over its contemporaneity with the Jeitun Culture and the subsequent Eneolithic (Chalcolithic) and Bronze Age cultures of the Kopetdag piedmont, but a widely accepted view (Korobkova 1975b; Masson and Sarianidi 1972:73) was that Keltiminar and Keltiminar-related sites spanned a long period from the Neolithic into the Bronze Age. Vinogradov (1968, 1981; Vinogradov and Mamedov 1975) assigned the Keltiminar sites to three phases: an early Neolithic Dariasai (Dar’ jasaj) phase in the 6th millennium, a developed Neolithic Janbas (Dzhanbas) phase in the 5th–4th millennia, and a recent Neolithic phase from the second half of the 3rd to the beginning of the 2nd millennia BCE. A general lack of radiocarbon dates from stratigraphically secure contexts means that the duration of the Keltiminar Culture cannot be precisely determined. However, preliminary results from research by the Polish-Uzbek team at the site of Ayakagytma (Fig. 6.2, and this volume, p. 48), together with reexamination of relative inter-site chronology inferred from archaeological and sedimentological evidence, led Brunet (2005:90–95) to suggest a revised threefold chronology: a 7th–6th millennium early Neolithic phase, a 5th–4th millennium later Neolithic phase, and a third phase from the mid 3rd to the beginning of the 2nd millennium BCE interpreted as Eneolithic (rather than Neolithic as Vinogradov had proposed) and which, Brunet argued, reflected contact with Bronze Age cultures of the Central Asian steppes. Brunet also re-examined regional variation among sites attributed to the Keltiminar Culture—the unity of which had previously been questioned by Kohl (1984:63)—and concluded that the lithic and ceramic materials found at several sites north and west of the Aral Sea in Uzbekistan, Kazakhstan, and northern Turkmenistan, previously classified as Keltiminar or Keltiminar-related, were so locally distinctive that they should not be attributed to the Keltiminar Culture (Brunet 2005: Fig. 2 and 95–96). Having summarized the problems of definition and chronology that have complicated the study of Keltiminar sites, we consider next what is known about settlement and subsistence in the Keltiminar-Culture zone. When the overall distribution of Neolithic sites is examined, a strong correlation is evident between site location and the former presence of freshwater lakes, deltas, rivers, and streams. There is abundant evidence in the landscape for the former existence of a much

more extensive and complex hydrological network than exists today in the Kyzylkum and along the lower courses of the Syrdarya, Zeravshan, and Amudarya, a network which included at various times connections between the Amudarya, the Sarykamysh depression, and the upper and lower channels of the Uzboi—a drainage pattern that reflects, at least in part, increased precipitation of the mid-Holocene Climatic Optimum (Brunet 2005:91; Dolukhanov 1986:127; Kohl 1984:61; this volume, pp. 25–26). The few stratified Keltiminar sites are located near former water bodies, and they provide some evidence, in the form of wood remains from such trees as alder, that implies the former presence of riparian tugai vegetation. The remains of wooden structures at some of these sites, such as Janbas 4 and Kavat 7, together with traces of hearths, pits, and postholes (Vinogradov 1981:148–54), have been interpreted as evidence of fully or partly sedentary occupation, but as Brunet (2005:96–97) points out, there is insufficient evidence to determine whether such sites were occupied continuously or repeatedly for shorter periods. Most of the unstratified sites are also located near former streams, rivers, and lakes, and their general shallowness and the lack of evidence of wooden structures suggests that they were occupied seasonally rather than year-round. The close association of settlements and water bodies is reflected in evidence of subsistence activities: remains of fish (principally pike and carp) and freshwater mollusks have been found at two of the excavated sites in the Akchadarya delta, Janbas 4 and Tolstov, where bones of several species of waterfowl were also found (Dolukhanov 1986:126–27; Vinogradov 1981:140–46). If small bones had been systematically retrieved by flotation and/or fine-mesh sieving, remains of fish and waterfowl would probably have been recovered at other sites. The importance of fishing and fowling is also attested by finds of fish hooks, net sinkers, and harpoon barbs (Masson and Sarianidi 1972:73). Mammal bones have been recovered at more sites, in the Akchadarya delta, the Zeravashan catchment, and the area of the Lyavlyakan lake, and Vinogradov (1981:139–41) reported the presence of a wide range of taxa of tugai, steppe, desert, and mountain habitats, including aurochs, wild boar, red and fallow deer, gazelle, onager, saiga, camel, and wild sheep, as well as smaller mammals such as fox, hare, and badger (Dolukhanov 1986:126 –27). Hunting evidently made a major contribution to subsistence,



the mesolithic and neolithic periods: sites, sequences, and subsistence

and the harvesting of grass seeds, roots and tubers, and other plant foods probably did so too, although there is only indirect evidence of this in the form of grindstones, mortars, pestles, and sickle blades. In her studies of use-wear traces on sickle blades, Korobkova (e.g., 1969, 1978) sought to distinguish between tools used to harvest wild and cultivated cereals and inferred that sickle blades found at sites in the lower Zeravshan valley had been used to cut domesticated cereals (Dolukhanov 1986:127; Korobkova 1969:186; 1992:275). However, there is as yet no conclusive evidence, in the form of charred grains and chaff, to support the idea that cereal cultivation was practiced at Keltiminar sites during the Neolithic, and Vinogradov (1981:139) maintained that the sickle blades found at the lower Zeravshan sites were evidence of the harvesting of wild plants. The question of whether stockbreeding was part of the Keltiminar subsistence economy during the Neolithic is also difficult to resolve because of the paucity of well preserved, identified, and analyzed assemblages from stratified sites. As mentioned earlier in this chapter (p. 59), Vinogradov (1979:60) suggested that in the early Neolithic the Keltiminar may have bred livestock as well as hunted, fished, and gathered, although no direct, well-attested osteoarchaeological evidence of domesticated herd animals at Neolithic Keltiminar sites had been published. Now the PolishUzbek team working at Ayakagytma in the southern Kyzylkum (this volume, pp. 48–49) has analyzed bone assemblages recovered there between 1996 and 2003. The bones are very poorly preserved and only 1,151 fragments out of a total of 7,092 could be identified to species or genus, but they came both from the upper stratigraphic units (2 and 4) in the Dzhanbas phase, radiocarbon dated to c. 3000–4000 cal. BCE, and from the lower unit (5) in the early Neolithic Dariasai phase, dated to c. 5500–6000 cal. BCE (Fontugne and Szymczak 2006; Lasota-Moskalewska et al. 2006). The taxa identified include a wide range of herd animals: cattle, ?bison, sheep/goat, horse/onager, camel, water buffalo, gazelle, and deer, as well as pig, dog, bird, fish, and tortoise. The specialists who undertook the analysis classified the cattle, sheep, ?goat, pig, and dog as domesticated, and camel, horse/onager, and water buffalo as possibly having been tamed, if not actually domesticated (Lasota-Moskalewska et al. 2006:208). The relative abundance of the taxa represented at Ayakagytma changes between units 2, 4, and 5 (and within sub-units of 5), but does not reveal an

67

overall trend from wild animals in the earliest levels to domesticates in the most recent levels. Animals regarded as domestic occur in all the units, and Lasota-Moskalewska et al. concluded that in both the earlier and the later occupation phases the economy was “nearly fully based” on domesticated and “most probably tamed” animals, with hunting, fishing, and gathering playing a “minor role” (ibid., p. 215). The main hunted animals are thought to have been gazelle, aurochs, and, less frequently, red and fallow deer. A remarkable feature of the remains is the presence of camel bones in the stratigraphic units of both the earlier and the later phases, with a pronounced increase starting in the middle of sub-unit 5b at c. 5600 cal. BCE, paralleled by a decrease in the representation of cattle and equids, which disappear completely during the final phase of the occupation (unit 2). It is suggested that these trends were mainly due to climatic changes (to greater aridity?) from the late Neolithic into the Bronze Age, and that the camels were probably tamed, if not domesticated. The inference about the status of the camels, together with the conclusion that domestic cattle, sheep, perhaps goat, pig, and dog were all kept and raised, focuses attention on how their domestic status was determined. That question is briefly discussed in the preliminary report, and it is stated (ibid., p. 206) that a particular bone fragment “was described as belonging to a domesticated individual only when diminution of skeleton, and narrowing of a compact bone layer were clearly marked.” It is not apparent whether “diminution” of the skeleton refers here to progressive reduction in bone sizes of a given taxon from the earlier to the later occupation phases, or to contrasts between larger and smaller bone sizes in the same taxon such as might, for example, be taken to indicate the presence of both wild and domestic cattle. If the latter meaning is intended, it is complicated by the possibility of sexual dimorphism, which is pronounced in wild cattle and would need to ruled out before “diminution” could be used as a criterion of domestication (see Russell et al. 2005 for a discussion of sexual dimorphism in relation to cattle remains from the Neolithic site of Çatalhöyük in Anatolia). The use of “a narrowing of a compact bone layer” (in the bone shaft wall?) as a criterion for domestication is also problematic because many variables could affect it, such as the size, sex, and age of the individual animal. The fact that the bone assemblages studied were small, poorly preserved, and contained a high propor-

68

origins of agriculture in western central asia

tion of unidentifiable bone fragments would of course have made it very difficult to use more rigorous morphological, metrical, and other criteria (cf. Meadow 1989b) to determine whether the cattle, sheep/goat, pig, and dog remains found at Ayakagytma were indeed from domesticated animals, but this lack means that at present the question of their status as domestic, tamed, or wild animals cannot be resolved. It is to be hoped that the analyses now underway of bone assemblages excavated in 2005, 2006, and 2007, and perhaps further on-site recovery of animal remains, will clarify this fundamental question. Meanwhile the preliminary analysis of the 1996–2003 bone assemblages raises interesting questions not only concerning domestication but also about possible relationships in the Neolithic with the agro-pastoral Jeitun-Culture settlements of the Kopetdag piedmont (discussed in Chapter 11). At present there is no evidence in the form of, for example, charred cereal remains or organic residues on stone blades (“sickle gloss”) to suggest that the Neolithic occupants of the Ayakagytma site engaged in any form of cultivation. However, some plant remains were recovered by flotation in 2008 (Karol Szymczak, pers. comm. 2009) and their analysis may provide such evidence, if they can be securely identified and dated by the AMS radiocarbon method. If evidence of grains or chaff is found and directly dated, the provisional conclusion that cereal cultivation was not part of the Keltiminar subsistence system will need to be revised. So, although the Keltiminar appear to have depended on a combination of hunting, fishing, and gathering with, perhaps, some form of livestock herding, it is also possible that grain crops were grown on a small scale by simple floodwater or décrue methods. By the early 2nd millennium BCE, during the so-called Steppe Bronze Age, there is evidence of domestic livestock (bones of goats, sheep, cattle, and asses) and also of domestic cereals (grain impressions of barley and wheat) at late Keltiminar sites, in the context of the Zaman-baba complex of the lower Zeravshan valley (Masson and Sarianidi 1972:125–28). These finds provide some evidence of a transition to a settled, agricultural way of life, which Masson and Sarianidi believe occurred as a result of close contacts with the long-established agro-pastoral communities of southern Turkmenistan. Okladnikov, too, assumed that the shift from hunting, fishing, and gathering to cultivation and stockbreeding took place among the Keltiminar groups of the Amudarya, Uzboi, and other

riverine areas during the Bronze Age as a result of contact with the Jeitun-Culture farmers of the southern zone (1956:216).

The Early Chalcolithic (Eneolithic) Anau IA Phase Although the main chronological focus in this volume is on the Neolithic period and its Mesolithic antecedents, a summary account is also needed (especially as a prelude to Part III) of the transition from the Jeitun Culture to the earliest part of the succeeding Chalcolithic (Eneolithic) period: Anau IA. The term Anau IA derives from Pumpelly’s excavation in 1904 of part of the north mound at Anau, when Schmidt (1908:130–32) described a distinctively tempered and decorated type of pottery found in the lowest layers and on that basis recognized the earliest (IA) phase of occupation at the site. In the 1930s Marushchenko found ceramics of Anau IA type at several other sites on the piedmont, but because he did not publish these finds the IA phase was long thought to be a development unique to Anau North, or even “an accidental phenomenon” (Khlopin 1963:21; Kohl 1984:65). After the Second World War Marushchenko conducted excavations at the eastern-piedmont site of Monjukli where he discovered Anau IA materials directly overlying Late Jeitun levels and interpreted this as evidence for a local evolution from Late Jeitun to Anau IA (Berdiev 1972c:11). Subsequently, Anau IA materials were found at a series of sites across the piedmont, with the greatest concentration in the central Akhal region, and also at two sites (Yam and XA6) in the upper Atrek valley in northeastern Iran (Berdiev 1974; Kohl 1984:65–67; Ricciardi 1980:53–57). Considerable controversy has surrounded the definition, duration, and subdivision of the Anau IA phase partly because, although most of the sites are located on the central piedmont, large-scale excavations have only been carried out at two of the eastern sites, Monjukli and Chakmakli, and only at Monjukli do Anau IA materials directly overlie Late-Jeitun levels. Also, until recently no IA levels had been radiocarbondated at any site. Despite the lack of absolute dates, a consensus eventually emerged that the phase probably dated to the late 6th and the beginning of the 5th millennia BCE (Kohl 1984:67–91). But it was not until Hiebert, Kurbansakhatov, and their colleagues undertook



the mesolithic and neolithic periods: sites, sequences, and subsistence

new excavations at Anau North in 1997 and obtained 19 radiocarbon dates for specific contexts in layers 3–20 that a well-founded chronology for Anau IA was established. Four dates were obtained for the Anau IA phase (represented by the lowest layers, 19 and 20), which (uncalibrated) broadly support the previous relative dating of the initial occupation of the site to the late 6th millennium BCE, and when calibrated, give a composite age of c. 4300 cal. BCE (Hiebert 2003:55–56). The cultural distinctiveness of the Anau IA phase is not limited to its ceramics. Other attributes that distinguish it from the Jeitun Culture include the occurrence of copper tools, non-local semi-precious stones such as lapis lazuli, stone hoe blades, and (as shown by the excavations at Monjukli and Chakmakli) larger, more internally complex settlements with multi-room buildings, storage areas, courtyards, and a division of the settlement into two parts divided by a street (Hiebert 2002a:33; Kohl 1984:67–71). There is less evidence in the Anau IA phase of innovation in subsistence activities than in material culture. Thus the economy continued to be based on a combination of cereal cultivation, livestock herding, and hunting. However, there is evidence from the IA layers at Anau North that bread wheat (Triticum aestivum s.l.) and six-row barley (Hordeum vulgare ssp. vulgare) were now being cultivated, and the relatively high water requirements of these plump-grain types of wheat and barley (compared with einkorn and two-row barley) suggest that small-scale irrigation may have been carried out on the piedmont near Anau North at this time (Miller 1999:15–17; 2003:137–38). This possibility receives some indirect support from the distribution of the Anau IA sites on the piedmont which, as Hiebert has pointed out (2002a:32), are all located between 200 and 500 m above sea level. In that zone, small-scale irrigation dependent on gravity flow from piedmont streams upstream of their terminal deltas at the edge of the Karakum would have been feasible, and this technologically simple form of irrigation may have preceded the more extensive and elaborate irrigation systems that developed in the Bronze Age in the large inland deltas of the Tedzhen and Murghab rivers. The pastoral component of the subsistence economy continued to be based mainly on domestic sheep and goats. It has been suggested that large cattle bones found in the early Chalcolithic IA layers at Anau North represent domesticated animals (Moore, Ermolova, and Forsten 2003:155–56), but Duerst (1908:365–69, and see this volume, p. 79) believed that they derived

69

from wild cattle and that the smaller bovid bones found in the upper layers were from domestic cattle. Remains of domestic pigs also occur in the upper (later Chalcolithic) layers at Anau North (Moore et al.: ibid.), as well as at the Chalcolithic site of Dashlidji in the Geoksyur oasis in the ancient delta of the Tedzhen river (Masson and Sarianidi 1972:61); and it is possible that pigs were raised at piedmont settlements in the Late Neolithic. The question of whether the material culture of the Anau IA phase represents a local development on the piedmont, as some Russian archaeologists have suggested, or was the result of external influences such as the movement of people from Iran into Turkmenistan cannot be resolved without more extensive excavation and accurate dating of piedmont sites that contain both Late-Jeitun and Anau IA levels. The resolution of this question is not directly relevant here, but it is highly desirable because it is critical to understanding the transition from small Neolithic village communities to larger, more complex Chalcolithic and Bronze Age urban settlements in the region.

note 1. As this book was going to press, Dr. Marjan Mashkour of the CNRS Archaeozoology Research Unit in Paris (UMR 7209 CNRS, France) kindly provided information on a French– Japanese initiative between the University of Tsukuba (Akira Tsuneki) and the CNRS (Marjan Mashkour and Jean-Denis Vigne) to re-date Sang-i Čakmaq. As a result, animal bones excavated by Masuda in the 1970s have now been directly dated by the AMS radiocarbon method. Two very similar dates show that the western mound was occupied by 7997 ± 42 BP (7059–6767 cal. BCE at 2 sigma or 95% confidence), whereas the new dates for the eastern mound range from 7182 ± 42 BP (6110–5985 cal. BCE at 2 sigma) and 6444 ± 42 BP (5479–5338 cal. BCE at 2 sigma). These results do not differ greatly from Masuda’s three dates for the western and eastern mounds as calibrated for this volume (see p. 63). They confirm that the western mound was occupied substantially earlier than Jeitun in contrast to the eastern mound, the earliest new date for which corresponds closely to the date of c. 6100 cal. BCE for the occupation of Jeitun. In 2009 a new stratigraphic investigation at Sang-i Čakmaq, combined with systematic recovery of organic remains, was undertaken by Kourosh Roustaei of the Iranian Center for Archaeological Research, Tehran. The samples obtained are now being analyzed and AMS dated, and the results promise to add significantly to our knowledge of the establishment of agriculture in northeastern Iran. I (D.R.H.) am most grateful to Dr. Mashkour for allowing me to refer here to the new radiocarbon dates and to the renewed research program being undertaken at this important Neolithic site.

part iii

Neolithic Crop Plants and Domestic Animals

7

Areas of Origin of the Crops and Domestic Animals

A

fundamental part of this enquiry into the transition from foraging to farming in western Central Asia is to try to determine whether agriculture began there independently, or whether the new way of life was introduced—wholly or in part—from elsewhere. One approach to this question is to ask whether any of the crops and domestic animals identified in the organic remains recovered at Jeitun and other early sites could have been domesticated locally from wild progenitors native to the region. To answer that question requires not only knowledge of the present physical environment and of environmental changes during the Pleistocene and Holocene (outlined in Chapters 1 and 2), but also familiarity with genetic studies of the species concerned and with archaeological evidence of their presence in the Neolithic. In recent years cytological, molecular, and archaeological investigation of the evolution and domestication of many crops and domestic animals has advanced rapidly, with the result that earlier assumptions about their ancestry and areas of origin are being radically re-examined (Brown et al. 2009; Bruford, Bradley, and Luikart 2003; Burger, Chapman, and Burke 2008; Doebley, Grant, and Smith 2006; Fuller 2007; Jones and Brown 2000; Zeder et al. 2006). In particular, analyses of modern and ancient DNA of several cereals and domestic herd animals that were an integral part of Neolithic agriculture in Eurasia— barley, wheat, goats, sheep, cattle, and pigs—suggest that they may have been domesticated several times in different areas (see below). There is also a growing realization that shortterm climatic and biotic changes during the Quaternary were more frequent and extreme, and more

profoundly affected the ranges of wild plants and animals, than previously thought. As a result, it has become more difficult to make sound inferences, from the present geographical distributions of their wild progenitors and other closely related (congeneric) wild species, about where particular crops and domestic animals originated. In this chapter, the principal crops and domestic animals that were present in Turkmenistan and adjacent areas during the Neolithic are reviewed, in order to assess whether they are likely to have been domesticated in the region or introduced, as domesticates, from elsewhere.

The Crops and Their Wild Progenitors Very little reliably identified and dated archaeobotanical evidence of domesticated plants from prehistoric sites has so far been obtained in western Central Asia. No systematic retrieval of macroscopic plant remains by flotation and/or fine sieving was undertaken in Turkmenistan until we carried it out at Jeitun between 1989 and 1997, and Naomi Miller (1999, 2003) did so at Anau North. Lisitsina (1978) had previously studied cereal grains and impressions in pottery from four Neolithic sites in southern Turkmenistan, but her identifications were insecure. Consequently only the crops we have definitively identified and dated from our excavations at Jeitun are discussed in this section. The cereal remains from Jeitun comprise charred grains and chaff of hulled and naked varieties of (probably six-row) cultivated barley (Hordeum

74

origins of agriculture in western central asia

vulgare L. [syn. sativum Pers.]) and two types of cultivated glume (hulled) wheat: einkorn (Triticum monococcum L., both the one-grain and the two-grain form) and another (emmer-like) type of uncertain origin (Jones, Valamoti, and Charles 2000), as well as three tentatively identified specimens of free-threshing wheat (of T. aestivum/durum type). Contrary to our earlier reports (Charles and Hillman 1992:83–94; Harris et al. 1993:332; Harris, Gosden, and Charles 1996:438), none of the remains examined can be definitely identified as cultivated emmer wheat (T. dicoccum Schübl.). Nor has any evidence of legume crops been found. These results add greatly to the earlier, less well documented reports of barley and wheat found at Jeitun, Chopan, Bami, and Chagylly (Lisitsina 1978:92; this volume, p. 62); and they raise the question of whether the wild progenitors of barley and the wheats found at Jeitun are likely to have been present in Turkmenistan in the Late Pleistocene and Early Holocene and might, therefore, have been domesticated locally.

Barley Wild barley (Hordeum spontaneum L. subsp. spontaneum C. Koch, or, according to von Bothmer et al. (1995), H. vulgare L. subsp. spontaneum C. Koch) is a predominantly self-pollinating annual diploid (chromosome number 2n=14) and the progenitor of domestic barley, with which it readily hybridizes and is fully fertile. In wild barley two rows of grain normally develop but under domestication six-row forms have been selected, as have naked-kernel forms (Harlan, de Wet, and Price 1973:317; Pourkheirandish and Komatsuda 2007). Wild barley is better adapted to aridity and more tolerant of poor, especially calcareous soils than wild wheats and it occurs more widely around the Southwest Asian Fertile Crescent, as Harlan and Zohary showed in 1966 when they published their wellknown distribution maps of the “primary habitats” of wild wheats and barley (updated in Zohary and Hopf 2000:37, 45, 66). Subsequently Zohary (1989:30) commented that “more isolated populations, usually of weedy forms” extend east across Central Asia (including Turkmenistan) as far as the western Himalayas and Tibet (see also Shao 1981; Witcombe 1978), and he later elaborated that comment, stating that “in north-east Iran, Central Asia, and Afghanistan wild spontaneum barley is. . .sporadic in its distribution; it

rarely builds large stands and seems to be completely restricted in most localities to segetal habitats, ruins, or to sites which have been drastically churned by human activity” (Zohary and Hopf 2000:67). When visiting the Badghyz Natural Reserve in southeastern Turkmenistan in 1992, two members of our research team (Hillman and Harris) observed apparently natural stands of wild barley growing in association with pistachio trees (Fig. 1.9, color), although it is possible that they are feral populations descended from domesticated barley. This possibility was also envisaged by Jan Valkoun (pers. comm. 1997) who noted that at least some of the “wild” barley populations of northeastern Iran and Turkmenistan have an upright growth habit, and synchronized tillering and seed maturation, implying that they are likely to be descended from weedy forms introduced with already domesticated barley during the spread of Neolithic agriculture. However, these observations do not preclude the possibility that genuinely wild barley grew in the intermontane valleys and on the piedmont slopes of the region when agricultural settlements were first established there early in the Neolithic, especially as the climate was warmer and wetter then, during the mid-Holocene Climatic Optimum, than it is today (this volume, pp. 25–26). Hillman had observed wild barley growing as a component of wild almond and pistachio woodlands in Southwest Asia, and he speculated (1996:188–89) that it might have survived the Late Pleistocene in refugia farther east and subsequently spread from them (as well as from mountain refugia in the northern Levant). Similar woodlands exist in southern Turkmenistan today (this volume, pp. 9–10), and it is quite possible that wild barley grew there in such habitats early in the Holocene. Genetic data derived from accessions of wild barley from west and east of the Zagros mountains collected at a wide range of locations between the Mediterranean and the Pamirs in Central Asia have shown that western and eastern barleys are part of an interbreeding population that shares many haplotypes, but that they often differ in the predominant haplotype present, which implies an ancient divergence between west and east and that all the eastern barleys are very unlikely to be feral rather than genuinely wild (Morrell, Lundy, and Clegg 2003, Morrell et al. 2005). This conclusion does not deny the possibility that some of the weedy forms referred to by Zohary and Valkoun may have derived from domesticated barley, but it supports the palaeoenvironmental inference that



areas of origin of the crops and domestic animals

wild barley may have been present in southern Turkmenistan in the Early Holocene. If so, its local domestication cannot be excluded on phytogeographical grounds. Other genetic studies of barley have concluded that it was domesticated only once or very few times. Thus Neale et al. (1988) investigated types of chloroplast DNA in a large collection of wild and cultivated barleys from Israel and Iran and demonstrated the existence of three lineages in wild barley, only one of which was (with two exceptions) represented in the cultivated barley, implying “a single or very few events of domestication” (Zohary 1999:137). This conclusion was supported by Badr et al. (2000) who analyzed AFLP (amplified-fragment length polymorphism) data from a large sample of wild and cultivated barleys and concluded that barley was domesticated only once, in southeastern Turkey. However, although the authors claimed that this result “closes the longlasting debate on the origin of barley” (ibid., p. 507), it should not be regarded as conclusive because AFLP data analyzed by cluster analysis (neighbor-joining trees) is not a genuinely phylogenetic method of inferring monophyletic origins for crops (or domestic animals), as has been demonstrated in simulations of cereal domestication (Allaby and Brown 2003, 2004; Allaby, Fuller, and Brown 2008; Allaby, Brown, and Fuller 2009; and see Salamini et al. 2004 and Honne and Heun 2009). The possibility that barley was domesticated more than once received further support from Morrell and Clegg’s (2007) comparison of the haplotype composition of wild and domesticated forms from a wide range of locations across southwestern, central, and eastern Asia. This reinforced their earlier evidence of an ancient divergence between western and eastern wild barleys, and showed that all accessions of cultivated (landrace) barleys from Asia east of the Zagros have substantial identity with eastern wild barleys and have been subject to later admixture from imported western landraces. This conclusion is supported by a phylogeographic analysis by Saisho and Purugganan (2007) of haplotype clusters of five genes in a large sample of barley cultivars from East and South Asia (including Iran, Afghanistan, and Pakistan), as well as from the Fertile Crescent, North Africa, and Europe, which revealed an area of genetic discontinuity between the landraces westward from the Fertile Crescent into Europe and the eastern ones from Iran to China, Korea, and Japan. Both studies imply that

75

barley was domesticated, independently of its domestication in the Fertile Crescent, somewhere east of the Zagros before western cultivars were introduced to central and eastern Asia; but the antiquity of such a putative domestication cannot be demonstrated conclusively from the genetic data alone. Morrell and Clegg (2007) interpreted archaeological finds of barley in Neolithic contexts at Jeitun, and at Mehrgarh in Baluchistan, as suggestive of areas where a second domestication of barley could have occurred. But at present there is insufficient archaeobotanical evidence to resolve the question of whether barley was domesticated locally in Turkmenistan and/ or in Pakistan or whether it was initially introduced to those sites as an already domesticated crop. For example, at Mehrgarh there is abundant evidence of barley in the form of plant impressions in mudbrick in the earliest (aceramic Neolithic Period I) levels, which, despite inconsistency in the radiocarbon chronology, probably date to between c. 7000 (or perhaps earlier) and c. 6000 cal. BCE (Jarrige 2000:278–83, 2007–08:151; Jarrige, Jarrige, and Quivron 2006). The barley consists mainly of the domesticated six-row hulled and naked forms but there is also some “wild” two-row barley (Costantini 2007–08:168–69; 1983:29– 31; Fuller 2006:22; Possehl 1999:459). The presence of both forms could be interpreted as evidence of initial local domestication, but it does not necessarily support that hypothesis because, as Pourkheirandish and Komatsuda (2007) have shown, six-row spikes and naked kernels (caryopses) were selected after wild two-row barley was domesticated, and furthermore, the apparently wild barley, which typically occurs around Mehrgarh in disturbed habitats, may have reached the region from farther west as a weed of cereal cultivation. Remains of barley recovered from Southwest Asian Neolithic sites show that its domestication was underway in the Levant and possibly also in the Zagros region by the end of the 9th millennium cal. BCE and that it spread as a domesticated crop throughout the Fertile Crescent during the Pre-Pottery Neolithic period1 (Charles 2007; Garrard 1999:77–79; Willcox 2005:535–38). As the initial domestication of barley took place more than two millennia before it was cultivated at Jeitun, there was ample time for it to be dispersed across northern Iran (together with einkorn wheat; see below) as one of the two founder crops of Neolithic cereal cultivation at the JeitunCulture sites—a circumstance that can be interpreted

76

origins of agriculture in western central asia

as evidence against an independent domestication of barley in Turkmenistan. However, lack at present of securely identified and well-dated remains of barley (and wheat) from Neolithic sites across northern Iran, from east of the Zagros mountains to the Kopetdag piedmont, means that the question of whether it was introduced to Turkmenistan as a crop from the Fertile Crescent, or locally domesticated, remains open. There is a similar lack of archaeobotanical evidence of barley (and wheat) at sites across Central Asia east of Turkmenistan, which makes it also impossible, at present, to do more than speculate about the route(s) by which these cereals first reached China. They were present by c. 2500 cal. BCE in northwestern China where seeds of barley and wheat have been found at the site of Xishanping near Tianshui in Gansu and AMS radiocarbon dated to c. 2600 cal. BCE (X. Li et al. 2007). Barley has also been found at Fengtai in the upper Huang-ho valley and wheat at Donghuishan in western Gansu, and both were being cultivated in the middle and lower Huang-ho valley by c. 2000 cal. BCE in the Late Longshan period (F. Li 1989; S. Li 2002:180; Zhao Zhijun, pers. comm. 2006). Remains of barley and wheat have also been found in South Korea and dated to c. 1000 cal. BCE (Crawford and Lee 2003).

The Glume Wheats Two types of glume wheat dominate the archaeobotanical samples recovered at Jeitun. Initially, c. 90% of the wheat remains were believed to consist of domestic einkorn (Harris et al. 1993:332), but further investigation has shown that some of the material closely resembles a type of glume wheat, morphologically distinct from einkorn, that has been reported from Neolithic and Bronze Age sites in Greece, Turkey, Hungary, Poland, Austria, and Germany (Jones, Valamoti, and Charles 2000; Köhler-Schneider 2003). Unlike einkorn, which is a diploid, the other glume wheat is probably a tetraploid and it may derive from wild Timopheev’s wheat (Triticum araraticum Jakubz. [syn. T. timopheevi Zhuk. subsp. araraticum]) or possibly from wild or cultivated emmer wheat (T. dicoccoides Körn and T. dicoccum). Wild T. dicoccoides occurs today throughout the Fertile Crescent, with a concentration in the central and southern Levant, whereas T. araraticum extends around the northern and eastern sectors of the crescent from southeastern Turkey to the southern

Zagros, with outlying populations in the Caucasus (Zohary and Hopf 2000:45), but neither species has been recorded east of the Caspian Sea. A domesticated form of Timopheev’s wheat (T. timopheevi Zhuk.) was cultivated until recently in Georgia, where it may have been domesticated from T. araraticum (Zohary and Hopf 2000:58), and it is also possible that the glume wheat now recognized in the Jeitun assemblage represents an extension far to the east of a formerly much wider Neolithic distribution of T. timopheevi cultivation. However, none of these possibilities implies that it might have been domesticated locally in Turkmenistan, and its presence there in the Neolithic is almost certainly the result of its introduction as a domestic cereal from somewhere west of the Caspian. In contrast to the uncertain origin of the newly recognized glume wheat from Jeitun, the ancestry of domestic einkorn is well understood. Its progenitor is wild einkorn (Triticum monococcum L. subsp. boeoticum [Bois.] A. et D. Löve [syn. T. boeoticum Boiss. emend. Schiem.]), a self-pollinating annual diploid grass (chromosome number 2n=14), with which domestic einkorn is fully interfertile. A second wild diploid wheat (T. urartu Tuman.) exists in Southwest Asia. It closely resembles two-grain wild einkorn but is intersterile with wild and domestic einkorn and is not implicated in the latter’s domestication (Waines and Barnhart 1992; Zohary and Hopf 2000:20–22, 36–38), although Heun, Haldorsen, and Vollan (2008) have speculated that finds of “einkorn” from Neolithic contexts at sites in the middle Euphrates valley may instead be domesticated T. urartu. Like other wheats, and unlike barley, wild einkorn generally avoids calcareous soils and occurs mainly on acid (often basaltic) soils. As a result, its distribution in Southwest Asia is less extensive and more patchy than wild barley. It is more cold- and droughttolerant than other wild wheats and its range extends in a broad band from the northeastern margins of the Fertile Crescent to western Turkey, with more isolated populations in the Caucasus, the central Levant, and in Greece and the Balkans where they are weedy forms that probably spread with domestic einkorn (Nesbitt 2001:46–48; Valkoun 2001; Valkoun, Waines, and Konopka 1998:295; Zohary and Hopf 2000:37). Heun et al. (1997, 2008) undertook AFLP analysis of DNA obtained from a large number of samples of wild and cultivated einkorn from sites between the Balkans and the Caucasus (mainly from within the northern Fertile Crescent), and concluded that einkorn was monophyletic and had probably been



areas of origin of the crops and domestic animals

domesticated in the Karacada ğ mountains a short distance east of those archaeological sites. A modification of this hypothesis was suggested by Kilian et al. (2007) who used AFPL and haplotype data to investigate einkorn domestication and proposed a “dispersedspecific model” of multiple “domestication events” in a larger area of southeastern Turkey. However, as was pointed out in the preceding section, the use of AFLP data to infer crop origins by constructing neighborjoining trees can produce misleading results. Nor did Heun et al. (1997) distinguish taxonomically between one-grain and two-grain forms of einkorn, which may have been separately domesticated. One- and two-grain einkorn have in the past been classified as separate subspecies of Triticum boeoticum: subsp. aegilopoides (Link) Schiem. (one-grain) and subsp. thaoudar Reuter ex Hausskn. (two-grain), or even as different species: Triticum aegilopoides (Link) Bal. and T. thaoudar Reuter. Whereas the one-grain form occurs mainly in the northwestern part of the range of wild einkorn in western Anatolia and around the Aegean, where it occupies what are probably secondary habitats, the more robust two-grain form is found throughout the northern Fertile Crescent, with intermediate forms also present, especially in central Anatolia, the Caucasus, and northwestern Iran (Kreuz and Boenke 2002:234; Zohary and Hopf 2000:36). The earliest records of domesticated one-grain einkorn come from sites in the northern Levant and southeastern Anatolia such as Tell el Kharkh, Nevalı Çori, and Cafer Höyük during the 9th millennium cal. BCE, in the Early PPNB (de Moulins 1997:52, 68; Pasternak 1998; Willcox 2005:536–37; 2007:24). In contrast, definite evidence of the domestic twograin form—which could derive either from Triticum boeoticum subsp. thaudar or possibly from T. urartu (Fuller 2007:8)—is restricted to two much later sites in northern Syria: Tell Sabi Abyad I, where the grain was found in a late Neolithic/early Chalcolithic context (van Zeist 1999), and Kosak Shamali, where the grain was recovered from burnt storage structures of the Chalcolithic (Ubaid) period (Willcox 2003). There is no evidence that two-grain einkorn continued to be cultivated at Sabi Abyad in later periods, and van Zeist suggested that it may have been domesticated in southeastern Anatolia, where the wild form is abundant, and introduced as a crop into northern Syria, where it was only cultivated for a short period and then abandoned. It (or domesticated T. urartu rather than two-grain einkorn according to Heun, Haldorsen, and

77

Vollan [2008:449]) also spread west into eastern and central Europe, where there is evidence of two-grain (as well as one-grain) einkorn in the Neolithic and the Bronze Age (Köhler-Schneider 2003:109; Kreuz and Boenke 2002:238), but it did not long survive as a cultivated crop in Southwest Asia or Europe. As it has been identified at Jeitun, it evidently also spread east into western Central Asia. These advances in understanding the beginnings of einkorn cultivation suggest that the one- and two-grain forms may have been separately domesticated in Southwest Asia, but much more genetic and archaeobotanical evidence is needed before the question of where and how frequently einkorn was domesticated in the region can be resolved. It is interesting that both forms were present at early Neolithic Jeitun, but highly improbable that either was independently domesticated in Central Asia. Wild einkorn has never been recorded east of the Caspian Sea, and it is unlikely that its range extended into Turkmenistan even during the climatic fluctuations of the Late Pleistocene and Early Holocene. Had it done so, it would probably have survived in isolated refugia in the intermontane valleys and piedmont of the Kopetdag, as a component of woodland and/or steppe vegetation. In its main area of distribution in Southwest Asia it commonly occurs as a component of oak woodlands; the extinction of oak in the Kopetdag during the Quaternary period (this volume, p. 19), and the absence of native species of Quercus in Turkmenistan today (Nikitin and Geldykhanov 1988:140), may partly account for the absence of wild einkorn east of the Caspian. The possibility that wild einkorn might have been present in southern Turkmenistan in the Early Holocene and have been domesticated locally cannot be totally excluded, but the weight of evidence is heavily against that hypothesis. Like the other type of glume wheat found at Jeitun, the presence there of einkorn in both the two-grain and the one-grain form is much more likely to be the result of their introduction, as domestic crops, from Southwest Asia. So too is the presence of the free-threshing Triticum aestivum/ durum wheat that has been tentatively identified in three samples recovered there (this volume, p. 153); and see Zohary and Hopf 2000:42–59 for a discussion of the distribution and origin in Southwest Asia of the tetraploid durum-type and hexaploid aestivum-type wheats). Finally, it is worth noting that the case for einkorn having been introduced during the Neolithic

78

origins of agriculture in western central asia

from Southwest Asia as an already domesticated crop is paralleled by evidence from Mehrgarh in Baluchistan. Domestic einkorn has been reported as present there (with the barley referred to above) in the earliest levels (Costantini 1983:31; Jarrige 2007–08:142), and it is most unlikely to have been domesticated locally because the site lies far to the southeast of the known range of wild einkorn in the northern Fertile Crescent and Turkey.

The Domestic Animals and Their Wild Progenitors The only domesticated animals whose remains have been recovered during excavations at Jeitun are the dog and two ungulate species, goat and sheep. The other domesticated ungulates associated with prehistoric agricultural and pastoral economies in Central Asia—pigs, cattle, horses, and camels—appear later in southern Turkmenistan than the Early Jeitun phase, although all of them were present as wild animals in the Late Pleistocene/Early Holocene. Evidence of them in the archaeological record, and the question of whether any of them may have been domesticated locally, is first reviewed, before the three domestic animals recorded at Jeitun are discussed.

Pigs Remains of pig have been recovered at Jeitun, but, on the grounds of their relative rarity and the large size of the molar teeth found, they are thought to derive from the wild boar (Sus scrofa), which probably inhabited thickets of tamarisk and reeds along the Kara Su (this volume, p. 177; Kasparov 1992:57; Shevchenko 1960:466, 470–73). Farther west in the Iranian Caspian lowland, bones of wild boar have been reported from Mesolithic levels at the sites of Ali Tappeh and Ghar-i Kamarband (Uerpmann and Frey 1981:148, 151). No definite evidence of domestic pigs has been reported from any Jeitun-Culture site, but by the Early Chalcolithic period domestic pigs—much smaller than modern wild boar—were evidently being raised at Anau North (during the IB1 and IB2 phases, c. 4000–3500 cal. BCE: Moore, Ermolova, and Forsten 2003:155–56). This accords with Duerst’s conclusion (1908:355–58), from his study of over 100 pig bones excavated at Anau North in 1904, that their generally

small size and the absence of pig bones in the lowest layers pointed to their being domesticated, although not at Anau itself; he thought that pigs might have been domesticated elsewhere in “Turkestan” or imported as tamed animals “from Iran or India” (ibid., p. 357). Much farther east at the Neolithic site of Ayakagytma in the Kyzylkum desert, a few pig bones have been recovered from both the older (c. 6000–5500 cal. BCE) and the more recent (c. 4000–3000 cal. BCE) levels (Lasota-Moskalewska et al. 2006:208, and this volume, p. 67). The bones are thought to be from domestic pigs, but in view of their poor preservation and the very small sample, this should be regarded as highly speculative. Thus, there is no conclusive evidence at present of domestic pigs in the region during the early Neolithic. Those reported from the IB and later levels at Anau and at other Late Neolithic/Chalcolithic sites in southern Turkmenistan, such as Dashlidji (Masson and Sarianidi 1972:61), are probably descendants of already domesticated pigs introduced from farther west, in and beyond Iran, although local interbreeding with wild boar may have occurred subsequently. All domestic pigs are descended from Sus scrofa, with the possible exception of the Sulawesi warty pig which according to Groves (2007:27–29) may have been domesticated from Sus celebensis. Local domestication of wild boar in western Central Asia in the Late Pleistocene/Early Holocene cannot be excluded on zoogeographical grounds, but without more definite evidence of domestic pigs at Mesolithic or early Neolithic sites in the region that hypothesis lacks support. This contrasts with early zooarchaeological evidence of pig domestication in the northern Levant and eastern Anatolia. There pigs appear to have been under some degree of human management by the Late/ Final PPNB1 at the sites of Çayönü (Ervynck et al. 2001; Hongo and Meadow 2000; Hongo et al. 2002:154–57) and Hallan Çemi (Redding 2005:43–44; Rosenberg et al. 1998), and perhaps already morphometrically domesticated at Hayaz, Halula, and Gürcütepe (Peters et al. 1999:41; Peters, von den Driesch, and Helmer 2005:113–14). By early in the Pottery Neolithic period (after c. 6500 cal. BCE) domestic pigs were present at many more sites around the Fertile Crescent, at least a millennium before there is evidence of them in western Central Asia, possibly in the early Neolithic at Ayakagytma and more definitely in the Chalcolithic at Anau North. Recent genetic research on the phylogeography of the genus Sus does not point to western Central Asia



areas of origin of the crops and domestic animals

as a probable area of pig domestication. Analyses of mitochondrial DNA (mt DNA) and nuclear genes of wild boar and domestic pig populations in Asia and Europe suggest that pigs were domesticated independently in eastern and western Eurasia, with the probability that several domestications from distinct wild lineages of Sus scrofa occurred in both regions (Albarella, Dobney, and Rowley-Conwy 2006:218–19; Giuffra et al. 2000; Kijas and Andersson 2001; Larson et al. 2005, 2007a, 2007b). But to confirm and elaborate this pattern, more extensive sampling of wild and domestic pig populations in Eurasia, including Central Asia, is needed, together with systematic examination and dating of existing zooarchaeological evidence (cf., for China, Flad, Yuan, and Li 2007:168–69, 192; Yuan and Flad 2002), and, where feasible, analyses of DNA from ancient specimens (Larson et al. 2007b).

Cattle No remains of domestic (or wild) cattle have been found at Jeitun. Nor have domestic cattle been reported from other Jeitun-Culture sites, with the exceptions of Chagylly in the eastern region and Chopan in the central region of the Kopetdag piedmont where they were found in Middle- and Late-Jeitun levels (Berdiev 1966:26–27). In the subsequent Early Chalcolithic period they are relatively abundant (remains of at least 22 individuals) at the small site of Dashlidji in the Geoksyur oasis, part of the ancient delta of the Tedzhen river (Masson and Sarianidi 1972:58–61), which may imply that domestic cattle played an increasing economic role as settlements based on irrigation agriculture developed in the oasis through the Chalcolithic. There is also evidence of cattle breeding at the large piedmont sites of Ilgynly-depe by the Middle Chalcolithic period and at Altyn-depe by the Early Bronze Age (Kasparov 1994:145–47). Bones identified as domestic cattle, and a smaller number probably from wild cattle (aurochsen), have recently been found at the Neolithic site of Ayakagytma in the Kyzylkum. They are more numerous in the small animal-bone assemblage so far analyzed than any of the other taxa identified (except camels), and more abundant in the earlier than the later phase of occupation (Lasota-Moskalewska et al. 2006:208). It is suggested that the wild cattle were regularly hunted only during the earliest occupation of the site (subunit 5c, c. 6000–5700 cal. BCE) and that some of them

79

may have been locally domesticated (an hypothesis that awaits further investigation through analysis of larger bone assemblages more recently excavated at the site). The presence of cattle identified as domestic diminishes rapidly from c. 5600 cal. BCE and ceases during the final phase of occupation, possibly as a result of climatic changes starting in the late Neolithic and continuing to the beginning of the Bronze Age (Lasota-Moskalewska et al. 2006:208–15, and see below, p. 81). The aurochs (Bos primigenius) is the wild progenitor of all domestic cattle. Three subspecies are recognized: Eurasian B. p. primigenius, South Asian B. p. namadicus, and North African B. p. africanus (syn. opisthonomous). Domestic cattle are conventionally divided into breeds with and without humps. The former (indicine or zebu) breeds are thought to descend from the South Asian subspecies, which is often referred to as Bos indicus, and the latter (taurine) breeds from the Eurasian subspecies, often referred to as Bos taurus (Clutton-Brock 1999:27, 84–85; Grigson 1985). Aurochsen are now extinct in the wild, but in the Late Pleistocene/Early Holocene they ranged across temperate and tropical Eurasia and North Africa where they browsed and grazed in forests, woodlands, and open-canopy shrub vegetation. Many aurochs bones were found in the Mesolithic levels at Ali Tappeh and Ghar-i Kamarband in the Iranian Caspian lowland (Coon 1951:44; McBurney 1968:396–97; Uerpmann and Frey 1981:148, 151), and although Coon speculated (ibid., p. 50) that the Bos bones he recovered from Neolithic levels at the latter site were from domesticated oxen, Uerpmann and Frey pointed out (ibid., p. 147) that the bones could not be distinguished from those of aurochs in the Mesolithic levels. Duerst concluded (1908:359–69), from his study of bovid bones from Anau North, that the cattle remains from the Early Chalcolithic IA stratum (now dated to c. 4500–4000 cal. BCE) derived from what he termed the Asiatic form of Bos primigenius, which he referred to as Bos namadicus (without implying that they were humped cattle of zebu type). He also showed that the smaller and more numerous bovid bones found in the upper layers of the site were from domestic cattle which, he hypothesized, might first have been domesticated locally as a long-horned breed and then underwent a reduction in size, or, alternatively, that the smaller animals “may have reached Anau with. . .other imported domestic animals” (ibid., p. 369). More recent examination of new bone assemblages from Anau North ex-

80

origins of agriculture in western central asia

cavated between 1978 and 1982 produced many bones of large long-horned cattle from the IA stratum and later layers, as well as evidence of some smaller cattle, all of which are interpreted as domestic (Ermolova 1985; Moore, Ermolova, and Forsten 2003:155–56). The prevalence of large cattle at Anau is attributed by Moore et al. to diversity in the domesticated population rather than to the occurrence of wild cattle, and there is no evidence of a progressive shift through time from larger to smaller bovid metatacarpal bones, such as would suggest local domestication. It is possible that the assemblage represents a mixed population of introduced (domestic) cattle and (hunted) wild aurochs, and, regardless of that possibility, it is probable that domestic cattle became an increasingly important component of the local agro-pastoral economy from early in the Chalcolithic period. If we turn from the very limited evidence of domestic cattle in Neolithic/Chalcolithic southern Turkmenistan to consider the results of recent genetic analyses of modern breeds in Eurasia and Africa, the likelihood of cattle having been independently domesticated in western Central Asia diminishes. Analyses of mtDNA, nuclear microsatellites, and Y-chromosome markers from present-day breeds in Africa, India, and Europe have revealed a very ancient divergence (c. 200,000 years ago) between taurine and indicine cattle, and significant clustering of variance into three continental groups. This implies that distinct ancestral populations of aurochsen were independently domesticated in two, possibly three regions: humped cattle in South Asia, humpless cattle in Southwest Asia (the principal source of European cattle), and perhaps also taurine cattle in North Africa where, however, there is evidence of admixture of taurine and indicine components (Bradley and Magee 2006; Bradley et al. 1996; Hanotte et al. 2002; Kim et al. 2003; Loftus et al. 1994; MacHugh et al. 1997; Troy et al. 2001). The genetic data presently available do not support a hypothesis of independent cattle domestication in Central Asia, but the possibility that aurochsen were domesticated in the eastern part of their range in the arid interior of Asia should not be completely excluded. In the Southwest Asian Fertile Crescent shifts toward increasing numbers of small cattle in bone assemblages at Neolithic sites, as well as evidence of earlier killing of the animals, suggest that the domestication process was underway by c. 8000 cal. BCE in the Middle PPNB and that by the Late/Final PPNB1 domestic cattle were present at many sites in the

northern Levant, including Çayönü, Hayaz, Gürcütepe, Halula, Tell es Sinn, Bouqras, and Ras Shamra (Helmer et al. 1998; Hongo et al. 2002:160–62; Öksüz 2000; Peters et al. 1999; Saña Seguί 2000). It is even possible that cattle were undergoing domestication by the Early PPNB at other sites in the northern Levant such as Dja’de (Helmer et al. 2005:92). The fact that domestic cattle are present at Middle and Late/Final PPNB sites in the northern Levant more than a millennium before they appear at Neolithic and Chalcolithic sites in southern Turkmenistan, coupled with the genetic evidence that identifies Southwest rather than Central Asia as a probable center of aurochs domestication, suggests that domestic cattle spread to Turkmenistan in the latter part of the Neolithic from the Fertile Crescent. If so, their introduction probably took place across northern Iran, but at present the absence of securely identified and well-dated finds of domestic cattle at Neolithic sites in northern Iran means that this supposition must remain tentative. The lack of any definite evidence that aurochsen were independently domesticated in western Central Asia contrasts with the situation south of the Hindu Kush and Afghan plateau at the site of Mehrgarh in Baluchistan. Analysis by Meadow (1993:304–13) of Bos remains from the aceramic Period I and ceramic Period II levels at Mehrgarh demonstrated a gradual increase between c. 7000 and c. 5000 cal. BCE in the proportion of cattle in the bone assemblages, and a decrease in their size. He interpreted this as evidence of a change from predominantly wild to predominantly domestic cattle of zebu type and thus of local domestication of “what is today called Bos indicus” (Meadow 1993:310), i.e., the South Asian subspecies Bos primigenius namadicus. Cattle of zebu type may also have been domesticated farther east in South Asia (Fuller 2006: 30), a possibility strongly reinforced by recent analyses of mtDNA diversity in zebu cattle that have identified two distinct haplogroups in the Indian subcontinent (Baig et al. 2005; Magee, Mannen, and Bradley 2007; Chen et al. 2010).

Equids and Camels No remains of wild or domestic equids or camels have been found at Jeitun. Shevchenko (1960:465) tentatively identified onager or kulan (Equus hemionus kulan, this volume, p. 14) from one fragment of



areas of origin of the crops and domestic animals

bone, but subsequent osteoarchaeological research at the site has not yielded any onager bones (Kasparov 1992:73; this volume, p. 176). Nor have remains of onager been found at any other Jeitun-Culture sites with the exception of Chagylly where they are represented by a few bones that increase in number from the Middle- to the Late-Jeitun levels (Berdiev 1966:27; Kasparov 1994:148). Onagers appear in Chalcolithic bone assemblages at the eastern piedmont sites of Ilgynly-depe and Altyn-depe, and Kasparov (1994:148) suggested that their earlier scarcity in the Neolithic may have been due to more humid climatic conditions in the 6th millennium BCE to which they were less well adapted than to the more arid climate of the Chalcolithic period when they were a major hunted prey. He implied too that this may account for the scarcity of onager bones in the Mesolithic and Neolithic levels at the Jebel and Dam Dam Cheshme rockshelters in the Bolshoi Balkhan massif reported by Tsalkin (1956:220). Onager bones were also present only in small numbers at the Mesolithic sites of Ghar-i Kamarband and Ali Tappeh in the Iranian Caspian lowland (Uerpmann and Frey 1981:148, 151). Equid and camel bones are represented in the animal-bone assemblage from Ayakagytma in the Kyzylkum in both the older and the more recent levels. Whether the equid remains are from horse or onager (“half-ass”), or both, could not be determined because of the poor preservation of the bones. They are most abundant in the oldest stratigraphic unit (5c, c. 6000–5700 cal. BCE), diminish thereafter, and are not represented in the final phase of occupation (unit 2). This decline parallels that of cattle (referred to above) and may be related to climatic changes. LasotaMoskalewska et al. (2006:215) infer that the horses (or half-asses) were “if not fully domesticated, then at least tamed.” However, given the very poor preservation of the remains and the relatively small number of equid bones identified (179), the inference seems very speculative. An alternative hypothesis that the equids were hunted appears more plausible. Camels comprise the largest taxon in the Ayakagytma animal-bone assemblage (490 identified bones) and they are present in all the stratigraphic units. In contrast to the cattle and equids, camel bones increase markedly from the middle of unit 5b at c. 5600 cal. BCE to over 60% of all animal remains by the end of the early Neolithic Dariasai phase and, after the settlement hiatus from c. 5400–4000 cal. BCE, to over

81

75% in units 4 and 2 (4000–3000 cal. BCE). It is suggested that the camels were “most probably if not fully domesticated” or “at least tamed,” and that the great increase in their number by the end of the occupation, when they had “completely supplanted cattle and horse,” was probably due to climatic changes (ibid., p. 215, and see below). In his study of equid and camel bones from Anau North on the Kopetdag piedmont, Duerst (1908:384– 99, 401–31) attributed the former to the domestic horse (Equus caballus) and the latter tentatively to the domestic Bactrian camel (Camelus bactrianus). Whereas horse bones were abundant and present in all the layers of the site, only two camel bones were found. They occurred in the highest layers, and Duerst assumed that the (probably Bactrian) camel was introduced late as a domestic animal “from Bactriana or the Iranian plateau” (ibid., p. 384). His identification of domestic horse at Anau was later disputed by several authors, most recently by Forsten (2000) who examined some of Duerst’s original specimens and concluded that the bones and teeth were from onager rather than horse (Moore et al. 2003:157–59). Since Duerst’s time, a few more camel bones have been recovered from excavations at Anau and identified by Ermolova (1985:86) as from the Bactrian rather than the dromedary, although it is unclear whether the remains represent hunted animals or isolated finds of domesticated camels (Moore et al. 2003:157). The paucity of archaeologically recovered equid and camel bones from Neolithic sites in western Central Asia may in part be due to the general lack of zooarchaeological research. But it may also imply that wild onagers, horses, and camels were not abundant or commonly hunted at that time, perhaps as a result of the onset of wetter and warmer conditions around 6000 cal. BCE at the beginning of the Climatic Optimum when populations of these wild herbivores, especially onagers and camels which are highly adapted to desert environments, probably declined. It is possible that changes in the relative abundance of cattle, equid, and particularly camel remains between c. 6000 and 3000 cal. BCE at Ayakagytma reflects oscillations between relatively warm, wet conditions during the Climatic Optimum and cooler, drier phases such as the one that occurred between c. 4800 and 4400 cal. BCE (this volume, p. 26), but without more detailed local palaeoenvironmental evidence for the mid Holocene than is available at present that speculation cannot be tested. In Turkmenistan, bones of

82

origins of agriculture in western central asia

what are believed to be wild horses do not appear in archaeological contexts until the Middle Chalcolithic, for example at sites in the Geoksyur oasis (Masson and Sarianidi 1972:69); and the earliest reliable evidence of domestic horses dates to the end of the 3rd millennium BCE (Ermolova 1983). West of Turkmenistan on the north Iranian central plateau, three sites on the Qazvin plain, Zaghe, Ghabristan, and Sagzabad, provide evidence of the exploitation of (mainly wild) equids from the Late Neolithic to the Iron Age (Mashkour 2002, 2003). Remains of onager dominate the equid-bone assemblages at all three sites, but another (now extinct) wild equid, Equus hydruntinus (Clutton-Brock 1999:115), which inhabited Southwest Asia and southern Europe from Palaeolithic to Neolithic times, has also been identified in the assemblages, thus extending its known prehistoric range farther east and its survival into the Iron Age. Some of the equid bones from the Neolithic and Chalcolithic levels in the Qazvin sites have “caballine” features which suggest that horses might possibly have been domesticated locally on the Iranian plateau. The domestic ass or donkey (Equus asinus) is also present by the Chalcolithic period, probably having been introduced from the Fertile Crescent. Far to the east in northern China, remains of wild horses have been found at 28 Palaeolithic sites, but there is very little evidence of horse bones at Chinese Neolithic sites, none of it incontestably of domestic horses, and it is not until c. 1400 BCE in the Late Shang dynasty that definitely domesticated horses are present in the middle and lower Huang-ho valley, probably introduced from the northwest out of Central Asia (Flad, Yuan, and Li 2007:169, 194; Yuan and Flad 2006). The loci of the initial domestications of horses and camels remain uncertain. Analyses of mtDNA sequences from samples of living and fossil horses have revealed a high degree of genetic variation in domestic breeds which could imply that wild horses were separately domesticated many times across an extensive area, but it also possible that initial domestication was more restricted in time and space, and that wild horses were later incorporated into domestic herds as the practice of horse breeding spread (Jansen et al. 2002; Levine 2006; Lister et al. 1998; Vilà et al. 2001, Vilà, Leonard, and Beja-Pereira 2006). In any case, it seems certain that horses were domesticated in the Eurasian steppe zone where the wild progenitor (Equus ferus) ranged in the Late

Pleistocene/Early Holocene. There is archaeological evidence that horses were being exploited intensively for food and other purposes at such sites as Dereivka in the Ukraine and Botai in Kazakhstan by c. 3500 cal. BCE, but whether the remains represent wild or domestic horses has been uncertain, largely because it is difficult to establish robust zooarchaeological criteria for distinguishing between wild, captive, tamed, and domesticated horses (Clutton-Brock 1992:54–55; Levine 1999a, 1999b; Olsen 2003, 2006a, 2006b). Now a fresh approach to the problem based on three independent lines of evidence—metrical analysis of metacarpal bones, examination of damage to mouth skeletal tissues caused by the use of bridles, and organic-residue analysis of potsherds indicating that both horse meat and milk were processed—has provided strong evidence that some at least of the Botai horses were closely managed, milked, and possibly ridden (Outram et al. 2009). In the Late Pleistocene/Early Holocene the range of wild horses probably extended south of the steppe zone into the deserts of Central Asia, where they would have overlapped with the range of the onager which is better adapted to arid conditions (Uerpmann 1987:34). However, horse populations would have been smaller there, especially in the desert lowlands, than farther north in the center of their range. In Central Asia they may have been more numerous and more intensively hunted in montane than in lowland desert habitats: for example, remains of wild horses dominate the bone assemblage in the Upper Palaeolithic horizons at the hunting-camp site of Shugnou in eastern Tajikistan (Davis and Ranov 1999:191; Vishnyatsky 1999:93–94; this volume, p. 54), and there are (inadequately documented) reports of Equus remains in the levels assigned to the Upper Palaeolithic at the Kara Kamar and Aq Kupruk II rockshelters in northern Afghanistan (Coon 1957:230; Dupree 1972:77). Little is definitely known about the ancestry and domestication of the two-humped and one-humped camels, the Bactrian and the dromedary. Small herds of wild camels with two humps that may represent the progenitor of the domestic Bactrian survive in the Gobi desert. Named Camelus ferus, they are often assumed to be a remnant population of the ancestral wild camel, but it has been suggested that they might be feral descendants of domestic camels (CluttonBrock 1999:156; Mason 1984:108). C. ferus probably inhabited most of Central Asia, including the Kara-



areas of origin of the crops and domestic animals

kum and Kyzylkum deserts, in the Late Pleistocene/ Early Holocene, and Mason states that its former distribution extended into southern Russia, Iran, and Afghanistan. Bulliet (1975:48–49) suggested that the Bactrian camel may have been domesticated on the borders of Iran and Turkmenistan several centuries prior to 2500 BCE. The earliest direct evidence of domesticated Bactrians (apart from the inconclusively identified remains from Anau in Turkmenistan and the as yet insufficiently documented camel bones from Ayakagytma in Uzbekistan, see above) comes from the large 3rdmillennium BCE site of Shahr-i Sokhta in east-central Iran, where remains were found of five camel bones, pieces of camel dung, and fibers of what was thought to be camel hair in fragments of woven cloth (Bökönyi and Bartosiewicz 2000:117; Compagnoni and Tosi 1978; Salvatori and Tosi 2006). Also, in the 3rd-millennium BCE levels of Sialk III near Esfahan, a representation of what may be a two-humped camel was found on a pottery sherd (Zeuner 1963:359). In Turkmenistan, clay models of 4-wheeled carts pulled by camels (and less frequently by horses) have been found at Early and Middle Bronze Age sites on the eastern piedmont, for example at Altyn-depe (Masson and Sarianidi 1972:109, 120, Plate 36), indicating that by c. 2000 BCE both had been incorporated as draft animals into systems of agricultural production. Also, by the end of the 3rd millennium, representations of domesticated Bactrian camels appear in the archaeological record of Margiana (the oasis settlements of the Murghab delta) in the form of terracotta figurines and an intricately carved steatite amulet from the site of Togolok 21 (Hiebert 1994b:378–79; Moore et al. 1994:425). The dromedary (Camelus dromedarius) is an important domestic animal in western Central Asia today, but there is no evidence that it was present in prehistoric times. Finer-limbed and faster than the Bactrian, it is well adapted to the hot deserts of Southwest Asia and North Africa, and although its area of origin remains uncertain, it was probably first domesticated in the Arabian peninsula (Bulliet 1975:42–48; Mason 1984:109; Uerpmann and Uerpmann 2002:258; Zeuner 1963:340–44). Dromedary bones have been found at several prehistoric sites in southeastern Arabia and analyzed by Uerpmann and Uerpmann (2002), who concluded that the earlier Neolithic and Bronze Age finds came from wild camels that were probably hunted, whereas remains of domestic camels did not appear until the Iron Age (at

83

the site of Tell Abraq, c. 900–800 BCE). The domestic dromedary subsequently spread across Southwest Asia, used mainly as a pack animal and in warfare, and may only have reached Central Asia in the first millennium AD, where it, and hybrids between it and the Bactrian, gradually displaced the latter as the preferred pack animal on the Silk Route to China (Bulliet 1975:168–71; Köhler-Rollefson 1996:287–88). It was also integrated into the agro-pastoral economy, and is still bred and milked in the Karakum desert. We can now turn from the domesticated ungulates whose presence in southern Turkmenistan postdates the Early Jeitun phase to the two herd animals, goat and sheep, whose remains have been recovered at Jeitun. Goat and sheep (caprine) bones are the most abundant class of animal remains at Jeitun (this volume, pp. 175–77; Kasparov 1992:51), and they are well represented at other Jeitun-Culture sites. Both played a major role in Neolithic subsistence, and the question of whether they may have been domesticated locally is critically important for our understanding of the origins of the agro-pastoral economy of the Jeitun Culture.

Goats Many living species of wild goats have been recognized taxonomically, but only three (interfertile) groups in the genus Capra are implicated in the ancestry of domestic goats—the bezoar or pasang (C. aegagrus), the markhor (C. falconeri), and the ibex (C. ibex). The bezoar (including subspecies) is recognized as the main progenitor, and the markhor and ibex may have contributed, after the domestication of the bezoar, to the ancestry of certain domestic breeds in, respectively, northwestern South Asia and northeastern Africa (Clutton-Brock 1999:76–78; Harris 1962; Manceau et al. 1999; Mannen, Nagata, and Tsuji 2001; Takada et al. 1997). All species of Capra are well adapted to rocky high-mountain habitats, and the bezoar is the most widely distributed of them, with a range in recent centuries that extended from Turkey to western Central Asia (Fig. 1.16). Although only small populations of the bezoar survive in Turkmenistan today (this volume, pp. 12–13), it was formerly more widespread there (and in northern Iran and northern Afghanistan), so the possibility that goats were domesticated locally cannot be excluded on zoogeographical grounds.

84

origins of agriculture in western central asia

There is no doubt that wild bezoar goats were hunted in western Central Asia prior to and during the Neolithic, but their presence has only been documented at a few sites. This may partly be due to the difficulty of distinguishing between the remains of wild and domestic goats. The customary morphological criteria for recognizing domestication are reductions in the size of the limb bones and the appearance of shorter, more rounded and sometimes twisted horn cores (Clutton-Brock 1999:78–79); and analysis of age and sex profiles can, when the evidence is adequate, demonstrate human management of herds preceding full domestication (Zeder 2001). Such criteria have not been systematically applied in the examination of goat bones from sites in the region and it is therefore difficult to determine the accuracy of reports of the presence of domestic and/or wild goats. This applies particularly to their reported presence in Mesolithic and Neolithic levels at the sites of Jebel and Dam Dam Cheshme 2 (DDC 2) in the Bolshoi Balkhan massif (Tsalkin 1956:221; Markov 1966a:123), Ghar-i Kamarband in the southern Caspian lowland (Coon 1951:44–50; Uerpmann and Frey 1981:148), and Aq Kupruk I and II in northern Afghanistan (Perkins 1972). These excavations were carried out before modern methods of osteoarchaeological analysis began to be applied to bone assemblages, so the interpretations of the data contained in the published reports must be treated with caution. In particular, the suggestion that goats were locally domesticated at Ghar-i Kamarband during the Mesolithic (Coon 1951:49–50) should be regarded with great skepticism, as should Perkins’ claim that both wild bezoar and domestic goats were present at the Aq Kupruk sites (Meadow 1989a:27–28). The goat and sheep bones excavated by Okladnikov at the Jebel rockshelter were examined by the zoologist V. I. Tsalkin who concluded that, because they were so poorly preserved, it was very difficult to judge whether they represented wild or domestic animals, although he tentatively concluded (1956:221), on the basis of the small sizes of the bones, that some of the specimens from levels 3 and 4 were from domestic goats. He also studied the bones from the DDC 2 rockshelter, and according to the excavator (Markov 1966a:123), he stated that some of the caprine bones from layer 4 were “indisputably” from domestic goats (and some of those from layer 3 were “indisputably” from domestic sheep). Although no radiocarbon dates were obtained for the archaeological sequence

at the site, Markov inferred that layer 4 was contemporaneous with the occupation of Jeitun and that it represented the beginning of the Neolithic. Assuming Tsalkin’s identifications were sound, this correlation, if correct, suggests that the presence of domestic goats in layer 4 may be the result of contact with goat (and sheep) herders of the Early Jeitun Culture rather than of local domestication of wild bezoar goats. However, Masson and Sarianidi (1972:170) disagreed with Markov’s correlation, believing instead that the (more recent) layer 3 was synchronous with Jeitun, and therefore that domestic goats were present at DDC 2 prior to the occupation of Jeitun and could have been domesticated there. But even if their correlation were correct, already domesticated goats could have reached the Bolshoi Balkhan massif from farther west earlier than the foundation of Jeitun. Without a radiocarbon chronology for the sequence at DDC 2, and more research on bone assemblages from other early sites in the region, these uncertainties cannot be resolved (and see this volume, pp. 217–18, for further discussion of the possible relations between the JeitunCulture and Bolshoi Balkhan sites). A more critical analysis of goat remains from Jeitun itself was undertaken by Kasparov (1992), using Tsalkin’s bone-size criteria for discriminating between wild and domestic caprines. On this basis, Kasparov (1992:51) concluded that some of the goat bones were from the wild bezoar, although most of them were the remains of domestic goats. If correct, this implies that the inhabitants hunted wild goats, probably in the Kopetdag and/or Bolshoi Balkhan mountains, as well as pasturing domestic herds closer to Jeitun. A few caprine bones of Neolithic age have also been excavated at Ayakagytma in the Kyzylkum, but their poor preservation and small number (19 fragments) prevented discrimination between goat and sheep. All but one of the bones came from the early Neolithic Dariasai phase (unit 5) dated to c. 6000–5500 cal. BCE, but none could be definitively identified as Capra. They were interpreted as remains of domesticated animals, except for one bone (in 5b) thought to be from a wild sheep (Lasota-Moskalewska et al. 2006:208–9). At the later site of Anau North on the Kopetdag piedmont small numbers of wild-goat bones have been identified in the Chalcolithic IB and IIA strata, and there too remains of domestic goats are much more abundant (Moore et al. 2003:155). Without more extensive and detailed osteoarchaeological evidence



areas of origin of the crops and domestic animals

from sites in western Central Asia, the question of whether goats may have been domesticated locally cannot be resolved, but consideration of wider archaeological and genetic evidence suggests otherwise. The earliest published zooarchaeological evidence for probable goat domestication comes at present (2008) from two PPNB sites in the northern and eastern Fertile Crescent: Nevalı Çori in the foothills of the southern Taurus mountains in southeastern Turkey and Ganj Dareh in the central Zagros mountains in western Iran, both within the range of the wild progenitor (the bezoar). The later PPNB site of Ali Kosh in the western foothills of the Zagros range south of Ganj Dareh provides early evidence for domestic goats at the margins of the range of the bezoar. At Nevalı Çori, an assemblage of goat (and sheep) bones, characterized by more females than males, a high proportion of immature animals, and a trend toward smaller individuals, has been recovered from levels that date to the Early PPNB,1 suggesting that goats and sheep may have been kept and bred in southeastern Turkey by that time (Peters et al. 1999:35–40). The assemblage of goat bones recovered at Ganj Dareh in the late 1960s and early 1970s has been re-analyzed and dated to c. 7900 cal. BCE by Zeder and Hesse (2000). They constructed sex-specific age profiles which showed that young male goats were selectively slaughtered and they concluded that the animals, although morphologically wild, were managed rather than hunted. They suggested that the Ganj Dareh assemblage provided evidence of initial human management of goats and a transition from hunting to herding, an interpretation originally proposed by Hesse (1978), with which Hole (1996) and Legge (1996) concurred. In contrast, the goat remains from Ali Kosh, which are chronologically at least 500 years more recent than those at Ganj Dareh (Zeder and Hesse 2000:2257), represent, as Flannery originally argued, a morphologically domesticated population characterized by changes in horn shape (Hole, Flannery, and Neely 1969:270–78). By the Late PPNB there is widespread evidence for domesticated goats (and sheep) at sites in the northern Fertile Crescent such as Çayönü, Hayaz, Gritille, and Gürcütepe, as indicated by both the diminution in size and the increased representation of female animals evident in the bone assemblages (Hongo et al. 2002:157–60; 2005; Peters, von den Driesch, and Helmer 2005:97–98). By the beginning of the Final PPNB domestic goats and sheep are present throughout the Fertile Crescent, integrated into the now established agro-pastoral economy (Harris 2002a:70–78).

85

The evidence that goats were domesticated in southeastern Turkey and western Iran and were being kept at sites around the Fertile Crescent by 7000 cal. BCE, long before they are present at Pottery Neolithic sites in southern Turkmenistan some 1000 km east of the Zagros, shows that there was ample time for domestic goats (and sheep) to have spread eastward across northern Iran during the millennium or more that separates the PPNB sites of the Fertile Crescent from the PN sites on the Kopetdag piedmont. But to reach a more definite answer to the question of whether goats are likely to have been independently domesticated in western Central Asia, we need to consider the results of recent genetic research on their origins and spread. Several research groups have investigated mtDNA in domestic and wild goat populations across much of Eurasia and northern Africa. Luikart et al. (2001) sampled 88 breeds of domestic goat, focusing on traditional “unimproved” breeds in remote rural areas. Their sampling included 406 domestic goats (none from Turkmenistan) and 14 wild goats of which 2 bezoars and 1 markhor were from Turkmenistan. The results revealed three highly divergent maternal lineages, the largest of which (A) was widespread in Southwest Asia, Europe, and Africa, whereas the other two much smaller lineages (B and C) were detected respectively in southern and eastern Asia and in central and eastern Europe. The authors inferred that the three lineages had expanded at different periods, and they suggested that A corresponded to the initial domestication of goats in the Fertile Crescent and their outward expansion from about 10,000 years ago, and that lineages C and B expanded about 6000 and 2000 years ago and represented secondary expansions long after the initial dispersal of domestic goats by people. Subsequently, the existence of three more lineages was recognized, now labeled D, F, and G and referred to as mitochondrial haplogroups by Naderi et al. (2007:2). Further research by Luikart et al. (2006) on mtDNA, Y-chromosome, and microsatellite markers for domestic goats and wild species of Capra supported the established view that the bezoar was the principal progenitor of domestic goats and reinforced the evidence for three main lineages, as well as the inference that goats were initially domesticated in the Fertile Crescent. Since then, Naderi et al. (2008) analyzed mtDNA from 473 wild-bezoar samples collected over the whole present range of the species and compared the results

86

origins of agriculture in western central asia

with the known genetic diversity of domestic goats. All the mtDNA haplogroups found in domestic goats were also found in the bezoar samples, and the geographical distribution of these haplogroups in the bezoars allowed the probable location of independent centers of goat domestication to be determined. Naderi et al. inferred that the most probable area of origin of the A haplogroup (to which 90% of domestic goat haplotypes belong) is eastern Anatolia, where it is common in wild bezoar populations, whereas the C haplogroup predominates in the southern Zagros and the central Iranian plateau (in Fars, Yadz, and Kerman provinces). Estimated bezoar population growth rates were found to be higher in the C haplogroup than in other haplogroups and Naderi et al. hypothesized that this reflected early “management” or “incipient domestication” of some bezoar populations in the central Zagros prior to their “true domestication”—which accords with Zeder and Hesse’s analysis of the goat bones from Ganj Dareh. Naderi et al. (2008) concluded that goats were probably domesticated independently in eastern Anatolia and the southern Zagros/central Iranian plateau, that the former region contributed more than the latter to the modern goat gene pool, and that, despite very extensive sampling of bezoars across Iran, no haplotypes were found that could have been domesticated in the eastern half of the central Iranian plateau or eastward of it—a conclusion that argues against a hypothesis of independent domestication in western Central Asia. Nor did Naderi et al. find any mtDNA evidence that would support an independent domestication of bezoars in the Indus valley region, for example at the site of Mehrgarh where remains of very small goats presumed to be domestic have been found in the earliest aceramic (Period I) levels, and where there is also evidence that young goats were buried with humans (Lechevallier, Meadow, and Quivron 1982; Meadow 1993:310). The site is on the margins of the recent range of the bezoar and Meadow reported (ibid.) the presence of bones of “some very large animals during Period IA [which suggests] that wild goats were also hunted at least until the end of that period.” Wild bezoars may well have been hunted and domestic goats raised at Mehrgarh (as evidently happened at Neolithic Jeitun), but the mtDNA evidence suggests that the domestic goats had spread to the area from the Fertile Crescent, rather than been locally domesticated. The recent genetic research on goat domestication offers no support for a hypothesis of independent

goat domestication in western Central Asia. Furthermore, the credibility of inferring past conditions from present-day data is strengthened by the observation that none of the bezoar haplogroups apparently underwent so much population reduction that the present genetic structures would not reflect those of the Early Holocene (Naderi et al. 2008:17,660). However, there are limits to the inferences that can plausibly be drawn from the genetic data, and what is now needed is to test and elaborate the data with analyses of ancient DNA (aDNA) recovered from goat bones in Neolithic (and Mesolithic) sites. No such investigations have as yet been undertaken in western Central Asia, but the feasibility of the technique has been demonstrated by analysis of aDNA from Neolithic and Chalcolithic/ Iron Age sites in southern France and northern Iran (Fernández et al. 2005, 2006; Mashkour, Fontugne, and Hatté 1999). In both investigations it proved possible to relate the aDNA sequences to corresponding haplogroups and to infer that in both cases the goats were descended from populations that originated in the Fertile Crescent.

Sheep All wild and domestic sheep are grouped in the genus Ovis but their classification is both confused and confusing. In recent decades taxonomists have tended to reduce the number of species and subspecies recognized, on the grounds that all extant wild sheep are capable of interbreeding and because genetic data have supplemented the morphological criteria on which many distinctions between taxa were initially based (Nadler et al. 1973). Wild sheep can be divided according to their diploid chromosome numbers into four groups and seven species, all but one of which (the American bighorn, O. canadensis) are native to western and central Asia (Clutton-Brock 1999:69–72). The Asiatic mouflon (O. orientalis, diploid chromosome number 2n=54) is generally believed to be the principal or only progenitor of all domestic sheep. The so-called European mouflon (2n=54), which was formerly referred to as O. musimon but is now usually classified as O. orientalis, survives in a feral state in the mountains of Corsica and Sardinia and is believed to be a descendant of domestic sheep originally introduced to the islands during the Neolithic (Poplin 1979). Two other species of Asiatic wild sheep, the argali (O. ammon, 2n=56) and the urial (O. vignei,



areas of origin of the crops and domestic animals

2n=58), have been regarded as possible contributors to the ancestry of domestic sheep (Ryder 1984:65–66; Zeuner 1963:159–61), but their higher chromosome numbers make this unlikely. Analyses of Ovis mtDNA by Hiendleder et al. (1998, 2002) also identify the Asiatic mouflon as the probable ancestor of domestic sheep, and provide no evidence of contributions from either the urial or the argali. However, two wild hybrid populations of mouflons and urials have been described, one in the Elburz mountains of northern Iran and the other in Iranian Baluchistan. Their existence has been attributed to the interbreeding of formerly separate populations that came into contact as a result either of the melting of montane glacial barriers early in the Holocene (Nadler et al. 1973:118) or, more probably, of a reduction of forest barriers caused by increasing aridity during the last two millennia coupled with more intensive exploitation of vegetation (Valdez, Nadler, and Bunch 1978:68). These cases of natural hybridization in the wild show that urials could have interbred with mouflons and contributed to the ancestry of some breeds of domestic sheep, although the available mtDNA data do not support that possibility. Today the Asiatic mouflon survives in small numbers in mountainous and hilly habitats from south-central Turkey to Armenia, south down the Zagros range, and east to the Elburz mountains south of the Caspian Sea (Fig. 1.15). In the Late Pleistocene/Early Holocene its range, attested by remains found in archaeological sites, extended farther west in central Turkey and farther southwest into the southern Levant, although apparently not east of the Elburz mountains into Turkmenistan (Uerpmann 1987:126–27). The range of the urial extends in mountain habitats from the eastern Elburz across Turkmenistan and farther east and south in western Central and South Asia (Fig. 1.15). The urial sheep of the Bolshoi Balkhan mountains in Turkmenistan have been classified as a steppe subspecies, O. v. arkal, that formerly ranged across northern Turkmenistan to the delta of the Amudarya (this volume, p. 12). Uerpmann (1987:130) argued that both the urial and the mouflon had more restricted ranges in the Late Pleistocene and were more separated geographically then than they are today—a view that accords with Valdez et al.’s (1978) suggestion that the urial-mouflon hybrid populations in Iran may be a recent phenomenon due to forest clearance having removed barriers to the extension and overlapping of their ranges. The

87

third species of wild sheep, the argali, is larger than the urial and the mouflon and lives at higher altitudes east of the range of the urial (Fig. 1.15). Its range extends eastward from beyond the upper Syrdarya valley through the Tien Shan and Altai mountains to Mongolia and south to the Tibetan plateau and the Himalayas. The geographical pattern of adjacent but spatially discreet ranges of the three main Asiatic species of wild sheep, from Turkey in the west to Tibet in the east, is borne out by such osteoarchaeological evidence as there is of the remains of wild sheep. Uerpmann (1987:126–32) documented the presence of the Asiatic mouflon at some 40 Late Palaeolithic, Mesolithic, and Neolithic sites around the Fertile Crescent and in central Turkey, and also of the urial at four such sites farther east in northern Iran (Ali Tappeh and Ghar-i Kamarband), southern Turkmenistan (Anau), and northern Afghanistan (Aq Kupruk). At the two Iranian sites fragments of horn cores typical of the urial were found in Mesolithic contexts (Uerpmann 1987:130; Uerpmann and Frey 1981:153–55). The small numbers of wild-sheep bones found at Anau came from the Chalcolithic IB and IIA strata and were greatly outnumbered by the bones of domestic sheep (Moore et al. 2003:155). Meadow (1989a) evaluated the sheep bone assemblages from Aq Kupruk I and II and concluded that Perkins’ (1972) identification of the remains of wild urial sheep in the Upper Palaeolithic (Kuprukian) levels was probably correct, but that his identification of domestic sheep in the “Neolithic” and later levels was not well founded. Remains of wild sheep (unspecified, but presumably urial) were also found by Coon in what he interpreted as Mesolithic deposits at the site of Kara Kamar in northern Afghanistan (Coon 1957:234). In Turkmenistan remains of wild and domestic sheep have been reported from Mesolithic and Neolithic levels at the Bolshoi Balkhan rockshelters of Jebel and Dam Dam Cheshme 2. Tsalkin (1956:221) was unable to determine, because of the poor preservation of the bones, whether any of the sheep bones excavated at Jebel were from domestic animals, but (as already mentioned) Markov (1966a:123) stated that Tsalkin reported some of the bones found in (Neolithic) layer 3 at DDC 2 to be “indisputably” from domestic sheep. Also, at Jeitun Kasparov (1992:51, 63–64) identified a few bones of wild urial sheep, all from adult animals, among the more abundant remains of domestic sheep and concluded that the urial was hunted in the moun-

88

origins of agriculture in western central asia

tains together with wild bezoar goats. In Uzbekistan, in the early Neolithic (Dariasai phase) levels at Ayakagytma in the Kyzylkum, a single bone was tentatively identified as from a wild sheep, but all the other caprine bones in the small assemblage were recorded as remains of domesticated sheep (and possibly goat). Lasota-Moskalewska et al. (2006:215) found the apparent lack of “interest in sheep (or goat) breeding. . .very strange” and concluded that sheep did not have “major economical importance” (and see Chapter 11 for general discussion of the possible role of stockbreeding in the subsistence economy at Ayakagytma). The genetic evidence for the ancestry of domestic sheep (Hiendleder et al. 1998, 2002), together with the distribution of their putative wild progenitors, identifies the Asiatic mouflon as the only, or at least the principal, ancestor, and the locus of domestication as west of Turkmenistan within the natural range of the mouflon. This general conclusion was supported by two further studies of ovine mtDNA. In an analysis of 836 samples from domestic breeds and wild sheep in Eurasia and Africa, Bruford and Townsend (2006) found no evidence of urial or argali contributions to present-day breeds, although they did emphasize the need for “further sampling of domestic populations in Central Asia” and elsewhere, as well as “a detailed study of O. orientalis” (Bruford and Townsend 2006:311, 314). This recommendation was partly addressed by an analysis by Tapio et al. (2006) that included data from Central Asia. They analyzed mt DNA from 48 breeds and local varieties of domestic sheep sampled across a wide area from northern Europe to east of the Caspian Sea, in a study that identified four divergent lineages (haplogroups A, B, C, and D) and thus confirmed the findings of Hiendleder et al. (1998, 2002), Pedrosa et al. (2005), and Guo et al. (2005). Groups A, B, and C are indicative of population expansions and group D was found only in a single sheep in the Caucasus region. Groups A and B are interpreted as representing expansions westward and eastward out of Southwest Asia and the Caucasus approximately 9000 years ago, which is broadly compatible with the archaeological evidence for initial sheep domestication in that region. Groups A, B, and C were all found in the two Central Asian sample areas (one south of the Aral Sea and the other in the Altai mountain region). The presence of A and B in Central Asia is attributed to a gradual spread of domesticated sheep from Southwest to Central Asia, and group C, which occurs mainly in Central Asia, is thought to

have emerged more recently there, two or three millennia after the domestication of sheep in Southwest Asia, possibly as a result of introducing wild mouflon ewes or lambs into local domesticated stock (Tapio et al. 2006:1781). This interpretation of the pattern of mtDNA diversity in Central Asian breeds should be regarded as provisional and it calls for further sampling in the region; but it does suggest that the presence of domestic sheep in Neolithic western Central Asia is attributable to their gradual spread from Southwest Asia rather than the result of local domestication of the wild mouflon, the range of which probably did not extend east of the Caspian. The zooarchaeological evidence already discussed demonstrates that wild urial sheep continued to be hunted during the Neolithic when the herding of domesticated sheep (and goats) became established as part of the agro-pastoral economy of the Jeitun Culture. This, together with the possibility (already mentioned) that sheep as well as goats were kept and bred in the Early PPNB at the site of Nevalı Çori, and definite evidence of domestic sheep at many sites around the Fertile Crescent by the Late PPNB, reinforces the genetic evidence and leads to the conclusion that the domestic sheep of Neolithic Turkmenistan were probably descendants of stock domesticated from the mouflon in Southwest Asia that had spread to western Central Asia during the Pottery Neolithic period. So too, perhaps, were the sheep recorded from the earliest aceramic and ceramic (Periods I and II) levels at Mehrgarh in Baluchistan, although, as Meadow suggested, it is possible that “local wild sheep. . . may have been kept and bred, only to be replaced later in the history of the region by western forms developed for their wool and fat production” (1993:311).

Dogs The dog is the third domestic animal, in addition to goats and sheep, whose remains have been found during excavations at Jeitun. Canid bones from the site were first reported, following Masson’s excavations in 1956–58, by Shevchenko (1960:473–75) who identified both wolf and domestic dog, but Kasparov, who analyzed the animal remains recovered during the 1989–91 excavations, identified all the (42) Canis bones he examined as domestic dog (Kasparov 1992:51, 71–73). Our excavations in 1993 and 1994 produced only two identifiable fragments of canid



areas of origin of the crops and domestic animals

remains which did not permit differentiation between wolf, jackal, and dog (this volume, p. 177). Kasparov (1992) reported that many of the goat, sheep, gazelle, and other bones from Jeitun show signs of having been gnawed by dogs and he assumed that the dogs were mainly used in hunting. During excavation in 1991, dog bones were found with a pottery vessel in a niche in a wall of one of the houses, suggesting deliberate burial of a domestic dog. Small clay figurines from Jeitun that represent animals provide further, tentative evidence of domestic dogs (Fig. 8.12, color). Kasparov (2000) analyzed similar animal figurines from Late Chalcolithic levels at the sites of Ilgynly-depe (70 figurines), Altyn-depe (13), and Kara-depe (13) on the Kopetdag piedmont east of Jeitun. Most of the identifiable figurines are of bulls, but Kasparov also identified 7 as dogs, which in style and proportions resemble the dog-like figurines found at Jeitun. Bones of domestic dogs were reported from Middle and Late Jeitun-Culture levels at Chagylly in the eastern piedmont by Berdiev (1966:27), and from the Chalcolithic IA and IB strata at Anau North where the remains varied in size from medium to large dogs (Moore et al. 2003:155, 157). Four bones identified as from domestic dog have also been found in the early Neolithic (Dariasai phase) levels at the site of Ayakagytma in the Kyzylkum (Lasota-Moskalewska et al. (2006:208). Although the bone and figurine evidence is sparse, it is sufficient to demonstrate the presence of domestic dogs at Jeitun and other sites in the Neolithic and to suggest that they were used in hunting wild game, for herding domestic sheep and goats, and perhaps as guard dogs and pets. Despite the existence today of over 400 distinct breeds, all domestic dogs are believed to be descendants of the wolf (Canis lupus) which ranged until recent times throughout Eurasia and North America. In Eurasia, both the larger subspecies of northern latitudes and the smaller wolves of Southwest, Central, and South Asia may have contributed to the ancestry of domestic dogs in a long process of taming and domestication that began at least 15,000 years ago (Clutton-Brock 1999:56–58). This view, which was originally based on anatomical and behavioral similarities between wolves and dogs and on archaeological evidence of dogs in Late Palaeolithic and Mesolithic contexts (see, for example, Clutton-Brock 1995), has since been borne out by genetic studies (Savolainen et al. 2002; Vilà et al. 1997; Vilà, Maldonado, and Wayne 1999; Wayne, Leonard, and Vilà 2006).

89

Results from an analysis of mtDNA from a sample of 162 wolves from 27 populations in Eurasia and North America and of 140 dogs representing 67 breeds by Vilà et al. (1997) strongly supported the view that the wolf was the exclusive ancestor of domestic dogs and showed that more than one wolf population had contributed to their ancestry, implying either that wolves were domesticated independently in several parts of the world at different times or that there was one initial domestication followed by episodes of admixture between dogs and wolves. Subsequently, a larger-scale analysis of mtDNA from a worldwide sample of 654 domestic dogs undertaken by Savolainen et al. (2002) found that mtDNA diversity present in East Asian dogs was much greater than in dogs from other areas, implying a common origin of all dog populations from a single gene pool. The greater genetic variation in East Asia and the pattern of phylogeographic variation suggested that dogs were domesticated approximately 15,000 years ago in East Asia, although a much earlier date of c. 40,000 years ago could also be inferred from the data. However, Wayne et al. (2006:283) subsequently argued that Savolainen et al. did not consider other factors that might have influenced the genetic diversity found in East Asian dogs and maintained that the evidence presented for the location and time of origin of domestic dogs was not conclusive— a conclusion since reinforced by an analysis of mtDNA diversity in 318 African village dogs, which was found to be similar to the diversity of the East Asian dogs sampled (Boyko et al. 2009). An East Asian locus of domestication is not at present supported by any definite archaeological evidence of domestic dogs in the Palaeolithic record of the region, although the remains of a possibly domestic dog have been found in the lowest cultural layers of a site in northern China (Nanzhuangtou) dated to approximately 12,000 BP (c. 14,000 cal. BP; Cohen 1998:25; Underhill 1997:113–14). Domestic dogs, sometimes buried in human graves, have been found in many Chinese and other East Asian Neolithic sites (Underhill 1997:121–28; Olsen 1985:48–70) and at the Neolithic site of Burzahom in Kashmir (Allchin and Allchin 1982:113). Also, at the large Chalcolithic site of Botai in northern Kazakhstan, where (as mentioned above) there is evidence of intensive exploitation of horses, dogs are the only other domestic animal found. Their bones comprise the second most abundant type of animal remains and most were found buried in pits in or close to houses. Olsen, who

90

origins of agriculture in western central asia

studied the bone assemblage (2000; 2003:99–100), showed that in their size and morphology the Botai dogs resemble the modern Samoyed breed, and she inferred from the placement of the burials in the foundations and at the thresholds of houses that dogs may have fulfilled a ritual role as guardians. The—very limited—archaeological evidence of domestic dogs in Palaeolithic contexts comes from much farther west, in Europe and the Levant. For example, two skulls of adult dogs were found at the Upper Palaeolithic site of Eliseevichi I on the central Russian plain west of Bryansk and radiocarbon dated to between c. 13,000 and c. 17,000 BP (c. 15,500– 20,000 cal. BP; Sablin and Khlopachev 2002); a dog mandible was found in a Late Palaeolithic human grave at Bonn-Oberkassel in Germany dated to c. 14,000 BP (c. 17,000 cal. BP; Nobis 1979); and canid remains identified as dogs have been reported from Epipalaeolithic (Natufian) contexts at several sites in Israel dated to c. 12,000–10,000 BP (c. 14,000–11,500 cal. BP), notably at Hayonim and Ein Mallaha where the dogs were buried with humans (Davis and Valla 1978; Dayan 1994; Tchernov and Valla 1997). Remains of domestic dogs have been found at several early Neolithic sites around the Fertile Crescent, for example at Jericho, Mureybet, Cayönü, and Jarmo (Clutton-Brock 1979:140–41; Lawrence 1967; Lawrence and Reed 1983; Peters et al. 1999:38). However, it should not be assumed that these early Neolithic dogs were necessarily descended from the Natufian dogs of the Levant because, as Tchernov and Valla point out (1997:66), it is possible that the Neolithic dogs of Southwest Asia were either domesticated anew in the region or introduced from elsewhere. As wolves were present in the Late Pleistocene/ Early Holocene across the whole of Asia, zoogeographically it is possible that dogs could have been domesticated anywhere between the Levant and China, including western Central Asia. There is, however, no conclusive evidence of domestic dogs at Late Palaeolithic/Mesolithic sites in the region. Coon (1957:155) found the muzzle of a very large dog in the lowest Mesolithic levels at the site of Ghar-i Kamarband, which he said resembled that of a St. Bernard dog and he inferred that domestic dogs were used in hunting during the Mesolithic. But H-P. Uerpmann, who analyzed the bone assemblage (Uerpmann and Frey 1981), thought the specimen more likely to be from a wolf (Uerpmann, pers. comm. 2007). No canid remains were identified at the nearby site of Ali Tappeh nor from the

Mesolithic (or Neolithic) levels at the Bolshoi Balkhan rockshelters of Jebel (Tsalkin 1956) and Dam Dam Cheshme, although according to Markov (1966a:123), Tsalkin identified several dog bones from the second (Bronze Age) layer at Dam Dam Cheshme 2. The apparent absence of remains of domestic dogs in the Caspian Mesolithic runs counter to a hypothesis of independent domestication from local wolf populations, but this conclusion must remain tentative in view of the relative lack of zooarchaeological research in western Central Asia. On present evidence, it seems probable that the existence of domesticated dogs at Jeitun-Culture sites (and at Ayakagytma in the Kyzylkum) is due to their having been introduced into the region during the Neolithic rather than the result of their local domestication from wolves.

Conclusion The foregoing review of archaeological, biogeographical, and genetic evidence concerning the areas of origin of the crops and domestic animals of Neolithic Turkmenistan and adjacent areas reveals how little we definitely know about how they were incorporated into the agricultural and pastoral systems that developed from c. 6000 cal. BCE (the beginning of the Early Jeitun phase). Without much more archaeobotanical and zooarchaeological research at Late Palaeolithic/Mesolithic and Neolithic sites in the region, together with accurate identification and direct (AMS) radiocarbon dating of the plant and animal remains found, it is difficult to reach well-founded conclusions about the central question of whether all the crops and domestic animals of the Jeitun Neolithic were introduced from elsewhere or whether some of them were independently domesticated in the region. The rapidly accruing genetic evidence on the origins of crops and domestic animals is beginning to provide valuable insights into the question, but close correlation between it and the biogeographical and archaeological evidence is difficult to achieve. Given the present inadequacy for western Central Asia of all three types of data, we cannot answer the question definitively, but by weighing the probabilities, already discussed, of local versus external domestications, it is possible to reach general, if tentative, conclusions. Regarding the cereal crops that were definitely present at Jeitun in the Early Neolithic, we can be confident that einkorn (and the tentatively identified free-



areas of origin of the crops and domestic animals

threshing wheat) were not domesticated locally, and although less is known about the ancestry of the other glume wheat, it almost certainly reached Jeitun as a domesticate from somewhere west of the Caspian. The phytogeographical and recently acquired genetic evidence that eastern forms of wild barley occur in Central Asia (as well as feral domestic barley) demonstrates that barley could have been domesticated independently in southern Turkmenistan, but its widespread distribution as a Neolithic crop around the Fertile Crescent long before it was present at Jeitun suggests that, like the wheats, it is more likely to have reached southern Turkmenistan as an already domesticated crop. The two domestic ungulates, sheep and goat, parallel wheat and barley in that the sheep appears to have been domesticated (from the mouflon) outside the region and subsequently dispersed eastward into Central Asia, whereas the distribution of the bezoar suggests that it could have been independently domesticated in the region. However, Naderi et al.’s (2008) analysis of bezoar mtDNA sampled across the wild goat’s present range found no support for possible domestication anywhere east of central Iran, so the domestic goats too are more likely to be descended from stock domesticated in Southwest Asia. The dog could have been domesticated locally from native wolves, but the lack of evidence of domestic dogs at Late Palaeolithic/Mesolithic sites in western Central Asia, compared with early evidence of them in Southwest Asia, tends to suggest otherwise. The evidence of the other domestic animals that were present in Turkmenistan and adjacent

91

areas in prehistoric and early historical times, but not found at Jeitun itself, reveals similar uncertainties. Pigs and cattle could have been domesticated locally from wild boar and aurochs, but the weight of archaeological and genetic evidence is against the possibility and in favor of both animals having been introduced as domesticates from Southwest Asia. Horses and Bactrian camels conclusively identified as domestic do not appear in the archaeological record until the Chalcolithic and the Bronze Age, but the wild progenitors of both are likely to have been present during the Late Pleistocene/Early Holocene and so might have been domesticated locally. Indeed, the discovery of horse and camel bones in early Neolithic and later levels at Ayakagytma in the Kyzylkum could be interpreted as evidence of local domestication, but the question of whether the remains derive from domesticated animals remains to be resolved. Finally, the dromedary appears only to have been introduced into Central Asia, as a domesticate, in the 1st millennium AD.

note 1. The Pre-Pottery Neolithic (PPN) period in the Levant lasted approximately three millennia and is conventionally divided into the PPNA and PPNB. The latter is subdivided into Early, Middle, Late, and Final PPNB (sometimes referred to as the PPNC). Approximate calibrated BCE dates for the divisions (after Peters, von den Driesch, and Helmer 2005:535) are as follows: PPNA 9500–8700, Early PPNB 8700–8200, Middle PPNB 8200–7500, Late PPNB 7500–7000, Final PPNB (PPNC) 7000–6500, succeeded by the Pottery Neolithic.

part iv

Archaeological–Environmental Investigations in Turkmenistan 1989–98

8

Jeitun, the Sumbar and Chandyr Valleys, and the Bolshoi Balkhan Region: Excavation and Survey with Chris Gosden

A

s explained in the Preface, the initial aim of our research in Turkmenistan was to build on the earlier work of the Russian and Turkmen archaeologists who discovered and described the Neolithic Jeitun Culture of the Kopetdag piedmont zone by undertaking archaeological-environmental investigations at the type site of Jeitun: the earliest known agricultural settlement in western Central Asia. As the project developed through the 1990s, its scope expanded to include exploratory survey and excavation in two other regions—the middle Sumbar and Chandyr valleys, and the Bolshoi Balkhan massif—which we hoped might help to elucidate the question of how agricultural communities originated on the piedmont. We undertook reconnaissance surveys and carried out small-scale excavations in both regions (described later in this chapter), and two members of our team (Coolidge and Kurbansakhatov) also reconnoitered most of the other Jeitun-Culture sites on the Kopetdag piedmont (this volume, pp. 60–61), but Jeitun itself was both the initial impetus for our work and the site we investigated most fully in our program of research. The earlier Russian-Turkmen excavations at Jeitun had concentrated on the overall layout and structures of the upper part of the settlement, and the main aim of our team was to complement this with fine-grained techniques of excavation and recovery in order to provide stratigraphically controlled and directly dated evidence of plant and animal remains, soils, and artifacts, and to investigate the settlement’s earlier history by excavating lower levels of the mound.

Consequently, our excavations were quite limited in extent, and in the final two seasons at Jeitun, 1993 and 1994, we concentrated on two houses in the lower layers of the site, the lower of which proved not to be underlain by other structures. We also excavated test pits across the site, partly to gain some idea of variability within it and also to sample the deeper layers at several places. Masson had based his conclusion that Jeitun was an agricultural settlement mainly on the large numbers of flint sickle blades found (studied principally by G. F. Korobkova, e.g., 1960), on the presence of animal bones that included sheep and goat (initially studied by A. I. Shevchenko, 1960), and on a small number of wheat and barley seeds and impressions of cereal grains in mudbrick (studied by G. N. Lisitsina, 1978:92). However, no systematic retrieval of plant remains by sieving and/or flotation was attempted and no radiocarbon dates on samples from the site were obtained. Two of our most important objectives were therefore to retrieve charred plant remains by systematic flotation of excavated samples, and to obtain a detailed radiocarbon chronology for Jeitun by using the AMS (accelerator mass-spectrometric) technique to date charred cereal grains and chaff (this volume, pp. 119–23). We achieved both these objectives, and our radiocarbon dates from Jeitun now form part of a broader effort to build radiocarbon chronologies for the Neolithic, Chalcolithic (Eneolithic), and Bronze Age in southern Turkmenistan (Hiebert 2002a:28–29; 2003:55–56).

96

origins of agriculture in western central asia

Excavations at Jeitun Jeitun is a small (0.7 ha) mound, with 3–4 m of cultural deposits, situated on sparsely vegetated windblown sand at the southern edge of the Karakum desert 28 km north-northwest of Ashgabat (Fig. 3.4, color). It is located at the margin of the alluvial fan of a river, the Kara Su, which (prior to construction of the Karakum Canal in the mid 20th century) traversed the piedmont, cut through the southernmost dune ridge of the Karakum, and dissipated in the desert close to the site (this volume, pp. 28–29, Fig. 3.1, and Fig. 3.2, color). The archaeological significance of Jeitun first became apparent in the 1950s and 1960s when excavations directed by V. M. Masson revealed a small settlement of rectangular mudbrick houses which he assigned, on the basis of similarities between the ceramic assemblages at Jeitun and at the western Iranian sites of Jarmo and Tepe Guran, to the 6th millennium BCE (Masson and Sarianidi 1972:36, 171). Masson’s main aim was to understand the social organization of the settlement and so his excavations focused particularly on the architecture and general layout of the site, which could be interpreted in social terms.

The first archaeological investigation at the site took place in 1952 when B. A. Kuftin and A. A. Maruschenko made several soundings in the mound, one of which penetrated 2 m of cultural deposits and revealed traces of five compacted floor levels, as well as flint blades and animal bones (Kuftin 1956:261, 271; Masson 1971:5–6). Although this was interpreted as evidence of five phases of settlement, each represented by a clay floor, the date and nature of the site remained in doubt. The first extensive excavation undertaken at Jeitun was directed by Masson in 1955 and 1956, when an area of 100 m2 was opened up (Masson 1957). This showed that the uppermost level of the site had been deflated, but that an earlier period of occupation (labeled phase 2 by Masson) was intact. In 1959 and again in 1962 and 1963 he uncovered almost the whole of the phase 2 settlement and showed it to be a small settlement of some 30 mudbrick houses, built as free-standing structures with courtyard areas and other ancillary features between them (Fig. 8.1; Masson 1971). The houses were very uniform in shape, close to square, with sides varying in length from c. 3.5 to c. 6.25 m. They consisted of single rooms and most of them contained large mudbrick ovens built against

Platform A

0

10

20

8.1 Plan of V. M. Masson’s excavation of the “second level” (phase 2) at Jeitun (based on Fig. 5 in Masson 1971).

30 Metres m



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

97

8.2 Plan of the Jeitun mound showing the locations of the trenches and test pits excavated between 1989 and 1994.

the northern or eastern walls, many with adjoining raised areas to the right of the ovens which have been interpreted as sleeping platforms. Excavation showed that some of the houses had single low entrances in one of their walls. Pottery was generally in the form of conical bowls with red painted patterns on a yellow background. Most stone tools were made of flint, with both blades and microliths reported, plus rarer occurrences of mortars and pestles and polished stone axes. Bone points and needles were also found, as were partially fired clay counters and animal and human figurines, and beads made of stone, bone, and

shell (Masson 1971:174–205; Masson and Sarianidi 1972:34–42). Although Masson concentrated on the phase 2 buildings, he also excavated several underlying structures which he designated phase 3. In a geometrical analysis of the published plans of nine excavated Neolithic–Bronze Age sites in southern Turkmenistan, including Jeitun, M. Ranieri (2000:593–601) demonstrated that architectural shapes were based on triads of small integers, probably used as rope lengths in the construction of what are almost-perfect orthogonal building segments. The regularity of house shape at Jeitun fits this model.

98

origins of agriculture in western central asia

Ranieri divided the houses of Masson’s phase 2 into three spatial groups according to their cardinal directions. He inferred that one group was likely to have been built before the other two—a hypothesis that contrasts with Masson’s assumption that all the upper surviving (phase 2) structures were contemporaneous. Ranieri’s theoretical analysis highlights the need for more extensive excavation of lower levels at Jeitun, and it was with the structures underlying the phase 2 level—which might provide clues to the earlier history of the site—that the recent excavations have been concerned. These started in 1987 when K. Kurbansakhatov (1992) opened a 10 x 10 m area at the eastern end of the mound which uncovered several houses thought to belong to phase 3.

Excavations in 1989 and 1990 Our involvement in work at the site began in 1989, in collaboration with Kurbansakhatov, Masson, and a Russian team from the Institute of the History of Material Culture of the Academy of Sciences in Leningrad (Masson and Harris 1994). The main excavations connected with our project took place in the center and at the northwestern edge of the mound (Fig. 8.2). Also, between 1990 and 1993 geoarchaeological investigations were conducted by the British team in the vicinity of Jeitun, leading to the discovery of several sequences of sand accumulation containing evidence of three levels of past soil development in the form of buried soils (this volume: Chapter 9, Sections 9.3 and 9.4). In 1989 a 10 x 11 m area was excavated in the center of the mound, under the direction of Berezkin (1992), with a small extension in the southwest corner from which the British participants took samples for flotation to determine whether charred remains of domesticated cereals were present that could be dated by the AMS technique. This proved successful and demonstrated conclusively that wheat and barley were cultivated at Jeitun during the early Neolithic. This excavation was continued in 1990 with the aim of uncovering structures from phase 3, together with any that might represent earlier periods of occupation. The central excavation, carried out in 1989 and 1990, uncovered a square house with sides approximately 5 m long, a hearth and raised (sleeping?) platform on the northeastern wall, holes in the floor that may have supported pottery vessels, and an entrance (doorway?) in the southeastern wall (Figs. 8.3,

8.4). There were two levels of lime-plaster floors and an upper floor of clay, indicating that the house was rebuilt several times; the two lower floors were separated by a layer of sand, perhaps deliberately placed to provide a level surface for the upper one. Three meters east of the doorway of the house there was a smaller structure without any internal features (Figs. 8.3, 8.4). The function of this outbuilding is unknown, but Berezkin (1992:29) found evidence of compacted animal dung on the upper of two floor levels and suggested that goats and sheep may have been penned in it, whereas Kasparov, who identified bones found inside the building, speculated (1992:75) that it may have been used for processing animal carcasses, including skins of fox and steppe cat. Remains of other structures were partially uncovered along the southwestern edge of the excavation. The yard area between the house and the other buildings contained layers, made up of ash and sand, mixed together with large amounts of animal bone, plant material, and artifacts.

N

Trench 1991a

0

5 Metres m

8.3 Simplified plan of the area excavated in 1989 and 1990 by Berezkin in the center of the Jeitun mound and the northeast extension (trench 1991a) excavated by the British team in 1991.



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

99

8.4 View of part of the central area excavated in 1989–90 and attributed to Masson’s third level, showing the main house, the internal oven with its adjacent raised platform on the northeast wall, and the entrance in the southeast wall, beyond which is a small outbuilding that may have been used as an animal pen or for processing animal carcasses, May 1989.

Many small hearths were found, with concentrations of animal bones around them. These finds as a whole suggest that domestic activities of many different kinds were carried out in the yard area. These buildings and yard layers were assigned to phase 3 in Masson’s scheme. However, there were two walls along the northeastern side of the excavation, one in the northern corner, the other near the southern corner, which were thought to lie below the other architectural features and to belong to phase 4. The wall near the southern corner was some 6 m long. It had a floor to the east and may have been part of an unusually large structure. The other wall was much shorter (1.5 m) and it was impossible to infer what sort of structure, if any, it belonged to. Little else was found below the phase 3 buildings, which led Berezkin to conclude that there had been a marked increase in the size of the settlement in phase 3. As all the phase 3-structures seemed to have been constructed at the same stratigraphic level, he thought that they had all

been built at the same time and possibly, therefore, by incomers rather than as a result of natural population growth. He believed that the phase 3 settlement was markedly larger than the preceding phase 4 settlement. The other excavation that Berezkin and the Russian team carried out was a long trench on the northwestern side of the mound in an area untouched by previous excavations (Fig. 8.2). It was 25 m long and 1.3 m wide and cut through the side of the mound from the top to sterile sand at the base. Several walls and floors thought to belong to both phases 2 and 3 were exposed along the trench. The probable existence of these phase 2 structures implies that the settlement was slightly larger than Masson originally thought and that the phase 3 remains may be as extensive as those of phase 2 (only large-scale excavation could test this possibility). The British excavations were much influenced by Berezkin’s work. It led us first to excavate in 1991 a small area in the baulk of

100

origins of agriculture in western central asia

the central excavation, and then, in 1993 and 1994, to carry out larger-scale (although still limited) excavations of houses at the southeastern end of Berezkin’s long trench.

Excavations in 1991 and 1992 During 1989 and 1990 British specialists worked at the site (Harris et al. 1993), but they did not direct the excavations. The main aims of our excavations from 1991 onward were to use fine-grained stratigraphic excavation techniques to distinguish individual features and fine layering on the site, and to undertake systematic sampling and fine sieving to retrieve charred plant remains, bones, and other small finds. This involved using techniques now standard in many parts of the world, but which were new in Turkmenistan. By using these methods we hoped to increase understanding of how the settlement originated and how its organization changed over time as different parts of the site were occupied. Standard British excavation techniques were used, whereby each individual feature is given a context number and is distinguished according to whether it is a built structure, a cut such a pit or posthole, or a fill. Excavated material was systematically sieved, flotation was employed (for the first time at Jeitun), and all sediment was passed through a 10-mm mesh. All levels measured in different years of excavation were tied into Masson’s original site datum (a metal post on top of the mound at 190 m above msl), and repeated measurement of the same feature in different years showed that the datum was stable. The question of levels is associated with the five phases of occupation into which the Russian excavators divided the site, based on general sets of floor levels. Our excavations suggest that this phasing is too general to convey the full complexity of the site’s use and its history of occupation. Nevertheless, the five-phase sequence does permit results from different seasons of excavation from the late 1950s onward to be compared. We were unable to replace it with a detailed alternative scheme, because for this to be done, more extensive excavation and dating would be required. We therefore continue to refer to the five phases as a convenient means of discussing different areas of the site, and in the conclusion to this section on the excavations at Jeitun we offer some tentative conclusions concerning the phasing and history of occupation of the site.

We carried out two excavations in 1991, but they were very limited in extent due to lack of time and personnel. First, a small (1.5 x 3 m) area, labeled trench 1991a, was excavated in the northeast corner of Berezkin’s central excavation (Fig. 8.2) in order to investigate how the deposits had built up and what kinds of structural evidence they contained, and also to obtain samples from secure stratigraphic contexts, including charred cereal grains for AMS radiocarbon dating. The excavation revealed a series of clay-lined hearths built up against two superimposed mudbrick structures, one assigned to phase 2 and the other to phase 3. The continuity from one phase to another in the construction and use of these hearths confirmed that Masson’s phases were useful as a starting point for interpreting the development of the site, but also that they do not encompass the straigraphical complexity that finer-scale excavation reveals. Despite its limited scale, such excavation can provide detailed evidence of human activities on the site, in this case in the form of hearths and their surrounding debris (Fig. 8.5 and Fig. 8.6, color). The contexts uncovered in this excavation consisted mainly of a series of hearths and layers of ash resulting from the cleaning of hearths and the dumping of rubbish. Layers of sand were also present, which may indicate periods of less intensive or different use of this part of the site. All the sand layers contained artifacts, charcoal, and bone, and there is no suggestion that the area was completely out of use for any period of time (see this volume, pp. 125–26, for a detailed description of the section shown in Fig. 8.5 and Fig. 8.6, color). The other major set of features consisted of mudbrick walls from at least two phases of construction. The individual bricks, although variable in size and shape, resembled long loaves. Some were almost a meter in length, and they tended to be more cylindrical than rectangular in cross section. Such “protobricks” distinguish buildings at Jeitun from those at later Jeitun-Culture sites where mudbricks are more rectangular in cross section and their shape is more standardized. Because the area excavated was so small, it was impossible to make any sense of the layout of the mudbrick and adjoining features, although it was possible to assign a relative date to the mudbrick features and note the occurrence of smaller features associated with them. This excavation is described in detail, with reference to numbered contexts, in Appendix 8.1 at the end of this account of the excavations at Jeitun.



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

101

14 22

12

26

22 27

28

30

Root channel Datum

33 38

38

42 42

42 41

41

43

44 44 45

45 46

0

1 Metre

8.5 The section through yard deposits in the northeast extension (trench 1991a) at Jeitun, sampled for particle-size, phytolith, and micromorphological analysis; context numbers shown; section datum at 188.6 m above msl; cf. Fig. 8.6 (color).

The second excavation we carried out in 1991 was even more restricted. It took place at the eastern side of the mound in a small (1.5 x 1.5 m) trench that had previously been excavated by Berezkin, following Kurbansakhatov’s excavations in the area in 1987. We labeled it trench 1991b, and deepened it in 1992 to become test pit 7 (Fig. 8.2). The main focus of our work here was on the sands underlying the complex of upper occupation layers. This new excavation revealed the existence of phase-3 yard layers, underlain by 40 cm of relatively sterile sand under which were earlier floor levels, each separated by layers of ash and sand. A depth of 4.70 m below datum was reached before excavation had to be halted because it was difficult and dangerous to dig any deeper. The lowest layers reached still contained some evidence of human occupation in the form of traces of burning and at 4.20 m below the datum there was a destruction level of mudbrick. A sample for dating was obtained from the lowest (4.70 m) level reached. It yielded a date (OxA-4695 at 95% confidence) of 6220–5900 cal. BCE (this volume, p. 120). During 1992 the British team, with Russian help, excavated six 1.5 m2 test pits elsewhere on the

site and also extended test pit 7 (Fig. 8.2). Most of the pits were dug across the southwestern part of the site, which Masson believed to have the deepest stratigraphy. They were designed as a rapid means of investigating the nature and distribution of the underlying deposits of phases 3 and 4. Test pits 5 and 6 came down on structures from phases 2 and 3, at which level we stopped excavating in order not to destroy these features. Test pit 2 revealed no features at all and may indicate that there was an open area in this part of the site. Test pits 1 and 4 went below features attributable to phases 2 and 3, but encountered no traces of earlier features. Test pit 4 was dug in the base of the central area that had been excavated in 1989 and 1990. The only features observed here below those of phase 3 were charcoal lenses accompanied by pottery and animal bone. A sample for dating was taken from the lowest charcoal lens, which gave a date (OxA-4693 at 95% confidence) of 6020– 5720 cal. BCE (this volume, p. 120). A further test pit (8) was dug in 1993 in an attempt to discover the limits of the site on its southwestern side, but no evidence of structures was found and only a few finds were recovered.

102

origins of agriculture in western central asia

Also in 1991 and 1992 a Russian team, under the direction of Berezkin, began an area excavation that extended out from the 25 m trench they had dug through the northwestern side of the mound in 1990 (Fig. 8.2). An area of c. 11 x 7 m was excavated and the remains of three houses were revealed. Two of them were thought to belong to phase 2 and the third underlying one was assigned to phase 3. One of the phase 2 houses was only partially excavated and the other was completely excavated. The latter was small (approximately 2.5 m2) and lay stratigraphically above and in the middle of the phase 3 house. Masson had not excavated it, and its discovery showed that, although his excavations were extensive, they did not include the entire phase 2 settlement. By the end of the 1992 season the first floor level in the phase 3 house, which had been overlain by several layers of sand and mudbrick destruction, had been reached. It was here that the British team started excavating in 1993.

ensured that all small fragments of bone and stone, as well as charred plant remains, were recovered. Any excavated deposit in excess of 50 l from a single context was sieved through a 10-mm sieve. In 1994 we altered this sieving strategy slightly so that, after the initial 50-l sample had been taken, the next 25 l of deposit from a context was passed through a 10-mm sieve and all subsequent deposit from each context went through a 15-mm mesh. Finds of bone and plain pottery were bagged by context and not located by square. Finds of diagnostic pottery (rims, bases, and decorated pieces), as well as clay figurines, bone implements, flint blades, and any other small finds were located to square. The methods used in 1994 were identical to those employed in 1993, except for the variation in sieving just described, and the 1994 grid was an extension of the one laid out in 1993.

Aims and Methods of the 1993 and 1994 Excavations

The main features encountered inside House A (Fig. 8.7 and Fig. 8.8, color) were an oven, a platform adjacent to it, and a series of floors interspersed with layers of mudbrick destruction and blown sand. A detailed context-by-context description of the excavation of House A is given in Appendix 8.2, so here we just summarize what we learned about its construction. The sequence of construction appears to have been as follows. The basal mudbrick material composing the floor (context 58) was laid down and the walls of the house (1) were built on top of this. The oven (5) and the adjacent platform (43) were probably built next, although we have no direct evidence for this seemingly logical sequence. The extensions to the oven (88 and 91) appear to be contemporary with the main oven construction. The gypsum cover (42) of the floor was then laid down and this evidently merged with the clay lining of the oven (52), suggesting that the oven lining and the gypsum cover were contemporary. Probably after the gypsum lining was laid down a small pit (87) was let into the floor in the southwest corner. The mudbrick feature (54/55) was laid down after the gypsum floor was applied to the mudbrick. Contexts 39 and 40, composed of mixed sand and gypsum, either represent later floor levels or the remains of organic floor coverings. Above these was the sand layer 37, which contained large numbers of finds including animal figurines and other clay artifacts (Fig. 8.9, color). It is not clear whether the series of small postholes (59–84) cut through context 37

The most extensive excavations reported in this volume are those carried out in 1993 and 1994 by the British team. Our main aim in 1993 was to excavate the house at the northwestern side of the mound attributed to phase 3 by Berezkin (and designated House A by us). At the end of the 1993 season it became obvious that there was another house beneath House A, as Berezkin had suspected. So, in 1994 (our final season at Jeitun) our main aim was to excavate this house (designated House B), the lowest architectural feature so far uncovered at Jeitun. It is the lowest of the three houses excavated on this part of the site, which spans phases 2 to 4, but whether an even higher phase 1 house ever existed above them is unknown. When we began excavation in 1993, a 1-m grid was laid out over the area to be excavated, with point 100/100 lying in the northwest corner to allow for extensions of the grid. Single-context methods of excavation were employed and all excavation was carried out with trowel and knife. From each context a standard 50-l sample of deposit was taken for flotation, with none of the finds removed, in order to gain a measure of the total amount of archaeological material per 50 l of deposit. The flotation samples were all wet-sieved through a series of sieves, the smallest of which had a 1-mm mesh. The fine sieve sizes used for flotation

The 1993 Excavation of House A



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

103

8.7 Plan of House A at Jeitun as excavated in 1993, showing the positions of the oven (1), the raised platform (2), the entrance (3), and, below, a cross section of the southeastern wall showing the positions of a window (1) and the entrance (2); previously published in Harris, Gosden, and Charles 1996; reproduced by courtesy of the Prehistoric Society.

into the gypsum floor relate to activities carried out in the house during its use, or were associated with people who may have camped in it during a period of disuse. At the conclusion of the excavation we removed

most of the northwestern wall of House A and dug a slot trench into the floor which uncovered part of the lowest house yet seen on the site. The northwestern wall of this lowest house had first been revealed in

104

origins of agriculture in western central asia

Berezkin’s 1992 excavation. He attributed it to phase 4, and we labeled the wall context 7. It had a mudbrick protrusion of unknown function on its northern side. The slot trench showed that the lowest house (House B) was approximately 4 m across and had an oven in the center of the north wall. The extent of the house toward the south could not be determined because it ran under an unexcavated baulk. The orientation of House A is identical to that of House B, although the upper house is larger and offset to the north. Excavation of House B was to be the main objective of our 1994 season at Jeitun.

The 1994 Excavations of House B and Associated Deposits Yard Deposits In 1994 we started by excavating layers of yard deposits from between the buildings previously excavated by Masson above Houses A and B. We removed the upper layers contemporary with Masson’s phase 2, which were approximately 2 m above the level of House B, in order to uncover yard deposits contemporary with Houses A and B, and to recover bones and plant remains from rubbish dumping and other activities that took place between the houses. In the event, once we reached levels contemporary with Houses A and B, the layers contained few artifacts, bones, and plant remains. Also, the excavations here were somewhat inconclusive because the top of the layers that we started excavating were at the point at which Masson had stopped digging, and they represented a surface which had been exposed and walked on for some time. They thus had no overlying archaeological context and were somewhat disturbed. So, because we had relatively little time on site we decided to stop the careful excavation of this area at an archaeologically arbitrary point and shoveled down to the level of House B. The layers we excavated in this area lack full contextual information and therefore cannot be fully interpreted. They are best regarded as a series of samples taken under carefully controlled conditions to recover organic remains and artifacts. Nevertheless, it is possible to make some general comments on the nature of the excavated layers. The area excavated was approximately 3 m x 7 m in extent and can be divided into northwestern and southeastern ends in terms of the features found. At the northwestern end there were several spreads of

sand, some of which were clean (contexts 135, 140, 145, 150, 152, 158) and some of which contained admixtures of charcoal and mudbrick and showed evidence of disturbance by roots and animals (137, 141, 163). They were adjacent to a mudbrick wall, presumably part of a house excavated by Masson, and we eventually came down on to a series of floor levels (which were only encountered during mass removal of the deposits by shoveling). This area had been kept relatively clean and may have been used for activities immediately adjacent to the house. At the southeast end there was evidence of hearths and small areas of ash and burning (143, 157, 169), disturbed by large animal burrows (151, 164). They were associated with a low mudbrick wall (167), which was only fully revealed when we were shoveling. It is impossible to deduce what function(s) these burnt areas represent due to the small size and truncated nature of the excavation, but it is probable that the two areas, only 2 or 3 m apart, had different functions. This reinforces the inference that there is much spatial differentiation within the site which reflects a variety of social and economic activities at the settlement.

House B As in House A, the main features found in House B were the remains of an oven with an adjacent platform (Fig. 8.10 and Fig. 8.11, color), and a series of floors interspersed with layers of sand and remnants of mudbrick. To uncover House B, we removed five spits of 20–30 cm depth by shovel. A sample was taken for wet sieving from each spit, and visible finds were collected and bagged. The northeastern and northwestern walls of the house had been uncovered by the slot trench dug at the end of the 1993 season, which had also revealed the oven attached to the northeastern wall. The southeastern wall lay below the floor of House A and the southwestern one below the area where we had excavated yard layers by shoveling. We removed the walls of House A (context 1) on its northwestern and southwestern sides down to the lowest (gypsum) floor level in the house, and then removed the gypsum floor (42). The floor consisted of a series of layers, some of which appeared to have been covered in red paint (samples were taken for analysis, but, unfortunately, they were subsequently lost in transit). The gypsum floor overlaid a sub-floor (176) made of solid mudbrick to a depth of about 8 cm that contained abundant plant remains, but otherwise was almost sterile. The sub-floor over-



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

105

8.10 Plan of House B at Jeitun as excavated in 1994, showing the positions of the oven (1), the raised platform (2), the entrance (3), and an incomplete tortoise carapace (4); previously published in Harris, Gosden, and Charles 1996, reproduced by courtesy of the Prehistoric Society.

laid a leveling or destruction layer (178) composed of lumps of clay (probably derived from the mudbrick of House B) and burnt clay, together with charcoal that underlaid the whole of House A. Next was a further leveling layer (181) that extended across the southern part of both houses. It was a thick (10–15 cm) layer

of mixed sand, charcoal, and pieces of mudbrick and it appeared to have been put across the whole area on which House A was subsequently built to provide level foundations for the house. A detailed context-bycontext description of the subsequent excavation of House B is given in Appendix 8.3. We also excavated

106

origins of agriculture in western central asia

several areas immediately outside the house, detailed descriptions of which are given in Appendix 8.4. The history of construction and use of House B seems to have been simpler than that of the overlying House A and can be summarized as follows. The lower house appears to have been constructed on clean sand (244) and thus provides evidence of the earliest use of this part of the site. Its solid mudbrick floor (249) was laid down first and the walls of the house were built at the same time. Presumably the oven and the platform were constructed at the same time as the floor and walls. Some of the lowest contexts we uncovered outside the house (234, 242) may be connected with its construction or they may represent floors that provided working areas outside it. It is uncertain at what point the buttressing at the corners of the house (129, 210, 215, 251) was built. The period during which the house was used is suggested by the gypsum floor (204, 225), which was renewed on several occasions. It appeared to be continuous with the lining of the oven and we assume that they were laid down at the same time and made of the same material. Contexts 199, 200, and 202, which lay on top of the floor and were all composed of clean sand and mudbrick, probably represent sand blown into the house while it was out of use. Context 196, which consisted of clay with gypsum patches, lay on top of these contexts, suggesting that it had been a temporary floor of some type, but whether it was laid while the house was still standing or whether it represents more ephemeral occupation after the house had collapsed cannot be determined. Two contexts (178, 181) on top of context 196 consisted mainly of mudbrick with some charcoal and sand and they were evidently designed to provide a foundation for House A. On top of them lay the solid mudbrick floor of House A (176) with its gypsum covering (42). The postholes in the southern corner of House A are probably evidence of some activity that post-dates the collapse of House B, although we could not infer what kind of activity they represent. The oven in House B revealed no in situ evidence of burnt material, but there was evidence of relining and the collapse of its roof. Fewer artifacts were found in House B than in House A, although one unusual find was made: an incomplete carapace of a steppe tortoise that had been (deliberately?) deposited close to the southwestern wall (Fig. 8.10 and Fig. 8.12, color). Overall, House B seems to have had a simpler history of use than House A, which reinforces the idea that occupation of the

lower level may have been less intensive than that of the upper levels of the mound.

Conclusion: Excavations at Jeitun and Their Interpretation Excavation at Jeitun has a long and complex history that falls into two main periods: the campaigns carried out by Masson between 1955 and 1963, and the excavations carried out by Kurbansakhatov, Berezkin, and the British team between 1987 and 1994. The campaigns had contrasted aims and employed a variety of different techniques, especially in the later period. Masson’s aims were to excavate as much as possible of the deflated remnants of phase 1, and especially to uncover as much of the phase 2 settlement as possible. He also made smaller soundings which encountered lower levels, mainly those attributed to phase 3. The excavations from 1987 onward concentrated on the lower levels, first Kurbansakhatov’s at the eastern end of the mound and then Berezkin’s in the central area. Berezkin paid more attention to the fine details of the stratigraphy than had earlier excavators, partly because by then a major aim was to recover samples for archaeobotanical and geoarchaeological analyses by the British specialists. Our excavations from 1990 onward were slow and painstaking by Russian standards because they were designed to elucidate some of the fine details of stratigraphy (necessarily over limited areas), to recover bulk samples for flotation, and to carry out systematic sieving. By combining the results of these later investigations we can gain some understanding of the broader history of the site (this volume, pp. 194–95 and 211–12), but for the purposes of this brief conclusion we retain Masson’s phases 1–4 (he mentions five phases, but it is not clear what the final phase refers to). At the same time, we suggest that, if fine-grained excavations were undertaken on a larger scale than we were able to achieve in the time and with the limited resources available, it would be possible to develop a more refined phasing of the occupation of Jeitun. Masson excavated some 30 phase 2 houses, which he interpreted as all contemporary with each other. He envisaged Jeitun as a permanent settlement sited on a low sand hill at the edge of the desert, occupied by some 30 nuclear families (Masson and Sarianidi 1972:33). Berezkin believed that there were stratigraphic grounds for concluding that the phase 3



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

houses he uncovered in the central area were likewise contemporary with each other and that they had been established within a very short period of time, probably by new settlers. However, he did distinguish an earlier and a later period within phase 3. The much sparser evidence both of architectural features and of yard deposits from phase 4 seemed to him to indicate a smaller population and a less intense form of occupation. This is confirmed to some extent by our small-scale excavations in the northeast extension, where upper layers, belonging to phases 2 and 3, appear to derive from more intensive activities surrounding hearths than the layers below them. Such a picture is further confirmed by the excavations of Houses A and B, the lower of which appeared to be surrounded by fairly ephemeral yard layers that produced smaller quantities of bones, plant remains, and artifacts than the layers above. The fact that our excavations were small scale and fine grain makes it difficult to draw broad stratigraphic conclusions. We emphasize the complexity of the use of the site and point out that the mudbrick architecture is only the most obvious evidence of occupation. For example, postholes and small hearths may indicate that people camped at the site between episodes of building. At several places, such as test pit 4 in the center of the mound and test pit 7 at its eastern end, there is evidence of hearths and charcoal lenses beneath the architectural levels. This indicates ephemeral use of the site prior to any evidence of buildings, although not pre-dating the architectural levels by any great period of time. There seems to be a definite change between phase 3, where large-scale settlement is in evidence, and the preceding phase 4, when there may have been fewer people and less intensive activity on the site. How far all the phase 3 houses were contemporary with each other, and likewise those of phase 2, is uncertain, and without finer grained stratigraphic evidence it is difficult to resolve that question. Looking back over the long history of excavation at Jeitun, it is clear that the aims of the excavators have changed markedly over the many years during which the site has been intermittently investigated. Between 1955 and 1963 Masson’s main aim was to uncover a fully functional Neolithic settlement revealed at a particular time in its existence, as represented by phase 2. When excavation was resumed between 1989 and 1994 objectives had changed with the advent of AMS radiocarbon dating and the development of rigorous bioarchaeological and geoarchaeological techniques. The main aim now was to investigate the longer-term

107

history of the settlement and to obtain more detailed evidence of the environment and the economy of its inhabitants by securing, dating, and analyzing samples retrieved from clearly defined stratigraphic contexts. Recent work at the site has yielded much new data, but more large-scale, fine-grained excavation of the substantial deposits that remain is needed if the full history of Jeitun is ever to be unraveled.

Fieldwork in the Middle Sumbar and Chandyr Valleys, 1996 In 1996 we carried out fieldwork in the middle Sumbar and Chandyr valleys in the western Kopetdag mountains during a three-week season of reconnaissance survey and trial excavations. We did so primarily because we thought that early Neolithic, Mesolithic, or Palaeolithic sites might exist there which antedated the Neolithic sites of the Kopetdag piedmont and might thus throw light on the origins of the Jeitun Culture and the beginnings of agriculture in Turkmenistan. One reason for regarding this as a possibility was that the valleys experience a Mediterranean-type climate with sufficient precipitation to sustain rainfed agriculture (Harris and Gosden 1996:384). More broadly, we hoped to find evidence of how, in these upland valleys well suited to agro-pastoralism, people subsisted in the early and mid Holocene, in contrast to the more arid environment of Jeitun.

The Regional Environment The Sumbar and Chandyr rivers drain the western end of the Kopetdag mountains in southwestern Turkmenistan, close to the present border with Iran. They occupy the two main intermontane valleys that trend east-west through the mountains, join the main Atrek drainage, and flow into the Caspian lowland along the Iranian frontier (Fig. 8.13). In their uppermost courses the Sumbar and Chandyr are mountain rivers, but their middle sections flow through wide valleys flanked by hills. Here the climate is relatively mild: in July average temperatures are between 28oC and 30oC, in January between 2oC and 4oC; there are long frost-free periods, high amounts of sunshine, and mean annual rainfall of between 300 and 250 mm (data from the Atlas of the Turkmenian SSR, Moscow, 1982). The valleys have deep fertile al-

108

origins of agriculture in western central asia

luvial soils and the native vegetation—much of which has long since been cleared for agriculture—contains many trees, shrubs, and herbs of the Mediterranean flora, including species with edible fruits such as pear, pomegranate, and pistachio. Today many fruits and vegetables, as well as cereals, are cultivated in the area. The valleys remained unglaciated during the Pleistocene and would have experienced higher temperatures and rainfall during the mid-Holocene Climatic Optimum (this volume, pp. 25–26). Cretaceous limestones form much of the Kopetdag mountains, which are seismically active and are today undergoing tectonic uplift. As a result, the landscape is intensively eroded, a natural process accentuated by crop cultivation and livestock grazing over many centuries.

Fieldwork Methodology Our fieldwork was concentrated in the middle Sumbar valley in the vicinity of Kara-Kala and along its

southern margin, with a shorter visit to the Chandyr, the southern side of which was inaccessible to us because it lies within the closed border zone along the Iranian frontier. In our surveys, we were following reconnaissance in 1980/81 by V. P. Lyubin and other members of a YuTAKE team who discovered 10 minor Palaeolithic sites and collected over 100 Palaeolithic stone tools, mainly on eroded slopes and river terraces along the south bank of the Sumbar near Kara-Kala and the north bank of the Chandyr near Kyzyl-Imam (Lyubin 1984:28–31). We focused on caves and rockshelters because, in the highly eroded landscape, such sites seemed to offer the best chance of finding early and/or pre-Neolithic deposits where late Pleistocene or early Holocene sediments might have been preserved. Initial reconnaissance was carried out on foot and this procedure located many shelters under limestone overhangs, together with a small number of caves. Sites were targeted for excavation following a dual strategy: on the one hand, to maximize our

km

8.13 Southwestern Turkmenistan showing the relationship of Jeitun to the Sumbar, Chandyr, and Atrek valleys, the Bolshoi Balkhan massif, and the Uzboi channel.



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

109

Parkhai

km

8.14 The middle Sumbar and Chandyr valleys showing the approximate locations of the archaeological sites visited and test-excavated in 1996; P=Parkhai, D=Dorian, PC=Parkhai Chalcolithic–Bronze Age, K=Karakesy, C=Chandyr.

chances of success, we wanted to examine sites in as many different parts of the Sumbar and Chandyr valleys as our limited time allowed; on the other, we wanted to look at the interrelationships of sites in one area to see how they varied in terms of their ages (dates), their types of sediment, and their artifacts. In total, we located and test-excavated eight rockshelter and cave sites and one open site, all in the Sumbar valley except for one in the Chandyr catchment (Fig. 8.14). Our basic unit of sampling was a 1-m2 trench and at most sites we only excavated a single trench. We dug mainly by individual contexts, although when time was short grosser forms of excavation were used. Dry sieving was employed throughout and 10-l samples of deposit were taken from the most promising contexts to ensure total recovery of bone, plant remains, and artifacts from them. The sites we investigated are described below: first, three close to Kara-Kala (Parkhai, Dorian, and the Parkhai Chalcolithic–Bronze Age site), next, a group of sites on the southern side of the Sumbar valley (Karakesy 1–5), and finally the only site we were able to visit in the Chandyr catchment, (Chandyr). In the event, our search for pre-Neolithic and Neolithic sites was unsuccessful. Possible reasons for this are discussed below in the conclusion that follows brief descriptions of the nine sites we testexcavated.

The site is situated northeast of Kara-Kala on the southern slopes of the Kara-il-chi mountains on the side of a ravine running approximately northsouth. At c. 800 m above msl, it is the highest site we excavated. The local bedrock is a mixture of sandstone and limestone, and around the site there is grassland with a scattering of trees and small shrubs, grazed by flocks of sheep and goats. The site consists of an outer rockshelter and an inner cave divided into two parts. The shelter is c. 20 m long with an overhang of several meters to the north and especially to the east. There is a large tumble of rocks near the cave’s entrance that forms the south end of the shelter. Lyubin excavated part of the shelter during his survey of the region. In 1994, during a preliminary visit, we removed spoil from what appeared to be Lyubin’s trench and excavated to a depth of 1 m where we encountered roof fall. In 1996 we excavated a 1-m 2 trench in the shelter 2 m west of Lyubin’s, which we labeled square 1. The stratigraphy consisted of alternating red-brown and yellow sandy layers. The former probably derive from humic material and the latter have weathered from bedrock. Our excavation reached a total depth of about 60 cm and produced few finds, the most notable being a possible Chalcolithic (Eneolithic) potsherd near the base of the excavation. The excavation was brought to an end by a large rock, which had probably fallen from the roof. We also extended Lyubin’s excavation 1 m to the west, where a similar sequence of layers was found, with some Bronze Age and later pottery in the upper layers and nothing in the lower ones. Inside the cave we opened a 2 x 1 m trench, labeling the eastern half square 2 and the western half square 3. Excavation was hampered in the early stages by bad light, as we were unable to make the generator we had brought with us work. The upper layers consisted of a series of small lenses of ash and produced one complete pot and a coin, both probably of Parthian date. There was also some bone and a few other sherds. Below this was a series of levels of exfoliated bedrock and sandy deposits derived from it. They were in a channel between two large rocks which may be either roof fall or bedrock. These layers, which lacked finds except for small amounts of bone and some charcoal, continued to a depth of 2 m below the surface. Bedrock was not reached, but we stopped excavating because the sides of the trench

110

origins of agriculture in western central asia

were extremely unstable and the gap between the two rocks was very small. Why the cave should contain so few remains is a mystery. Below the Parthian material there was little in the way of finds and no evidence of hearths or burning. There was certainly no sign of Neolithic or earlier material and also little in the way of natural bone accumulation (at least bone that has survived). Yet both the cave and the shelter are large, easily accessible, and ideally sited for the exploitation of upland plant and animal resources, as well as having ready access to the Sumbar with its valley-floor resources. We cannot offer any convincing explanation for why there is so little archaeological material in the cave.

Dorian Situated at the western end of the Kara-il-chi range northwest of Kara-Kala at c. 650 m above msl, Dorian is a large cave some 17 m wide and 7 m deep. We excavated a 1-m2 trench on the western side of the entrance. The deposit consisted of a series of ashy lenses and small hearths to a depth of 95 cm (contexts 1–12). The basal layer was a green sandy deposit, similar to the base of Karakesy 4 (see below). A considerable quantity of bone was recovered and two apparently recent sherds from the upper part of the deposits, but no artifacts were found lower down. We thus have no indication of the date of the site, but also no reason to believe that the deposits are particularly ancient.

The Parkhai Chalcolithic (Eneolithic)– Bronze Age Site This is a large open site composed of a settlement and cemetery, the latter having been excavated previously by I. N. Khlopin (1981, 1989). It is located close to Kara-Kala on the north side of the road out of the town to the west. Houses have recently been built on the site and building was continuing in 1996. The site extends over at least 500 x 300 m and its surface is covered with a variable density of bone and pottery. We hoped to find Neolithic deposits beneath the Chalcolithic layers, but, given the size of the site and the limited time available to us, the search for earlier material was likely to be fruitless unless underlying

Neolithic deposits were extensive. As time was short, we made use of an existing hole dug by one of the land owners and excavated a 2.2 x 1 m trench. The hole was 1.8 m deep before we started excavation and we removed another 80 cm of deposit, making a total depth of 2.6 m. The excavation encountered a small stream channel, and two more were noted in section above the level at which our excavation had started. Above these stream channels in the section there was a thick mudbrick foundation. It was apparent that the Chalcolithic habitation in this part of the site had been built on an area which had had some stream activity at an earlier time, making it unlikely that Neolithic deposits would be encountered here. Our excavation came down on an old surface composed of weathered bedrock. We dug into this for another 30 cm but no earlier features or artifacts were encountered. We were thus unsuccessful in finding any pre-Chalcolithic deposits, but this does not necessarily mean that they do not exist in other parts of the site.

The Karakesy Sites These sites are in a series of rockshelters on the southern side of the Sumbar valley c. 8 km due south of Kara-Kala, west of the road to the Chandyr valley. The sites are located at an average altitude of c. 550 m. in shallow limestone ravines aligned north-south and parallel to the road. They run down from a ridge to the south to a low valley in the north, and only reach a maximum depth of 10 m below the average height of the ground surface. The local rock consists of limestone interbedded with layers of clayshale. These layers are relatively soft and weather easily, causing collapses of the limestone, which may be precipitated by earthquakes. This process has resulted in a series of limestone overhangs that form rockshelters of varying sizes, many although not all of which contain archaeological deposits. We excavated five of these shelters, numbering them Karakesy 1–5 in the sequence in which we excavated them. They exhibited considerable variation in the nature of deposits and finds.

Karakesy 1 This site, at the lowest end of the ravine that is nearest to the road, is a large (c. 10 x 4 m) shelter with a relatively stable roof. The deposit was not more than



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

1.5 m above the stream which still flows through the ravine in wet weather. The upper layers consisted of sand and decayed rock, together with small hearths and ash (contexts 1–7). Context 8 was a thin (5–8 cm) layer of light grey sandy soil with ash and some charcoal. It yielded three sherds of black pottery of unknown type with geometric incised decoration. Below this context there was a series of layers (9–15) of sand and charcoal with some bone, but no other artifacts. A maximum depth of c. 70 cm was reached. We found no parallel for the pottery in any of the other sites we excavated in the Sumbar-Chandyr region, suggesting that Karakesy 1 may be earlier than the other sites we investigated.

Karakesy 2 Located higher up in the same ravine as Karakesy 1 and some 3 m above the present stream bed, Karakesy 2 contained deposits to a maximum depth of c. 70 cm. They consisted of humic layers and clays washed out of the clay-shale. Very few artifacts were found, but a bone assemblage was recovered which may, however, have derived from natural processes.

111

but that is improbable as there is no known evidence of relatively recent volcanic activity that could have deposited such ash). Context 10 separated the Parthian level from the lower earlier levels, but we have no indication of the date(s) of these earlier levels due to a lack of artifacts. Below context 10 there was 40 cm of small hearths and ashy soil lenses (11–16) that contained bone and charcoal, but no artifacts. Beneath these deposits there was a green sandy layer mixed with broken bedrock, sloping from the southeast down to the west (17 and 18). Under this was another series of small ash lenses (excavated as one context (19) due to lack of time). Finally there was a gray stony layer (20) sitting on bedrock and almost certainly a product of weathering from the bedrock. This was the deepest set of deposits in any of the Karakesy sites and probably represents a series of relatively short-term occupations that produced hearths, burnt bone, and very few artifacts. There is no indication that even the basal deposits are very old. We also found some sherds of possible Neolithic age eroding out of a bank in this ravine, but lack of time prevented us from investigating further by excavating a section in the bank.

Karakesy 3 This small shelter is located in the second ravine west of the road, several hundred meters upslope from the first ravine. Sixty cm of deposit was exposed in a 1-m2 trench, ranging from recent dung-rich layers on the surface to clayey layers weathered from the bedrock, and containing large pieces of rock lower down. Small amounts of charcoal and bone were recovered throughout, and one possible hearth was found in the basal layers at the southeast corner of the trench. The site was probably used only sporadically.

Karakesy 4 This shelter is also located in the second ravine west of the road, some 200 m north of Karakesy 3. The shelter appears small, but this is because it is almost completely full of sediment. It stretches back at least 4 m from the present drip line and its entrance is about 10 m wide, partly obscured by roof fall. The deposits reached a maximum depth of 1.70 m. The upper 80 cm consisted of soil and ash layers (contexts 1–9), at the bottom of which we found what may be a Parthian pot and two iron rings, probably also Parthian. Context 10 was a 6–10 cm thick layer of light-brown material, of fine vesicular texture and light in weight, which may be eroded bedrock (it looked like volcanic ash,

Karakesy 5 This is a small shelter, c. 2 m deep and 15 m wide, in the third ravine west of the road. We excavated a 1-m2 trench at the north end of the shelter in an area relatively free of roof fall. A maximum of 60 cm of deposit was encountered. It comprised a layer of earth and broken rock (context 2), a layer of predominantly broken rock (3), a brown clay soil (4), a gray clay (5), and a light clayey deposit (6) on bedrock. A small amount of bone was found, but no pottery. There was no definite evidence here of any human use of the shelter and the bone could be a natural accumulation. We dug a second 1-m2 trench c. 4 m out from the shelter, down the slope toward the base of the ravine. Here again a series of layers was found composed of clayey deposit and rock to a depth of 80 cm, with no indication of human presence.

Chandyr This combined rockshelter and cave is the only site we excavated in the Chandyr catchment. It is located at c. 650 m above msl some 100 m west of the road that leads south to the valley from Kara-Kala

112

origins of agriculture in western central asia

(Fig. 8.14).The shelter is approximately 8 m wide and 2 m deep, and at its southern end the cave stretches back a further 4 m. We excavated two 1-m2 trenches under the drip line of the shelter at the point where the cave starts. Trench 1 was excavated first, but when a depth of 2 m had been reached trench 2 was opened to the north to enable excavation to continue to a greater depth. Bedrock was reached at 2.5 m. The site is very similar to Karakesy 4 in that it is composed of many small hearths and ashy lenses. It was the deepest of all the sites excavated and one of the richest in terms of the bone material recovered. Pottery that may be Parthian or from a more recent period was found on bedrock indicating that the site was relatively recent.

Conclusion Despite the fact that we test excavated eight rockshelter/cave sites and one open site (Parkhai Chalcolithic) at varying locations and altitudes in the middle Sumbar and Chandyr valleys, no pre-Chalcolithic artifacts were found in the course of excavation. We also carried out fairly extensive surface surveys in the region and the only materials found that might be preChalcolithic were the few putative Neolithic sherds (mentioned above) that were observed near Karakesy 4. In all these surveys we found no stone tools of any period with definite signs of having been worked, apart from a very small number of pieces that may have been flaked. This is at variance with the reports of Lyubin (1984:28–31) and Vishnyatsky (1996:37–43; 1999:77–78), who reported surface finds of Palaeolithic pebble cores, choppers, side scrapers and crude flakes in the Sumbar and Chandyr valleys. Our failure to find any Palaeolithic, Mesolithic, or Neolithic material was both surprising and disappointing, and we are at a loss to explain why we recovered no early material. Three possible explanations present themselves, none of which is wholly convincing. The first and least likely is that the areas we surveyed were unoccupied prior to the Chalcolithic period. Given that many other parts of Turkmenistan were occupied in the Palaeolithic, Mesolithic, and Neolithic periods, as well as neighboring areas in Iran and Afghanistan (this volume: Chapter 6), it seems highly improbable that there were no pre-Chalcolithic settlements in the areas we examined, especially as the climate is relatively mild (and was probably more

so in the mid-Holocene) and the diversity of natural resources would have been greater here than in most of the rest of Turkmenistan. Nor, as the middle Sumbar and Chandyr valleys were not glaciated, would occupation have been precluded during Pleistocene glaciations. The second possibility is that the rockshelters and caves in the region were formed in post-Neolithic times. Given the amount of tectonic uplift and seismic activity in the Kopetdag, it is quite possible that small rockshelters and caves are relatively rapidly created and destroyed and that the present ones are not more than a few millennia, or even only a few centuries, old. Added to this is the probability that rates of erosion and redeposition are high due both to the tectonic uplift and to the susceptibility to erosion of some of the sediments, particularly the loess that occurs as a superficial deposit throughout the area. This could have resulted in early open sites—if they existed in the past—being deeply buried by recent deposits. The third possibility is that earlier sites do exist but that we failed to find them, either because we looked in the wrong places or because our excavations were on too small a scale. This remains a possibility, but one that is difficult to assess without more fieldwork in the region over longer periods of time. In addition to the need for more extensive surveys and further test excavations, there is the problem that stone tools may be very difficult to recognize owing to the nature of the local rock, much of which is soft limestone that tends not to have good flaking properties. It readily erodes, so what may once have been well-flaked stone tools may now resemble natural stones, having over time lost all indications that they had been struck and used. As already suggested, none of these hypotheses is entirely convincing. It seems most probable that a combination of the rockshelters and caves being recent features, and our sampling being limited, provides the best explanation. If the shelters and caves are rapidly created and destroyed, ancient ones may be scarce or non-existent and our chances of discovering them therefore low. Many of the sites that we excavated looked very similar in character, with a series of small hearths and ash lenses indicating low-level, perhaps intermittent, occupation over some period of time. We can only hope that future archaeological research in the region will throw more light on these conundrums. It was disappointing not to obtain new evidence of prehistoric settlement in the Sumbar and



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

Chandyr valleys, but their possible significance in relation to the beginnings of agriculture in the region is nevertheless discussed in Chapter 11.

The Bolshoi Balkhan Region, 1997 The field season we conducted in March and April 1997 in the Bolshoi Balkhan region (Fig. 8.13) had two main aims: to recover archaeological and environmental evidence pertaining to the Mesolithic– Neolithic transition in the region, and to assess the possibilities for future work in the area. The local environment of the Bolshoi Balkhan massif has already been described in Chapter 4, and the history and results of previous investigations there by Soviet archaeologists, principally A. P. Okladnikov and G. E. Markov, have been outlined in Chapters 5 and 6. Three main rockshelter sites were excavated and the results reported by these pioneers: Jebel (Djebel), which was excavated by Okladnikov between 1947 and 1950 and which established the first Mesolithic–Neolithic sequence for Turkmenistan (Fig. 8.15, color); and the Dam Dam Cheshme (DDC) rockshelters 1 and 2 some 20 km to the southeast of Jebel (Fig. 4.1). DDC 1 and 2 were excavated by Markov in the 1960s following preliminary investigations by Okladnikov in the late 1940s and early 1950s, and they yielded Mesolithic–Neolithic sequences comparable to the one at Jebel. In 1997 our initial aim was to recover direct evidence of past subsistence at these sites in the form of plant and animal remains that could provide AMS dates and establish a secure chronology for the Mesolithic–Neolithic transition. However, we had to modify this strategy in the field once we saw how extensive the Russian excavations had been and how much material had already been removed from these key sites. Nevertheless, we were able to carry out small-scale excavations at DDC 1 and 2 and at two smaller rockshelters nearby, DDC 3 and 4. We also located and excavated test pits at three other sites in the region and visited a fourth (see the end of this chapter for brief descriptions of them).

Excavations at the Dam Dam Cheshme Rockshelters Given that our primary aim was to recover and date plant and animal remains that would illuminate

113

the Mesolithic–Neolithic transition, we needed to locate areas of cultural deposit left unexcavated by Okladnikov and Markov at DDC 1 and 2, and, ideally, also find and excavate pristine occupation layers at DDC 3 and 4. By excavating deposits remaining in DDC 1 and 2, and recovering stone tools and potsherds as well as organic remains, we hoped also to be able to link our new data to Markov’s published stratigraphies. However, as the following descriptions of our investigations show, we failed to achieve these objectives, although our excavations at the DDC sites and our reconnaissance surveys in the area did add to the archaeological record of the Bolshoi Balkhan region.

Dam Dam Cheshme 1 DDC 1 is a large rockshelter situated at the base of a limestone cliff that is part of the dissected southwestern escarpment of the Bolshoi Balkhan massif (Fig. 4.1). The entrance is visible from a distance and commands an extensive view over the piedmont that slopes down to the channel of the former Uzboi river (Figs. 8.16, 4.3, color). The shelter consists of a semicircular recess under the overhang of the cliff, c. 50 m wide at the entrance and extending inward c. 15 m to the apex of the semicircle. Today its floor consists mainly of exposed and weathered bedrock and there are two pools near the back of the shelter that are fed by seepage and by occasional waterfalls from an eroded nick in the cliff face above (Figs. 8.17, 8.18, color). DDC 1 was first investigated by the geologist V. V. Shumov who dug a test pit in it. In 1947 Okladnikov excavated a 10 x 1 m trench next to Shumov’s pit across the shelter and identified six cultural layers (Okladnikov 1951:97–100). Markov subsequently excavated a much larger area in the shelter and found only five layers in the center. He inferred, from the typology of the stone tools and pottery, that the cultural sequence as a whole extended from the Upper Palaeolithic to the early Bronze Age, and that the transition from the Mesolithic to the Neolithic occurred in layer 4 (Markov 1981). In 1997, we surveyed the interior of the shelter and excavated a first (2 x 1 m) test pit at the rear where the roof was close to the present surface. The test pit was dug to a depth of 70 cm. In it, superficial dusty deposits gave way to compacted damp gray clay, with occasional streaks of yellow mineralized material that proved to be weathered bedrock. The same situation

114

origins of agriculture in western central asia

was found in a second (1 m2) test pit that we dug 1 m to the southwest of test pit 1. There, after excavating through a meter of weathered rock in various states of decomposition, we augered to a total depth of 2.70 m through weathered rock without encountering any archaeological deposits. We dug a third (2 x 1 m) test pit, between the two pools at the rear of the shelter, into a deposit that proved to be backfill from previous excavation, as was the deposit in a fourth (1 m2) test pit that we excavated at the eastern edge of the shelter. Thus all four test pits demonstrated, as Markov’s brief account (1981) of his excavation had suggested, that almost all the cultural deposits in DDC 1 had been excavated prior to our visit. We therefore decided to abandon the site and concentrate our efforts on the other DDC rockshelters.

Dam Dam Cheshme 2 DDC 2 is a deep rockshelter a few minutes’ walk northeast of DDC 1 (Fig. 8.19, color). It is situated at the base of a cliff on the eastern side of a channel cut by an intermittently flowing stream that emerges from a canyon in the mountains. A waterfall that flows after rain spills from the overhanging cliff onto the floor of the shelter, which is approximately 50 x 50 m in extent and occasionally used by shepherds as a pen for livestock. Between 1949 and 1952 Okladnikov (1953:30) excavated several test pits, including a 40-m2 trench in 1952, but the main investigations were undertaken in 1963 and 1964 by Markov (1966a, 1966b) who excavated over 250 m2 at the rear of the shelter (Fig. 8.20). He reached bedrock at a depth of 3.2 m in the central area and over 5.0 m in the southwestern part. Markov recognized 23 cultural sub-layers which he grouped into 9 layers on the basis of changes in the color and texture of the sediments, the presence of undisturbed sterile and mineralized layers, and differences in the nature of the archaeological finds. In total, the finds amounted to 11,060 stone artifacts, several thousand fragments of bone, 609 fragments of pottery, a considerable quantity of perforated shells, and 2 bone needles. Pottery was restricted to the upper four layers, with just a single sherd in layer 5 (this volume, pp. 206–7). Markov ascribed the lower layers, devoid of pottery, to the Mesolithic, except for the lowest, layer 9, which he regarded as possibly Upper Palaeolithic because it lacked geometric stone tools. He ascribed the upper layers to the early Neolithic (layer 4), the late Neolithic and Eneolithic (layer 3), and the Bronze Age (layers 2 and 1).

Area 2

Area 1 0

10

20 Metres m

8.20 Dam Dam Cheshme 2 showing the locations of Areas 1 and 2 partially excavated in 1997, and (in gray) the approximate combined extent of the areas previously excavated by Okladnikov, Markov, and Khamrakuliev; the dotted line shows the outer limit of the overhang at the rear of the rockshelter.

Okladnikov’s and Markov’s excavations, and a small trench dug in 1978 by S. Khamrakuliev (1979), resulted in the removal of a massive amount of sediment from the shelter and the formation of spoil heaps at its entrance. As a result, little undisturbed archaeological deposit remains. It is concentrated at the rear of the shelter under an overhang where its surface is only a few cm below the roof. Most of the remaining deposit forms a north-south baulk c. 20 m in length and from 0.5 m to 1.5 m in height. It is divided into two parts by a former excavation trench that cut inward to where bedrock reaches the roof. We decided to excavate two areas on either side of the gap formed by the trench, which we labeled Areas 1 and 2 (Fig. 8.20). Area 1 (Fig. 8.21) was longer (c. 5.8 m) than area 2 (c. 2.4 m), and because both areas consisted of parts of the deposit that had originally extended farther across the back of the shelter, they were difficult to interpret, but the deposits we excavated clearly preserved the general stratigraphy of this part of the site. The layers in both areas were composed of silts and sands, the former probably derived from wind action and the latter from the waterfall above the shelter. Small quantities of artifacts had been incorporated



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

115

well sorted, others less so. No fragments of charcoal or gypsum were present in these layers. There were few finds and none in the lowest four contexts. As in Area 1, there was little evidence in Area 2 of human activity. Many of the deposits were washed in or had weathered from the surrounding walls of the shelter and mixed with occasional artifacts that found their way to the back of the shelter. Most of the deposits are probably relatively recent, as is suggested by the two AMS radiocarbon dates we obtained: one (OxA–12548) on sheep/goat bone from context 3 and one (Beta–172095) from a charred barley grain from context 102. They yielded dates respectively of 3191 ± 35 BP or between 1530 and 1400 cal. BCE, and 2850 ± 50 BP or between 1190 and 900 cal. BCE (and see this volume, p. 198 and p. 201).

8.21 The south-facing section of Area 1 in Dam Dam Cheshme 2 prior to excavation, April 1997; the base of the deposit can be seen tapering upward on bedrock toward the back of the rockshelter. (Photo by Patrick Blackman)

Dam Dam Cheshme 3

into the deposits, presumably as a result of people engaged in activities in the front of the shelter throwing rubbish toward the rear. All the material excavated was dry-sieved through a 5-mm mesh and samples for wet-sieving were taken from most contexts. In Area 1 we encountered a series of yellowbrown silts, fine in texture and containing few finds (contexts 2–5). Underlying these were yellow sands which graded into clay (6–9) and these were overlaid by sands with a greater admixture of gravel (10–13). We then encountered a series of clays, often finely laminated, that probably derived from weathering of the walls of the shelter (14–22) and which were underlain by yellowish sands with some silt (23–26). At the base of the deposit there was a lens of fine greenish silt (27) overlying compact gray sand (28) that lay on bedrock. Sporadic finds of flint, pottery, and bone were made throughout the deposit, and, with the exception of the stone artifacts most of which came from context 22 (this volume, p. 204), there were no concentrations of finds. Nor was there direct evidence of human occupation, in the form of hearths or ashy layers, in this part of the shelter. In Area 2 similar sets of layers were encountered, but they were not as deep. The uppermost layers were composed of brown and yellow sands with inclusions of limestone and some gypsum and charcoal flecks (contexts 101–109). Below these the deposits became stonier and more silty, with increased numbers of limestone fragments (109–114). Some deposits were

DDC 3 (Fig. 8.22, color) is a small rockshelter located a short distance north of DDC 2 on the western side of the same canyon. It had previously been test excavated, probably by Markov. We excavated two trenches in the shelter in the hope of finding evidence of, and dating, its occupation. Trench 1 was excavated adjacent to a former trench (perhaps dug by Markov), and trench 2 was placed slightly farther to the northwest and excavated through undisturbed sediment. Both trenches ran from the mouth of the shelter toward the rear. Trench 1 cut through a series of very complex deposits made up of small hearths interspersed with rocks, which were especially prevalent at the front of the shelter and are probably the result of roof fall. Trench 2 revealed a series of layers running horizontally across the shelter. The main contexts recognized in our excavations of trenches 1 and 2 are described in Appendix 8.5. The excavation evidence, especially the hearths, indicates that the shelter was frequently but not intensively used. Apart from the hearths, there was very little evidence of occupation: only pottery (including two sherds from wheel-turned vessels probably of Iron Age or later date), 10 stone artifacts, and very small quantities of charred plant remains, wood charcoal, and animal bone (see this volume: Chapter 10, Sections 1–5). We obtained one AMS date (OxA–12546) on a fragment of sheep/goat bone from context 52: 2978 ± 27 BP or between 1320 and 1110 cal. BCE (and see this volume, p. 201). This date, and the presence of potsherds throughout the deposits, suggests relatively recent use of the shelter. We found no evidence

116

origins of agriculture in western central asia

of Mesolithic occupation, and the general lack of archaeological material implies that it was not occupied continuously for long periods. Instead, it may have functioned as a campsite for hunters, pastoralists, and other people who migrated between the piedmont and the mountains.

Dam Dam Cheshme 4 DDC 4 is a small rockshelter at the base of a cliff at the head of the canyon in which DDC 2 and 3 are located. A test excavation revealed a series of sands and gravels that appear to have been deposited as a result of water intermittently flowing down the face of the cliff from a waterfall above the cave. No artifacts were recovered and we found no other evidence of human use of this shelter, which appeared to have been test excavated previously, possibly by Markov.

Reconnaissance Survey for Other Sites Following the generally disappointing results of our excavations at the DDC sites, we devoted the rest of the 1997 field season to searching for other sites that might offer better prospects for future research on the Mesolithic–Neolithic transition in the region. Previous surveys by Russian and Turkmen archaeologists had demonstrated the existence of several rockshelter/cave and open sites around the margins of the Bolshoi Balkhan, and in the time available to us we located and briefly investigated three sites along the southern flank of the massif and one north of the mountains (Fig. 4.1). Two proved to be rockshelter (Bashkovdan) and cave (Charla’uk) sites and two were open-air sites (Joyruk and Oyukli). An attempt was made to visit a fifth site located close to the village of Adjikuli on the eastern side of the massif (Fig. 4.1), where Vishnyatsky found three deflated Mesolithic/ Neolithic sites in the 1980s, but we were unable, for lack of time, to examine them.

Bashkovdan The name Bashkovdan (“the five pools” in Turkmeni) refers to a channel in a canyon that penetrates the southern escarpment of the massif for some 2 km. About 200 m from the head of the canyon there are two rockshelters, one above the other. The upper shelter is impressively large: approximately 50 m wide, 35 m deep and 35 m high. Water seeps over its inner walls and is channeled into several pools surrounded by fig trees.

Most of the floor consists of bedrock but there are also several patches of sediment on it. A 2 x 1 m test pit was dug in one of the patches in the southern part of the shelter. At the top of the deposit light-gray silt was underlain by a silty clay with occasional pebbles that became more abundant in the next 8–12 cm, where some specks of charcoal were encountered. At 40 cm the stones, including angular rock fragments, were still more abundant, and at 65 cm waterlogged clay weathered from the bedrock was reached. Two more test pits were excavated in the center of the shelter revealing gray sand to a depth of 25 cm, where bedrock was reached. No archaeological material was found in any of the test pits or in our exploration of the cave. The lower shelter is more exposed, smaller (approximately 15 m wide, 10 m deep, and 10 m high) and also contains a pool. Bedrock forms most of its floor, but a 1.0 x 0.5 m test pit dug in an area of sediment went through a light-brown silt with gravel and rock fragments to a depth of 80 cm, where excavation was ended. At a depth of 30 cm several potsherds were found which probably came from a single, possibly medieval, vessel (this volume, p. 207). A third shelter was located some 300 m farther down the canyon on a meander of the stream channel. It is about 25 m wide, resembles DDC 2 in appearance, and is used as a sheep pen. There were signs, in the form of what appeared to be two partly backfilled pits, that the sediment on the floor of the shelter had previously been text-excavated. A small test pit dug by us in another patch of sediment went through sheep droppings and 15 cm of gray sand on top of bedrock, but no archaeological remains were found.

Charla’uk This site is a cave situated about 1 km up a short canyon in the escarpment east of Bashkovdan close to a waterfall at the head of the canyon. The cave entrance is a horizontal fissure in the cliff some 3 m wide and 0.5 m high. The cave extends from the fissure into a circular chamber c. 30 m deep and 25 m wide, with a roof that reaches a maximum height of 4 m. A spring at the northern end of the cave flows into a shallow pool in the center. The cave is totally dark and inhabited by bats, and we could only excavate with light from a generator. Two test pits were dug. The first was a 1-m2 pit near the eastern edge of the pool. At the surface there was a 2–3 cm layer of yellow sand underlain by a 5-cm layer of red-brown material identified as bat droppings. Below this was a 10–15



jeitun, sumbar and chandyr valleys, and bolshoi balkhan region: excavation & survey

cm layer of grayish sand, which gave way to yellower sand containing pieces of stone that had probably fallen from the roof. Bedrock was reached at a depth of 70 cm. The second test pit (2 x 1 m) was dug about 4 m east of the first. Its stratigraphy closely resembled that of the first pit, except that the layer of yellow sand containing rocks was underlain by some 2 m of sand. At a depth of 2.5 m we decided to end the excavation. Regrettably, we found no traces of ancient occupation in either of the test pits or in our exploration of the cave and its surroundings.

Oyukli Oyukli, situated north of the massif (Fig. 4.1), is an open-air site which had previously been visited by several Russian archaeologists, notably Markov (1961; Markov and Khamrakuliev 1980). When we visited the site it extended over an area of about 50 x 50 m on a deflated sand dune. Two 1-m2 test pits revealed sand to a depth of 1 m. The site appears to have no stratigraphy and all the archaeological material we observed—scatters of stone tools and potsherds (this volume, p. 207)—was on the surface, which accords with Markov’s report. Oyukli is probably a deflated Neolithic site.

Joyruk Joyruk (Djoyruk) is an open-air site located on the top of a high (15–30 m) sand cliff north of a large meander of the Uzboi channel close to the southeast corner of the massif. It consists of a series of smaller sites that extend for several hundred meters along the cliff top. For lack of time, we were only able to visit part of the site, which appeared to be deflated, and we were

117

unable to carry out any excavations. A rapid surface collection yielded 10 flint artifacts and a few potsherds (this volume, p. 207) . No structures or stratified deposits were observed, but this large and complex site would probably repay further investigation.

Conclusion Two factors made our investigations in the Bolshoi Balkhan mountains less productive than we had hoped. First, the scale of earlier Russian excavations at both DDC 1 and 2 had left little unexcavated deposit. The deposit that remained was outside the main areas of occupation of these rockshelters and yielded little evidence of human activity. Our excavations at DDC 3 produced archaeological material that was more recent than the Mesolithic–Neolithic transition, which may be why it had not been fully excavated by the Russian archaeologists, whose interest, like ours, also focused on the Mesolithic and Neolithic periods. The second factor is that there may be relatively few undisturbed early deposits in this climatically and geomorphologically volatile landscape, as seems to be indicated by our brief investigations of other sites in the Bolshoi Balkhan region. Many caves and rockshelters either never accumulated deposits or have had them subsequently eroded. It is also evident that Russian archaeologists had carried out extensive surveys in the area, and if they had found sites with early stratified deposits, they would have reported them. The best hope for future excavations of Mesolithic and Neolithic sites appears to lie in the open sites that are known to exist along the former course of the Uzboi.

9

Jeitun: Dating and Analysis of Excavated Materials

I

n this chapter, the materials excavated and sampled at Jeitun by the British team are described and the results of their analyses are presented. These reports combine presentation of technical data (some of which is contained in appendices) with interpretation of the results, and they form the basis for the conclusions about the nature of the site and the history of its occupation that are summarized at the end of the chapter. The radiocarbon chronology that we established for Jeitun is described first, followed by sections on sediments and soils, plant and animal remains, stone tools, and pottery.

Section 9.1 Dating the Site: Radiocarbon Chronology with Chris Gosden and John Meadows Prior to our excavations no radiocarbon dates had been determined for Jeitun, and obtaining an AMS radiocarbon chronology was therefore one of our principal aims when we began research at the site. Masson had inferred, mainly on the basis of similarities between the ceramic assemblages he excavated at Jeitun and those found at the Southwest Asian sites of Jarmo and Tepe Guran, that Jeitun had been occupied in the 6th millennium BCE (Masson and Sarianidi 1972:36, 171), and his division of its stratigraphy into five phases implied a long-lasting occupation. Expectations of longevity for Jeitun-Culture sites were reinforced by the investigations by Berdiev and others between 1963 and 1973 of substantial tell sites on the Kopetdag piedmont, two of which, Chopan and Togolok (although not adequately dated), were reported to have Early Jeitun levels (this volume, pp. 60–61). Such expectations were also derived from excavations at Neolithic sites in the Levant and Turkey, many of which had demonstrated long-term (although discontinuous) occupation starting in the

late Pleistocene or early Holocene. From the outset of our research at Jeitun we regarded AMS radiocarbon determinations as key to understanding the site itself and when agriculture was practiced there. In 1990, we submitted 5 samples for dating, followed in 1992 and 1993 by a further 6, all taken within the site (Harris et al. 1993:330; Harris, Gosden, and Charles 1996:437). The samples came from a variety of locations and deposits, partly from features associated with buildings and yard areas and partly from our test pits. All 11 dates were obtained from individual cereal and weed seeds or chaff fragments, which established incontrovertibly that wheat and barley were being cultivated at Jeitun by c. 6000 cal. BCE. In 1994 we obtained 3 more AMS dates from charcoal in three buried soils (palaeosols) that revealed traces of two artificially cut ditch-like features, one above the other, exposed in the side of a modern irrigation ditch adjacent to the site (see below and this chapter: Sections 9.3 and 9.4). All 14 samples were processed at the Oxford Radiocarbon Accelerator Unit, University of Oxford, UK.

120

origins of agriculture in western central asia

Table 9.1 AMS radiocarbon results obtained on-site at Jeitun. All samples consisted of charred remains of cereal or weed seeds and/or chaff, identified where possible. The second column shows inferred relationships of the 1990 samples to Masson’s building levels/phases of occupation (I–IV) and of the 1992/93 samples to numbered contexts identified by the British team. Uncalibrated (BP) dates are conventional radiocarbon ages (Stuiver and Polach 1977). Calibrated date ranges were obtained by the maximum intercept method (Stuiver and Reimer 1986), using OxCAL 3.10 software (Bronk Ramsey 1995, 1998, 2001, 2005) and the INTCAL 04 calibration data (Reimer et al. 2004). On-site Samples

Calibrated Dates (cal. BCE)

Laboratory no.

Probable phases (I–IV) and context nos.

Material dated

δ13C (‰)

Radiocarbon ages (BP)

1 sigma (68% confidence)

2 sigma (95% confidence)

OxA-2912

I

T. monococcum

-24.1

7100 ± 90

6060–5890

6210–5770

OxA-2913

II/III

T. monococcum

-25.6

7180 ± 90

6100–5980

6230–5880

OxA-2914

III/IV

T. monococcum

-25.2

7270 ± 100

6240–6020

6380–5920

OxA-2915

IV

T. monococcum

-24.3

7200 ± 90

6210–5990

6240–5890

OxA-2916

below IV

T. monococcum

-25.7

7190 ± 90

6210–5990

6240–5890

OxA-4690

12

Aegilops sp.

-25.1

7035 ± 65

6000–5840

6030–5740

OxA-4691

17

indet. seeds

-20.7

6850 ± 65

5780–5660

5890–5630

OxA-4692

37

Aegilops sp.

-24.5

7025 ± 70

5990–5840

6030–5730

OxA-4693

TP4

Aegilops sp.

-25.9

7000 ± 70

5990–5790

6020–5720

OxA-4694

111

Aegilops sp.

-24.4

7125 ± 70

6060–5920

6100–5880

OxA-4695

TP7

indet. cereal grain and chaff

-24.7

7170 ± 70

6080–5990

6220–5900

9.1 Calibration of the Jeitun radiocarbon results (Table 9.1) by the probability method (Stuiver and Reimer 1993), using the INTCAL 04 data (Reimer et al. 2004).



jeitun: dating and analysis of excavated materials

121

9.2 Simple Bayesian model of on-site radiocarbon dates from Jeitun, implemented in OxCAL 3.10 software (Bronk Ramsey 1995, 1998, 2001, 2005). Distributions in outline represent simple calibrations of the radiocarbon dates by the probability method, as shown in Figure 9.1. The solid distributions are “posterior density estimates” generated by the model that indicate the probable calendar dates of samples and events obtained by combining the radiocarbon dates with the relative dating information incorporated in the model structure. Note the satisfactory overall index of agreement (A=91.8%, the critical value of A being 60.0%), which indicates that the radiocarbon results are consistent with the relative dating of samples implied by the model structure.

The On-site Dates The 11 on-site dates are given in Table 9.1 with the calibrations shown at one and two standard deviations. Their interpretation can be further refined by looking at the probability that the true age of a sample falls on a particular date (Fig. 9.1). Table 9.1 and Figure 9.1 show that the dates of all the samples probably fall between 6300 and 5600 cal. BCE, but it is likely that the overall timespan represented by these samples is relatively short. In an attempt to define more precisely the period during which Jeitun is likely to have been occupied, we have used Bayesian statistical methods to model the

probable duration of occupation. The model makes no assumption about the age order of the 11 samples dated, but it does assume that they are representative of a single, continuous period of activity. This assumption, implemented in the OxCAL 3.10 computer program using phase boundaries (Fig. 9.2; Bronk Ramsey 2000), reduces the dispersion of the calibrated dates, some of which is caused by the statistical scatter in radiocarbon measurements. The satisfactory overall index of agreement obtained (A=91.8%; the critical value of A, 60%, is analogous to the 5% significance level in a χ 2 test: Bronk Ramsey 1995) indicates that the radiocarbon results are consistent with the model structure, and do not suggest discontinuity of occupa-

122

origins of agriculture in western central asia

that we hoped would provide a phased chronological sequence, which differentiated the two superimposed houses and also levels within them, and which might test Masson’s sequence of five discrete building levels and inferred phases of occupation. The 1992/93 samples came from the following numbered test pits and contexts.

OxA-4695 This sample was obtained from a burnt area beneath remains of mudbrick architecture in test pit 7 at the eastern end of the site. It was taken at c. 4.70 m below 9.3 Duration of site occupation, derived from the model shown in Figure 9.2. datum in one of the lowest levels reached in any excavation at Jeitun and it may therefore date some of the tion of the site. The model estimates that occupation earliest use of the site. spanned 100–340 years at 68% probability or 30–440 years at 95% probability (modeled date estimates are conventionally italicized) (Fig. 9.3). OxA-4693 Although Masson (following Kuftin’s and MarThis sample was taken from the lowest charcoal uschenko’s soundings of the Jeitun mound in 1952, lens of a small hearth encountered in test pit 4 in the this volume, p. 196) recognized five phases of occentral area of the site. The hearth was below remains cupation based on what were interpreted as succesof mudbrick architecture attributable to Masson’s sive levels of building construction, the radiocarbon phases 2 and 3 and it probably represents an earlier dates do not indicate either a long or a discontinuous phase in the life of the settlement. period of time. In fact if OxA-4691, which is slightly later than the other dates, is omitted, the results are OxA-4694 statistically consistent with a single radiocarbon date This sample came from context 111, which was (T’=11.0, T’(5%)=16.9, ν=9); Ward and Wilson 1978). the lowest deposit in the oven of House A and conAlthough we can be confident that the radiocarbon sisted of a layer of sand containing lumps of mudbrick, results accurately date the cereal and weed seeds and burnt clay, and charcoal. chaff fragments themselves, the taphonomy of these samples is inevitably somewhat uncertain. We can OxA-4692 use the results to infer the dates of the depositional The source of this sample was context 37, a contexts from which they were recovered, but strictly sand layer near the oven and the adjacent platform of speaking the dates only provide maximum ages for House A, which contained many figurines, potsherds, those deposits. and flint blades. The sand may represent a period of The five samples (OxA-2912–6) obtained for datdisuse in the life of the house after the initial floor ing in 1990 (before the British team undertook their layers were laid down. own excavations) lack context descriptions, and given their uncertain relationship to Masson’s phases, they OxA-4691 are not discussed further. In selecting samples during This sample was taken from context 17, which our excavations in 1992 and 1993 we chose contexts lay immediately beneath context 11 and consisted of a



jeitun: dating and analysis of excavated materials

layer of brown clay and sand. The layer contained several potsherds and flint blades and had several holes cut into it, some of which may have been postholes remaining from an ephemeral structure built after House A went out of use.

OxA-4690 This sample came from a small hearth (context 12) within one of the upper mudbrick destruction layers (11) of House A. Context 12 was one of three ash lenses within context 11 that derived from small fires. The fires were probably lit while the building was out of use and they may be the result of ephemeral activity that took place soon after the destruction of House A. The main conclusion to be drawn from the onsite dates is that Jeitun was occupied for, at most, 300 or 400 years, and possibly for only 100 or 200 years. The dates suggest that very little time elapsed between the first indications of human activity on the site— the hearth encountered at c. 4.70 m below datum in test pit 7 which appears to pre-date any mudbrick architecture—and the construction of the uppermost buildings, probably soon after 6000 cal. BCE. The results do not allow discontinuous occupation phases (or discrete building levels) to be distinguished.

Interpretation of the On-site Chronology When we consider the layout and structure of the site as a whole in relation to the radiocarbon dates, it is clear that a large number of houses were built over a short period of time, with many superimposed on each other; but it is impossible, from the results of our admittedly limited excavations, to provide firm evidence for discrete levels or phases of building or of changes in artifact forms (and see Section 9.12). The superimposed houses that we investigated are separated by layers of mudbrick debris, presumably derived from the decay or destruction of the buildings, and also by layers of windblown sand. The sand layers could have accumulated very rapidly, but they nevertheless indicate that occupation on any one area of the site was not entirely continuous. This inference is reinforced by the fact that houses were generally built on top of one another, although the walls of new houses were not aligned directly with or built on top of the walls of earlier ones, as was done for example at Neolithic Çatalhöyük in Turkey (Hodder and Cessford 2004). There are also indications of more ephemeral use of the site, such as sets

123

of postholes (e.g., those cut into context 17 mentioned above in relation to OxA-4691) that may signify shortlived structures erected and used after the destruction of one house and before the construction of another. These types of evidence suggest very complicated patterns of on-site activity that only further fine-grained, large-scale excavations could more fully reveal. Several possibilities arise from these observations. First, that the occupants of Jeitun lived there continually throughout the site’s occupation, but rebuilt abandoned or collapsed houses rapidly. There are ethnographic examples of such behavior, where a house might be abandoned on the death of important occupants and then rebuilt, or where the occupancy of a house may be tied to the life-cycle of the group living in it (Boivin 2000). A second possibility (which is not however supported by the modeled radiocarbon dates) is that Jeitun was occupied, abandoned, and re-occupied by sedentary populations at successive intervals throughout the history of the site. Third, it may have been occupied only intermittently as one place visited in a seasonal subsistence cycle that encompassed other sites, or, in a variation of that pattern, part of the population may have lived there fulltime while others followed a more mobile existence for part of the year. The question of the temporal nature of occupation at the site—in particular whether Jeitun was a fully sedentary settlement, as has generally been assumed—is discussed more fully in the final section of this chapter and summarized in Chapter 11, but there is no doubt that the 14 radiocarbon determinations we obtained enable Jeitun to be placed for the first time within an absolute chronology, and that this allows comparisons to be more securely made with other Neolithic sites in Central and Southwest Asia and elsewhere.

The Off-site Dates Table 9.2 shows the calibrated BCE and uncalibrated BP values of the three off-site AMS dates we obtained (OxA-4914, OxA-4915, OxA-4916). They came from charcoal in buried-soil horizons exposed in the side of the modern irrigation drainage ditch close to the eastern edge of the site (Fig. 3.1), and they date the three poorly developed (immature) palaeosols and the two associated ditch-like features already mentioned. The lower “ditch” was cut from the A horizon of the middle palaeosol (Soil II), and the upper one, which

124

origins of agriculture in western central asia

Table 9.2 AMS radiocarbon dates from charcoal recovered from three palaeosols exposed in the side of a modern irrigation drainage ditch close to the eastern edge of the Jeitun mound. Section datum (present ground surface) at 182.7 m above msl, 7.3 m below the site datum. Calibrations according to the INTCAL 04 curve (Reimer et al. 2004) calculated using OxCAL 3.10 software (Bronk Ramsey 1995, 1998, 2001, 2005). Off-site Samples

Calibrated Dates (cal. BCE)

Laboratory no.

Depth in m below section datum

Uncalibrated dates BP

1 sigma (68% confidence)

2 sigma (95% confidence)

OxA-4914

0.16–0.20

6940 ± 100

5970–5720

6010–5650

OxA-4915

0.51–0.65

7080 ± 65

6020–5890

6070–5800

OxA-4916

0.85–1.00

7140 ± 220

6230–5800

6450–5600

was shallower and wider, was cut from the Ag horizon of the upper palaeosol (Soil I). In addition to the cut features, all three soils contain evidence of human activity in the form of scattered charcoal. Potsherds and fragments of bone were also present and very small pieces of mudbrick were observed (in thin sections). These materials are likely to have been deposited during cultivation of the soils. However, Wilkinson’s measurements of magnetic susceptibility (this chapter: Section 9.4) do not provide evidence of burning or other intensive human activity, and he infers that the charcoal (including the samples dated) was not produced in situ by the combustion of organic material but originated in the settlement and was re-deposited in the palaeosols by wind blow and/or overland flow during intense rain storms. The discovery of the palaeosol sequence and especially of the ditch-like features is important for our understanding of the site and the agricultural use of areas around it. It seems likely that the ditch-like features were used for purposes of water management, possibly irrigation if precipitation was insufficient or too irregular for rainfed cereal cultivation. Unfortunately the archaeobotanical evidence from Jeitun is inconclusive on the question of whether the cereals were rainfed, irrigated, or grown on areas of high water table, or by some combination of these methods (see Section 9.6). Nor could stratigraphic connection between the upper ditch and the Jeitun mound be definitely established because a large ridge of spoil from the modern irrigation ditch covered the area between them, and to determine whether it did connect with the settlement would have required much more extensive excavation than we were able to undertake.

Despite these difficulties, the first step in assessing the relationship between the ditch-like features and the site is to determine whether they were contemporary. Although the relatively low amounts of carbon present in two of the samples that were dated resulted in larger error ranges, the fact that the dates accord with the stratigraphic sequence, from the oldest at the bottom to the youngest at the top, and even more significantly that they all fall within the date ranges for the site itself, indicates that the paleosols and the ditch-like features are indeed contemporary with the occupation of Jeitun. Furthermore, the immaturity of the soils and their separation by deposits of windblown sand suggests that areas of cultivable land around Jeitun suffered sand encroachment and that cereal cultivation took place in an environment of short-term environmental instability. This phenomenon can be seen as part of the wider context of the location of Jeitun in an unstable environment subject to sudden f loods and sand storms where the alluvium of the terminal fan of the Kara Su is overlain by the sands of the southern margin of the Karakum. The palaeosol sequence shows that local accumulation of windblown sand was a discontinuous process, interrupted by pauses sufficiently long for some soil development to occur. Evidently this happened at least three times within the brief period during which Jeitun was occupied, and it is likely that the topography and vegetation around the site also altered in other ways that affected its inhabitants, such as changes in the deposition and erosion of sediments caused by short- and long-term fluctuations in the discharge of the Kara Su, and perhaps as a result of overgrazing by domestic sheep and goats near the settlement.



jeitun: dating and analysis of excavated materials

125

D

uring the 1990–1994 field seasons at Jeitun, Susan Limbrey undertook geoarchaeological investigations of sediments and soils both at the site and in its environs. She closely examined a sequence of deposits exposed in an excavated section near the center of the mound; sampled fragments of mudbrick and mortar from the remains of a wall of one of the Neolithic buildings on the site and from a house floor; studied a sequence of deposits exposed in the side of a modern irrigation drainage ditch close to the eastern edge of the mound where she found evidence of three buried-soil (palaeosol) horizons; and took samples for phytolith, particle-size and chemical analysis, and micromorphological examination. In Sections 9.2 and 9.3 she reports first on the on-site deposits, including the samples of brick, mortar, and floor material, and second on the off-site sequences. Also, in 1993 Keith Wilkinson took samples for magnetic-susceptibility measurement from the off-site sequence of palaeosols and excavated part of one of two ditch-like features that Limbrey discovered in the sequence. He describes his investigation in Section 9.4.

Section 9.2 Yard Deposits and Building Materials at Jeitun Susan Limbrey Deposits were examined and sampled down a section in the north face of an area excavated earlier by V. I. Masson and identified as a yard between buildings. This face formed one side of the northeast extension, which was excavated back from it in 1991 (Fig. 8.2, trench 1991a), and context numbers were assigned. Not all the contexts identified in excavation reached the section, and some contexts of lesser extent occur within others, so for the section the number sequence has gaps and overlaps. Study of the deposits was aimed at understanding them in terms of materials and mode of deposition, and was complementary to phytolith analysis and the study of charred plant remains and animal bones (Sections 9.5, 9.6, 9.7, and 9.9). Particle-size analysis was used to characterize the mineral components, and micromorphological study to examine microstructure and the condition and organization of the materials. Building materials were also studied: samples of brick and mortar from a wall in direct relationship to the yard deposits, and floor material from a nearby house. The aim was to examine how the materials were made, to identify the raw materials, and to be able to identify fragments of them occurring in the yard deposits.

Field Description of the Section through Yard Deposits The deposits exposed in the section (Fig. 8.5 and Fig. 8.6, color) fell broadly into three parts, which

in earlier excavations had been correlated by Masson (1971) with his building phases 2, 3, and 4.

The Upper Part of the Section This corresponds to Masson’s phase 2 and is dominated by dark gray ash, with varying content of charcoal, gypsum, sand, humic material, and burnt and unburnt brick and mortar. There is an overall slope up and a thickening to the west, but between three predominantly ash layers (contexts 22, 26, and 30), contexts that are predominantly sandy (24 which contains ash lenses and 27 which contains mudbrick fragments) thin westward. Toward the west, the lower boundaries of the ashes are clear, the upper ones diffuse; toward the east, the distinction between ashy and sandy contexts is less well defined as the ashes become thinner and poorer in charcoal. Pale gray ash lenses taper in across the section from the west, above context 26 and below context 30. Within the deposits, the ashy materials lie in lenses up to 2 m long. Within context 30, the material parted readily along a plane on which lay entire silica skeletons of graminaceous leaf blades. Gypsum occurs as nodules, from more than 1 cm in diameter down to small flecks. Burnt and unburnt brick and mortar occur, from pieces a few cm across down to small fragments. Lumps, often flattened, of yellow concretionary material with embedded sharp bone fragments, and other concretionary or compacted masses occur: the former are carnivore excreta, the latter are possibly also faecal. Faunal burrows, from the size of rodent or lizard down to small arthropod, are common, but bioturbation has not obscured stratigraphy, even fine laminae being clearly defined. Much bioturbation appears to be penecontemporaneous with deposition, in addition to that which has clearly penetrated from later surfaces.

126

origins of agriculture in western central asia

In context 26 there are clusters of gourd-shaped cells composed of concentric layers of fine material which form the nests of solitary bees.

sands with thin layers and lenses of ash, gypsum, and charcoal, with occasional cultural material, occur. A darker zone some 30 cm down within the sands may be a buried soil.

The Middle Part of the Section This corresponds to Masson’s phase 3 and is predominately sandy. Masson placed his 3/4 boundary between two sand layers, but it was found that these converged and both passed beneath the phase 3 building to the west of the section studied. Ash layers and lenses taper in at first from the west, then from the east, and range from very pale gray to black, charcoal-rich deposits, and each is distinct and clearly differentiated. The sands (33 and 42) are sometimes themselves ashy and lie in finely laminated form in units up to 10 cm thick between the ashes. The sand is fine, well sorted, and colors are pale brown (Munsell color notation 10Y R 6/3) to brown (10Y R 5/3) and grayish brown (10Y R 5/2) where more ashy. Laminar structure is common. Small white concretions and reddish brown laminae and staining occur. A sandy ash (38) lies between them. Within context 42, context 41, a prominent charcoal-rich ash, deepens at intervals into pockets cutting into underlying sands and tapers out half way across the section, with a fine pale gray ash on its surface extending a little farther west. Context 43, a thin, charcoal-rich black ash, also tapers out across the section within 42. The lowest 10 cm of context 42 includes laminae of distinctly reddened sand.

The Lower Part of the Section This corresponds to Masson’s phase 4. Contexts 44 and 45 form a complex of charcoal- and gypsumrich materials, with thin sands between them. Context 44 includes a gypsum-filled hollow with charcoal in its base and a layer of charcoal spreading right across the section below the reddened sands at the base of context 43, and gypsum spreading westward from the hollow in an uneven layer, mixed with ash and sand. A wedge of sand comes in from the west between the charcoal and the gypsum. A similar “gypsum hearth” was noted at the same level beyond the section studied in detail. The deposits of context 45 consist of thin ashes and sands lying approximately horizontally and extending right across the whole section. Context 46 is a distinctive brown humic material laminated with gypsum and gray ash, continuous across the section and found in excavation to lap up against the stump of a building wall. Below this, predominantly clean

Sampling A column of samples was taken for phytolith analysis at a point 70 cm west along the section. Sampling was spaced to include each distinctive lamina or lens, taking the middle part of the thicker ones, so the gaps between samples were not uniform and areas where materials merged were avoided. A sample from the present-day surface near the site was taken for comparison. These samples were also used for particle-size analysis. Kubiena boxes (9 x 5 x 4 cm) for micromorphological analysis, with bag samples adjacent to them, were taken of a group of contexts (26, 27, and 30) at approximately 220 cm west along the section. These contexts, in the upper part of the deposits, were sufficiently coherent to maintain their structure under drying and impregnation. The middle, sand-dominated part of the section could not be sampled successfully for micromorphology because it was too incoherent. In the lower part, contexts of particular interest (44 containing the gypsum hearth, and 46, the brown humic material) were selected for sampling. Several other small samples were taken of components of the deposits occurring in coherent and distinctive form. From the wall revealed in excavation (context 18) a block sample was taken to include both brick and mortar, so that contact between them could be examined as well as each type of material. Although the generally ashy or sandy nature of the contexts at the points where micromorphology samples were taken, and at the position of the column chosen for phytolith analysis, are consistent, details differ on account of lateral variation within each context and the disposition of lenticular and taperingout laminae. For example, context 26 was described as “gray ash with charcoal” at the micromorphology sampling point, but as “gray, ashy/sandy” at the phytolith column. Brief descriptions of the contexts, giving texture and color as they were represented at the position of the phytolith-sample column, are provided in Table 9.3. The micromorphology samples were numbered and relate to their contexts as follows:



jeitun: dating and analysis of excavated materials

J1: this sample was taken in 1990 and is not correlated with contexts identified in 1991; it consists of dark gray ash, laminated sand, and light gray ash, with burnt material including goat droppings; it is described in Limbrey (1992b:95). J2: context 46, brown humic material and gypsum (and see Fig. 9.6, color). J3: context 44, gypsum hearth. J4: contexts 26, 27, and 30; the sample was taken with the intention of capturing the structure of the solitary bees’ nest, which appeared to lie within a disturbance of 27 and 30, but the slide cut from the block missed it. J5: context 30; gray ash with charcoal and mudbrick fragments (and see Fig. 4a,b, color). J6: context 26, with part of context 27; gray ashy sand J9: from context 30, a flattened mass, thought in the field to be possibly faecal. J11: context 18; brick and mortar from wall. J12: no context number, material from a house floor excavated by the Russian team (and see Fig. 9.5, color). J7 was a sample of the predominantly sandy context 42, and was not sufficiently cohesive to maintain its coherence under resin impregnation; J8 was a small cluster of bee nest cells; and J10 was a piece of yellow, bone-rich carnivore excreta; none of these samples was sectioned.

Analysis Methods Partial particle-size analysis was carried out on samples from the phytolith column by dry sieving. The samples were small, and the very low silt plus clay content meant that, for many of them, much larger samples would have been needed for full analysis. A modern surface sample was treated in the same way. Samples of brick and mortar were analyzed by wet sieving and determination of silts and clay by SediGraph X-ray particle-size analyzer. pH in water at 1:2.5 dilution and in CaCl2 buffer was determined on the samples taken from selected contexts for macrodescription of micromorphology samples (J2, J3, J5, J6) and on the brick and mortar (J11). Field descriptions were amplified by bench examination of micromorphology blocks and bag samples taken adjacent to them, using a low-power

127

microscope. Colors are given in Munsell notation in dry condition. Micromorphology blocks were dried through acetone replacement of water, impregnated with Crystic resin by progressive concentration from solution in acetone, and cured under stepwise increase and decrease of temperature. Thin sections were prepared in the soil micromorphology workshop at the University of Newcastle upon Tyne, UK.

Results Particle-size analysis of the depositional contexts of the phytolith samples (Table 9.3) gives a very consistent strong dominance of fine sand, the samples below context 46 consisting of over 90% fine sand with the remainder being predominantly coarse silt very similar to the blown sand of the present-day surface. Ash admixture introduces a wider size distribution, with a marked silt increase which can be attributed to the silica skeleton and phytolith content. Building materials contribute clay, whether from fragments in the deposits or from the abrasion of structures in the course of use and contact with people and domestic animals. Where gypsum content has a high proportion of lenticular sparitic grains, the coarse and medium sand content rises. Because the samples were taken for phytolith analysis, included material was avoided; particles above sand grade were fragments of brick and mortar, pottery, bone, and charcoal, with a few small angular fossiliferous limestone fragments. Soil concretions account for much of the small coarse sand component and, in sample P18, aggregated gypsum. The results of the particle-size analysis of the samples of mudbrick and mortar are shown in Table 9.4. As already mentioned, pH values were determined for the mudbrick and mortar and also for samples from contexts 26, 30, 44, and 46 (Table 9.5). The pH values are surprisingly low, and show that the soluble-salt concentration is low. The deposits within the site are more strongly leached than the alkaline soils of the buried-soil sequence (this chapter: Section 9.3), the site itself having provided a specialized environment. The finer texture and lower porosity of the mudbrick has resulted in resistance to leaching. It should be noted, however, that the section studied in 1991 was in the area where Berezkin had started a new excavation in 1989 (this volume, p. 98) and where Masson had previously removed the uppermost two levels, so chemical conditions are probably not representative of those that prevailed through the greater part of the site’s history.

128

origins of agriculture in western central asia

Table 9.3 List of phytolith samples P1–P33 showing depths in cm below the section datum at 188.6 m above msl, context numbers, the results of particle-size analysis, and brief descriptions of the deposits. Silt + clay Fine Medium Phytolith Depth below Context Coarse % below sand % sand % no. sand % sample section 2.0–0.6mm 0.6–0.2mm 0.2–0.06mm 0.06mm datum no.

Characteristics of the deposits; colors, apart from P23, are all hue 10YR, value 5 or 6, chroma 1 to 4, ashes being in the lower and sands in the higher chromas

P1 P2 P3 P4 P5 P6 P7

0.0–2.0 2.0–4.5 4.5–6.5 6.5–10.0 10.0–15.5 15.5–24.0 24.0–29.0

14 14 14 22 24 26 30

3.0 0.1 5.6 3.6 5.0 3.2 7.4

8.0 16.0 12.8 10.4 9.6 5.0 20.2

77.6 67.2 69.2 76.0 79.4 83.2 69.6

11.4 16.7 12.4 10.0 5.4 8.6 2.8

P8 P9 P10

29.0–39.0 39.0–43.0 43.0–45.0

33 33 33

4.2 2.0 4.8

9.0 5.0 14.8

82.6 84.2 70.8

4.2 8.8 9.6

P11 P12 P13

45.0–47.0 47.0–56.0 56.0–59.0

38 42 –

2.6 0.6 4.8

7.0 4.0 18.0

85.6 83.6 51.4

4.8 12.4 25.8

P14

59.0–63.0

41

1.3

15.6

50.6

22.5

P15 P16

63.0–72.0 72.0–76.0

42 43

2.4 1.0

4.8 9.6

79.6 80.2

13.2 9.2

P17 P18

76.0–79.0 79.0–84.0

44 44

1.4 20.2

0.8 22.8

70.6 38.6

27.2 18.4

P19 P20 P21 P22 P23 P24

81.0–84.0 84.0–91.0 91.0–94.0 94.0–96.0 96.0–99.0 99.0–101.0

45 45 45 45 46 –

0.8 6.2 4.8 3.6 3.0 6.8

4.2 14.0 10.8 9.4 14.0 18.0

83.0 67.6 66.8 81.2 57.8 68.4

12.0 12.2 17.6 5.8 25.2 4.8

P25 P26 P27 P28 P29 P30

101.0–107.0 107.0–109.0 109.0–113.0 113.0–116.0 116.0–122.0 127.0–129.0

– – – – – –

0.2 4.4 4.0 1.2 0.4 1.0

0.6 0.6 2.0 8.4 1.2 1.4

93.2 35.6 90.0 81.8 95.2 95.8

6.0 46.4 4.0 8.6 3.2 1.8

P31 P32 P33 Modern surface

132.5–135.0 141.0–143.0 c. 158.0

– – –

0.2 0.2 0.4 0.2

0.6 0.6 0.8 0.2

96.6 95.6 95.2 96.2

2.6 3.6 3.6 3.8

pale brown fine ashy light brownish gray fine ashy grayish brown fine ashy gray ash with charcoal, mudbrick brown, sandy gray, ashy sand gray ash with charcoal and mudbrick fragments pale brown slightly ashy sand brown sand, but top is thin white ash laminated brownish gray ash, charcoal, gypsum grayish brown sandy, ashy brown sand grayish brown ash with charcoal, in context 42 pale gray ash with charcoal, wihin context 42 brown sand laminated sand with slightly burnt sand and ash mixed dark ash and sand gypsum-filled hollow with charcoal at its base ash, sand, stained by adjacent “hearth” black/gray ash, with sand laminae gray ashy, sandy, slightly burnt yellowish brown sand brown humic material, 7.5YR 5/4 gray sand, laminae within brown humic material yellowish brown sand white ash lens brown sand with white concretions light gray, ashy darker yellowish brown sand with charcoal yellowish brown sand with charcoal and concretions yellowish brown sand yellowish brown sand yellowish brown sand pale brown sand

Detailed descriptions of the micromorphology are given in Appendix 9.1 at the end of this Section. The main components of the deposits are represented in the slides of samples J4, J5, and J6, and micromorphological descriptions of them are given by type of fabric rather than by context or by separate description of each slide,

which would be repetitive because common elements occur throughout. The building materials are next described and these descriptions include fragments within the other fabrics. Then, the gypsum hearth in context 44, the humic layer, context 46, and the putative faecal mass from context 30 are described.



jeitun: dating and analysis of excavated materials

129

Table 9.4 Particle-size analysis of mudbrick and mortar. Size grades of particles (British Standard 1377): coarse sand 2.0–0.6 mm, medium sand 0.6–0.02 mm, fine sand 0.02–0.06 mm, coarse silt 0.06–0.02 mm, medium silt 0.02–0.006 mm, fine silt 0.006–0.002 mm, clay 30 SI units x10 -6) readings in the upper parts of the A horizons and relatively low (1 mm) and fine (10% of samples) were targeted for identification. Plant nomenclature and ecological information for the species were obtained from the Flora of Turkey (Davis, Cullen, and Coode 1965–88), the Flora of Iraq (Townsend and Guest 1966–85), and Identification of the Plants of Turkmenistan (Nikitin and Geldykhanov 1988).

into one of the four groups. This classification is based on similarity in wild seed-type composition between the Jeitun samples and the ethnographic samples from Amorgos. In the discriminant analysis described above, each Jeitun sample can only be classified into one of the Amorgos processing groups. In order to determine whether the Jeitun samples represent a coherent group distinct from the Amorgos groups, a second discriminant analysis was carried out in which the Jeitun samples were entered as a separate group to be discriminated together with the four other (processing) groups (cf. Charles 1998). In this analysis the Jeitun samples could be classified into the Jeitun group or into one of the Amorgos groups. The ordination technique of correspondence analysis was used to explore internal variation among the samples (Jones 1991; Jongman et al. 1987; Lange 1990). Correspondence analysis arranges samples along axes on the basis of species composition and vice-versa. CANOCO for Windows (ter Braak and Smilauer 1997–99) and CANODRAW (Smilauer 1992) were used to conduct the analysis and plot the results, respectively. In all correspondence-analysis diagrams, axis 1 was plotted horizontally and axis 2 vertically. Species present in at least 10% of samples were included; rare species tend to skew the plots and obscure major trends in the data (Gauch 1982:213–14; Jongman et al. 1987:109–11).

Results

Statistical Methods Using the method demonstrated by Jones (1984, 1987), the Jeitun samples were compared with ethnographic data relating to traditional cereal processing in Amorgos, Greece, in order to determine which processing stages were represented in the samples. The seeds of the wild plant taxa were categorized according to attributes determining their behavior during the crop-processing sequence: their size, headedness, and aerodynamic properties. A discriminant analysis was performed (using the “direct” method from SPSS: Norusis 1992) on the dataset comprising the Jeitun samples and the Amorgos samples from the different crop-processing stages. The discriminant analysis maximizes the separation between the four known groups representing the four major by-products/ products of crop processing (winnowing by-product, coarse sieve by-product, fine sieve by-product, fine sieve product) and classifies each of the Jeitun samples

151

Thirty-nine samples in total were analyzed. The archaeobotanical composition of common taxa present in the samples is summarized in Appendix 9.6, together with the densities of identifiable plant items per liter of soil processed. The samples exhibit a large range of densities (counts of identifiable plant items per liter of soil processed), from 3 items per liter to 1,682 items per liter, indicating different depositional histories. Eighteen samples contain at least 100 items per liter—a high density of remains that could reflect single depositional events rather than the accumulation of plant remains over time.

Cultivated Plants The dominant crop type, in terms of both presence in samples (38 out of 39) and relative frequency within samples, is glume wheat. This is chiefly repre-

152

origins of agriculture in western central asia

A

A

B

B

C

C

9.14 Glume wheat spikelet forks from Jeitun: (a) einkorn, (b) emmer, (c) “new” type. (Drawings by Jane Goddard)

sented as chaff (spikelet forks and individual glume bases, 38 samples), with grain present in most samples (35) but in small numbers. Initial reports (Charles and Hillman 1992; Harris et al. 1993:330–33; Harris, Gosden, and Charles 1996:436–39) identified the glume wheat material as predominantly Triticum monococcum (einkorn wheat), with a small proportion of T. dicoccum (emmer) type. Subsequently, a new type of glume wheat has been recognized (Jones, Valamoti, and Charles 2000) and the Jeitun material was recon-

9.15 Glume wheat grain from Jeitun: (a) two-grain einkorn, (b) one-grain einkorn, (c) possible (cf.) emmer. (Drawings by Jane Goddard)

sidered in the light of this work. Spikelet forks resembling einkorn (Fig. 9.14a), emmer (Fig. 9.14b), and the new type of glume wheat (Fig. 9.14c) have now been positively identified at Jeitun. Most of the well-preserved chaff is of the einkorn type, but many specimens were poorly preserved and hence identified as “glume wheat indeterminate”. Furthermore, terminal glume bases (with glumes at a 90° angle to the rachis, cf. Cappers, Van Thuyne, and Sikking 2004) occur in 15 samples and presumably derive from emmer or the



jeitun: dating and analysis of excavated materials

A

B

C 9.16 Barley material from Jeitun: (a) naked barley grain, (b) naked 6-row barley rachis, (c) hulled barley grain. (Drawings by Jane Goddard)

153

are no grains typical of emmer, although five poorly preserved specimens have tentatively (“cf.”) been identified (4 samples, Fig. 9.15c). None of the identifiable grain resembles the “new” type described by Köhler-Schneider (2001:119, Plate 3; 2003: Fig. 5a). A few rachis internodes (present in three samples) were tentatively identified as free-threshing wheat (T. aestivum/ durum). Other cereal crops present are naked and hulled varieties of barley (Hordeum sativum, terminology after Charles 1984). Naked barley is represented as both grain (17 samples, Fig. 9.16a) and rachis internodes (11 samples, Fig. 9.16b). The majority of the barley rachis internodes have robust and elongated lateral floret “stalks” with raised rear glume insertion points. They correspond, therefore, to the naked form of lax-eared six-row barley, also known as four-row barley (Jacomet 1987). A further indication that these rachis internodes derived from six-row barley is that the glume bases bracketing the central grain have a vertical orientation, rather than diverging orientations as found in the two-row form (Fig. 9.16b). While the naked barley rachis internodes appear to be of the six-row type, no definite asymmetrical grains of naked barley were observed. Hulled barley grains were observed in 11 samples (Fig. 9.16c). None of the barley rachis lacking thickened lateral floret bases, and hence potentially deriving from hulled barley, was sufficiently well preserved to be identified as two- or six-row. No definite asymmetrical grains of hulled barley were observed.

Wild Taxa and Dung new type of glume wheat, because the terminal spikelet in einkorn is infertile (Percival 1974:172) and unlikely to be preserved by charring. In this report, counts of terminal spikelet forks are given, but the remainder of the chaff is referred to simply as “glume wheat” (Appendix 9.6). The most frequently occurring glume-wheat grain type derives from two-grain spikelets and resembles two grain einkorn (32 samples, Fig. 9.15a); einkorn from one-grain spikelets is also common (23 samples, Fig. 9.15b). There

The coarse flots were characterized by two types of wild plant: goat-face grass (Aegilops sp.) in 34 samples and caper (Capparis sp.) in 32 samples. Both the grain and chaff of Aegilops were present (Appendix 9.6). The relative abundance of Capparis seeds in the samples might indicate that the fruits were collected for human food (Blakelock and Townsend 1980:140; Rivera et al. 2002:308–9), and the seeds have also been reported from Neolithic sites in Turkey, Jordan, and

154

origins of agriculture in western central asia

Iran (Helbaek 1970:230). Other common wild plants include the grasses Bromus spp. and Eremopyrum sp., alison (Alyssum sp.), and club rush (Scirpus maritimus syn. Bolboschoenus maritimus). Analysis of the fine flots (0.99). Three samples are classified as other categories but all with low probability (0.75 or lower): two samples (8911 and 9447) are classified as fine-sieved by-products and one sample (9034) as a coarse-sieved by-product. As noted above, the Jeitun samples tend to be dominated by glume-wheat glume bases. Other applications of Jones’ (1984, 1987) method of identifying crop-processing stages from archaeobotanical samples dominated by glume-wheat glume bases have suggested that they represent the by-products of fine sieving, not of winnowing (cf. Bogaard 2004:68). The classification of glume base samples as fine sieve



jeitun: dating and analysis of excavated materials

Table 9.10 Relative frequency (percentage) of glume-wheat grain and glume bases in samples containing at least 50 crop items. Glume-wheat glume base to grain ratios are given in the final column. Sample Glume-Wheat Glume-Wheat Glume-Wheat No. Grain (%) Chaff (%) Chaff : Grain Ratio

Table 9.11 The classification of wild taxa according to their size, headedness, and aerodynamic properties (after Jones 1984). Seeds smaller than 2 mm in diameter are considered “small.” Small, Free, and Light Aeluropus sp

9238

1

99

197

9320

1

99

143

9237

1

99

89

9111

2

98

64

9217.2

2

98

61

9107

2

98

41

9121

3

98

39

9044

3

97

30

8911

3

97

29

9228

4

96

27

9217.1

4

96

26

9423

4

96

22

9225

4

96

22

9032

5

95

20

9118

6

94

17

9132

6

94

16

9447

7

93

14

9117

8

92

12

Eremopyrum sp

9442

8

92

12

Galium type 1

9126

8

92

11

Galium type 2

9025

9

91

10

Heliotropium sp

9444

11

89

8

Melilotus/Trifolium indet

9042

11

89

8

Polygonum cf aviculare

9227

12

88

7

Scirpus maritimus

9023

13

87

7

Big, Free, and Heavy

Stipagrostis sp Small, Headed, and Light Erodium type 1 Suaeda type 2 Small, Headed, and Heavy Alyssum sp Suaeda type 1 Big, Headed, and Heavy Capparis sp. Small, Free, and Heavy Astragalus/Trigonella indet Bromus type 1 Bromus type 2 Centaurea sp Chenopodium album CYPERACEAE

9232

14

86

6

Aegilops sp

9122

16

84

5

Asparagus sp

9034

18

82

4

9125

19

81

4

9052

34

66

2

9201

0

100

by-products is in agreement with the ethnographic observations of Hillman (1981, 1984a, 1984b), which indicate that glume material is separated from grains by fine sieving (following spikelet pounding). However, experimental work by Küster (1985) and Meurers-Balke and Lüning (1992) shows that winnowing is capable of separating grain and glume material.

155

To summarize thus far, the results of the cropprocessing analysis (Fig. 9.18a, Table 9.12a) indicate that the wild plant seeds in the Jeitun samples resemble those in ethnographic winnowing by-products (i.e., rich in taxa with small free light seeds), but the crop component of these samples (mainly glume wheat glume bases) could be interpreted as fine sieve by-product material. The classification of most Jeitun samples as winnowing by-products, therefore, suggests that the crop and wild plant material may not in general derive from the same source.

156

origins of agriculture in western central asia

Table 9.12a The classification of samples by the discriminant function extracted to distinguish the four groups of ethnographic samples from Amorgos. Crop-processing Stage (or predicted group membership) Winnowing by-product

Sample

Probability

9023

1.00

9025

1.00

9032

1.00

9042

1.00

9044

0.96

9052

0.99

9107

1.00

9117

1.00

9118

1.00

9121

1.00

9122

1.00

9125

1.00

9126

1.00

9129

0.94

9131

1.00

9132

1.00

9201

0.97

9212

1.00

9214

1.00

9217.1

1.00

9217.2

1.00

9225

1.00

9227

1.00

9228

1.00

9232

0.55

9237

0.99

9238

1.00

9304

1.00

9320

1.00

9323

1.00

9324

1.00

9423

1.00

9442

1.00

9444

0.99

9448

1.00

Coarse-sieve by-product

9034

0.75

Fine-sieve by-product

8911

0.75

9447

0.54

9.18a Scatter plot of the archaeological samples from Jeitun and the ethnographic samples from Amorgos on the first two discriminant functions.

9.18b Scatter plot of the archaeological samples from Jeitun and the ethnographic samples from Amorgos on the first two discriminant functions. In this analysis, the Jeitun samples are entered as a separate group to be discriminated.

As described above in Methods, another way of comparing the Jeitun and Amorgos samples is to enter the Jeitun samples as a separate group in discriminant analysis. In this analysis, the discriminant functions are extracted to distinguish five groups (i.e., the four processing groups plus the Jeitun sample group). In the resulting plot of samples on the first two discriminant functions (Fig. 9.18b), it is evident that the arrangement of the crop-processing groups is changed considerably in comparison with the previous analysis (Fig. 9.18a) and that the Jeitun samples form a distinct cluster. This result suggests that the Jeitun samples are a coherent group that differs in wild seed-type composition from all the ethnographic processing groups. The classification results (Table 9.12b) show that all but one of the archaeological samples are placed in the Jeitun group rather than in any of the ethnographic processing groups. Analogous results were obtained at Abu Salabikh in southern Iraq, where it was concluded that many of the wild plant seeds in



jeitun: dating and analysis of excavated materials

Table 9.12b The classification of samples by the discriminant function extracted to distinguish the four groups of ethnographic samples from Amorgos plus the Jeitun group. Crop-processing Stage (or predicted group membership) Jeitun group

Fine-sieve by-product

Sample

Probability

8911

0.00

9023

0.21

9025

0.22

9032

0.45

9034

0.00

9042

0.75

9044

0.01

9052

0.06

9107

0.03

9117

0.70

9118

0.50

9121

0.19

9122

0.56

9125

0.87

9126

0.38

9129

0.05

9131

0.67

9132

0.12

9201

0.02

9212

0.65

9214

0.46

9217.1

0.15

9217.2

0.97

9225

0.74

9227

0.80

9228

0.05

9232

0.00

9237

0.25

9238

0.26

9304

0.58

9320

0.94

9323

0.94

9324

0.86

9423

0.13

9442

0.39

9444

0.06

9448

0.61

9447

0.07

157

the assemblage originated from a source other than crop processing (Charles 1998). The classification of the Jeitun samples in the two discriminant analyses (Figs. 9.18a, b, Tables 9.12a, b)— combined with their crop composition (Fig. 9.17, Table 9.10)—suggests that at least some of the wild plant seeds were not harvested and processed with the cereal material in the samples. As noted above, wild plant seeds were observed in the matrix of charred dung pellets in some samples. A possible source of charred plant remains other than crop processing, therefore, is animal dung, which may have been burned as fuel. This hypothesis is considered further below.

Exploring Internal Variation in Sample Composition A correspondence analysis was performed on all 39 samples; taxa present in at least 10% of samples (30 in total) were included. These taxa are listed in Table 9.13a, and Table 9.13b shows how similar categories of taxa were amalgamated in some cases in order to simplify the analysis. In the correspondence-analysis plot of samples from the initial analysis (not shown), seven samples emerged as outliers: 9117, 9126, 9227, 9304, 9323, and 9324, which contain unusually high quantities of Aeluropus sp., and 9232, which contains unusually high quantities of Suaeda spp. (see Appendix 9.6). The decision was taken to remove these samples from the analysis in order to explore variation among the remaining samples. A correspondence analysis, therefore, was run with the remaining 32 samples and 30 taxa. Both the sample (Fig. 9.19a) and the species (Fig. 9.19b) plots from this analysis show an even spread of points, with no strong outliers. In Figure 9.19b, the crop taxa (in particular, barley rachis and glume-wheat glume bases) and certain wild plant taxa (especially Aegilops sp.) are located toward the right (positive) end of axis 1, whereas other wild taxa are located toward the left (negative) end. This suggests that some wild taxa are more closely associated with crop material than others.

Investigating the Sources of Wild Plant Seeds in the Samples In order to investigate the origin of the Jeitun wild taxa in more detail, it is necessary to consider their ecology. Wild taxa that did not arrive on site as harvested crop weeds may potentially be distinguished by their pattern of growth or seasonality. The flowering/

158

origins of agriculture in western central asia

Table 9.13a Codes for taxa included in the correspondence analysis.

Table 9.13b List of the taxa included in the amalgamated groups used in the correspondence analysis. Name of Amalgamated Group

Taxa Included in Group

Glume wheat grain

Emmer grain, cf.

Code

Type Name

Bar_G

Barley grain

Bar_R

Barley rachis internodes

culm_no

Culm nodes (>1 mm diameter)

Glwht_C

Glume wheat glume bases

Glwht_G

Glume wheat grain

Wht_Br

Basal wheat rachis

Ce-Gr_G

Cereal/large Gramineae

Einkorn indeterminate grain Glume wheat indeterminate grain

Aegi_Chf

Aegilops glume bases

Emmer grain, cf.

Aegi_sp

Aegilops sp.

Aelu_sp

Aeluropus sp.

Alys_sp

Alyssum sp.

Aspa_sp

Asparagus sp.

Brom_1

Bromus type 1

Brom_2

Bromus type 2

Naked barley grain

Capp_spi

Capparis sp.

Cent_sp

Centaurea sp.

Barley indeterminate grain

Chen_alb

Chenopodium album

Cruc_1

CRUCIFERAE type 1

Cype_F

CYPERACEAE indet

Erem_sp

Eremopyrum sp.

Erod_1

Erodium type 1

Gali_1

Galium type 1

Gali_2

Galium type 2

Heli_sp

Heliotropium sp.

Poly_avi

Polygonum cf. aviculare

Reed_Cu

reed culm nodes (Phragmites/Arundo)

Scir_Mar

Scirpus maritimus

sleg

small-seeded LEGUMINOSAE

Stip_sp

Stipagrostis sp.

Suaeda

Suaeda

fruiting times of specialized crop weeds tend to coincide with that of the crop, while the fruiting times of other wild plants may encompass a much wider time period. In present-day Turkmenistan, along the foothills of the Kopetdag, the cereal harvest typically occurs in early May. Using flowering/fruiting data from relevant floras (see Methods, above), the archaeobotanical taxa were classified as early, intermediate, or late in their

Einkorn 1-grain Einkorn 2-grain

Glume wheat glume bases

Terminal glume bases Other glume bases

Barley grain

Barley rachis internodes

Small-seeded LEGUMINOSAE

Suaeda

Hulled barley grain

Naked barley rachis Barley indeterminate rachis Melilotus/Trifolium indeterminate Trigonella/Astragalus indeterminate Suaeda type 1 Suaeda type 2 Suaeda indeterminate

seasonality relative to the time of harvest (Table 9.14). Early taxa are those whose f lowering/fruiting periods include harvest time (May) but do not extend beyond it. They may represent arable weeds that were harvested with the cereal crops when in fruit. Intermediate taxa begin flowering/fruiting in April before the cereal harvest but continue beyond it, until June or July. They may have been harvested with the cereal crops when in fruit but their origin is uncertain because they continue fruiting after the harvest. Late taxa begin flowering/fruiting in May



jeitun: dating and analysis of excavated materials

or June and continue until August or later. They are unlikely to have been harvested with the cereal crops when in fruit. The ability of wild-plant seasonality to explain the spread of species points in the correspondence-analysis plot was explored by coding the wild taxa as early, intermediate, or late (Fig. 9.20); a few taxa could not be classified due to their ambiguous level of identification. In Figure 9.20, there is a separation of early and late taxa along axis 1: the few early taxa tend to occur toward the positive (right) end of the axis, together with the cereal components, while late taxa tend to occur toward the negative (left) end. Intermediate taxa are generally distributed along axis 1. These patterns are consistent with the hypothesis that early taxa were harvested with the cereal crops, whereas late taxa derive from some other source(s), possibly charred sheep/goat dung. The meaning of the first axis is further clarified in Figure 9.21, where sample points are shown as pie-charts indicating the proportions of different seasonality categories, crop material, and unclassified taxa (including culm nodes, which may derive from cultivated or wild grasses) in each sample. There is a clear trend along axis 1 in the relative amounts of late taxa versus crop material: late taxa predominate in samples toward the negative (left) end, whereas crop material predominates in samples toward the positive (right) end. Early taxa tend, like crop material, to increase toward the positive (right) end of axis 1, although proportions of early taxa are highest in two samples located toward the positive (top) end of axis 2. These two samples also contain significant proportions of late taxa and relatively little crop material. Overall, however, .

9.19a (top) Correspondence analysis plot of 32 samples. Species codes are given in Table 9.19b 9.19b Correspondence analysis plot of 30 taxa (present in >10% of samples)..

159

160

origins of agriculture in western central asia

Table 9.14 Seasonality of the Jeitun wild plants (Nikitin and Geldykhanov 1988; Townsend and Guest 1966–85). The month of typical winter cereal harvest in the Kopetdag foothills is indicated by shading. Seasonality Class Early

Months of Flowering/Fruiting Mar

Aegilops sp. Alyssum sp.

X

Eremopyrum sp.

Apr

May

X

X

X

X

X

X

Intermediate

Jun

Jul

(X)

(X)

Aug

Oct

(X)

(X)

 

Asparagus sp.

X

X

X

(X)

Bromus type 1

X

X

X

(X)

Bromus type 2

X

X

X

(X)

Centaurea sp.

X

X

X

X

Chenopodium album

(X)

X

X

X

Erodium type 1

X

X

X

(X)

Galium type 1

X

X

X

(X)

(X)

Galium type 2

X

Polygonum cf. aviculare small LEGUMINOSAE

Sep

(X)

X

Late

(X)

(X)

X

X

(X)

(X)

X

X

X

(X)

X

(X)

(X)

(X)

 

Aeluropus sp.

 

X

X

X

X

Capparis sp.

X

X

X

X

X

Heliotropium sp.

 

X

X

X

X

X

Scirpus maritimus

X

X

X

X

X

X

Stipagrostis sp.

X

X

X

X

Suaeda spp.

 

X

X

X

X

(X)

( ) = rarely

it appears that early taxa are more closely associated with crop material than late taxa. The seven outlier samples excluded from the main correspondence analysis (see above) contain little crop material and are dominated by one or both of the late taxa, Suaeda spp. and Aeluropus sp. These samples, therefore, like those located toward the negative (left) end of axis 1 in Figure 9.21, are rich in taxa unlikely to derive from harvested crops/crop processing. It was suggested above that the presence of dung-derived wild taxa in the Jeitun samples relates to the classification of most samples as winnowing byproducts, characterized by a predominance of taxa with light seeds separated off by winnowing (Jones 1984, 1987). Of the four taxa with light seeds (i.e., small free light and small-headed light) in the Jeitun

assemblage (Table 9.11), three are late taxa and the fourth is intermediate in flowering/fruiting times (Table 9.14). The classification of the samples as winnowing by-products, therefore, is largely caused by the presence of probable dung-derived taxa. Finally, in order to investigate the possibility that sample composition relates to archaeological context, the correspondence analysis plot of samples was coded to show the major context types represented (ash dumps, yard deposits, floor deposits, hearths, unknown) (Fig. 9.22). There is no obvious relationship between archaeological context and sample content, with the possible exception of samples from floor deposits, all three of which occur toward the positive end of axis 1, although not at the extreme positive end (Fig. 9.22). Overall, therefore, it appears that the trend along axis 1 noted above (in the relative amounts of



jeitun: dating and analysis of excavated materials

161

9.21 Correspondence analysis plot of samples shown as pie charts indicating proportions of different seasonality categories, crop material, and unclassified taxa.

9.20 Correspondence analysis plot of taxa coded according to seasonality categories (see Table 9.15).

late taxa versus crop material) does not relate directly to context type. Similarly, coding of the correspondence analysis plot of samples according to density (number of items per liter soil processed) does not show any clear patterning (plot not shown).

Discussion Assessing the Role of Animal Dung as a Source of Charred Plant Remains There are three pieces of evidence which suggest that charred animal dung is a major source of plant remains in the Jeitun assemblage: the presence of late taxa (Aeluropus sp., Suaeda spp.) in the matrix of fragmented dung pellets in a few samples, the dominance of late taxa in some samples, and the “anomalous” classification of the Jeitun samples in crop-processing analyses. It should be emphasized that late taxa could have reached the site by a variety of routes; however, the fact that some late taxa have been found in dung indicates that sheep and goats ate the seeds of these late-maturing wild plants. The three characteristics of the Jeitun assemblage listed above fulfill three criteria proposed by Charles (1998) for recognizing dung-derived mate-

9.22 Correspondence analysis plot of samples coded according to context type.

rial in archaeobotanical assemblages: the presence of recognizable animal dung (and especially the observation of charred plant remains within dung), incompatibility between archaeobotanical samples and ethnobotanical crop-processing products or byproducts, and the presence of late-maturing wild plants that are unlikely to have been harvested with

162

origins of agriculture in western central asia

crops when in fruit. A fourth and final criterion is the mixing of crops which, for practical reasons, are unlikely to have been processed together (e.g., glume wheat and free-threshing cereals). In crop terms, the Jeitun samples are heavily dominated by glume wheat, with only low levels of free-threshing cereals (barley and free-threshing wheat) present. There is no firm evidence, therefore, for deliberate and systematic mixing of crops in these samples. While it appears likely that the seeds of latematuring wild plants derive from charred animal dung, it remains to determine how this material came to be mixed with the crop material. There are three possibilities: 1. both late taxa and crop material (together with early taxa/crop weeds) are present in the same dung pellets. For this to occur, livestock would need to be eating plant material from these two sources at more or less the same time, for example by grazing late-summer plants supplemented by eating cereal material, or consuming a mixed fodder containing such material; 2. late taxa and crop material (crop weeds) derive from two different categories of dung, one containing late taxa, the other containing cereal material and early taxa; 3. only late taxa are observed in dung pellets. Such dung-derived material would have been mixed with cereal material either deliberately (to make dung cakes for fuel) prior to deposition or by accident following deposition (e.g., cereal residues discarded into dung fires, spent fuel and processing waste mixed in midden deposits). The negative relationship between late taxa and crop material observed in the correspondence analysis could accommodate all three of these possibilities. Because these two types of material derive from two different sources originally, they could have been combined in varying proportions in the same dung pellets or archaeological deposits. There is some negative evidence to suggest that sheep and goats were not fed cereal material as none was observed in the matrix of fragmented pellets, but more experimental work is needed to investigate the ability of different cereal components to survive livestock digestion (see e.g., Valamoti and Charles 2005). It is not possible on present evidence to decide which of the three possibilities is best represented in the Jeitun assemblage. The most secure conclusion is that the late taxa, at least, derive from charred animal dung, and this observation has important implications for the reconstruction

of animal and crop husbandry at Neolithic Jeitun (see below).

The Crop Spectrum and Its Interpretation Previous archaeobotanical work at Neolithic sites in southern Turkmenistan has been restricted to the examination of cereal remains and impressions in mudbrick at Jeitun itself and three other JeitunCulture sites, Bami, Chopan, and Chagylly (Lisitsina 1978:92; Masson 1971:79). Barley and wheat were reported as present at Jeitun, Chopan, and Bami (identified just as Hordeum sp. and Triticum sp.), and two-row barley and bread and club wheat at Chagylly (identified as H. distichum L., T. aestivum L., and T. compactum Host.). By comparison with these earlier reports, our investigations at Jeitun have confirmed the presence there of (six-row) barley and of only three specimens tentatively identified as free-threshing (T. aestivum/T. durum) wheat. The plant remains now recovered at Jeitun and reported here represent a substantial addition to the body of evidence for plant use in Neolithic western Central Asia. In terms of the range and relative abundance of seed crops at Jeitun, the results of the present analysis reinforce those previously reported (Charles and Hillman 1992; Harris et al. 1993:330–33; Harris, Gosden, and Charles 1996:436–39), namely that glume wheats constituted the main crop in the assemblage. The other definitely attested cereal crops present, although at low levels, are naked and hulled barley. There is no evidence to suggest that pulses were cultivated. As already stated, einkorn appears to have been the dominant glume-wheat crop at Jeitun (see also Charles 2007:40, 47). While it must be stressed that a restricted range of deposits is represented by the samples under discussion, the minor role of emmer is striking, given its ubiquity in Neolithic sites in Southwest Asia and Europe and its reputation as “the main crop in the spread of the Neolithic agricultural technology from the Near East nuclear area” (Zohary and Hopf 2000:48). The crop spectrum at Jeitun, and to a lesser extent at the approximately contemporary site of Ali Kosh in the Deh Luran plain of Iran (Helbaek 1969), is very narrow compared with the diverse cereal and pulse assemblages known from sites in the Southwest Asian Fertile Crescent and southeastern Europe (Charles 2007). Assuming that the limited crop spectrum at Jeitun is not an artifact of differential preservation, it may imply that einkorn and barley were selected as crops due to their tolerance of low soil fertility and



jeitun: dating and analysis of excavated materials

low water availability (Percival 1974; Carter and Stoker 1985), although such extrapolations to Jeitun based on modern crop ecology may be misleading. Southern Turkmenistan is not conventionally included in maps indicating the natural distribution (“primary habitat”) of wild einkorn, wild emmer, or wild barley (Zohary and Hopf 2000: Maps 1, 3, 5), but the ecological requirements of at least some of these wild cereals do not exclude western Central Asia as a natural home. Indeed, Harris and Gosden (1996:372) have pointed out that southern Turkmenistan is part of the larger biogeographical region in which most of the crops of Neolithic Southwest Asian agriculture originated. Furthermore, Hillman (1996:188–89) has speculated that wild barley might have spread westward in early postglacial times from refugia in this region, if such existed, as well as from other refugia farther west.

Rainfed Agriculture versus Cultivation of Irrigated and/or Flooded Areas A distinction has been made based on flowering/fruiting times between wild plant seeds likely to have arrived on site as arable weeds harvested with the cereal crops, and those coming into fruit after the harvest and likely to have arrived on site by other routes, such as in animal dung. The ecological characteristics of potential weed taxa of arable cultivation can be used to investigate crop-growing conditions. The most probable weeds of crops grown at Jeitun are Aegilops and Eremopyrum (Table 9.15a); the other early taxon in the assemblage, Alyssum sp., is less closely associated with crop material in the correspondence analysis (Fig. 9.19b) and so its identification as a crop weed is less certain. It was previously argued that rainfed winter cultivation of cereals was unlikely at Neolithic Jeitun, on the grounds that: (a) climatic conditions in the past were not substantially different from the present, with average annual precipitation too low (at 4 and >2 mm with a brass receiver at the bottom for retaining the 4 and >2 mm fractions were then sorted in their entirety with the aid of a low-power binocular microscope in order to separate charcoal, seeds, mollusk shell, bone, fruits, and fragments of burnt dung. Each class of material was bagged separately and materials other than charcoal were returned to Sheffield. Depending on their size, charcoal fragments were either hand- or pressure-fractured with a carbon-steel razor blade in order to produce clean surfaces, when possible in all three anatomical planes (transverse, radial longitudinal, and tangential). The surfaces were then examined under a highpower, epi-illuminating Olympus BHMJ microscope at magnifications of x50, x100, x200, and x500. Identifications were made by comparing the charcoal fragments with charred specimens and thin sections of fresh wood in the C. A. Western wood reference collection held at the Institute of Archaeology, UCL, and with anatomical descriptions of wood in Fahn, Werker, and Baas (1986), Greguss (1959), and Schweingruber (1990). Individual specimens were photographed using the SEM (scanning electron

167

microscope) facilities at the Institute of Archaeology, UCL.

Results Sample Composition and Taphonomic Observations The charcoal material from Jeitun that was examined came from 24 flotation samples (split samples were combined for the purpose of this study) amounting in total to 1,799 charcoal fragments (Table 9.16). Six taxa were identified: tamarisk (Tamarix), poplar (Populus), willow (Salix), alder (Alnus), undifferentiated chenopods (Chenopodiaceae), and a monocotyledonous plant closely resembling common reed (Phragmites). SEM microphotographs of selected specimens are reproduced in Figure 9.23. Identification was possible only to the generic level because preservation of the charcoal fragments was very poor and because of a lack of detailed reference material and studies of wood anatomy for this part of Central Asia. Chenopodiaceae could only be identified to family level. The differences between individual genera in the Chenopodiaceae are frequently concentrated in the presence or absence of rays and the occurrence of helical thickenings (Schweingruber 1990), which are very difficult to trace in small fragments of badly distorted charred twig and/or small round wood (as characterize most chenopod shrubs). Table 9.16 shows that there is very little variation in sample composition that can be attributed to differences between domestic and non-domestic contexts. Tamarisk is the dominant taxon in terms both of presence in samples and numbers of fragments, followed by chenopods. The remaining taxa appear more erratically in the samples, which could suggest that tamarisk and chenopods were collected more intensively by the Neolithic inhabitants of Jeitun than the other taxa represented. However, due to the accumulation of mineral precipitates, charcoal preservation is poor, as is shown by the very high proportion of fragments that are indeterminate (52%), which suggests that the Jeitun charcoal assemblages have undergone substantial post-depositional reworking and deterioration. This may explain both the low number of taxa identified and why very few samples contain relatively well preserved charcoal fragments, free of excessive accumulation of mineral inclusions, such as the yard sample 34 which contains a large concentration of alder charcoal (Table 9.16).

168

origins of agriculture in western central asia

Table 9.16 Charcoal samples and botanical identifications from material excavated at Jeitun in 1990, 1993, and 1994. The numbers recorded in the rows for each taxon refer to charcoal fragments per sample. Sample no.

8

16

Context

17

18

118

19

24

28

yard

33

yard

34

38

yard

ext. hearth

40

41

Year excavated

1990

1990

1990

1990

1990

1990

1990

1990

1990

1990

1990

1990

>4 (weight /g) >2 (weight /g) Total weight

0.5 1.1 1.6

8.48 11.16 19.64

1.11 1.55 2.66

42.44 36.13 78.57

2.02 4.28 6.3

1.24 1.72 2.96

4.43 3.82 8.25

0.34 0.18 0.52

0.55 0.91 1.46

0.85 1.52 2.37

0.39 0.16 0.55

0.71 1.88 2.59

Tamarix

19

36

18

54

10

13

10

3

6

4

5

3

cf. Tamarix

26

15

11

19

5

7

13

3

8

7

5

12

2

4

Salicaceae Populus

30

6

Salicaceae - Salix

2

Alnus

69

cf. Alnus

1

2

cf. Phragmites Chenopodiaceae

2

4

6

4

3

1

19

7

2

2 4

3

13

8

Indet.

49

41

45

23

44

58

39

18

72

24

13

73

Total

96

96

80

100

60

100

100

28

170

50

25

100

4

9

25

25

27

34

44

45

175

178

195

195

235

Sample no.

43

45

46

56

Context

yard

yard?

yard?

yard

Year excavated

1990

1990

1990

1990

1993

1993

1993

1994

1994

1994

1994

1994

>4 (weight /g) >2 (weight /g) Total weight

0.31 0.38 0.69

0.75 0.72 1.47

0.12 0.43 0.55

8.52 0.82 9.34

2.85 2.9 5.75

5.08 2.14 7.22

5.01 5.16 10.17

3.73 1.05 4.78

0.28 0.52 0.8

2.071 2.071

0.32 0.56 0.88

0.24 0.45 0.69

9

12

3

7

9

21

21

47

4

7

4

4

20

7

18

7

5

25

16

26

5

1

20

16

Tamarix cf. Tamarix

hearth oven fill threshold

Salicaceae Populus

1

Salicaceae - Salix Alnus cf. Alnus cf. Phragmites Chenopodiaceae

2

5

4

3

11

1

7

2

1

1

4

1

Indet.

30

31

44

41

52

51

52

26

24

27

30

29

Total

59

50

67

60

70

100

100

100

40

38

60

50

Interpretation Local Vegetation and Firewood Collection in the Early Neolithic Despite the poor preservation of most of the charcoal, and the probable biases in the representation of taxa, it is reasonable to infer that tugai vegetation

existed along piedmont rivers in the vicinity of the site, such as the local channel of the Kara Su. The clear predominance of tamarisk and the regular occurrence of chenopods in the charcoal samples support this inference. Both flourish on saline, sandy-loamy depressions and tolerate all but extreme seasonal flooding. The presence of reeds and especially alder further



jeitun: dating and analysis of excavated materials

169

9.23 SEM micrographs of identified charcoal specimens from Jeitun; TS=Transverse Section, RLS=Radial Longitudinal Section. Upper row, left, and center: Alnus TS, RLS, right: Tamarix TS. Middle row, left: Populus TS, center: Salix TS, right: cf. Phragmites TS. Lower row, left, center and right: Chenopodiaceae TS. (Photos by Eleni Asouti)

suggests that more continuously inundated surfaces such as marshes and swamps may have existed near the site. The low frequencies of willow and poplar in the samples could indicate that alluvial surfaces such as levees were relatively rare in the vicinity, because neither tree can withstand prolonged flooding, but structurally light woods, such as those of the Salicaceae (willows and poplars), are more prone to post-depositional

deterioration than denser woods such as tamarisk and most of the chenopods, and are therefore less likely to be preserved in identifiable form. It is also reasonable to infer that the inhabitants of Jeitun collected most of their firewood in the near environs of the site. However, given the poor preservation of most of the charcoal, it is not possible to reach secure conclusions as to the intensity of firewood col-

170

origins of agriculture in western central asia

lection, and the relative contributions of firewood and dung to fuel consumption (cf. Section 9.6). Riparian vegetation would have been available and accessible year-round in the vicinity of the settlement, although its abundance may have varied with the frequency and intensity of flooding following seasonal increases in the discharge of piedmont streams, especially in the local channel of the Kara Su. An attempt was made to identify microscopic traces of fungal hyphae in the charcoal fragments in order to evaluate whether deadwood may have been an important source of fuel (for details of the methodology used, see Asouti 2005). However, this did not produce positive results, mainly because of the poor preservation of the charcoal fragments, although a specimen of Rhamnaceae charcoal from the Dam Dam Cheshme 1 rockshelter in the Bolshoi Balkhan massif provides a rare example of deadwood,

identified by the presence of fungal hyphae (see Fig. 10.1). Nevertheless, it remains possible that firewood collection was conditioned to some degree by the seasonal abundance of readily collectable, and therefore less labor-demanding, fallen branchwood, as has been inferred for the Neolithic sites of Çatalhöyük (Asouti 2005) and Pinarba şı in central Anatolia, Turkey (Asouti 2003). Modern ecological observations have recorded massive natural pruning of branches and fruits of saksaul shrubs (Haloxylon spp.) in winter when average daily temperatures fall below 5° C, and in spring occasional late frosts can be sufficiently severe to kill the one-year shoots of saksauls and other woody plants (Orlovsky 1994:39, 41). It is thus possible that the Neolithic inhabitants of Jeitun took advantage of such seasonal abundances of easily collected deadwood, although this supposition cannot be supported by the charcoal evidence currently available.



jeitun: dating and analysis of excavated materials

171

Section 9.8 Pollen and Charcoal-particle Analysis: Sampling Off-site Deposits at Jeitun David Harris

O

ne of the objectives of the preliminary investigations carried out during our first visit to Jeitun in 1989 was to determine whether pollen was sufficiently well preserved in sedimentary contexts near the site to allow changes in vegetation to be studied by standard palynological techniques. We were also interested in the possibility of analyzing charcoal particles preserved in the sediments as a means of investigating fire history (Harris 1992). The only contexts that appeared to merit investigation were swampy depressions that exist among the dune sands in the vicinity of the site. Silty clays accumulate in the depressions as a result of intermittent flooding and, at the time of our visit in April, some of them contained standing water that had been discharged via the channel of the local river (the Kara Su) from the area of irrigated grapevine cultivation north of the site (see Chapter 3). Small stands of tamarisk trees, reeds, and sedges grow around and in the depressions. We expected that the alkaline and aerobic conditions of the sediments, and the arid climate in which annual potential evapotranspiration greatly exceeds precipitation, were unlikely to favor the preservation of pollen. Nevertheless, we decided to dig a trial pit at the margin of one of the depressions and take samples to determine whether pollen was preserved to any depth. The results of this investigation were, as expected, largely negative, but they are of some interest and are presented here. A colleague at the UCL Institute of Archaeology, Patricia E.J. Wiltshire, subsequently analyzed the samples in the Institute’s pollen laboratory, and the results of her analysis are incorporated in this Section. I am indebted to her for assisting our investigation in this way.

trees intermixed with reeds grew in the swamp, and salt encrustations lined the water’s edge. A 1-m2 trial pit was dug 5 m from the southern edge of the swamp where the soil surface was dry and lacked any efflorescence of salt. Water was encountered at depths of 25 cm on the northwestern side and at 45 cm on the southeastern side of the pit. Samples for pollen and charcoal-particle analysis were taken at 5 cm intervals from the surface to -25 cm in the northwestern corner and from -30 cm to -45 cm in the southeastern corner of the pit.

Laboratory Methods and Results Pollen Analysis Approximately 2.0 g of sediment from each sample were subjected to standard acetolysis and hydrofluoric acid digestion. Samples were stained with 1.0% aqueous safranine and examined with a phase-contrast microscope of x400 magnification. Where necessary, oil-immersion at x1000 magnification was used. Identifiable pollen was only present in the sample taken at 1 cm below the surface. A pollen count of 150 grains was made on that sample and six taxa were identified (Table 9.17).

Charcoal-particle Analysis The method adopted in this experiment was to count the number of charcoal particles with a long axis of more than 5 microns for five traverses of each slide-mounted pollen sample at depths below the surface of 1, 10, 20, 30, and 40 cm. This was considered acceptable because the amount of sediment used in each sample was approximately the same. Table 9.17 shows the average values of the five counts per sample.

Sampling The depression chosen for the investigation was located approximately 5 km south-south-east of Jeitun and some 200 m east of the road from Ashgabat. In April 1989 its center was filled with standing water to a depth of 3 mm=fragments recovered from dry sieving through 3–4 mm-aperture mesh; W/S>3 mm=fragments recovered from the 3–4 mm-mesh residues from wet sieving; W/S3mm W/S>3mm W/S14 mm) and smaller blades (widths 10 microns) and saddles associated with Phragmites. C3 pooid rondels dominate the slide (8,607,000 per gm), followed by platey cells from woody plants. There are many dendritics, bulliforms, and smooth long cells (leaf/stem cells). Rods and cones from sedges are also present, as are single polyhedron cells and elongates from woody plants. C ontext 26 (sample P6): a grayish-brown ashy layer. P6 analysis: JE-01-15. This slide contains leaf/ stem, barley, and wild grass silica skeletons, with more smooth long cells than any other type, cones from sedges, and saddles possibly from Phragmites. As with the other samples, the greatest number of cells are from the C3 pooid rondels (11,300 per gm). Many of the cells appear pitted and the pH value of the sedimentary matrix increases the likelihood that the deposit has been chemically weathered. There are dendritic cells from grass inflorescences and, from woody plants, platey cells as well as single polyhedron cells and scalloped cells. Context 30 (sample P7): gray ash with fine white flecks and fragments of charcoal and mudbrick; laminar structure, parting readily along planes on which lie silica skeletons of graminaceous plant material. P7 analysis: JE-01-14. The silica skeletons counted were of leaf/stem, wild grass, barley (some greater than, or equal to, 100 contiguous cells), Bromus sp. type, Phragmites leaves (14,000 per gm) and stems (11,000 per gm), cereal straw, and wheat type. As in the other samples, C3 pooid rondels (13,806,000 per gm) outnumber panicoid C4 bilobes (894,000 per gm). Rods,

252

appendix 9.4

cones from sedges, and elongates from woody plants are present and many platey cells from wood were counted. There are also bulliforms (leaves), trichomes (awns and leaves), smooth long cells from leaf and stem plant parts, and dendritics from inflorescences. Large keystones (>10 microns) and saddles associated with C4 panicoid plants, including Phragmites, are present. C ontext 33 (sample P10): light brown-gray laminated ash with charcoal and gypsum. P10 analysis: JE-01-7. This slide contains cells of wheat type (Fig. 9.13c), barley, wild grass, and Phragmites. There are large keystones (>10 microns) and saddle shaped cells also associated with Phragmites. Dendritics are abundant and trichomes (from awns and leaves) are also present, suggesting derivation from grain-processing. Starch grains were observed, but no spherulites from dung. Context 41 (sample P14): grayish-brown charcoalrich ash within context 42. P14 analysis: JE-01-8. Large silica skeletons (some greater than, or equal to, 100 contiguous cells) of wheat type, barley, and cereal straw are present. Identifiable multi-cells of wheat type outnumber those of barley, but this sample contains many other silica skeletons that lack the diagnostic attributes necessary for positive identification. Wild grass and Phragmites (both stems and leaves) are present. C3 rondels dominate the sample (5,597,000 per gm) and there are also C4 bilobed cells, as well as small numbers of cones from sedges and elongates from woody plants. C ontext 44 (sample P18): a gypsum-filled hollow with charcoal in its base, referred to by Limbrey as a “gypsum hearth.” P18 analysis: JE-01-12. Only four barley skeletons, two of cereal straw and six of cereal leaf/stem were observed in this slide. Amorphous silica aggregates from wood were noted but not counted. Platey cells from woody plants dominate the sample (7,106 per gm) followed by C3 pooid rondels (6,560 per gm). Sedge cones are also present, as are single polyhedron cells from woody plants. There are a few bulliforms on the slide, as well as some dendritics and smooth long leaf/stem cells. Context 45 (sample P19): reddish-brown/dark brown heat-discolored sand and ash adjacent to the gypsum hearth (context 44). P 19 analysis: JE-01-13. Two slides were made from this sample. The first was used to count phytoliths and was found to contain the largest number of silica skeletons of both wild and domestic taxa, including

numerous wild grass, wheat type, cereal straw, and barley multi-cells. Large wheat type and leaf/stem multicells (some greater than, or equal to, 100 contiguous cells) were noted, as well as many platey cells from wood. Many Phragmites leaf skeletons as well as large keystones (>10 microns) and saddles were counted, but no stem skeletons of reed were found. The large number of Phragmites leaf skeletons present suggested that there could be a dung component in the sample, because reed was commonly used as fodder in ancient Southwest and Central Asia (Rosen, pers. comm. 2003). However, no faecal spherulites were observed when the slide was examined under cross-polarized light. A second slide was therefore prepared (by the process described above under Methods). It showed aggregates of spherulites that would have been processed out of the first sample because of their specific gravity, or dissolved by HCl. Overall, the slide gives the impression that both wood and dung were used as fuel in the adjacent hearth. This impression is reinforced by the presence of charcoal particles, which were noted but not counted. C ontext 46 (sample P23): brown humic material laminated with gypsum and gray ash forming a continuous layer across the section, with a very high concentration of faecal spherulites indicative of dung (see Section 9.2). P23 analysis: JE-01-9. Two slides were made from the sample. The first was counted for phytoliths. It did not contain any cereal skeletons and only a few leaf skeletons of wild grass and Phragmites. Many C3 and C4 single-celled phytoliths, including dendritics, were present, as well as cones from sedges and leaf cells (bulliforms and Phragmites leaf skeletons). Platey cells from woody plants were also present. However, faecal spherulites were absent, and because they were very abundant in a micromorphological slide from a similar sample (J2, Appendix 9.1), a second slide was prepared (by the same procedure used to make the second slide of sample P19; see above). The second slide contained aggregates of spherulites that had been processed out of the first slide by the application of HCl and SPT. Calcitic druses from wood also appeared to be present in this second slide, although not in the first (but see Appendix 9.1 for comment by Limbrey on the possible presence of druses). Sample P28. No context number was assigned to the source of sample P28. It came from a light-gray ashy lens at a depth of 113.0–116.0 cm below the section datum in the deposits of predominantly fine sands



appendix 9.4

of typical windblown lenticular structure containing some charcoal and occasional fragments of mudbrick that underlie the deepest context (46) and which were sounded to a maximum depth of c.158 cm below the section datum. P28 analysis: JE-01-10. A few barley and wild grass silica skeletons, as well as some unidentified skeletons, were observed. C3 rondels dominate this sample (5,055,000 per gm), which also includes bulliforms (leaf cells), platey cells from woody plants, and cones from sedges. Large keystones (>10 microns), often associated with Phragmites, are present, although there are no Phragmites skeletons either from the leaves or from the stems. There are many trichomes (awns and leaves). Charcoal particles were noted in this sample, but not counted.

Mudbrick and Mortar: Samples JE-01-1 and JE-01-2 The mudbrick consisted of very pale brown loamy sand with unevenly distributed coarse sand and occasional lumps of silt/clay and nodules of gypsum. In her micromorphological thin section (sample J11), Limbrey observed occasional fragments of shell, charcoal, and bone. Mudbrick analysis: JE-01-1. This sample is dominated by platey cells from woody plants (1,339,762 per gm), but it also contained a significant number of C3 pooid rondels (347, 024 per gm). Sedge cones, bulliforms, C4 bilobes and saddles, large keystones (>10 microns, as found in Phragmites), scalloped cells of woody plants, dendritics, and smooth long leaf/ stem cells are all present. The only silica skeletons are smooth long cells from leaf/stem plant parts. The presence of spherulites (Fig. 9.13b) suggests that dung may have been added to the mudbrick to increase its tensile strength (Rosen, pers. comm. 2003). The mortar consisted of light brownish-gray loamy sand with more fine gypsum but fewer coarse nodules than the mudbrick and abundant fragments of shell. Limbrey noted charcoal, rounded clay pellets, and calcite-cemented sand pellets in the micromorphological thin section (sample J11). Mortar analysis: JE-01-2. This sample is dominated by C3 rondels (3,483,000 per gm) and platey cells from woody plants (2,563,000 per gm). There are also many bulliforms. Smooth long cells and leaf/stem silica skeletons are also present, as are sedge cones. Charcoal particles were noted but not counted.

253

Off-site Dune Sands and Buried Soils: The S Samples Sample S2 (laboratory number JE-01-3): from the upper buried soil at 14–20 cm below section datum; brown loamy sand (10Y R 5/3) with faint gray and yellow mottling; pH 8.5. S2 analysis: JE-01-3. The main components of this sample are rondels from pooid C3 grasses (5,087,000 per gm), as well as keystones and nonkeystone bulliforms found in leaves. There are single polyhedrons from dicotyledonous plants, but no smooth-walled long cells from leaves and stems, or dendritic cells from inflorescences, although rodshaped long cells are present in smaller numbers (51,000 per gm). Many of the cells appear to be pitted, indicating that they may have been subjected to chemical weathering. Sample S7 (laboratory number JE-01-4): from the middle buried soil at 45–51 below section datum; yellowish brown sand (10Y R5/4) with faint gray mottling; pH 8.6. S7 analysis: JE-01-4. Few phytoliths were observed on this slide and no silica skeletons. Bulliforms appear in small numbers, as do keystone bulliforms, smooth long cells (from leaves and stems), and trichomes. Rondels from pooid C3 grasses are present in greater numbers than any other cell type (85,000 per gm). One sedge cone, one C4 bilobe, and one scalloped cell from a woody plant were counted. Starch grains and spherulites were visible in cross-polarized light. Sample S11 (laboratory number JE-01-5): from the lower buried soil at 80–100 cm below section datum; gray and grayish brown (10Y R 5/1, 5/2) and yellowish brown (10Y R 5/4) sandy loam with yellowish staining around gray patches (10Y R 5/8); pH 9.0. S11 analysis: JE-01-5. Two slides were made from this sample. The appearance of the first suggested that the fine fraction was missing from the original subsample, and so a second slide was made. However, the second slide showed that the appearance of the first was characteristic of the sandy loam matrix of the sample, and so the first was analyzed. It contained no silica skeletons. Rondels from pooid C3 grasses were the most common cell type (191,000 per gm), but smooth long leaf/stem cells, rods, bilobes from C4 grasses, cones from sedges, and elongates from woody plants were also present. Amorphous silica aggregates from wood were noted but not counted.

254

appendix 9.5

Appendix 9.5: Jeitun: Numbers of Types of Phytolith per Gram for Each Sample Analyzed ➾ Sample

P4

P5

P6

P7

P10

P14

P18

P19

P23

2339000 540000 1320000 1559000 0 301000 1139000 1139000 0 240000 119000 0 7016000 480000 61000 0 61000 0 0 0 480000 0 0 0 1859000 480000 0 0 0

427000 98000 228000 589000 0 623000 1444000 3021000 0 560000 589000 0 8607000 755000 361000 162000 100000 0 0 0 33000 0 0 328000 1905000 0 888000 0 0

607 87 260 520 87 347 867 3042 0 1301 0 0 11300 434 260 347 0 0 0 347 0 260 0 867 3909 0 260 87 0

2679000 1580000 1854000 825000 0 2266000 3503000 7075000 0 2817000 206000 894000 13806000 1031000 894000 137000 755000 69000 275000 0 549000 0 69000 275000 2885000 0 1168000 137000 686000

1239000 315000 440000 1805000 84000 357000 1637000 2792000 0 461000 231000 0 1176000 84000 252000 42000 0 0 0 0 482000 84000 0 440000 713000 189000 210000 84000 0

1847000 788000 598000 5624000 109000 625000 3532000 5923000 109000 761000 1005000 82000 5597000 408000 897000 82000 136000 0 0 188000 788000 109000 0 1032000 1847000 679000 489000 109000 0

328 55 382 164 0 546 819 4649 0 874 55 0 6560 0 164 0 0 55 0 0 109 0 0 55 7106 205 2130 0 874

277714 98739 86400 203658 24686 117258 141943 129600 0 117258 43200 37029 2048914 43200 12343 24686 12343 6171 0 0 24686 24686 0 49371 771429 154286 74057 0 166629

3245000 1508000 2087000 1624000 0 1855000 5222000 12518000 0 1158000 1508000 232000 24340000 1158000 463000 0 0 232000 232000 116000 1158000 232000 116000 5795000 5447000 1274000 1508000 0 9387000

253000 54000 4000 0 0 277000 117000 0 127000 0 4000

71000 149000 8000 21000 0 23000 3000 0 7000 0 11000

145 29 0 36 0 7 29 0 7 0 0

379000 322000 6000 86000 0 110000 11000 24000 14000 0 9000

48000 97000 39000 66000 0 8000 5000 0 7000 0 54000

231000 573000 93000 88000 0 129000 50000 9 29000 0 188000

41 14 0 27 0 0 0 0 0 0 14

672686 271543 49371 37029 0 55543 43200 0 425829 0 37029

176000 236000 10000 27000 0 22000 0 0 7000 0 17000

SINGLE-CELL Long (Smooth) Long (Sinuate) Long (Rods) Long (Dendritic) Papillae Hairs Trichomes Bulliforms Ovals Keystones Crenates Bilobes Rondels Saddles Cones Flat Tower Horned Tower Rugulose Spheroid Smooth Spheroid Bottle-shaped Elongate Tracheids Two-tiered Blocks Platey Sheet Single Polyhedron Scalloped Single Jigsaw Puzzle MULTI-CELL Leaf/Stem Unidentified Husks Wheat Husks Barley Husks Aegilops Wild Grass Husks Phragmites Stem Bromus-type Stem Phragmites Leaf Cyperaceae Cereal Straw



appendix 9.5

255

Appendix 9.5 (cont’d.)

Sample

P28

Brick

Mortar

S2

S7

S11

183000 386000 82000 202000 61000 426000 2436000 2537000 0 873000 386000 61000 5055000 202000 386000 40000 0 21000 21000 40000 224000 40094000 264000 2212000 101000 527000 0 0 0

109065 19830 29745 39660 0 0 0 198300 0 0 39660 19830 347024 49575 69405 9915 0 19830 0 59490 19830 0 0 29745 1339762 19830 0 19830 0

362000 28000 28000 0 0 0 0 1727000 0 1031000 111000 28000 3483000 0 139000 28000 0 0 0 0 84000 0 84000 28000 2563000 56000 28000 56000 557000

0 103000 51000 0 0 0 0 5911000 0 1491000 411000 0 5087000 47000 205000 154000 0 0 51000 0 51000 0 0 51000 0 0 926000 0 0

7000 0 0 0 0 0 0 20000 0 10000 1000 1000 85000 0 1000 0 0 0 0 0 0 0 2000 0 0 3000 0 1000 0

10000 0 7000 0 0 0 0 0 0 0 0 4000 191000 0 6000 0 0 0 0 0 2000 0 0 0 0 0 0 0 0

21000 25000 0 6000 0 7000 0 0 0 0 0

32224 1239 0 0 0 0 0 0 0 0 0

24000 0 0 0 0 0 0 0 0 1000 0

17000 0 0 0 2000 0 0 0 0 2000 0

0 0 0 0 0 0 0 0 0 0 0

0 0 0 0 0 0 0 0 0 0 0

SINGLE-CELL Long (Smooth) Long (Sinuate) Long (Rods) Long (Dendritic) Papillae Hairs Trichomes Bulliforms Ovals Keystones Crenates Bilobes Rondels Saddles Cones Flat Tower Horned Tower Rugulose Spheroid Smooth Spheroid Bottle-shaped Elongate Tracheids Two-tiered Blocks Platey Sheet Single Polyhedron Scalloped Single Jigsaw Puzzle MULTI-CELL Leaf/Stem Unidentified Husks Wheat Husks Barley Husks Aegilops Wild Grass Husks Phragmites Stem Bromus-type Stem Phragmites Leaf Cyperaceae Cereal Straw

256

appendix 9.6

Appendix 9.6: Jeitun: Summary Table of the Archaeobotanical Composition of the Samples (only wild plant types occurring in five or more samples are included)

Michael Charles and Amy Bogaard ➾ Soil Sample # Context Type Vol. Floated (L) No. of Items/L Cereal Grain

8911 9023 9025 9032 9034 9042 9044 9052 9107

9111

9117

9118 9121 9122 9125 9126 9129 9131 9132 ashy ashy ashy ashy floor ctyd. ctyd. ctyd. ctyd. ctyd. ctyd. floor other hearth hearth dump hearth dump hearth dump hearth other dump 5 88

30 62

15 124

30 54

7.5 57

5 229

10 142

10 24

5.0 66

11 25

5.5 247

5 100

5 155

3.75 301

6 268

8 252

7.5 12

7.5 199

1.875 462

Einkorn 1-grained

2

6

8

14

3

5

14

7

0

0

2

4

2

4

7

7

9

2

2

Einkorn 2-grained

2

7

5

2

8

7

13

15

0

1

2

7

1

40

43

19

8

2

3

Einkorn indeterminate

0

5

1

4

0

0

0

1

2

2

0

5

1

25

8

2

0

1

1

Emmer grain, cf.

0

0

0

0

0

0

1

0

0

0

0

0

0

0

0

1

0

0

0

Glume wheat indeterminate

3

2

8

5

5

0

11

15

2

1

0

0

0

3

6

0

6

0

3

Hulled barley grain

0

0

0

0

0

0

0

0

0

0

0

0

0

1

1

0

1

0

0

Naked barley grain

1

1

1

1

3

2

4

2

0

0

2

2

1

1

3

2

0

0

1

Barley indeterminate grain

1

0

1

5

2

0

0

18

0

0

2

0

0

0

0

0

0

3

3

Cereal/large grass indeterminate grain

0

9

4

2

15

1

0

0

0

3

2

0

0

0

1

3

0

0

2

4

0

0

2

12

0

40

4

2

0

0

4

2

0

0

0

0

0

0

196

134

209

486

59

98

1124

71

162

256

46

250

115

331

254

249

24

30

143

cf. Free-threshing wheat rachis

0

0

0

0

0

0

0

0

0

1

0

0

4

0

0

0

0

0

0

Basal wheat rachis

0

0

0

0

0

0

0

0

0

0

0

0

4

0

0

0

2

0

0

Naked barley rachis

0

0

0

0

0

0

4

0

0

0

2

0

0

0

0

0

0

0

1

Barley indeterminate rachis

0

0

0

0

0

0

0

0

1

1

0

0

0

0

0

0

0

0

0

Basal barley rachis

0

0

0

0

0

0

2

0

0

0

0

0

0

0

0

0

0

0

0

Culm nodes (>1 mm diam.)

0

4

1

1

0

0

0

1

0

0

0

2

0

1

4

0

0

0

1

Culm nodes (1 mm diam.)

0

0

0

0

0

0

1

3

2

2

1

0

0

0

0

0

0

0

0

0

Culm nodes (3mm fraction

lyb

12.6

12.6

2.3

distal blade fragment

No

42

>3mm fraction

lyb

broken flake

No

127

>3mm fraction

lyb

32.5

13

4.3

complete blade

Yes

131

>3mm fraction

g

chip

No

133

dry sieve

lyb

37.5

13.9

3.1

proximal blade fragment

No

135

>3mm fraction

vdb

chip

No

135

>3mm fraction

lyb

13.5

11.2

1.6

distal blade fragment

No

136

dry sieve

lyb

18.2

8.5

1.4

medial blade fragment

No

137

dry sieve

g

11.0

14.8

6.3

burnt shatter

No

137

dry sieve

lyb

16.2

13.4

2.1

distal blade fragment

Yes

137

dry sieve

myb

13.0

15.3

2.7

broken flake

No

10.6 1–33 %

139

dry sieve

g

35.5

14.5

2.4

medial blade fragment

No

142

dry sieve

lyb

48.1

11.1

4.1

proximal blade fragment

Yes

144

>3mm fraction

lyb

8.1

complete flake

No

144

>3mm fraction

lyb

4.9

proximal blade fragment

No

144

>3mm fraction

rb

chip

No

144

>3mm fraction

myb

chip

No

145

dry sieve

myb

proximal blade fragment

No

146

>3mm fraction

g

chip

No

146

>3mm fraction

g

chip

No



52.4

9.4

2.6



appendix 9.11

267

Appendix 9.11 (cont’d.) Length (mm) Width (mm) Thickness (mm)

Context

Recovery

Color (see Table 9.21)

146

>3mm fraction

pr

chip

No

146

>3mm fraction

rb

chip

No

146

>3mm fraction

g

chip

No

146

>3mm fraction

g

chip

No

146

>3mm fraction

w

chip

No

146

>3mm fraction

pr

broken flake

No

146

>3mm fraction

myb

chip

No

146

>3mm fraction

g

distal blade fragment

No

146

>3mm fraction

g

chip

No

146

>3mm fraction

lyb

chip

No

147

dry sieve

dyb

proximal blade fragment

No

148

>1mm

lyb

chip

No

148

>1mm

lyb

1–33 %

chip

No

148

>1mm

rb

1–33 %

chip

No

148

>3mm fraction

lyb

chip

No

148

>3mm fraction

lyb

chip

No

148

>3mm fraction

g

chip

No

148

>3mm fraction

g

chip

No

148

>3mm fraction

myb

148

>3mm fraction

myb

148

>3mm fraction

148

>3mm fraction

152

>3mm fraction

lyb

16.1

11.8

152

>3mm fraction

myb

16.7

9 9.8

Cortex

Technological Category

6.0

7.8

1.2

10.8

5.5

1.4

26.6

11.1

2.6

Used?

chip

No

6.5

proximal blade fragment

No

g

5.9

complete flake

No

myb

3.5

complete flake

No

2.5

medial blade fragment

Yes

2.9

medial blade fragment

Yes

2.0

152

dry sieve

lyb

7.5

complete flake

No

154

>3mm fraction

myb

10.3

broken flake

No

154

>3mm fraction

myb

4.3

broken flake

No

154

>3mm fraction

lyb

5.7

complete flake

No

154

>3mm fraction

lyb

11.5

flake fragment

No

154

>3mm fraction

lyb

12.4

flake fragment

No

154

>3mm fraction

lyb

19.3

flake fragment

No

154

>3mm fraction

dyb

11.3

flake fragment

No

155

dry sieve

myb

56.5

12

2.5

proximal blade fragment

Yes

155

dry sieve

lyb

38.0

9

1.8

proximal blade fragment

No

155

dry sieve

myb

18.8

10

2.5

distal blade fragment

No

158

dry sieve

lyb

22.7

11.4

2.7

proximal blade fragment

Yes

158

dry sieve

rb

36.7

10.2

2.5

proximal blade fragment

Yes

158

dry sieve

rb

32.1

14.1

2.5

proximal blade fragment

Yes

161

>3mm fraction

myb

46.2

11.5

5.0

proximal blade fragment

Yes

162

>3mm fraction

lyb

chip

No

162

dry sieve

lyb

proximal blade fragment

Yes

1–33 %

29.4

9.6

2.8



268

appendix 9.11

Appendix 9.11 (cont’d.) Length (mm) Width (mm) Thickness (mm)

Context

Recovery

Color (see Table 9.21)

162

dry sieve

myb

164

>3mm fraction

rb

164

>3mm fraction

w

11.4

16.5

166

dry sieve

vdb

14.0

9.2

166

>3mm fraction

myb

21.5

166

dry sieve

lyb

9.8

Cortex 1–33 %

16.7

11.5

Technological Category 3.1

Used?

flake fragment

No

chip

No

3.2

medial blade fragment

No

2.2

medial blade fragment

Yes

13.6

4.2

complete flake

No

9.4

2.3

medial blade fragment

Yes

167

dry sieve

lyb

28.4

10.6

2.7

proximal blade fragment

No

170

dry sieve

dyb

22.9

12.5

3.6

proximal blade fragment

Yes

170

dry sieve

g

7.1

11.9

3.6

f­lake fragment

Yes

171

dry sieve

dyb

26.6

12.8

3.0

medial blade fragment

No

172

>3mm fraction

lyb

3.8

complete flake

No

172

dry sieve

rb

22.5

10.9

2.6

proximal blade fragment

Yes

172

dry sieve

myb

172

dry sieve

lyb

1–33 %

16.9

14.3

3.3

proximal blade fragment

Yes

46.7

10.6

3.2

complete blade

Yes

32.8

10.9

2.3

proximal blade fragment

Yes

27.5

13.5

3.1

proximal blade fragment

No

18.3

10.1

2.7

medial blade fragment

No

13.4

10.8

2.6

flake fragment

No

66.6

14.9

5.5

complete blade

No

173

dry sieve

dyb

173

dry sieve

dyb

178

dry sieve

vdb

179

dry sieve

lyb

179

>3mm fraction

lyb

179

>3mm fraction

myb

chip

No

179

>3mm fraction

rb

chip

No

179

dry sieve

myb

23.1

11.4

4.1

medial blade fragment

Yes

181

dry sieve

myb

57.7

9.9

2.7

complete blade

No

181

dry sieve

dyb

20.1

10

3.8

medial blade fragment

Yes

184

lyb

62.5

11.7

3.7

proximal blade fragment

Yes

chip

No

219

dry sieve >3mm fraction dry sieve

223

dry sieve

lyb

223

dry sieve

dyb

228

dry sieve

dyb

14.2

229

dry sieve

dyb

37.4

195

33–66 % 1–33 %

g pr

33.4 1–33 %

9.8

3.0

medial blade fragment

Yes

46.0

9.8

2.7

proximal blade fragment

Yes

23.8

11.9

3.1

medial blade fragment

Yes

12.1

1.6

medial blade fragment

No

13.2

3.6

distal blade fragment

No

231

>3mm fraction

pr

7.5

10.9

2.1

medial blade fragment

No

231

>3mm fraction

lyb

13.9

8.5

2.1

medial blade fragment

Yes

239

>3mm fraction

lyb

8.9

broken flake

No

241

dry sieve

myb

28.8

11.4

2.0

distal blade fragment

Yes

243

dry sieve

dyb

8.8

8.8

3.1

medial blade fragment

Yes

245

dry sieve

dyb

10.3

8.9

1.4

medial blade fragment

No

Bibliography Aitken, M. J. 1990. Science-based Dating in Archaeology. London: Longman. Aladin, N. V., I. S. Plotnikov, M. I. Orlova, A. A. Filipov, A. O. Smurov, D. D. Pirulin, , O. M. Rusakova, and L. V. Zhakova. 1996. Changes in the Form and Biota of the Aral Sea over Time. In The Aral Sea Basin, ed. P. P. Micklin and W. D. Williams, pp. 33–55. Heidelberg: Springer-Verlag. Albarella, U., K. Dobney, and P. Rowley-Conwy. 2006. The Domestication of the Pig (Sus scrofa): New Challenges and Approaches. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 209–27. Berkeley: University of California Press. Albert, R. M., and S. Weiner. 2001. Study of Phytoliths in Prehistoric Ash Layers from Kebara and Tabun Caves Using a Quantitative Approach. In Phytoliths: Applications in Earth Sciences and Human History, ed. J. D. Meunier and F. Colin, pp. 251– 66. Lisse: A. A. Balkema. Alessio, M., F. Bachechi, and C. Cortesi. 1967. University of Rome Carbon-14 Dates. Radiocarbon 9:360–62. Allaby, R. G., and T. A. Brown. 2003. AFLP Data and the Origins of Domesticated Crops. Genome 46:448–53. ———. 2004. Reply to the Comment by Salamini et al. on “AFLP Data and the Origins of Domesticated Crops.” Genome 47:621–22. Allaby, R. G., T. A. Brown, and D. Q. Fuller. 2009. A Simulation of the Effect of Inbreeding on Crop Domestication Genetics with Comments on the Integration of Archaeobotany and Genetics: A Reply to Honne and Heun, Vegetation History and Archaeobotany 19:151–58.

Allaby, R. G., D. Q. Fuller, and T. A. Brown. 2008. The Genetic Expectations of a Protracted Model for the Origins of Domesticated Crops. Proceedings of the National Academy of Sciences, USA 105:13,982–86. Allchin, B., and R. Allchin. 1982. The Rise of Civilization in India and Pakistan. Cambridge: Cambridge University Press. Alley, R. B., and A. M. Ágústsdóttir. 2005. The 8.2 Event: Cause and Consequences of a Major Holocene Abrupt Climate Change. Quaternary Science Reviews 24:1,123–49. Alley, R. B., P. A. Mayewski, T. Sowers, M. Stuiver, K. C. Taylor, and P. U. Clark. 1997. Holocene Climatic Instability: A Prominent Widespread Event 8200 Years Ago. Geology 25:483–86. Ammerman, A. J., and P. Biagi, eds. 2003. The Widening Harvest: The Neolithic Transition in Europe, Looking Back, Looking Forward. Boston: Archaeological Institute of America. Anderberg, A-L. 1994. Atlas of Seeds and Small Fruits of Northwest-European Plant Species, with Morphological Descriptions. Part 4, Resedaceae-Umbelliferae. Stockholm: Swedish Natural Science Research Council. Angel, J. L. 1972. A Middle Palaeolithic Temporal Bone from Darra-I Kur, Afghanistan. In Prehistoric Research in Afghanistan (1959–1966) by L. Dupree, pp. 54−56. Transactions of the American Philosophical Society 62(4). Philadelphia. Arne, T. J. 1945. Excavations at Shah Tepe, Iran. Publication 27, VII, Archaeology 5, of the Sino–Swedish Expedition. Stockholm. Asouti, E. 2003. Woodland Vegetation and Fuel Exploitation at the Prehistoric Campsite of Pinarba şı, South-Central Anatolia, Turkey: The Evidence

270

bibliography

from the Wood Charcoal Macro-remains. Journal of Archaeological Science 30:1185–1201. ———. 2005. Woodland Vegetation and the Exploitation of Fuel and Timber at Neolithic Çatalhöyük: Report on the Wood Charcoal Macro-remains. In Inhabiting Çatalhöyük: Reports from the 1995–99 Seasons, ed. I. Hodder, pp. 459–67. Çatalhöyük Research Project Volume 4. Cambridge: McDonald Institute for Archaeological Research. Asouti, E., and P. Austin. 2005. Reconstructing Woodland Vegetation and Its Exploitation by Past Societies, Based on the Analysis and Interpretation of Archaeological Wood Charcoal Macro-remains. Environmental Archaeology 10:1–18. Ataev, C., A. K. Rustamov, and S. Shammakov. 1994. Reptiles of Kopetdagh. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 329–50. Dordrecht: Kluwer. Atagarryev, E., and O. K. Berdiev. 1967. Arkheologicheskoye izuchenie Turkmenistana za gody Sovietskoi vlasti’ [The archaeological study of Turkmenistan during the Soviet period]. Sovietskaya Arkheologiya (3): 124–41. ———. 1970. The Archaeological Exploration of Turkmenistan in the Years of Soviet Power. East and West 20:285–306 (English version of Atagarryev and Berdiev 1967). Atamuradov, K. I. 1994. Paleogeography of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 49–64. Dordrecht: Kluwer. Aurenche, O. 1985. La tradition architecturale à l’est de la Mésopotamie (Iran–Turkménistan) du 10e au 4e millénaires. In L’Archéologie de la Bactriane Ancienne, Actes du Colloque franco–soviétique, Dushanbe (U.R.S.S.), 27 octobre à 3 novembre 1982, pp. 235–39 and Figs. 1–12 (on unnumbered pages). Paris: Éditions du Centre Nationale de la Recherche Scientifique. Azarnoush, M., and B. Helwing. 2005. Recent Archaeological Research in Iran—Prehistory to Iron Age. Archäologische Mitteilungen aus Iran und Turan 37:189–246. Babaev, A. G. 1994. Landscapes of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 5–22. Dordrecht: Kluwer. Badr, A., K. Müller, R. Schäfer-Pregl, H. El Rabey, S. Effgen, H. H. Ibrahim, C. Pozzi, W. Rohde, and F. Salamini. 2000. On the Origin and Domestica-

tion History of Barley (Hordeum vulgare). Molecular Biology and Evolution 17:499–510. Baig, M., A. Beja-Pereira, R. Mohammad, K. Kulkarni, S. Farah, and G. Luikart. 2005. Phylogeography and Origin of Indian Domestic Cattle. Current Science 89:38–40. Baird, D. 2005. Pınarba şı. Anatolian Archaeology 11:12– 13. ———. 2006. A Report on Excavations at Pınarba şı, Karaman Province, 2003. http://www.liv.ac.uk/ sacos/research/projects/pinarba ş i/index.html [accessed 8 June 2009]. Barker, G. 1985. Prehistoric Farming in Europe. Cambridge: Cambridge University Press. ———. 2006. The Agricultural Revolution in Prehistory: Why Did Foragers Become Farmers? Oxford: Oxford University Press. Bar-Yosef, O., and A. Belfer-Cohen. 2002. Facing Environmental Crisis: Societal and Cultural Changes at the Transition from the Younger Dryas to the Holocene in the Levant. In The Dawn of Farming in the Near East, ed. R. T. J. Cappers and S. Bottema, pp. 55–66. Studies in Early Near Eastern Production, Subsistence and Environment 6. Berlin: Ex Oriente. Beijerinck, W. 1947. Zadenatlas der Nederlandsche Flora. Wageningen, Gelderland: Veenman. Berdiev, O. K. 1963a. Yuzhnaya Turkmenia v epokhu Neolita [Southern Turkmenia in the Neolithic period]. Moscow: Academy of Sciences, Institut Etnografii. ———. 1963b. Stratigrafiya Bamiskogo poselenia [Stratigraphy of the Bami settlement]. Sovietskaya Arkheologiya (4): 188–94. ———. 1964a. Chagylly-depe. Izvestiya Akademii nauk Turmenskoi SSR (1): 89–95. ———. 1964b. Stratigrafiya Togolok-depe v svyazi s poseleniem plemen Dzheitunskoi kultury [Stratigraphy of Togolok-depe in conjunction with the settlements of the Jeitun-Culture tribes]. Sovietskaya Arkheologiya (3): 271–77. ———. 1964c. Razvitie arkheologicheskoi nauki v Turkmenistane za gody Sovietskoi vlasti [The advance of archaeological science in Turkmenistan during the years of Soviet power]. Izvestiya Akademii nauk Turkmenskoi SSR (5): 38–45. ———. 1965. Neolit Novoi Nisy [The Neolithic of New Nissa]. Sovietskaya Arkheologiya (2): 237–42. ———. 1966. Chagylly-depe—novyi pamyatnik Neoliticheskoi Dzheitunskoi kultury [Chagylly-depe—a



bibliography

new site of the Neolithic Jeitun Culture]. In Materialnaya kultura Narodov Srednei Azii i Kazakhstana, ed. N. A. Kislyakov and M. G. Vorobyova, pp. 3–28. Moscow: Akademiya nauk. ———. 1968a. Chakmakli-depe—novii pamyatnik vremeni Anau IA [Chakmakli-depe—new site of the Anau IA period]. In Istoriya, Arkheologiya i Etnografiya Srednei Azii, ed. V. M. Masson, pp. 26–34. Moscow: Nauka. ———. 1968b. Novye raskopki na poseleniyakh Pessedjik-depe i Chakmakli-depe [New excavations at the sites of Pessedjik-depe and Chakmakli-depe]. Karakumskie Drevnosti (2): 10–17. ———. 1968c. Izuchenie pamyatnikov epokhi Neolita [A study of sites of the Neolithic period]. Karakumskie Drevnosti (1): 9–17. ———. 1969. Drevneishie zemledeltsy Yuzhnogo Turkmenistana [The Most Ancient Agriculturalists of Southern Turkmenistan]. Ashgabat: Y lym. ———. 1970. Nekotorye rezultaty izucheniya drev­ nezemledelcheskikh poselenii [Some results from the study of ancient agricultural settlements]. Karakumskie Drevnosti (3): 14–31. ———. 1971a. Yuzhnaya Turkmenia i Iran v epokhu kamnya i rannego metalla [Southern Turkmenia and Iran in the stone and early metal ages]. In Istoriya Iranskogo gosudarstva i kultury, ed. B. G. Gafurov, E. A. Grantovskii, and M. S. Ivanov, pp. 164–69. Moscow: Glavnaya redaktsiya vostochnoi literatury. ———. 1971b. Neopublikovannye materialy iz neoliticheskikh pamyatnikov yuzhnogo Turkmenistana [Unpublished materials from Neolithic sites in southern Turkmenistan]. Izvestia Akademii nauk Turkmenskoi SSR (3): 76–81. ———. 1971c. Nekotorye rezultaty arkheologicheskikh rabot Instituta istorii imeni S. Batyrova Akademii nauk TSSR (1959–1966) [Some results from archaeological work by the Batyrov Institute of History of the Academy of Sciences of the Turkmenian SSR (1959–1966)]. In Materialnaya kultura Turkmenistana 1, ed. V. M. Masson, pp. 7–22. Ashgabat: Ylym. ———. 1972a. Chopan-depe—pamyatnik Neolitiches­ koi Dzheitunskoi kultury [Chopan-depe—a site of the Neolithic Jeitun Culture]. In Materialnaya kultura Turkmenistana 1, ed. V. M. Masson, pp. 60–86. Ashgabat: Y lym. ———. 1972b. Mondzhukli-depe—mnogoslo­inoye pose­­­ l­enie neolita i rannego eneolita v Yuzhnom Turk­

271

menistane [Monjukli-depe—a multi-level settlement of the Neolithic and Early Eneolithic in southern Turkmenistan]. Karakumskie Drevnosti (4): 11–34. ———. 1973. Nekotorie itogi izucheniya poseleniya neoliticheskoi Dzheitunskoi kultury Pessedjikdepe [Some results of the study of the Pessedjikdepe settlement of the Neolithic Jeitun Culture]. TD Spipai 1972 [Abstracts of papers for the session devoted to the results of archaeological field investigations in 1972 in the USSR]: 211–13. Tashkent: FAN. ———. 1974. Keramika vremeni Anau 1A [Ceramics of the Anau IA period]. In Materialnaya kultura Turkmenistana 2, ed. V. M. Masson, pp. 5–38. Ashgabat: Y lym. ———. 1976. Materialnaya kultura Turkmenistana v periode neolita i rannego eneolita [Material culture of Turkmenistan in the Neolithic and Early Eneolithic]. In Pervobytnyi Turkmenistan, ed. V. M. Masson, pp. 14–82. Ashgabat: Ylym. Berezkin, Y. E. 1970. Opyt klassifikatsii Dzheitunskoi keramiki s ispolzovaniem statisticheskikh metodov [Attempt at a classification of Jeitun pottery using statistical methods]. Karakumski Drevnosti (3): 143–50. ———. 1992. Raskopki na poselenii Dzheitun v 1989– 1991 gg. Predvaritelnye rezultaty [Excavations at Jeitun during the years 1989–1991. Preliminary results]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 13–30. Ashgabat: Ylym. Berggren, G. 1969. Atlas of Seeds and Small Fruits of Northwest-European Plant Species, with Morphological Descriptions. Part 2, Cyperaceae. Stockholm: Swedish Natural Science Research Council. ———. 1981. Atlas of Seeds and Small Fruits of Northwest-European Plant Species, with Morphological Descriptions. Part 3, Salicaceae-Cruciferae. Stockholm: Swedish Natural Science Research Council. Binford, L. R. 1968. Post-Pleistocene Adaptations. In New Perspectives in Archeology, ed. S. R. Binford and L. R. Binford, pp. 313–41. Chicago: Aldine. Blakelock, R. A., and C. C. Townsend. 1980. Capparidaceae. In Flora of Iraq, Vol. 4, Part 1, ed. C. C. Townsend and E. Guest, pp. 139–45. Baghdad: Ministry of Agriculture and Agrarian Reform. Boessneck, J. 1969. Osteological Differences between Sheep (Ovies aries Linné) and Goat (Capra hircus Linné). In Science in Archaeology, ed. D. Brothwell

272

bibliography

and E. S. Higgs, pp. 331–58. London: Thames and Hudson. Bogaard, A. 2004. Neolithic Farming in Central Europe. London: Routledge. Bogucki, P. 1996. The Spread of Early Farming in Europe. American Scientist 84:242–53. Boivin, N. 2000. Life Rhythms and Floor Sequences: Excavating Time in Rural Rajasthan and Neolithic Çatalhöyük. World Archaeology 31(3): 367–88. Bökönyi, S., and L. Bartosiewicz. 2000. A Review of Animal Remains from Shahr-i Sokhta (Eastern Iran). In Archaeozoology of the Near East IVB, ed. M. Mashkour, A. M. Choyke, H. Buitenhuis, and F. Poplin, pp. 116–52. Groningen: Centre for Archaeological Research and Consultancy. Boomer, I., N. Aladin, I. Plotnikov, and R. Whatley. 2000. The Palaeolimnology of the Aral Sea: A Review. Quaternary Science Reviews 19:1259–78. Bowes, B. G. 1996. A Colour Atlas of Plant Structure. London: Manson. Boyko, A. R., R. H. Boyko, C. M. Boyko, H. G. Parker, M. Castelhano, L. Corey, J. D. Degenhardt, A. Auton, M. Hedimbi, R. Kityo, E. A. Ostrander, J. Schoenebeck, R. J. Todhunter, P. Jones, and C. D. Bustamente. 2009. Complex Population Structure in African Village Dog Domestication History. Proceedings of the National Academy of Sciences, USA 106:13,903–8. Bozarth, S. R. 1992. Classification of Opal Phytoliths Formed in Selected Dicotyledons Native to the Great Plains. In Phytolith Systematics: Emerging Issues, ed. G. Rapp and S. C. Mulholland, pp. 193– 214. New York: Plenum Press. Bracegirdle, B., and P. H. Miles. 1971. An Atlas of Plant Structure, Vol. 1. London: Heinemann Educational Books. Bradley, D. G., D. E. MacHugh, P. Cunningham, and R. T. Loftus. 1996. Mitochondrial Diversity and the Origins of African and European Cattle. Proceedings of the National Academy of Science, USA 93:5131–35. Bradley, D. G., and D. A. Magee. 2006. Genetics and the Origins of Domestic Cattle. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 317–28. Berkeley: University of California Press. Braidwood, L. S. 1983. Additional Remarks on the Jarmo Obsidian. In Prehistoric Archeology along the Zagros Flanks, ed. L. S. Braidwood, R. J. Braid-

wood, B. Howe, C. A. Reed, and P. J. Watson, pp. 285–88. Oriental Institute Publications 105. Chicago: University of Chicago. Braidwood, R. J., and B. Howe. 1960. Prehistoric Investigations in Iraqi Kurdistan. Studies in Ancient Oriental Civilization 31. Chicago: University of Chicago Press. Bridges, E. M. 1997. World Soils. 3rd ed. Cambridge: Cambridge University Press. Bronger, A., R. Winer, O. Derevjanko, and S. Aldag. 1995. Loess-Palaeosol Sequences in Tadjikistan as a Palaeoclimatic Record of the Quaternary of Central Asia. In Wind Blown Sediments in the Quaternary Record, ed. E. Derbyshire, pp. 69–81. Quaternary Proceedings 4. West Sussex: Quaternary Research Association and J. Wiley. Bronk Ramsey, C. 1995. Radiocarbon Calibration and Analysis of Stratigraphy: The OxCAL program. Radiocarbon 37:425–30. ———. 1998. Probability and Dating. Radiocarbon 40: 461–74. ———. 2000. Comment on “The Use of Bayesian Statistics for 14C Dates of Chronologically Ordered Samples: A Critical Analysis.” Radiocarbon 42:199–202. ———. 2001. Development of the Radiocarbon Calibration Program. Radiocarbon 43:355–63. ———. 2002. OxCAL 3.8. Oxford: University of Oxford Research Laboratory for Archaeology and the History of Art. ———. 2005. OxCAL. 3.10. Oxford: University of Oxford Research Laboratory for Archaeology and the History of Art. Brown, T. A., M. K. Jones, W. Powell, and R. G. Allaby. 2009. The Complex Origins of Domesticated Crops in the Fertile Crescent. Trends in Ecology and Evolution 24:103–9. Bruford, M. W., D. G. Bradley, and G. Luikart. 2003. DNA Markers Reveal the Complexity of Livestock Domestication. Nature Reviews Genetics 4:900–910. Bruford, M. W., and S. L. Townsend. 2006. Mitochondrial DNA Diversity in Modern Sheep. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 306–16. Berkeley: University of California Press. Brun, P. 2005. From Arrows to Bullets: The Fortifications of Abdullah Khan Kala (Merv, Turkmenistan). Antiquity 79:616–24.



bibliography

Brunet, F. 1999. La néolithisation en Asie centrale: un état de la question. Paléorient 24(2): 27–48. ———. 2002. Asie centrale: vers une redéfinition des complexes culturels de la fin du Pléistocène et des débuts de l’Holocène. Paléorient 28(2): 9–24. ———. 2003. Du Mésolithique en Asie centrale. Addenda. Paléorient 29(1): 167–69. ———. 2005. Pour une nouvelle étude de la culture néolithique de Kel’teminar, Ouzbékistan. Paléorient 31(2): 87–105. Bukinich, D. D. 1924. Istoriya pervotbytnogo oroshaemogo zem­ledeliya v Zakaspiiskoi oblasti v svyazi s voprosom o proiskhozhdenii zemledeliya i skvotovodstva [History of primitive irrigation agriculture in the Transcaspian region and its connection with questions concerning the origin of agriculture and herding]. Khlopkovoe Delo 3–4:92–135. Bulliet, R. W. 1975. The Camel and the Wheel. Cambridge, Mass.: Harvard University Press. Bullock, P., N. Federoff, A. Jongeruis, G. Stoops, and T. Tursina. 1985. The Handbook for Soil Thin Section Description. Albrighton, Shropshire: Waine Research Publications. Burger, J. C., M. A. Chapman, and J. M. Burke. 2008. Molecular Insights into the Evolution of Crop Plants. American Journal of Botany 95:113–22. Canti, M. G. 1997. An Investigation of Microscopic Calcareous Spherulites from Herbivore Dungs. Journal of Archaeological Science 24:219–31. ———. 1998. The Micromorphological Identification of Faecal Spherulites from Archaeological and Modern Materials. Journal of Archaeological Science 25:435–44. ———. 1999. The Production and Preservation of Faecal Spherulites: Animals, Environment and Taphonomy. Journal of Archaeological Science 26: 251–58. Cappers, R. T. J., T. Van Thuyne, and L. Sikking. 2004. Plant Remains from Predynastic El Abadiya 2 (Naqada area, Upper Egypt). In Egypt at Its Origins: Studies in Memory of Barbara Adams, ed. S. Hendrickx, R. F. Friedman, R. M. Cialowicz, and M. Chlodnicki, pp. 277–93. Orientalia Lovaniensia Analecta 138. Leuven: Uitgeverij Peeters en Departement Oosterse Studies. Carter, K. E., and R. Stoker. 1985. Effects of Irrigation and Sowing Date on Yield and Quality of Barley and Wheat. New Zealand Journal of Experimental Agriculture 13:77–83. Charles, M. 1984. Introductory Remarks on the Cere-

273

als. Bulletin of Sumerian Agriculture 1:17–31. ———. 1998. Fodder from Dung: The Recognition and Interpretation of Dung-derived Plant Material from Archaeological Sites. Environmental Archaeology 1:111–22. ———. 2007. East of Eden? A Consideration of Neolithic Crop Spectra in the Eastern Fertile Crescent and Beyond. In The Origins and Spread of Domestic Plants in Southwest Asia and Europe, ed. S. Colledge and J. Conolly, pp. 37–51. Walnut Creek, CA: Left Coast Press. Charles, M., and G. Hillman. 1992. Vyrashchivanie selsko-khozyaist vennykh kultur v pustynnoi zone (po rezultatam izucheniya karbonizirovannykh rastitelnykh makroostatkov) [Crop husbandry in a desert environment (evidence from the charred plant macro-remains)]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 83–94. Ashgabat: Ylym. Charles, M., C. Hoppé, A. Bogaard, G. Jones, and J. G. Hodgson. 2003. Using Weed Functional Attributes for the Identification of Irrigation Regimes in Jordan. Journal of Archaeological Science 30:1429–41. Chen, F., Z. Yu, M. Yang, E. Ito, S. Wang, D. B. Madsen, X. Huang, Y. Zhao, T. Sato, H. J. B. Birks, I. Boomer, J. Chen, C. An, and B. Wünnemann. 2008. Holocene Moisture Evolution in Arid Central Asia and Its Out-of-Phase Relationship with Asian Monsoon History. Quaternary Science Reviews 27:351–64. Chen, S., B-Z. Lin, M. Baig, B. Mitra, R. J. Lopes, A. M. Santos, D. A. Magee, M. Azevedo, P. Tarroso, S. Sasazaki, S. Ostrowski, O. Mahgoub, T K. Chaudhuri, Y-p Zhang, V. Costa, L. J. Royo, F. Goyache, G. Luikart, N. Boivin, D. Q. Fuller, H. Mannen, D. G. Bradley and A. Beja-Pereira. 2010. Zebu Cattle Are an Exclusive Legacy of the South Asian Neolithic. Molecular Biology and Evolution 27:1−6. Chepalyga, A. L. 2007. The Late Glacial Great Flood in the Ponto–Caspian Basin. In The Black Sea Flood Question: Changes in Coastline, Climate, and Human Settlement, ed. Y. Yanko-Hombach, A. S. Gilbert, N. Panin, and P. M. Dolukhanov, pp. 119–48. Dordrecht: Springer. Childe, V. G. 1928. The Most Ancient East: The Oriental Prelude to European Prehistory. London: Kegan Paul, Trench, Trubner. ———. 1934. New Light on the Most Ancient East. Lon-

274

bibliography

don: Routledge & Kegan Paul. Clutton-Brock, J. 1979. The Mammalian Remains from the Jericho Tell. Proceedings of the Prehistoric Society 45:135–57. ———. 1992. Horse Power: A History of the Horse and the Donkey in Human Societies. London: Natural History Museum Publications. ———. 1995. Origins of the Dog: Domestication and Early History. In The Domestic Dog, Its Evolution and Behaviour and Interactions with People, ed. J. Serpell, pp. 7–20. Cambridge: Cambridge University Press. ———. 1999. A Natural History of Domesticated Mammals. Cambridge: Cambridge University Press. Cohen, D. 1998. The Origins of Domesticated Cereals and the Pleistocene–Holocene Transition in China. The Review of Archaeology 19(2): 22–29. COHMAP Members. 1988. Climatic Changes of the Last 18,000 Years: Observations and Model Simulations. Science 241:1043–52. Colledge, S., and J. Conolly, eds. 2007. The Origins and Spread of Domestic Plants in Southwest Asia and Europe. Walnut Creek, CA: Left Coast Press. Compagnoni, B., and M. Tosi. 1978. The Camel: Its Distribution and State of Domestication in the Middle East during the Third Millennium B.C. in Light of Finds from Shahr-i Sokhta. In Approaches to Faunal Analysis in the Middle East, ed. R. H. Meadow and M. A. Zeder, pp. 91–103. Peabody Museum Bulletin 2. Cambridge, MA: Peabody Museum of Archaeology and Ethnology, Harvard University. Coningham, R. A. E., H. Fazeli, R. L. Young, and R. E. Donahue. 2004. Location, Location, Location: A Pilot Survey of the Tehran Plain in 2003. Iran 42:1–12. Coningham, R. A. E., H. Fazeli, R. L. Young, G. K. Gilmore, H. Karamian, M. Maghsoudi, R. E. Donahue, and C. M. Batt. 2006. Socio-economic Transformations: Settlement Survey in the Tehran Plain and Excavations at Tepe Pardis. Iran 44:33–62. Conolly, J. 1999. The Çatalhöyük Flint and Obsidian Industry. British Archaeological Reports International Series 787. Oxford: Archaeopress. Coolidge, J. 2005. Southern Turkmenistan in the Neolithic: A Petrographic Case Study. British Archaeological Reports International Series 1423. Oxford: Archaeopress. Coon, C. S. 1951. Cave Explorations in Iran 1949. Philadelphia: University of Pennsylvania Museum. ———. 1952. Excavations in Hotu Cave, Iran: A Pre-

liminary Report. Proceedings of the American Philosophical Society 96(3): 231–49. ———. 1957. Seven Caves: Archaeological Explorations in the Middle East. London: Jonathan Cape. Coon, C. S., and E. K. Ralph. 1955. Radiocarbon Dates for Kara Kamar, Afghanistan, U. of Pa. II. Science 122:921–22. Costantini, L. 1983. The Beginnings of Agriculture in the Kachi Plain: The Evidence of Mehrgarh. In South Asian Archaeology 1981, ed. B. Allchin, pp. 29–33. Cambridge: Cambridge University Press. ———. 2007–08. The First Farmers in Western Pakistan: The Evidence of the Neolithic Agro-Pastoral Settlement at Mehrgarh. Proceedings of the International Seminar “First Farmers in Global Perspective” (Lucknow 18–20 January 2006), ed. R. Tewari, K. S. Saraswat, and G. C. Singh. Prāgdhārā: Journal of the U. P. State Archaeology Department 18:167–78. Crawford, G. W., and G-A. Lee. 2003. Agricultural Origins in the Korean Peninsula. Antiquity 77:87–95. Crawford, V. 1963. Beside the Kara Su. Bulletin of the Metropolitan Museum of Art 22:263–73. Crowfoot Payne, J. 1983. Appendix C: The Flint Industries of Jericho. In Excavations at Jericho. Vol. 5, The Pottery Phases of the Tell and Other Finds, ed. K. M. Kenyon and T. A. Holland, pp. 622–758. London: British School of Archaeology in Jerusalem. Daher, A. G. 1969. A Study of Grave Groups from Tureng Tepe in Late Third and Early Second Millennia B.C. in North-East Iran as Reconstructed from the Excavations during 1931–32. MA thesis, Dept. of Anthropology, University of Pennsylvania. Philadelphia. Davis, P. H., J. Cullen, and M. J. E. Coode. 1965–88. Flora of Turkey. 10 vols. Edinburgh: Edinburgh University Press. Davis, R. S. 1978. The Paleolithic of Afghanistan. In The Archaeology of Afghanistan from Earliest Times to the Timurid Period, ed. N. Hammond and R. Allchin, pp. 37–70. London: Academic Press. ———. 1990. Central Asian Hunter-Gatherers at the Last Glacial Maximum. In The World at 18,000 BP. Vol. 1, High Latitudes, ed. O. Soffer and C. Gamble, pp. 266–75. London: Unwin Hyman. Davis, R. S., and V. A. Ranov. 1999. Recent Work on the Paleolithic of Central Asia. Evolutionary Anthropology 8(5): 186–93. Davis, S. J. M., and F. R. Valla. 1978. Evidence for Domestication of the Dog 12,000 Years Ago in the



bibliography

Natufian of Israel. Nature 276:608–10. Davis, W. M. 1905. A Journey across Turkestan. In Explorations in Turkestan with an Account of the Basin of Eastern Persia and Sistan, Expedition of 1903, ed. R. Pumpelly, pp. 23–119. Carnegie Institution of Washington Publication No. 26. Washington, DC. Dayan, T. 1994. Early Domesticated Dogs of the Near East. Journal of Archaeological Science 21:633–40. de Moulins, D. 1997. Agricultural Changes at Euphrates and Steppe Sites in the Mid-8th to the 6th Millennium B.C. British Archaeological Reports International Series 683. Oxford: Archaeopress. Dennell, R. W. 2009. The Early Palaeolithic Settlement of Asia. Cambridge: Cambridge University Press. Derevyanko, A. P., and E. L. Zun. 1992. Upper Palaeolithic Cultures. In History of Civilizations of Central Asia, Vol. 1, ed. A. H. Dani and V. M. Masson, pp. 89–108. Paris: UNESCO. Deshayes, J. 1963. Rapport préliminaire sur les deux premières campagnes de fouille à Tureng Tépé. Syria: Revue d’Art oriental et d’Archéologie 40:85–99. Ding, Z. L., V. Ranov, S. L. Yang, A. Finaev, J. M. Han, and G. A. Wang. 2002. The Loess Record in Southern Tajikistan and Correlation with the Chinese Loess. Earth and Planetary Science Letters 200:387–400. Dobney, K. 2007. Appendix 1. Recording Protocol for Mammal and Bird Remains. In Farmers, Monks and Aristocrats: The Environmental Archaeology of Anglo-Saxon Flixborough by K. Dobney, D. Jaques, J. Barrett, and C. Johnstone, pp. 248–59. Oxford: Oxbow. Dobney, K., M. Beech, and D. Jaques. 1999. Hunting the “Broad Spectrum Revolution” in Northern Mesopotamia: The Characterisation of Proto/ Early Neolithic Animal Exploitation at Qermez Dere. In Zooarchaeology of the Pleistocene/Holocene Boundary. Proceedings of a Symposium Held at the 8th Congress of the International Council for Archaeozoology (ICAZ), ed. J. Driver, pp. 47–57. British Archaeological Reports International Series 800. Oxford: Archaeopress. Dobney, K., and K. Rielly. 1988. A Method for Recording Archaeological Animal Bones: The Use of Diagnostic Zones. Circaea 5:79–96. Dodonov, A. E., and L. L. Baiguzina. 1995. Loess Stratigraphy of Central Asia: Palaeoclimatic and Palaeoenvironmental Aspects. Quaternary Science Reviews 14:707–20. Dodonov, A. E, T. A. Sadchikova, S. N. Sedov, A. N. Si-

275

makova, and L. P. Zhou. 2006. Multidisciplinary Approach for Paleoenvironmental Reconstruction in Loess–paleosol Studies of the Darai Kalon Section, Southern Tajikistan. Quaternary International 152/153:59–69. Doebley, J. F., B. S. Grant, and B. D. Smith. 2006. The Molecular Genetics of Crop Domestication. Cell 127:1,309–21. Dolukhanov, P. M. 1981. The Ecological Prerequisites for Early Farming in Southern Turkmenia. In The Bronze Age Civilization of Central Asia: Recent Discoveries, ed. P. L. Kohl, pp. 359–85. New York: M. E. Sharpe. ———. 1986. Foragers and Farmers in West-Central Asia. In Hunters in Transition: Mesolithic Societies of Temperate Eurasia and Their Transition to Farming, ed. M. Zvelebil, pp. 121–32. Cambridge: Cambridge University Press. ———. 1994. Environment and Ethnicity in the Middle East. Aldershot: Avebury. ———. 1997. The Pleistocene–Holocene Transition in Northern Eurasia: Environmental Changes and Human Adaptations. Quaternary International 41/42:181–91. Dolukhanov, P. M., A. L. Chepalyga, V. K. Shkatova, and N. V. Lavrentiev. 2009. Late Quaternary Caspian: Sea Levels, Environments and Human Settlement. Open Geography Journal 1:33–47. Duerst, J. U. 1908. Animal Remains from the Excavations at Anau and the Horse of Anau in Its Relation to the History and to the Races of Domesticated Horses. In Explorations in Turkestan, Expedition of 1904, Vol. 2, ed. R. Pumpelly, pp. 341–442. Washington, DC: Carnegie Institution of Washington. Dupree, L. 1952. The Pleistocene Artifacts of Hotu Cave, Iran. Proceedings of the American Philosophical Society 96(3): 250–57. ———. 1963. Deh Morasi Ghundai: A Chalcolithic Site in South-Central Afghanistan, pp. 57–136. Anthropological Papers of the American Museum of Natural History, Vol. 50, No. 2. New York. ———. 1964. Prehistoric Archaeological Surveys and Excavations in Afghanistan: 1959–1960 and 1961–1963. Science 146:638–40. ———. 1967. The Prehistoric Period of Afghanistan. Afghanistan Historical and Cultural Quarterly (Kabul) 20(3): 8–27. ———. 1972. Prehistoric Research in Afghanistan (1959– 1966). Transactions of the American Philosophical Society 62(4). Philadelphia.

276

bibliography

Dupree, L., and B. Howe. 1963. Results of an Archaeological Survey for Stone Age Sites in North Afghanistan. Afghanistan (Kabul) 8(2): 1–15. Dyson, R. H., Jr., and S. M. Howard, eds. 1989. Tappeh Hesār: Reports of the Restudy Project, 1976. Monografie di Mesopotamia II. Florence: Casa Edi­trice Le Lettere. Eastwood, A., G. Lazkov, and A. Newton. 2009. The Red List of Trees of Central Asia. Cambridge, UK: Fauna and Flora International. Ermolova, N. M. 1983. Skotovodstvo i okhota v tsentralnoi Azii v epokhu Eneolita i bronzy (po dannym paleozoologii) [Herding and hunting in Central Asia in the Eneolithic and Bronze Age (on the basis of palaeozoology)]. In Arkheologiya Srednei Azii i Blizhnevo Vostoka: II Sovetsko–Amerikanksii sympozium, pp. 44–48. Tashkent: FAN. ———. 1985. Novye issledovaniya ostatkov mlekopitayushchikh iz eneoliticheskogo poseleniya Anau [New studies of mammalian remains from the Eneolithic settlement of Anau], Izvestiya Akademii nauk Turkmenskoi SSR: Seriya Obshchestvennykh nauk 1:85–87. Ervynck, A., K. Dobney, H. Hongo, and R. Meadow. 2001. Born Free? New Evidence for the Status of Sus scrofa at Neolithic Çayönü Tepesi (Southeastern Anatolia, Turkey). Paléorient 27(2): 47–73. Esau, K. 1965. Plant Anatomy. 2nd ed. New York: John Wiley. Evershed, R. P., S. Payne, A. G. Sherratt, M. S. Copley, J. Coolidge, D. Urem-Kotsu, K. Kotsakis, A. E. Özdoğan, O. Nieuwenhuyse, P. M. M. G. Akkermans, D. Bailey, R-R. Andeescu, S. Campbell, S. Farid, I. Hodder, N. Yalman, M. Özba şaran, E. Biçakci, Y. Garfinkel, T. Levy, and M. M. Burton. 2008. Earliest Date for Milk Use in the Near East and Southeastern Europe Linked to Cattle Herding. Nature 455:528–31. Fahn, A. 1990. Plant Anatomy. 4th ed. Oxford: Pergamon Press. Fahn, A., E. Werker, and P. Baas. 1986. Wood Anatomy and Identification of Trees and Shrubs from Israel and Adjacent Regions. Jerusalem: The Israel Academy of Sciences and Humanities. Fazeli, H., R. A. E. Coningham, and C. M. Batt. 2004. Cheshmeh-Ali Revisited: Towards an Absolute Dating of the Late Neolithic and Chalcolithic of Iran’s Tehran Plain. Iran 42:13–23. Fazeli, H., R. A. E. Coningham, R. L. Young, G. K. Gilmore, M. Maghsoudi, and H. Raza. 2007.

Socio-economic Transformations in the Tehran Plain: Final Season of Settlement Survey and Excavations at Tepe Pardis. Iran 45:267–85. Fazeli, H., E. H. Wong, and D. T. Potts. 2005. The Qazvin Plain Revisited: A Reappraisal of the Chronology of Northwestern Central Plateau, Iran, in the 6th to the 4th Millennium BC. Ancient Near Eastern Studies 42:3–82. Feng, Z-D., C. B. An, and H. B. Wang. 2006. Holocene Climatic and Environmental Changes in the Arid and Semi-arid Areas of China: A Review. The Holocene 16:119–30. Fernández, H., S. Hughes, J-D. Vigne, D. Helmer, G. Hodgins, C. Miquel, C. Hänni, G. Luikart, and P. Taberlet. 2006. Divergent mtDNA Lineages of Goats in an Early Neolithic Site, Far from the Initial Domestication Areas. Proceedings of the National Academy of Sciences, USA 103:15,375–79. Fernández, H., P. Taberlet, M. Mashkour, J-D. Vigne, and G. Luikart. 2005. Assessing the Origin and Diffusion of Domestic Goats Using Ancient DNA. In First Steps of Animal Domestication: New Archaeozoological Approaches, ed. J-D. Vigne, J. Peters, and D. Helmer, pp. 50–54. Oxford: Oxbow. Fet, G. N. 1994. Vegetation of Southwest Kopetdagh. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 149–72. Dordrecht, South Holland: Kluwer. Fet, V., and K. I. Atamuradov, eds. 1994. Biogeography and Ecology of Turkmenistan. Dordrecht, South Holland: Kluwer. Flad, R. K., J. Yuan, and S. Li. 2007. Zooarchaeological Evidence for Animal Domestication in China. In Late Quaternary Climate Change and Human Adaptation in Arid China, ed. D. B. Madsen, F-H. Chen, and X. Gao, pp. 167–203. Amsterdam: Elsevier. Flannery, K. V. 1965. The Ecology of Early Food Production in Mesopotamia. Science 147:1,247–56. ———. 1968. Archeological Systems Theory and Early Mesoamerica. In Anthropological Archeology in the Americas, ed. B. J. Meggers, pp. 67–87. Washington, DC: Anthropological Society of Washington. Fontugne, M., and K. Szymczak. 2006. Radiocarbon Dating of Ayakagytma “The Site.” In Exploring the Neolithic of the Kyzyl-Kums: Ayakagytma “The Site” and Other Collections by K. Szymczak and M. Khud­ zhanazarov, pp. 26–27. Warsaw: Institute of Archaeology, Warsaw University. Forsten, A. 2000. A Note on the Equid from Anau, Turkestan, “Equus caballus pumpellii” Duerst. In



bibliography

Archaeozoology of the Near East IVB, ed. M. Mashkour, A. M. Choyke, H. Buitenhuis, and F. Poplin, pp. 153–55. Groningen: Centre for Archaeological Research and Consultancy. Forster, T., and K. Heller. 1994. Loess Deposits from the Tadjik Depression (Central Asia): Magnetic Properties and Paleoclimate. Earth and Planetary Science Letters 128:501–12. Fuller, D. Q. 2006. Agricultural Origins and Frontiers in South Asia: A Working Synthesis. Journal of World Prehistory 20:1–86. ———. 2007. Contrasting Patterns in Crop Domestication and Domestication Rates: Recent Archaeobotanical Insights from the Old World. Annals of Botany 100:903–9. Gale, A. J., and P. G. Hoare. 1991. Quaternary Sediments: Petrographic Methods for the Study of Unlithified Rocks. London: Belhaven Press. Ganyalin, A. F. 1959. Kholm Ilgynly-depe [The Ilgynlydepe mound]. Trudy Instituta Istorii, Arkheologii i Etnografii Akademii nauk Turkmenskoi SSR 15:15–29. ———. 1967. Raskopki v 1959–1961 gg. na Altyn-depe [Excavations in 1959–1961 at Altyn-depe]. Sovietskaya Arkheologiya (4): 207–18. Garazhian, O. 2006. Some Traces of Turkmenistan Prehistoric Cultures in Khorassan Province, Iran. In Second Symposium of Iranian Young Archaeologists, ed. S. Zare, pp. 69–99. Tehran: Iranian Center for Archaeological Research [in Persian]. ———. In press. Report of the First Season of Stratigraphic Sounding at Tapeh Borj, Neyshabour. In Third Symposium of Iranian Young Archaeologists, ed. S. Zare. Tehran: Iranian Center for Archaeological Research [in Persian]. Gardin, J-C., and B. Lyonnet. 1978–79. La prospection archéologique de la Bactriane Orientale (1974– 1978: premiers resultats). Mesopotamia 13–14: 99–154. Garrard, A. 1999. Charting the Emergence of Cereal and Pulse Domestication in South-West Asia. Environmental Archaeology 4:67–86. Gauch, H. G. 1982. Multivariate Analysis in Community Ecology. Cambridge: Cambridge University Press. Ghirshman, R. 1938. Fouilles de Sialk près de Kashan 1933, 1934, 1937. 2 vols. Paris: Librairie Orientaliste Paul Geuthner. Giuffra, E., J. M. H. Kijas, V. Amarger, Ö. Carlborg, J-T. Jeon, and L. Andersson. 2000. The Origin of the Domestic Pig: Independent Domestication and Subsequent Introgression. Genetics 154:1,785–91.

277

Goren, Y. 1999. On Determining Use of Pastoral Cave Sites: A Critical Assessment of Spherulites in Archaeology. Mitekufat Haeven: Journal of the Israel Prehistoric Society 29:123–28. Grace, R. 1989. Interpreting the Function of Stone Tools: The Quantification and Computerisation of Microwear Analysis. British Archaeological Reports International Series 474. Oxford. Greguss, P. 1959. Holzanatomie der Europäischen Laubholzer und Sträucher, Budapest: Akademiai Kiadó. Grigson, C. 1985. Bos indicus and Bos namadicus and the Problem of Autochthonous Domestication in India. In Recent Advances in Indo-Pacific Prehistory, ed. V. N. Misra and P. Bellwood, pp. 425–28. New Delhi: Oxford & IBH. Grosswald, M. G. 1980. Late Weichselian Ice Sheet of Northern Eurasia. Quaternary Research 13:1–32. Groves, C. 1981. Ancestors for the Pigs: Taxonomy and Phylogeny of the Genus Sus. Technical Bulletin No. 3, Dept. of Prehistory, Research School of Pacific Studies, Australian National University. Canberra. ———. 2007. Current Views on Taxonomy and Zoogeography of the Genus Sus. In Pigs and Humans: 10,000 Years of Interaction, ed. U. Albarella, K. Dobney, A. Ervynck, and P. Rowley-Conwy, pp. 14–41. Oxford: Oxford University Press. Gubaev, A., G. Koshelenko, and M. Tosi, eds. 1998. The Archaeological Map of the Murghab Delta: Preliminary Reports 1990–95. Rome: IsIAO (Istituto Italiano per l’Africa e l’Oriente). Guo, J., L-X. Du, Y-H. Ma, W-J. Guan, H-B. Li, Q-J. Zhao, X. Li, and S-Q. Rao. 2005. A Novel Maternal Lineage Revealed in Sheep (Ovis aries). Animal Genetics 36:331–36. Gupta, S. P. 1979. Archaeology of Soviet Central Asia and the Iranian Borderlands. 2 vols. Delhi: B R Publishing Corp. Halstead, P. 1989. Like Rising Damp? An Ecological Approach to the Spread of Farming in South East and Central Europe. In The Beginnings of Agriculture, ed. A. Milles, D. Williams, and N. Gardner, pp. 23–53. British Archaeological Reports International Series 49. Oxford. Hanotte, O., D. G. Bradley, J. W. Ochieng, Y. Verjee, E. W. Hill, and J. E. Rege. 2002. African Pastoralism: Genetic Imprints of Origins and Migrations. Science 296:336–39. Harlan, J. R, J. M. J. de Wet, and E. G. Price. 1973. Comparative Evolution of Cereals. Evolution 27: 311–25.

278

bibliography

Harlan, J. R., and D. Zohary. 1966. Distribution of Wild Wheats and Barley. Science 153:1,074–80. Harris, D. R. 1962. The Distribution and Ancestry of the Domestic Goat. Proceedings of the Linnean Society of London 173:79–91. ———. 1969. Agricultural Systems, Ecosystems and the Origins of Agriculture. In The Domestication and Exploitation of Plants and Animals, ed. P. J. Ucko and G. W. Dimbleby, pp. 3–15. London: Duckworth. ———. 1972. Swidden Systems and Settlement. In Man, Settlement and Urbanism, ed. P. J. Ucko, R. Tringham, and G. W. Dimbleby, pp. 245–62. London: Duckworth. ———. 1989. An Evolutionary Continuum of PeoplePlant Interaction. In Foraging and Farming: The Evolution of Plant Exploitation, ed. D. R. Harris and G. C. Hillman, pp. 11–26. London: Unwin Hyman. ———. 1990a. Settling Down and Breaking Ground: Rethinking the Neolithic Revolution. Amsterdam: Stichting Nederlands Museum voor Anthropologie en Praehistorie, Twaalfde Kroon-Voordracht. ———. 1990b. Vavilov’s Concept of Centres of Origin of Cultivated Plants: Its Genesis and Its Influence on the Study of Agricultural Origins. Biological Journal of the Linnean Society 39:7–16. ———. 1992. Analiz pyltsy i kusochkov uglya [Pollen and charcoal-particle analysis]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 104–6. Ashgabat: Y lym. ———. 1996. Domesticatory Relationships of People, Plants and Animals. In Redefining Nature: Ecology, Culture and Domestication, ed. R. Ellen and K. Fukui, pp. 437–63. Oxford: Berg. ———. 1998. The Spread of Neolithic Agriculture from the Levant to Western Central Asia. In The Origins of Agriculture and Crop Domestication, ed. A. B. Damania, J. Valkoun, G. Willcox, and C. O. Qualset, pp. 65–82. Aleppo, Syria: ICARDA (International Center for Agricultural Research in Dry Areas). ———. 2002a. Development of the Agro-Pastoral Economy of the Fertile Crescent during the PrePottery Neolithic Period. In The Dawn of Farming in the Near East, ed. R. T. J. Cappers and S. Bottema, pp. 67–83. Studies in Early Near Eastern Production, Subsistence and Environment 6. Berlin: Ex Oriente. ———. 2002b. The Expansion Capacity of Early Agri-

cultural Systems: A Comparative Perspective on the Spread of Agriculture. In Examining the Farming/Language Dispersal Hypothesis, ed. P. Bellwood and C. Renfrew, pp. 31–39. Cambridge: McDonald Institute for Archaeological Research. ———. 2003. Climatic Change and the Beginnings of Agriculture: The Case of the Younger Dryas. In Evolution on Planet Earth: The Impact of the Physical Environment, ed. L. J. Rothschild and A. M. Lister, pp. 379–94. London: Academic Press. ———. 2007. Agriculture, Cultivation and Domestication: Exploring the Conceptual Framework of Early Food Production. In Rethinking Agriculture: Archaeological and Ethnoarchaeological Perspectives, ed. T. Denham, J. Iriarte, and L. Vrydaghs, pp. 16–35. Walnut Creek, CA: Left Coast Press. Harris, D. R., and C. Gosden. 1996. The Beginnings of Agriculture in Western Central Asia. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, ed. D. R. Harris, pp. 370–89. London: UCL Press; Washington, DC: Smithsonian Institution Press. Harris, D. R., C. Gosden, and M. P. Charles. 1996. Jeitun: Recent Excavations at an Early Neolithic Site in Southern Turkmenistan. Proceedings of the Prehistoric Society 62:423–42. Harris, D. R., and S. Limbrey. 1992. Sovremennoye prirodnoye okryzhenie Dzheituna [The presentday environmental setting of Jeitun]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 7–13. Ashgabat: Ylym. Harris, D. R., V. M. Masson, Y. E. Berezkin, M. P. Charles, G. F. Korobkova, K. Kurbansakhatov, A. J. Legge, and S. Limbrey. 1993. Investigating Early Agriculture in Central Asia: New Research at Jeitun, Turkmenistan. Antiquity 67:324–38. Hedges, R. E. M., R. A. Housely, C. Bronk Ramsey, and G. J. van Klinken. 1994. Radiocarbon Dates from the Oxford AMS System: Archaeometry Datelist 18. Archaeometry 36:337–74. Helbaek, H. 1969. Plant Collecting, Dry Farming, and Irrigation Agriculture in Prehistoric Deh Luran. In Prehistory and Human Ecology of the Deh Luran Plain: An Early Village Sequence from Khuzistan, Iran, ed. F. Hole, K. V. Flannery, and J. A. Neely, pp. 383–426. University of Michigan Museum of Anthropology Memoir 1. Ann Arbor. ———. 1970. The Plant Husbandry of Hacilar. In Excavations at Hacilar, Vol. 1, ed. J. Mellaart, pp. 189–



bibliography

244. Edinburgh: Edinburgh University Press. Helmer, D., L. Gourichon, H. Monchot, J. Peters, and M. Saña Segui. 2005. Identifying Early Domestic Cattle from Pre-Pottery Neolithic Sites on the Middle Euphrates Using Sexual Dimorphism. In First Steps of Animal Domestication: New Archaeozoological Approaches, ed. J-D. Vigne, J. Peters, and D. Helmer, pp. 86–95. Oxford: Oxbow. Helmer, D., V. Roitel, M. Saña, and G. Willcox. 1998. Interprétations environmentales des données archéozoologiques et archéobotaniques en Syrie du nord de 16 000 BP à 7 000 BP, et les débuts de la domestication des plantes et des animaux. In Natural Space, Inhabited Space in Northern Syria 10th–2nd Millennium B.C., ed. M. Fortin and O. Aurenche, pp. 9–33. Québec-Lyon: Canadian Society for Mesopotamian Studies and Maison de l’Orient Méditerranéen. Helms, S., and A. Betts. 1987. The Desert “Kites” of the Badiyat Esh-Sham and North Arabia. Paléorient 13(1): 41–67. Heptner, V. G., and N. P. Naumov, eds. 1989, 1992, 1998, 2002. Mammals of the Soviet Union. Vol. I (1989), Ungulates. Leiden: Brill. Vol. II, Part 1a (1998), Sirenia and Carnivora. Part 1b (2002), Carnivora. Part 2 (1992), Carnivora. Enfield, NH: Science Publishers. Herrmann, G. 1999. Monuments of Merv: Traditional Buildings of the Karakum. London: Society of Antiquaries of London. Herzschuh, U., P. Tarasov, B. Wünnemann, and K. Hartmann. 2004. Holocene Vegetation and Climate of the Alashan Plateau, NW China, Reconstructed from Pollen Data. Palaeogeography, Palaeoclimatology, Palaeoecology 211:1–17. Hesse, B. 1978. Evidence for Husbandry from the Early Neolithic Sites of Ganj Dareh in Western Iran. Ph.D. dissertation, Columbia University. Ann Arbor: University Microfilms. Heun, M., S. Haldorsen, and K. Vollan. 2008. Reassessing Domestication Events in the Near East: Einkorn and Triticum urartu. Genome 51:444–51. Heun, M., R. Schäfer-Pregl, D. Klawan, R. Castagna, M. Accerbi, B. Borghi, and F. Salamini. 1997. Site of Einkorn Wheat Domestication Identified by DNA Fingerprinting. Science 278:1,312–14. Hiebert, F. T. 1994a. Origins of the Bronze Age Civilization of Central Asia. Peabody Museum of Archaeology and Ethnology, Harvard University, American School of Prehistoric Research Bulle-

279

tin 42. Cambridge, MA. ———. 1994b. Production Evidence for the Origins of the Oxus Civilization. Antiquity 68:372–87. ———. 2002a. The Kopet Dag Sequence of Early Villages in Central Asia. Paléorient 28(2): 25–41. ———. 2002b. Bronze Age Interaction between the Eurasian Steppe and Central Asia. In Ancient Interactions: East and West in Eurasia, ed. K. Boyle, C. Renfrew, and M. Levine, pp. 237–48. Cambridge: McDonald Institute for Archaeological Research. ———. 2003. A Central Asian Village at the Dawn of Civilization. Excavations at Anau, Turkmenistan. University Museum Monograph 116. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Hiendleder, S., B. Kaupe, R. Wassmuth, and A. Janke. 2002. Molecular Analysis of Wild and Domestic Sheep Questions Current Nomenclature and Provides Evidence for Domestication from Two Different Subspecies. Proceedings of the Royal Society London, Biological Sciences 269:893–904. Hiendleder, S., K. Mainz, Y. Plante, and H. Lewalski. 1998. Analysis of Mitochondrial DNA Indicates that Domestic Sheep Are Derived from Two Different Ancestral Maternal Sources: No Evidence for Contributions from Urial and Argali Sheep. Journal of Heredity 89:113–20. Higgs, E. S., and M. R. Jarman. 1969. The Origins of Agriculture: A Reconsideration. Antiquity 43:31– 41. Hillman, G. C. 1981. Reconstructing Crop Husbandry Practices from Charred Remains of Crops. In Farming Practice in British Prehistory, ed. R. Mercer, pp. 123–62. Edinburgh: Edinburgh University Press. ———. 1984a. Interpretation of Archaeological Plant Remains: The Application of Ethnographic Models from Turkey. In Plants and Ancient Man: Studies in Palaeoethnobotany, ed. W. van Zeist and W. A. Casparie, pp. 1–42. Rotterdam: Balkema. ———. 1984b. Traditional Husbandry and Processing of Archaic Cereals in Modern Times. Part 1, The Glume-Wheats. Bulletin of Sumerian Agriculture 1:114–52. ———. 1996. Late Pleistocene Changes in Wild PlantFoods Available to Hunter-Gatherers of the Northern Fertile Crescent: Possible Preludes to Cultivation. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, ed. D. R. Harris,

280

bibliography

pp. 159–203. London: UCL Press; Washington, DC: Smithsonian Institution Press. ———. 2000. The Plant Food Economy of Abu Hureyra 1: The Epipalaeolithic. In Village on the Euphrates: From Foraging to Farming at Abu Hureyra by A. M. T. Moore, G. C. Hillman, and A. J. Legge, pp. 327–99. New York: Oxford University Press. Hillman, G., R. Hedges, A. Moore, S. Colledge, and P. Pettitt. 2001. New Evidence of Lateglacial Cereal Cultivation at Abu Hureyra on the Euphrates. The Holocene 11:383–93. Hodder, I., and C. Cessford. 2004. Daily Practice and Social Memory at Çatalhöyük. American Antiquity 69:17–40. Hole, F. 1996. The Context of Caprine Domestication in the Zagros Region. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, ed. D. R. Harris, pp. 263–81. London: UCL Press; Washington, DC: Smithsonian Institution Press. Hole, F., K. V. Flannery, and J. A. Neely. 1969. Prehistory and Human Ecology of the Deh Luran Plain: An Early Village Sequence from Khuzistan, Iran. University of Michigan, Museum of Anthropology Memoir No. 1. Ann Arbor. Hongo, H., and R. H. Meadow. 2000. Faunal Remains from Prepottery Neolithic Levels at Çayönü, Southeastern Turkey: A Preliminary Report Focusing on Pigs (Sus sp.). In Archaeozoology of the Near East IVA, ed. A. M. Choyke, M. Mashkour, H. Buitenhuis, and F. Poplin, pp. 121–40. Groningen: Centre for Archaeological Research and Consultancy. Hongo, H., R. H. Meadow, B. Öksüz, and G. İlgezdi. 2002. The Process of Ungulate Domestication in Prepottery Neolithic Çayönü, Southeastern Turkey. In Archaeozoology of the Near East V, ed. H. Buitenhuis, A. M. Choyke, M. Mashkour, and A. H. Al-Shiyab, pp. 153–65. Groningen: Centre for Archaeological Research and Consultancy. ———. 2005. Sheep and Goat Remains from Çayönü Tepesi, Southeastern Turkey. In Archaeozoology of the Near East VI, ed. H. Buitenhuis, A. M. Choyke, L. Martin, L. Bartosiewicz, and M. Mashkour, pp. 112–23. Groningen: Centre for Archaeological Research and Consultancy. Honne, B. J., and M. Heun. 2009. On the Domestication Genetics of Self-fertilizing Plants. Vegetation History and Archaeobotany 18:269–72. Hoppé, C. 2001. A Thousand Years of Farming: Agricultural Practices from the Byzantine to Early Ot-

toman Period at Khirbet Faris, the Kerak Plateau, Jordan. Ph.D. thesis, University of Sheffield, UK. Huntington, E. 1915. Civilization and Climate. London: H. Milford. ———. 1919. The Pulse of Asia: A Journey in Central Asia Illustrating the Geographic Basis of History. New York: Houghton Mifflin. Hurst, B. J., and B. Lawn. 1984. University of Pennsylvania Radiocarbon Dates XXII. Radiocarbon 26:212–40. Islamov, Y. I. 1990. Sel’oungour, un nouveau site du paléolithique inférieur en Asie centrale. L’Anthropologie 94:675–88. Jacomet, S. 1987. Prähistorische Getridefunde: Eine Anleitung zur Bestimmung prähistorischer Gersten- and Weizen- Funde. Basel: Botanisches Institut der Universität, Abteilung Pflanzensystematik und Geobotanik. Jansen, T. P., P. Forster, M. A. Levine, H. Oelke, M. Hurles, C. Renfrew, J. Weber, and K. Olek. 2002. Mitochondrial DNA and the Origins of the Domestic Horse. Proceedings of the National Academy of Sciences, USA 99:10,905–10. Jarrige, J-F. 2000. Mehrgarh Neolithic: New Excavations. In South Asian Archaeology 1997, ed. M. Taddei and G. De Marco, pp. 259–83. Rome: Istituto Italiano per l’Africa e ’Oriente. ———. 2007–08. Mehrgarh Neolithic. Proceedings of the International Seminar “First Farmers in Global Perspective” (Lucknow 18–20 January 2006), ed. R. Tewari, K. S. Saraswat, and G. C. Singh. Prāgdhārā: Journal of the U. P. State Archaeology Department 18:135–54. Jarrige, J-F., C. Jarrige, and G. Quivron. 2006. Mehrgarh Neolithic: The Updated Sequence. In South Asian Archaeology 2001. Vol. I, Prehistory, ed. C. Jarrige, and V. Lefèvre, pp. 129–41. Paris: Éditions Recherche sur les Civilisations. Jones, G. 1984. Interpretation of Archaeological Plant Remains: Ethnographic Models from Greece. In Plants and Ancient Man: Studies in Palaeo­ ethnobotany, ed. W. van Zeist and W. A. Casparie, pp. 43–61. Rotterdam: Balkema. ———. 1987. A Statistical Approach to the Archaeological Identification of Crop Processing. Journal of Archaeological Science 14:311–23. ———. 1991. Numerical Analysis in Archaeobotany. In Progress in Old World Palaeoethnobotany, ed. W. van Zeist, K. Wasylikowa, and K-E. Behre, pp. 63–80. Rotterdam: Balkema.



bibliography

Jones, G., S. Valamoti, and M. Charles. 2000. Early Crop Diversity: A “New” Glume Wheat from Northern Greece. Vegetation History and Archaeobotany 9:133–46. Jones, M., and T. Brown. 2000. Agricultural Origins: The Evidence of Modern and Ancient DNA. The Holocene 10:769–76. Jongman, R. H. G., C. J. F. ter Braak, and O. F. R. van Tongeren. 1987. Data Analysis in Community and Landscape Ecology. Wageningen: Centre for Biometry. Kamakhina, G. L. 1994. Kopetdagh-Khorassan Flora: Regional Features of Central Kopetdagh. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 129–48. Dordrecht: Kluwer. Karkanas, P., J-P. Rigaud, J. F. Simek, R. M. Albert, and S. Weiner. 2002. Ash Bones and Guano: A Study of the Minerals and Phytoliths in the Sediments of Grotte XVI, Dordogne, France. Journal of Archaeological Science 29:721–32. Kasparov, A. K. 1992. Kostnye ostatki poseleniya Dzheitun v Yuzhnom Turkmenistane [Faunal remains from the Jeitun settlement, southern Turkmenistan]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 50–76. Ashgabat: Y lym. ———. 1994. Environmental Conditions and Farming Strategy of the Protohistoric Inhabitants of South-Central Asia. Paléorient 20(2): 143–49. ———. 2000. Zoomorphological Statuettes from Eneolithic Layers at Ilgynly-depe and Altyn-depe in South Turkmeniya. In Archaeozoology of the Near East IVB, ed. M. Mashkour, A. M. Choyke, H. Buitenhuis, and F. Poplin, pp. 156–63. Groningen: Centre for Archaeological Research and Consultancy. Khamrakuliev, S. 1979. Novye issledovaniya v grote Dam-Dam Cheshme 2 [New research in the Dam-Dam Cheshme 2 cave]. Izucheniya Akademii nauk Turkmenskoi SSR 5:81–85. Kharin, N. G. 1994. Desertification of the Arid Lands of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 65–76. Dordrecht: Kluwer. Khlopin, I. N. 1963. Eneolit yuzhnykh oblastei Srednei Azii [The Eneolithic of the southern regions of Central Asia]. Arkheologiya SSSR–Svod arkheologicheskikh istochnikov B 3–8 Chast [Part] 1. Moscow: Akademiya nauk.

281

———. 1981. The Early Bronze Age Cemetery of Parkhai II: The First Two Seasons of Excavations: 1977–78. In The Bronze Age Civilization of Central Asia: Recent Discoveries, ed. P. L. Kohl, pp. 3–34. Armonk, NY: M. E. Sharpe. ———. 1989. Mogilnik Parkhai II (nekotorye itogi issledovaniya) [The Parkhai II Burial Ground (Some Research Results)]. Sovietskaya Arkheologiya (3): 113–30. Kijas, J. M. H., and L. Andersson. 2001. A Phylogenetic Study of the Origin of the Domestic Pig Estimated from the Near-complete mtDNA Genome. Journal of Molecular Evolution 52:302–8. Kilian, B., H. Ökan, A. Walther, J. Kohl, T. Dagan, F. Salamini, and W. Martin. 2007. Molecular Diversity at 18 Loci in 321 Wild and 92 Domesticate Lines Reveal No Reduction of Nucleotide Diversity during Triticum monococcum (einkorn) Domestication: Implications for the Origins of Agriculture. Molecular Biology and Evolution 24: 2,657–68. Kim, K-I., J-H. Lee, S-S. Lee, and Y-H. Yang. 2003. Phylogenetic Relationships of Northeast Asian Cattle to Other Cattle Populations Determined Using Mitochondrial DNA D-Loop Sequence Polymorphism. Biochemical Genetics 41:91–98. Kobashi, T., J. P. Sevringhaus, E. J. Brook, J-M. Barnola, and A. M. Grachev. 2007. Precise Timing and Characterization of Abrupt Climate Change 8,200 Years Ago from Air Trapped in Polar Ice. Quaternary Science Reviews 26:1,212–22. Kohl, P. L., ed. 1981. The Bronze Age Civilization of Central Asia: Recent Soviet Discoveries. Armonk, NY: M. E. Sharpe. ———. 1984. Central Asia: Palaeolithic Beginnings to the Iron Age. Éditions Recherche sur les Civilisations, Synthèse 14. Paris. ———. 1992. Central Asia (Western Turkmenistan). In Chronologies in Old World Archaeology, Vol. 2., ed. R. W. Ehrich, pp. 154–62. 3rd ed. Chicago: University of Chicago Press. Kohl, P. L., and D. L. Heskel. 1980. Archaeological Reconnaissances in the Darreh Gaz Plain: A Short Report. Iran 18:160–72. Köhler-Rollefson, I. 1996. The One-Humped Camel in Asia: Origin, Utilization and Mechanisms of Dispersal. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, ed. D. R. Harris, pp. 282–94. London: UCL Press; Washington, DC: Smithsonian Institution Press.

282

bibliography

Köhler-Schneider, M. 2001. Verkohlte Kultur- und Wild­ pflanzen aus Stillfried an der March als Spiegel spätbronzezeitlicher Landwirtschaft im Weinviertel. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. ———. 2003. Contents of a Storage Pit from Late Bronze Age Stillfried, Austria: Another Record of the “New” Glume Wheat. Vegetation History and Archaeobotany 12:105–11. Korobkova, G. F. 1960. Opredelenie funktsii kamennykh i kostyanykh orudii s poselniya Dhzeitun po sledam raboty [The determination of the function of stone and bone implements at the Jeitun site on the basis of working traces]. Trudy YuTAKE 10:110–33. ———. 1969. Orudiya truda i khozyaistvo neoliticheskikh plemen Srednei Azii [The tools and economy of the Neolithic tribes of Central Asia]. Materialy i issledovaniya po arkheologii SSSR, No. 158. Leningrad: Nauka. ———. 1975a. Poselenie neoliticheskikh zemledeltsev: Gievdzhik-depe v Yuzhnom Turkmenistane [A site of Neolithic agriculturalists: Gievdzhikdepe in southern Turkmenistan]. Uspekhi sredneaziatskoi arkheologii (3): 71–73. ———. 1975b. Kultury i lokalnye varianty mezolita i neolita Srednei Azii (po materialam kamennoi industrii) [Cultures and local variants of the Mesolithic and Neolithic in Central Asia (based on lithic industries)]. Sovietskaya Arkheologiya (3): 8–27. ———. 1978. Drevneishie zhatvennye orudiya i ikh proizvoditelonost (v svete eksperimentalno-trasologicheskovo izucheniya) [The most ancient sickles and their productivity (in the light of experimental use-wear study)]. Sovietskaya Arkheologiya (4): 36–52. ———. 1992. Les cycles de l’économie agricole (d’après l’outillage): resultants des analyses expérimentales et tracéologiques. In Préhistoire de l’Agriculture: Nouvelles Approches Expérimentales et Ethnographiques, ed. P. C. Anderson, pp. 271–80. CNRS, Monographie du Centre Nationale de la Recherché Scientifique no. 6. Paris. ———. 1996. Srednyaya Aziya i Kazakhstan [Central Asia and Kazakhstan]. In Neolit Severnoi Evrazii [The Neolithic of Northern Eurasia], ed. B. A. Rybakov, pp. 87–126. Moscow: Nauka. Korobkova, G. F., O. Lollekova, and T. A. Sharovskaya. 1992. Trasologicheskii analiz orudii truda

iz tretiego sloya poseleniya Dzheitun [Use-wear analysis of tools from the third level at the Jeitun settlement]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 34–49. Ashgabat: Ylym. Korobkova, G. F., and V. T. Volovik. 1972. Gadymidepe: novyi pamyatnik Dzheitunskoi kultur [Ga­ dymi-depe: a new site of the Jeitun Culture]. Uspekhi sredneaziatskoi arkheologii 2:41–42. Korobkova , G. F., and K. Y. Yusupov. 1979. Otkrytie novoi neoliticheskoi kultury na Murgabe. [Discovery of a new Neolithic culture on the Murghab river]. Uspekhi sredneaziatskoi arkheologii 4:80. Korshunov, V. M. 1994. Ecology of the Bearded Goat (Capra aegagrus Ersleben 1777) in Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 231–46. Dordrecht: Kluwer. Kreuz, A., and N. Boenke. 2002. The Presence of Two-grained Einkorn at the Time of the Bandkeramik Culture. Vegetation History and Archaeobotany 11:233–40. Kuftin, B. A. 1956. Polevoi otchet o rabote XIV otryada YuTAKE po izucheniyu kultury pervobytnoobshchinykh osedlo-zemledelcheskikh poselenii epokhi medi i bronzy v 1952 g. [Field report: on the work of the XIV brigade of YuTAKE in 1952 concerning the study of primitive sedentary agricultural settlements of the Copper and Bronze ages]. Trudy YuTAKE 2:260–90. Kurbanov, D. 1994. Flora of Kopetdagh. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 105–28. Dordrecht: Kluwer. Kurbansakhatov, K. K. 1987. Eneolit Anau [Eneolithic Anau]. Ashgabat: Y lym. ———. 1992. Raskopki tretiego sloya poseleniya Dzheitun [Excavation of the third level at the Jeitun settlement]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietskoangliiskoi ekspeditsii), ed. V. M. Masson, pp. 30–34. Ashgabat: Ylym. Küster, H. 1985. Neolitische Pflanzenreste aus Hochdorf, Gemeinde Eberdingen (Kreis Ludwigsburg). In Horchdorf I, Forschungen und Berichte zur Vor- und Frühgeschichte in Baden-Württemberg 19, ed. H. Küster and U. Körber-Grohne, pp. 13–83. Stuttgart: Konrad Theiss Verlag. Kutzbach, P. J., P. J. Behling, and R. Selin. 1993. Simulated Climatic Changes: Results of the COHMAP



bibliography

Climate-Model Experiments. In Global Climates since the Last Glacial Maximum, ed. H. E. Wright, Jr., J. E. Kutzbach, T. Webb III, W. F. Ruddiman, F. A. Street-Perrott, and P. J. Bartlein, pp. 24–93. Minneapolis: University of Minnesota Press. Kuznetsova, M. V., M. V. Kholodova, and A. A. Luschek­ ina. 2002. Phylogenetic Analysis of Sequences of the 12S and 16S rRNA Mitochondrial Genes in the Family Bovidae: New Evidence. Russian Journal of Genetics 38:942–50. Lamb, H. F., F. Gasse, A. Benkaddour, N. el-Hamouti, S. van der Kaars, W. T. Perkins, N. J. Pearce, and C. N. Roberts. 1995. Reaction between Centuryscale Holocene Arid Intervals in Tropical and Temperate Zones. Nature 373:134–37. Lamberg-Karlovsky, C. C. 1975. Third Millennium Modes of Exchange and Modes of Production. In Ancient Civilization and Trade, ed. J. A. Sabloff and C. C. Lamberg-Karlovsky, pp. 341–68. Albuquerque: University of New Mexico Press. ———. 1994. Foreword. Initiating an Archaeological Dialogue: The USA-USSR Archaeological Exchange. In Origins of the Bronze Age Civilization of Central Asia by F. T. Hiebert, pp. xvii–xxx. Peabody Museum of Archaeology and Ethnology, Harvard University, American School of Prehistoric Research Bulletin 42. Cambridge, MA. Lamberg-Karlovsky, C. C., and M. Tosi. 1973. Shahri Sokhta and Tepe Yaha: Tracks on the Earliest History of the Iranian Plateau. East and West 23:21–57. Lange, A. G. 1990. De Horden near Wijk bij Duurstede. Plant Remains from a Native Settlement at the Roman Frontier: A Numerical Approach. Amersfoort: ROB (Rijksinstituut voor Oudheidkundig Bodemonderzoek). Larson, G., U. Albarella, K. Dobney, and P. RowleyConwy. 2007a. Current Views on Sus Phylogeography and Pig Domestication as Seen through Modern mtDNA Studies. In Pigs and Humans: 10,000 Years of Interaction, ed. U. Albarella, K. Dobney, A. Ervynck, and P. Rowley-Conwy, pp. 30–41. Oxford: Oxford University Press. Larson, G., U. Albarella, K. Dobney, P. Rowley-Conwy, J. Schibler, A. Tresset, J-D. Vigne, C. J. Edwards, A. Schlumbaum, A. Dinu, A. Bălăçsescu, G. Dolman, A. Tagliacozzo, N. Manaseryan, P. Miracle, L. van Wijngaarden-Bakker, M. Masseti, D. G. Bradley, and A. Cooper. 2007b. Ancient DNA, Pig Domestication, and the Spread of the

283

Neolithic into Europe. Proceedings of the National Academy of Sciences, USA 104:15,276–81. Larson, G., K. Dobney, U. Albarella, M. Fang, E. Matisoo-Smith, J. Robins, S. Lowden, H. Finlayson, T. Brand, E. Willerslev, P. Rowley-Conwy, L. Andersson, and A. Cooper. 2005. Worldwide Phylogeography of Wild Boar Reveals Multiple Centers of Pig Domestication. Science 307:1,618–21. Lasota-Moskalewska, A., J. Piątkowska-Małecka, A. Grę­zak, and K. Szymczak. 2006. Animal Bone Remains from Ayakagytma “The Site.” In Exploring the Neolithic of the Kyzyl-Kums: Ayakagytma “The Site” and Other Collections by K. Szymczak and M. Khudzhanazarov, pp. 206–17. Warsaw: Institute of Archaeology, Warsaw University. Lawrence, B. 1967. Early Domestication of Dogs. Zeitschrift für Säugetierkunde 32(1): 44–59. Lawrence, B., and C. A. Reed. 1983. The Dogs of Jarmo. In Prehistoric Archaeology along the Zagros Flanks, ed. L. S. Braidwood, R. J. Braidwood, B. Howe, C. A. Reed, and P. J. Watson, pp. 485–89. University of Chicago, Oriental Institute Publications Vol. 105. Chicago. Lechevallier, M., R. H. Meadow, and G. Quivron. 1982. Dépôts d’animaux dans les sépultures néolithiques de Mehrgarh, Pakistan. Paléorient 8(1): 99–106. Legge, A. J. 1992. Soderzhanie ovets i koz na Dzheitune [Herding sheep and goats at Jeitun]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 77–83. Ashgabat: Y lym. ———. 1996. The Beginning of Caprine Domestication in Southwest Asia. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, ed. D. R. Harris, pp. 238–62. London: UCL Press; Washington, DC: Smithsonian Institution Press. Legge, A. J., and P. A. Rowley-Conwy. 1987. Gazelle Killing in Stone Age Syria. Scientific American 257(2): 76–83. Levine, M. 1999a. The Origins of Horse Husbandry on the Eurasian Steppe. In Late Prehistoric Exploitation of the Eurasian Steppe by M. Levine, Y. Rassamakin, A. Kislenko, and N. Tatarintseva, pp. 5–58. Cambridge: McDonald Institute for Archaeological Research. ———. 1999b. Botai and the Origins of Horse Domestication. Journal of Anthropological Archaeology 18(1): 29–78.

284

bibliography

———. 2006. mtDNA and Horse Domestication: The Archaeologist’s Cut. In Equids in Time and Space, ed. M. Mashkour, pp. 192–201. Oxford: Oxbow. Li, F. 1989. A New Discovery of Ancient Agricultural Remains in the Neolithic Sites at Donghuishan, Minle County, Gansu. Nongye Kaogu [Archaeology of Agriculture] 1:56–65 [in Chinese]. Li, S. 2002. The Interaction between Northwest China and Central Asia during the Second Millennium BC: An Archaeological Perspective. In Ancient Interactions: East and West in Eurasia, ed. K. Boyle, C. Renfrew, and M. Levine, pp. 171–82. Cambridge: McDonald Institute for Archaeological Research. Li, X., J. Dodson, X. Zhou, H. Zhang, and R. Masutomoto. 2007. Early Cultivated Wheat and Broadening of Agriculture in Neolithic China. The Holocene 17:555–60. Limbrey, S. 1992a. Obsledovanie allyuvialnykh otlo­ zhenii i pochv [Study of alluvial deposits and soils]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietskoangliiskoi ekspeditsii), ed. V. M. Masson, pp. 99– 103. Ashgabat: Y lym. ———. 1992b. Mikromorfologicheskoye izuchenie dvorovykh otlozhenii [Micromorphological study of yard deposits]. In Novye issledovaniya na poselenii Dzheitun (Predvaritelnye soobshcheniya o rabotakh sovietsko-angliiskoi ekspeditsii), ed. V. M. Masson, pp. 94–98. Ashgabat: Y lym. Lindqvist, C., and G. Possnert. 1997. The Subsistence Economy and Diet at Jakobs/Ajvide and Stora Förvar, Eksta Parish and Other Prehistoric Dwellings and Burial Sites on Gotland in Longterm Perspective. In Remote Sensing. Vol. 1, Applied Techniques for the Study of Cultural Resources and the Localization, Identification and Documentation of Sub-surface Prehistoric Remains in Swedish Archaeology, ed. G. Burenhult, pp. 29–90. Stockholm: Department of Archaeology, University of Stockholm. Linnard, W. 1963. The Saiga. Oryx 7:30–33. Lisitsina, G. N. 1965. Orosayemoye zemledeliye epokhi eneolitica na yuge Turkmenii [Irrigation Agriculture of the Eneolithic Period in southern Turkmenistan]. Moscow: Nauka. ———. 1969. The Earliest Irrigation in Turkmenia. Antiquity 43:279–88. ———. 1978. Stanovlenie i razvitie oroshaemogo zemledeliya v yuzhnoi Turkmenii [Evolution and Development of Irrigation Agriculture in southern Turkmenistan].

Moscow: Nauka. Lister, A. M., M. Kadwell, L. M. Kaagen, W. C. Jordan, M. B. Richards, and H. F. Stanley. 1998. Ancient and Modern DNA from a Variety of Sources in a Study of Horse Domestication. Ancient Biomolecules 2:267–80. Liu, C. L., M. L. Wang, P. C. Jiao, S. Li, and Y. Z. Chen. 2003. Holocene Yellow Silt-layers and the Paleoclimate Event of 8200 ka BP in Lop Nur, Xinjang, NW China. Acta Geologica Sinica–English Edition 77:514–18. Liu, X., H. V. Hunt, and M. K. Jones. 2009. River Valleys and Foothills: Changing Archaeological Perceptions of North China’s Earliest Farmers. Antiquity 83:82–95. Liu, X. Q., J. Shen, S. M. Wang, X. D. Yang, G. B. Tong, and E. L. Zhang. 2002. A 16,000-year Pollen Record of Qinghai Lake and Its Paleoclimate and Paleoenvironment. Chinese Science Bulletin 47:1,931–36. Lobova, E. V. [1960] 1967. Pochvy Postynnoi Zony SSSR. Moscow: Akademiya Nauk. Translated by N. Kaner as Soils of the Desert Zone of the USSR. Jerusalem: IPST (Israel Programme of Scientific Translation), 1967. Loftus, R. T., D. E. MacHugh, D. G. Bradley, P. M. Sharp, and P. Cunningham. 1994. Evidence for Two Independent Domestications of Cattle. Proceedings of the National Academy of Science, USA 91:2,757–61. Lollekova, O. 1982. Gadymi-depe—novyi pamyatnik vostochnoi gruppy Dzheitunskikh poselenii [Gadymi-depe—a new site in the eastern group of Jeitun-Culture settlements]. In Novye arkheologicheskie otkrytiya v Turkmenistane, ed. O. Babaev and A. I. Shchetenko, pp. 3–10. Ashgabat: Ylym. Luikart, G., H. Fernández, M. Mashkour, P. R. England, and P. Tarbelet. 2006. Origins and Diffusion of Domestic Goats Inferred from DNA Markers: Example Analyses of mtDNA, Y Chromosome, and Microsatellites. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 294–305. Berkeley: University of California Press. Luikart, G., L. Gielly, L. Excoffier, J-D. Vigne, J. Bouvet, and P. Tarbelet. 2001. Multiple Maternal Origins and Weak Phylogeographic Structure in Domestic Goats. Proceedings of the National Academy of Science, USA 98:5,927–32. Lyubin, V. P. 1984. Paleolit Turkmenii (istoriya issle-



bibliography

dovaniya, novye materialy, blizhaishie zadachi) [The Turkmenian Palaeolithic (history of research, new materials, priority tasks)]. Sovietskaya Arkheologiya (1): 26–45. Machalett, B., M. Frechen, U. Hambach, E. A. Oches, L. Zöller, and S. B. Marković. 2006. The Loess Sequence from Remiskowka (Northern Boundary of the Tien Shan Mountains, Kakakhstan). Part 1: Luminescence Dating. Quaternary International 152/153:202–12. MacHugh, D. E., M. D. Shriver, R. T. Loftus, P. Cunningham, and D. G. Bradley. 1997. Microsatellite DNA Variation and the Evolution, Domestication and Phylogeography of Taurine and Zebu Cattle (Bos taurus and Bos indicus). Genetics 146:1,071–86. MacNeish, R. S. 1967. An Interdisciplinary Approach to an Archaeological Problem. In The Prehistory of the Tehuacan Valley. Vol. 1, Environment and Subsistence, ed. D. S. Byers, pp. 14–24. Austin: University of Texas Press. Magee, D. A., H. Mannen, and D. G. Bradley. 2007. Duality in Bos indicus mtDNA Diversity: Support for Geographical Complexity in Zebu Domestication. In The Evolution and History of Human Populations in South Asia: Inter-disciplinary Studies in Archaeology, Biological Anthropology, Linguistics and Genetics, ed. M. D. Petraglia and B. Allchin, pp. 385–91. Dordrecht: Springer. Mahfroozi, A. 2004. Preliminary Report on Archaeological Surveys and Excavations in Eastern Mazandaran. Gozārešhā-ye bāstān šenāsi [Archaeological Reports] 2:263–78 [in Persian]. Mamedov, A. V. 1997. The Late Pleistocene–Holocene History of the Caspian Sea. Quaternary International 41/42:161–66. Manceau, V., L. Després, J. Bouvet, and P. Tarbelet. 1999. Systematics of the Capra Genus Inferred from Mitochondrial DNA Sequence Data. Molecular Phylogenetics and Evolution 13:504–10. Mangerud, J., M. Jakobsson, H. Alexanderson, V. Astakhov, G. K. C. Clarke, M. Henrikson, C. Hjort, G. Krinner, J-P. Lunkka, P. Möller, A. Murray, O. Nikolskaya, M. Saarnisto, and J. I. Svendsen. 2004. Ice-dammed Lakes and Rerouting of the Drainage of Northern Eurasia during the Last Glaciation. Quaternary Science Reviews 23:1,313–32. Mannen, H., Y. Nagata, and S. Tsuji. 2001. Mitochondrial DNA Reveal that Domestic Goat (Capra hircus) Are Genetically Affected by Two Subspecies

285

of Bezoar (Capra aegagrus). Biochemical Genetics 39:145–54. Marcolongo, B., and P. Mozzi. 1998. Outline of Recent Geological History of the Kopet-Dagh Mountains and the Southern Kara-Kum. In The Archaeological Map of the Murghab Delta: Preliminary Reports 1990–95, ed. A. Gubaev, G. Koshelenko, and M. Tosi, pp. 1–13. Rome: IsIAO (Istituto Italiano per l’Africa e l’Oriente). Markov, G. E. 1961. Raskopki pervobytnoi stoyanki Oyukli [Excavations of the primitive site of Oyukli]. Vestnik Moskovskogo gosudarstvennogo universiteta 9(3): 67–82. ———. 1966a. Grot Dam-Dam Cheshme 2 v vostochnom Prikaspii [Dam-Dam Cheshme cave 2 east of the Caspian]. Sovietskaya Arkheologiya (2): 104–25. ———. 1966b. Raskopki pamyatnika pervobytnoi kul­ tury v zapadnom Turkmenistane: grota Dam-Dam Cheshme 2 [Excavations of a prehistoric site in western Turkmenistan: Dam-Dam Cheshme cave 2]. Nauchnye Soobshcheniya 5:85–91. ———. 1981. Pamyatnik pervobytnosti v vostochnom Prikaspii: grot Dam-Dam Cheshme 1 [A prehistoric site east of the Caspian: Dam-Dam Cheshme cave 1]. Vestnik Moskovskogo gosudarstvennogo universiteta 8(3): 41–55. Markov, G. E., and S. Khamrakuliev. 1980. Novaya stoyanka ojuklinskoi kultury v severnom Pribalkhane [The new site of the Ojukli culture in the northern Pre-Balkhan]. Vestnik Moskovskogo gosudarstvennogo universiteta 4:67–76. Mashkour, M. 2002. Chasse et élevage au nord du plateau central Iranien entre le Néolithique et l’Âge du Fer. Paléorient 28(1): 27–42. ———. 2003. Equids in the Northern Part of the Iranian Central Plateau from the Neolithic to Iron Age: New Zoogeographic Evidence. In Prehistoric Steppe Adaptation and the Horse, ed. M. Levine, C. Renfrew, and K. Boyle, pp. 129–38. Cambridge: McDonald Institute for Archaeological Research. Mashkour, M., M. Fontugne, and C. Hatté. 1999. Investigations on the Evolution of Subsistence Economy in the Qazvin Plain (Iran) from the Neolithic to Iron Age: New Zoogeographic Evidence. In Late Prehistoric Exploitation of the Eurasian Steppe by M. Levine, A. Kislenko, and N. Tatarintseva, pp. 129–38. Cambridge: McDonald Institute for Archaeological Research. Mason, I. L., ed. 1984. Evolution of Domesticated Animals. London: Longman.

286

bibliography

Masson, V. M. 1957. Dzheitun i Kara-depe. Predvaritelnoye soobshchenie o rabotakh 1955g. [Jeitun and Kara-depe. Preliminary report of work carried out in 1955]. Sovietskaya Arkheologiya (1): 143–60. ———. 1961. The First Farmers of Turkmenia. Antiquity 35:203–13. ———. 1964. Srednyaya Aziya i Drevnii Vostok [Central Asia and the Ancient East]. Moscow/Leningrad: Nauka. ———. 1971. Poselenie Dzheitun: Problema Stanovleniya Proisvodyashchei Ekonomiki [The Jeitun Settlement: The Question as to the Emergence of a Productive Economy]. Materialy i issledovaniya po arkheologii SSR, 180. Leningrad: Nauka. ———. 1988. Altyn-depe. Trans. H. N. Michael. University Museum Monograph 55. Philadelphia: University of Pennsylvania Museum. Original in Russian, Altyn-depe. Leningrad: Nauka, 1981. ———. 1996. The Cultural Zones and Variations of the Late Chipped Stone Industries in Central Asia. In Neolithic Chipped Stone Industries of the Fertile Crescent and Their Contemporaries in Adjacent Regions, ed. S. K. Kozlowski and H. G. K. Gebel, pp. 91–103. Studies in Early Near Eastern Production, Subsistence and Environment 3. Berlin: Ex Oriente. Masson, V. M., and D. R. Harris. 1994. New Excavations at Jeitun, Turkmenistan: The First Five Years. In New Archaeological Discoveries in Asiatic Russia and Central Asia, ed. V. M. Masson, pp. 14–16. St. Petersburg: Institute of History of Material Culture, Russian Academy of Sciences. Masson, V. M., and V. I. Sarianidi. 1972. Central Asia: Turkmenia before the Achaemenids. London: Thames and Hudson. Masuda, S. 1972. Excavation at Tappe Sang-e Cagmaq. In Proceedings of the Ist Annual Symposium of Archaeological Research in Iran, 1972, ed. F. Bagherzadeh, pp. 1–5. Tehran: Ministry of Culture and Art. ———. 1974a. Exacavations at Tepe Sang-e Čaxamaq. In Proceedings of the IInd Annual Symposium of Archaeological Research in Iran, ed. F. Bagherzadeh, pp. 23–33. Tehran: Iranian Centre for Archaeological Research. ———. 1974b. Tepe Sang-e Čaxamaq. Iran 12:222–23. ———. 1976. Report of the Archaeological Investigations at Šahrud, 1975. In Proceedings of the IVth Annual Symposium of Archaeological Research in Iran, ed. F. Bagherzadeh, pp. 63–70. Tehran: Iranian

Centre for Archaeological Research. ———. 1977. Tappeh Sang-i Čaxmaq: Investigations of a Neolithic Site in Northeastern Iran. Tokyo: University of Hiroshima [in Japanese]. ———. 1984. The Excavation at Tappeh Sangh-e Caxmaq. Archiv für Orientforschung Internationale Zeitschrift für die Wissenschaft vom Vorderen Orient 31:209–12. McBurney, C. B. M. 1964. Preliminary Report on Stone Age Reconnaissance in North-Eastern Iran. Proceedings of the Prehistoric Society 30:382–99. ———. 1968. The Cave of Ali Tappeh and the Epi-Palaeolithic of N.E. Iran. Proceedings of the Prehistoric Society 34:385–411. Meadow, R. H. 1989a. Prehistoric Wild Sheep and Sheep Domestication on the Eastern Margin of the Middle East. In Early Animal Domestication and Its Cultural Context, ed. P. J. Crabtree, D. Campana, and K. Ryan, pp. 25–36. MASCA Research Papers in Science and Archaeology, Supplement to Vol. 6. Philadelphia: MASCA, University Museum of Archaeology and Anthropology, University of Pennsylvania. ———. 1989b. Osteological Evidence for the Process of Animal Domestication. In The Walking Larder: Patterns of Domestication, Pastoralism and Predation, ed. J. Clutton-Brock, pp. 80–90. London: Unwin Hyman. ———. 1993. Animal Domestication in the Middle East: A Revised View from the Eastern Margin. In Harappan Civilization: A Recent Perspective, ed. G. L. Possehl, pp. 295–320. 2nd rev. ed. New Delhi: Oxford & IBH. Metcalfe, C. R. 1960. Anatomy of the Monocotyledons. Vol. I, The Gramineae. Oxford: Clarendon Press. Meurers-Balke, J., and J. Lüning. 1992. Some Aspects and Experiments Concerning the Processing of Glume Wheats. In Préhistoire de l’Agriculture: Nouvelles Approches Expérimentales et Ethnographiques, ed. P. C. Anderson, pp. 341–62. CNRS, Monographie du Centre Nationale de la Recherché Scientifique 6. Paris. Miller, N. F. 1999. Agricultural Development in Western Central Asia in the Chalcolithic and Bronze Ages. Vegetation History and Archaebotany 8:13–19. ———. 2003. The Use of Plants at Anau North. In A Central Asian Village at the Dawn of Civilization. Excavations at Anau, Turkmenistan by F. T. Hiebert, pp. 127–38. University Museum Monograph 116. Philadelphia: University of Pennsylvania Muse-



bibliography

um of Archaeology and Anthropology. Milner-Gulland, E. J. 2004. The Saiga Antelope— Teetering on the Brink but Still Cause for Hope. Oryx 38(3): 250–51. Milner-Gulland, E. J., M. V. Kholodova, A. Bekenov, O. M. Bukreeva, Iu A. Grachev, L. Amgalan, and A. A. Lushchekina. 2001. Dramatic Declines in Saiga Antelope Populations. Oryx 35(4): 340–45. Moloney, N., S. L. Olsen, and V. Voloshin. 2001. Lower and Middle Palaeolithic Occupation in Central Kazakhstan: The Batpak Valley and Environs. In A Very Remote Period Indeed: Papers on the Palaeolithic Presented to Derek Roe, ed. S. Milliken and J. Cook, pp. 138–43. Oxford: Oxbow Books. Moore, A. M. T., and G. C. Hillman. 1992. The Pleistocene to Holocene Transition and Human Economy in Southwest Asia: The Impact of the Younger Dryas. American Antiquity 57:482–94. Moore, K. M., N. M. Ermolova, and A. Forsten. 2003. Animal Herding, Hunting and the History of Animal Domestication at Anau depe. In A Central Asian Village at the Dawn of Civilization. Excavations at Anau, Turkmenistan by F. T. Hiebert, pp. 154– 59. University Museum Monograph 116. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Moore, K. M., N. F. Miller, F. T. Hiebert, and R. H. Meadow. 1994. Agriculture and Herding in the Early Oasis Settlements of the Oxus Civilization. Antiquity 68:418–27. Morrell, P. L., and M. T. Clegg. 2007. Genetic Evidence for a Second Domestication of Barley (Hordeum vulgare) East of the Fertile Crescent. Proceedings of the National Academy of Sciences, USA 104:3,289–94. Morrell, P. L., K. E. Lundy, and M. T. Clegg. 2003. Distinct Geographic Patterns of Genetic Diversity Are Maintained in Wild Barley (Hordeum vulgare ssp. spontaneum) despite Migration. Proceedings of the National Academy of Sciences, USA 100:10,812–17. Morrell, P. L., D. T. Toleno, K. E. Lundy, and M. T. Clegg. 2005. Low Levels of Linkage Disequilibrium in Wild Barley (Hordeum vulgare ssp. spontaneum) despite High Rates of Self-fertilization. Proceedings of the National Academy of Sciences, USA 102:2,442–47. Movius, H. L., Jr. 1953. Palaeolithic and Mesolithic Sites in Soviet Central Asia. Proceedings of the American Philosophical Society 97:383–421. Naderi, S., H-R. Rezaei, F. Pompanon, M. G. B. Blum,

287

R. Negrini, H-R. Naghash, Ö. Balkiz, M. Mashkour, O. E. Gaggiotti, P. Ajmone-Marsan, A. Kence, J-D. Vigne, and P. Taberlet. 2008. The Goat Domestication Process Inferred from Large-scale Mitochondrial DNA Analysis of Wild and Domestic Individuals. Proceedings of the National Academy of Sciences, USA 105:17,659–64. Naderi, S., H-R. Rezaei, P. Taberlet, S. Zundel, S-A. Rafat, H-R. Naghash, M. A. A. El-Barody, O. Ertugrul, and F. Pompanon. 2007. Large-scale Mitochondrial DNA Analysis of the Domestic Goat Reveals Six Haplogroups with High Diversity. PLoS One 10:1–12. Nadler, C. F., K. V. Korobitsina, R. S. Hoffman, and N. N. Vorontsov. 1973. Cytogenetic Differentiation, Geographic Distribution, and Domestication in Palearctic Sheep (Ovis). Zeitschrift für Säugetierkunde 38:109–25. Neale, D. B., M. A. Saghai-Maroof, R. W. Allard, Q. Zhang, and R. A. Jorgensen. 1988. Choroplast DNA Diversity in Populations of Wild and Cultivated Barley. Genetics 120:1,105–10. Negahban, E. O. 1977. Preliminary Report of the Qazvin Expedition: Excavation of Zaghe, Qabristan, Sagzabad, 1971–1772. Marlik (University of Tehran) 2:26–44. ———. 1979. A Brief Report on the Painted Building at Zaghe (Late 7th–Early 6th Millennium B.C.). Paléorient 5:239–50. Nesbitt, M. 2001. Wheat Evolution: Integrating Archaeological and Biological Evidence. In Wheat Taxonomy: The Legacy of John Percival, ed. P. D. S. Caligari and P. E. Brandham, pp. 37–59. Linnean Society of London Special Issue 3. London. Nesbitt, M., and S. O’Hara. 2000. Irrigation in Central Asia: A Long-term Perspective from Turkmenistan. In The Archaeology of Drylands: Living at the Margin, ed. G. Barker and D. Gilbertson, pp. 102–22. London: Routledge. Nikitin, V. V., and A. M. Geldykhanov. 1988. Opredelitel rastenii Turkmenistana [Identification of the Plants of Turkmenistan]. Leningrad: Nauka. Nishiaki, Y. 2000. Lithic Technology of Neolithic Syria. British Archaeological Reports International Series 840. Oxford: Archaeopress. Nobis, G. 1979. Der älteste Haushund lebte vor 14,000 Jahren. UMSHAU 19:610. Norusis, M. J. 1992. SPSS for Windows: Base System User’s Guide Release 6.0. Chicago: SPSS Inc. Okladnikov, A. P. 1951. Drevneishie arkheologiche­

288

bibliography

skie pamyatniki Krasnovodskogo poluostrova [Earliest archaeological sites of the Krasnovodsk peninsula]. Trudy YuTAKE 2:73–104. ———. 1953. Izuchenie pamyatnikov kamennogo veka v Turkmenii (Po dannym rabot IX otryada YuTAKE v 1952 g.) [A study of Stone Age sites in Turkmenia (using data from the work undertaken by the 9th Brigade of YuTAKE in 1952)]. Izvestiya Akademii nauk Turkmenskoi SSR (2): 28–32. ———. 1956. Peshchera Djebel-pamyatnik drevnei kultury prikaspiiskikh plemen Turkmnenii [The Jebel cave-site of the ancient culture of the Caspian tribes of Turkmenia]. Trudy YuTAKE 7:11–219. Öksüz, B. 2000. Analysis of the Cattle Bones of the Prepottery Settlement of Çayönü, Southeastern Turkey. In Archaeozoology of the Near East IVA, ed. M. Mashkour, A. M. Choyke, H. Buitenhuis, and F. Poplin, pp. 154–63. Groningen: Centre for Archaeological Research and Consultancy. Olsen, S. J. 1985. Origins of the Domestic Dog: The Fossil Record. Tucson: University of Arizona Press. Olsen, S. L. 2000. The Secular and Sacred Roles of Dogs at Botai, North Kazakhstan. In Dogs through Time: An Archaeological Perspective, ed. S. J. Crockford, pp. 71–92. British Archaeological Reports International Series 889. Oxford: Archaeopress. ———. 2003. The Exploitation of Horses at Botai, Kazakhstan. In Prehistoric Steppe Adaptation and the Horse, ed. M. Levine, C. Renfrew, and K. Boyle, pp. 82–103. Cambridge: McDonald Institute for Archaeological Research. ———. 2006a. Early Horse Domestication on the Eurasian Steppe. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 245–69. Berkeley: University of California Press. ———. 2006b. Early Horse Domestication: Weighing the Evidence. In Horses and Humans: The Evolution of Human-Equine Relationships, ed. S. L. Olsen, S. Grant, A. M. Choyke, and L. Bartosiewicz, pp. 81–113. British Archaeological Reports International Series 1560. Oxford: Archaeopress. Orlovsky, N. S. 1994. Climate of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 23–48. Dordrecht: Kluwer. Outram, A. K., N. A. Stear, R. Bendrey, S. Olsen, A. Kasparov, V. Zaibert, N. Thorpe, and R. P. Evershed. 2009. The Earliest Horse Harnessing and

Milking. Science 323:1,332–35. Palmer, C. 1998. The Role of Fodder in the Farming System: A Case Study from Northern Jordan. Environmental Archaeology 1:1–10. Pasternak, R. 1998. Investigations of Botanical Remains from Nevali Çori PPNB, Turkey: A Short Report. In The Origins of Agriculture and Crop Domestication, ed. A. B. Damania, J. Valkoun, G. Willcox, and C. O. Qualset, pp. 170–77. Aleppo: ICARDA (International Center for Agricultural Research in Dry Areas). Payne, S. 1973. Kill-off Patterns in Sheep and Goats: The Mandibles from Asvan Kale. Anatolian Studies 23:281–303. ———. 1987. Reference Codes for the Wear State in the Mandibular Cheek Teeth of Sheep and Goats. Journal of Archaeological Science 14:609–14. Pearsall, D. M. 2000. Paleoethnobotany: A Handbook of Procedures. 2nd ed. San Diego: Academic Press. Pedrosa, S., M. Uzun, J. J. Arranz, B. Gutiérrez-Gill, F. San Primitivo, and Y. Bayón. 2005. Evidence of Three Maternal Lineages in Near Eastern Sheep Supporting Multiple Domestication Events. Proceeding of the Royal Society, London, Biological Sciences 272:2,211–17. Percival, J. 1974. The Wheat Plant. London: Duckworth. Perkins, D., Jr. 1972. The Fauna of the Ap Kupruk Caves: A Brief Note. In Prehistoric Research in Afghanistan (1959–1966) by L. Dupree, p. 73. Transactions of the American Philosophical Society 62(4). Philadelphia. Peteet, D. M., ed. 1993. Global Younger Dryas? Quaternary Science Reviews 12(5): 277–355. Peters, J., D. Helmer, A. von den Driesch, and M. Saña Seguί. 1999. Early Animal Husbandry in the Northern Levant. Paléorient 25(2): 27–48. Peters, J., A. von den Driesch, and D. Helmer. 2005. The Upper Euphrates-Tigris Basin: Cradle of Agro-pastoralism? In The First Steps of Animal Domestication: New Archaeozoological Approaches, ed. J-D. Vigne, J. Peters, and D. Helmer, pp. 96–124. Oxford: Oxbow. Pillans, B., and T. Naish. 2004. Defining the Quaternary. Quaternary Science Reviews 23:2,271–82. Piperno, D. R. 2006. Phytoliths: A Comprehensive Guide for Archaeologists and Paleoecologists. Langham, MD: Altamira Press. Poplin, F. 1979. Origine du mouflon de Corse dans une nouvelle perspective paléontologique: par



bibliography

marronage. Annales de Génétique et de Sélection Animale 11:133–43. Popov, K. P. 1994. Trees, Shrubs, and Semishrubs in the Mountains of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 173–86. Dordrecht: Kluwer. Possehl, G. L. 1999. Indus Age: The Beginnings. Philadelphia: University of Pennsylvania Press. Pourkheirandish M., and T. Komatsuda. 2007. The Importance of Barley Genetics and Domestication in a Global Perspective. Annals of Botany 100:999–1,008. Protopopov, K. V., and S. V. Butomo. 1959. Razvitie tekhniki zhidkikh stsintillyatorov i ee primenenie dlya datirovki po radiouglerodu (C14) [The development of the technique using liquid scintillators and its application for radiocarbon dating (C14)]. Sovietskaya Arkheologiya (2): 7–34. Pumpelly, R., ed. 1905. Explorations in Turkestan with an Account of the Basin of Eastern Persia and Sistan, Expedition of 1903. Carnegie Institution of Washington, Publication No. 26. Washington, DC. ———. 1906. Interdependent Evolution of Oases and Civilizations. Bulletin of the Geological Society of America 17:637–70. ———. 1908. Ancient Anau and the Oasis-World. In Explorations in Turkestan, Expedition of 1904. Prehistoric Civilizations of Anau, Origins, Growth, and Influence of Environment, ed. R. Pumpelly, pp. 1–80. Carnegie Institution of Washington, Publication No. 73, Vol. I. Washington, DC. Puschnigg, G. 2006. Ceramics of the Merv Oasis: Recycling the City. Walnut Creek, CA: Left Coast Press. Ralph, E. K. 1955. University of Pennsylvania Radiocarbon Dates, I. Science 121:149–51. Ranieri, M. 2000. The Development of Architectural Forms of Prehistoric Turkmenistan, a Perspective for the Evaluation of Social Complexity: Geometrical Aspects. In South Asian Archaeology 1997, ed. M. Taddei and G. de Marco, pp. 593–615. Rome: Istituto per l’Africa e l’Oriente. Ranov, V. A. 2001. Loess–paleosoil formation of southern Tajikistan and the loess Palaeolithic. Praehistoria 2:7–27. ———. 2003. De l’ancien et du neuf dans l’étude du Mésolithique en Asie moyenne. Paléorient 29(1): 157–65. Ranov, V. A., and R. S. Davis. 1979. Toward a New Outline of the Soviet Central Asian Paleolithic. Current Anthropology 20:249–70.

289

Ranov, V. A., and A. E. Dodonov. 2003. Small Instruments of the Lower Palaeolithic Site Kuldara and Their Geoarchaeological Meaning. In Lower Palaeolithic Small Tools in Europe and Asia, ed. J. M. Burdukiewicz and A. Ronen, pp. 133–47. British Archaeological Reports International Series 1115. Oxford: Archaeopress. Redding, R. W. 1981. Decision Making in Subsistence Herding of Sheep and Goats in the Middle East. Ph.D. dissertation, Dept. of Anthropology, University of Michigan. Ann Arbor: University Microfilms. ———. 2005. Breaking the Mold: A Consideration of Variation in the Evolution of Animal Domestication. In First Steps in Animal Domestication: New Archaeozoological Approaches, ed. J-D. Vigne, J. Peters, and D. Helmer, pp. 41–48. Oxford: Oxbow. Reimer, P. J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, C. Bertrand, P. G. Blackwell, C. E. Buck, G. Burr, K. B. Cutler, P. E. Damon, R. L. Edwards, R. G. Fairbanks, M. Friedrich, T. P. Guilderson, K. A. Hughen, B. Kromer, F. G. McCormac, S. Manning, C. Bronk Ramsey, R. W. Reimer, S. Remmele, J. R. Southon, M. Stuiver, S. Talamo, F. W. Taylor, J. van der Plicht, and C. E. Weyhenmeyer. 2004. IntCal04 Terrestrial Radiocarbon Age Calibration, 0–26 cal. kyr. BP. Radiocarbon 46:1,029–58. Rezvani, H. 1999. Prehistoric Settlement Patterns and Cultures in Semnan Province, Central Plateau, Iran. In The Iranian World: Essays on Iranian Art and Archaeology Presented to Ezat O. Negahban, ed. A. Alizadeh, Y. Majidzadeh, and S. M. Shahmirzadi, pp. 220–21. Tehran: Iran University Press [in Persian with an English summary]. Ricciardi, R. V. 1980. Archaeological Survey in the Upper Atrek Valley (Khorassan, Iran): Preliminary Report. Mesopotamia 15:51–72. Rivera, D., C. Inocencio, C. Obón, E. Carreño, A. Reales, and F. Alcaraz. 2002. Archaeobotany of Capers (Capparis) (Capparaceae). Vegetation History and Archaeobotany 11:295–313. Roberts, N. 1998. The Holocene: An Environmental History. 2nd ed. Oxford: Blackwell. Rosen, A. M. 1992. Preliminary Identification of Silica Skeletons from Near Eastern Archaeological Sites: An Anatomical Approach. In Phytolith Systematics: Emerging Issues, ed. G. Rapp and S. C. Mulholland, pp. 129–47. New York: Plenum Press. ———. 1993. Phytolith Evidence for Early Cereal Exploitation in the Levant. In Current Research in

290

bibliography

Phy­tolith Analysis: Applications in Archaeology and Paleoecology, ed. D. M. Pearsall and D. R. Piperno, pp. 160–71. MASCA Research Papers in Science and Archaeology, Vol. 10. Philadelphia: MASCA, University Museum of Archaeology and Anthropology, University of Pennsylvania. ———. 1999. Phytolith Analysis in Near Eastern Archaeology. In The Practical Impact of Science on Near Eastern and Aegean Archaeology, ed. S. Pike and S. Gitin, pp. 9–15. Weiner Laboratory Monograph 3. London: Archetype. Rosen, A., and S. Weiner. 1994. Identifying Ancient Irrigation: A New Method Using Opaline Phytoliths from Emmer Wheat. Journal of Archaeological Science 21:125–32. Rosenberg, M., M. Nesbitt, R. W. Redding, and B. L. Peasnall. 1998. Hallan Çemi, Pig Husbandry, and Post-Pleistocene Adaptations along the Taurus– Zagros Arc (Turkey). Paléorient 24(1): 24–41. Roustaei, K. In press. Archaeological Landscape Survey of the Shahroud Area. In Proceedings of the International Seminar on the Archaeology of Iran: North and Northeastern Regions, Tehran, November 2005. Tehran: Iranian Center for Archaeological Research [in Persian]. Russell, N., L. Martin, and H. Buitenhuis. 2005. Cattle Domestication at Çatalhöyük Revisited. Current Anthropology 46:S101–8. Rustamov, A. K., and O. Sopyev. 1994. Vertebrates in the Red Data Book of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 205–30. Dordrecht: Kluwer. Rustamov, I. G. 1994. Vegetation of the Deserts of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 77– 104. Dordrecht: Kluwer. Rychagov, G. I. 1997. Holocene Oscillations of the Caspian Sea, and Forecasts Based on Palaeogeographical Reconstructions. Quaternary International 41/42:167–72. Ryder, M. L. 1984. Sheep. In Evolution of Domestic Animals, ed. I. L. Mason, pp. 63–84. London: Longman. Sablin, M. V., and G. A. Khlopachev. 2002. The Earliest Ice Age Dogs: Evidence from Eliseevichi I. Current Anthropology 43:795–99. Saisho, D., and M. D. Purugganan. 2007. Molecular Phylogeography of Domesticated Barley Traces Expansion of Agriculture in the Old World. Genetics 177:1,765–76.

Salamini, F., M. Heun, A. Brandolini, H. Özkan, and J. Wunder. 2004. Comment on “AFLP data and the origins of domesticated crops.” Genome 47:615–20. Salnikov, V. B. 1994. Formation of the Fish Population in the Artificial Hydrographic Network of Turkmenistan (the Amudarya River Basin). In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 365–87. Dordrecht: Kluwer. Salvatori, S., and M. Tosi. 2006. Shar-i Sokhta Revised Sequence. In South Asian Archaeology 2001. Vol. I, Prehistory, ed. C. Jarrige and V. Lefèvre, pp. 281–92. Paris: Éditions Recherche sur les Civilisations. Saña Seguί, M. 2000. Animal Resource Management and the Process of Animal Domestication at Tell Halula (Euphrates Valley–Syria) from 8800 BP to 7800 BP. In Archaeozoology of the Near East IVA, ed. M. Mashkour, A. M. Choyke, H. Buitenhuis, and F. Poplin, pp. 242–56. Groningen: Centre for Archaeological Research and Consultancy. Sapamuradov, J. 2005. Saiga in Turkmenistan: Past and Present. Saiga News 1:7. Sarianidi, V. 1992. Food-producing and Other Neolithic Communities in Khorasan and Transoxiana: Eastern Iran, Soviet Central Asia and Afghanistan. In History of Civilizations of Central Asia, Vol. 1, ed. A. H. Dani and V. M. Masson, pp. 109–26. Paris: UNESCO. Savolainen, P., Y. Zhang, J. Luo, J. Lundeberg, and T. Leitner. 2002. Genetic Evidence for an East Asian Origin of Domestic Dogs. Science 298:1,610–13. Schmidt, E. F. 1937. Excavations at Tepe Hissar, Damghan. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Schmidt, H. 1908. Archaeological Excavations in Anau and Old Merv. In Explorations in Turkestan, Expedition of 1904, Vol. 1, ed. R. Pumpelly, pp. 81–210. Washington, DC: Carnegie Institution of Washington. Schweingruber, F. H. 1990. Anatomy of European Woods. Stuttgart: Haupt. Shackleton, N. J., Z. An, A. E. Dodonov, J. Gavin, G. J. Kukla, V. A. Ranov, and L. P. Zhou. 1995. Accumulation Rate of Loess in Tadjikistan and China: Relationship with Global Ice Volume Cycles. In Wind Blown Sediments in the Quaternary Record, ed. E. Derbyshire, pp. 1–6. Quaternary Proceedings 4. West Sussex, England: Quaternary Research Association and J. Wiley.



bibliography

Shaffer, J. G. 1978. The Later Prehistoric Periods. In The Archaeology of Afghanistan from Earliest Times to the Timurid Period, ed. F. R. Allchin and N. Hammond, pp. 71–186. London: Academic Press. Shahmirizade, S. M. 1977. Tepe Zaghe: A Sixth Millennium BC Village in the Qazvin Plain of the Central Iranian Plateau. Ph.D. dissertation, Dept. of Anthropology, University of Pennsylvania. Phildelphia. Shao, Q. 1981. The Evolution of Cultivated Barley. Barley Genetics 4:22–25. Shcherbak, N. N. 1994. Zoogeographic Analysis of the Reptiles of Turkmenistan. In Biogeography and Ecology of Turkmenistan, ed. V. Fet and K. I. Atamuradov, pp. 307–28. Dordrecht: Kluwer. Sherratt, A. 2007. Diverse Origins: Regional Contributions to the Genesis of Farming. In The Origins and Spread of Domestic Plants in Southwest Asia and Europe, ed. S. Colledge and J. Conolly, pp. 1–20. Walnut Creek, CA: Left Coast Press. Shevchenko, A. I. 1960. K istorii domashnikh zhivotnykh Yuzhnogo Turkmenistana [Towards a history of domestic animals in southern Turkmenistan]. Trudy YuTAKE 10:464–77. Shidrang, S. 2004. Preliminary Report on the Hand Industries of Komishan Cave, Collected from Deposits in an Illegally Excavated Pit. Gozārešhā-ye bāstān šenāsi [Archaeological Reports] 2:279–84 [in Persian]. Smilauer, P. 1992. CANODRAW 3.0 User’s Guide. New York: Microcomputer Power. Stein, J. K., and P. A. Teltser. 1989. Size Distributions in Artefact Classes: Combining Macro- and Microfractions. Geoarchaeology 1:1–30. Stronach, D. 1972. Yarim Tepe. In Excavations in Iran: The British Contribution, ed. P. R. S. Moorey, pp. 21–23. Oxford: Organizing Committee of the Sixth International Congress of Iranian Art and Archaeology. Stuiver, M., and H. A. Polach. 1977. Reporting of 14C data. Radiocarbon 19:355–63. Stuiver, M., and P. J. Reimer. 1986. A Computer Program for Radiocarbon Age Calculation. Radiocarbon 28:1,022–30. ———. 1993. Extended 14C Data Base and Revised CALIB 3.0 14C Calibration Program. Radiocarbon 35:215–30. Stuiver, M., P. J. Reimer, E. Bard, J. W. Beck, G. S. Burr, K. A. Hughen, B. Kromer, F. G. McCormac, J. van der Plicht, and M. Spurk. 1998. INTCAL98 Radiocarbon Age Calibration 24,000–0 cal. BP.

291

Radiocarbon 40:1,041–84. Szymczak, K., and M. Khudzhanazarov. 2006a. The Stone Age of the Turanian Lowland: A Problem of Settlement Discontinuity. In Suyanggae and Her Neighbours, ed. L. Domañska and Y. Lee, pp. 123–37. Łódź: Institute of Archaeology, University of Łódź, and Institute of Korean Prehistory, Chungbuk University. ———. 2006b. Exploring the Neolithic of the Kyzyl-Kums: Ayakagytma “The Site” and Other Collections. Warsaw: Institute of Archaeology, Warsaw University. Takada, T., Y. Kikkawa, H. Yonekawa, S. Kawakami, and T. Amano. 1997. Bezoar (Capra aegagrus) Is a Matriarchal Candidate for Ancestor of the Domestic Goat (Capra hircus): Evidence from the Mitochondrial DNA Diversity. Biochemical Genetics 35:315–26. Tapio, M., N. Marzanov, M. Ozerov, M. Ćinkulov, G. Gonzarenko, T. Kiselyova, M. Murawski, H. Viinalass, and J. Kantanen. 2006. Sheep Mitochondrial DNA Variation in European, Caucasian, and Central Asian Areas. Molecular Biology and Evolution 23:1776–83. Tchernov, E., and F. F. Valla. 1997. Two New Dogs, and Other Natufian Dogs, from the Southern Levant. Journal of Archaeological Science 24:65–95. ter Braak, C. F. J., and P. Smilauer. 1997–99. CANOCO for Windows Version 4.02. Wageningen: Centre for Biometry. Tixier, J. 1963. Typologie de l’Epipaléolithique du Maghreb. Paris: Arts et Metier Graphiques. Tolstov, S. P. 1946. The Early Culture of Khwarizm. Antiquity 20:92–99. ———. 1958. The Prehistoric Cultures and Primitive Irrigation Systems of Ancient Chorasmia. In Thirteenth Annual Report and Bulletin for 1955–56 [University of London Institute of Archaeology]: 8–36. Tolstov, S. P., and A. S. Kes. 1960. Nizovya Amudari, Sarykamysh, Uzboi: istoriya formirovaniya i zaseleniya [The Lower Reaches of the Amudarya, Sarykamysh, Uzboi: history of their formation and settlement]. Materialy Khorezmskoi Ekspeditsii, Vol. 3. Moscow: Nauka. Tosi, M. 1973–74. The Northeastern Frontier of the Ancient Near East. Mesopotamia 8–9:21–76. Townsend, C. C., and E. Guest. 1966–85. Flora of Iraq. 9 vols. Baghdad: Ministry of Agriculture and Agrarian Reform. Trinkaus, K. M. 1981. The Partho-Sassanian Northeast Frontier: Settlement in the Damghan Plain,

292

bibliography

Iran. Ph.D. dissertation, Dept. of Anthropology, University of Pennsylvania. Philadelphia. Troy, C. S., D. E. MacHugh, J. F. Bailey, D. A. Magee, R. T. Loftus, P. Cunningham, A. T. Chamberlain, B. C. Sykes, and D. G. Bradley. 2001. Genetic Evidence for Near-Eastern Origins of European Cattle. Nature 410:1,088–91. Tsalkin, V. I. 1956. Predvaritelnye rezultaty izucheniya faunisticheskogo materiala iz raskopok Dzhebela, proizvedennykh A. P. Okladnikovym [Preliminary results from the study of faunal remains found in the excavations at Djebel undertaken by A. P. Okladnikov]. Trudy YuTAKE 7:220–21. Twiss, P. C. 1992. Predicted World Distribution of C3 and C4 Grass Phytoliths. In Phytolith Systematics: Emerging Issues, ed. G. Rapp and S. C. Mulholland, pp. 113–28. New York: Plenum Press. Uerpmann, H-P. 1987. The Ancient Distribution of Ungulate Mammals in the Middle East. Beihefte zum Tübinger Atlas des Vorderen Orients, Reihe A. (Naturwissenschaften) 27. Wiesbaden: Dr. L. Reichert. Uerpmann, H-P., and W. Frey. 1981. Die Umbegung von Ġār-e Kamarband (Belt Cave) und Ġār-e ‘Al ī Tappe (Beh-Šahr, Mā zandarān, N-Iran) heute und im Spätpleistozän. In Beiträge zur Umweltgeschichte des Vorderen Orients [Contributions to the environmental history of Southwest Asia], ed. W. Frey and H-P. Uerpmann, pp. 134–90. Beihefte zum Tübinger Atlas des Vorderen Orients, Reihe A. (Naturwissenschaften) 8. Wiesbaden: Ludwig Reichart Verlag. Uerpmann, H-P., and M. Uerpmann. 2002. The Appearance of the Domestic Camel in South-East Arabia. Journal of Oman Studies 12:235–60. Underhill, A. P. 1997. Current Issues in Chinese Neolithic Archaeology. Journal of World Prehistory 11:103–60. Valamoti, S., and M. Charles. 2005. Distinguishing Food from Fodder through the Study of Charred Plant Remains: An Experimental Approach to Dung-derived Chaff. Vegetation History and Archaeobotany 14:528–33. Valdez, R., C. F. Nadler, and T. D. Bunch. 1978. Evolution of Wild Sheep in Iran. Evolution 32:56–72. Valkoun, J. 2001. The Geographical Distribution of Wild Wheats in Their Historical Setting and Current Context. In Wheat Taxonomy: The Legacy of John Percival, ed. P. D. S. Caligari and P. E. Brandham, pp. 81–94. Linnean Society of London Spe-

cial Issue 3. London. Valkoun, J., J. G. Waines, and J. Konopka. 1998. Current Geographical Distribution and Habitat of Wild Wheats and Barley. In The Origins of Agriculture and Crop Domestication, ed. A. B. Damania, J. Valkoun, G. Willcox, and C. O. Qualset, pp. 293–99. Aleppo: ICARDA (International Center for Agricultural research in Dry Areas). van Berg, P-L., M. van der Linden, S. Lemaitre, N. Cauwe, and V. Picalause. 2004. Desert-Kites of the Hemma Plateau (Hassake, Syria). Paléorient 30(1): 89–99. van der Veen, M., and N. Fieller. 1982. Sampling Seeds. Journal of Archaeological Science 9:287–98. van Zeist, W. 1999. Evidence for Agricultural Change in the Balikh Basin, Northern Syria. In The Prehistory of Food: Appetites for Change, ed. C. Gosden and J. Hather, pp. 350–73. London: Routledge. Vavilov, N. I. 1992a [1926]. Centers of Origin of Cultivated Plants. In Origin and Geography of Cultivated Plants, ed. V. F. Dorofeyev, pp. 22–143. Cambridge: Cambridge University Press. Original Russian version published in 1926. ———. 1992b [1935]. The Phyto-geographical Basis for Plant Breeding. In Origin and Geography of Cultivated Plants, ed. V. F. Dorofeyev, pp. 316–66. Cambridge: Cambridge University Press. Russian original published in Leningrad in 1935. ———. 1992c [1940]. The Theory of the Origin of Cultivated Plants after Darwin. In Origin and Geography of Cultivated Plants, ed. V. F. Dorofeyev, pp. 421–42. Cambridge: Cambridge University Press. Russian original published in Leningrad in 1940. Vavilov, N. I., and L. E. Rodin. 1997. Five Continents. Rome: International Plant Genetic Resources Institute. (English version of a book originally planned and partly completed by Vavilov, published on the 110th anniversary of his birth.) Velichko, A., and E. I. Kurenkova. 1990. Environmental Conditions and Human Occupation of Northern Eurasia during the Late Valdai. In The World at 18,000 BP. Vol. 1, High Latitudes, ed. O. Soffer and C. Gamble, pp. 255–65. London: Unwin Hyman. Vilà, C., J. A. Leonard, and A. Beja-Pereira. 2006. Gen­ etic Documentation of Horse and Donkey Domestication. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 342–53. Berkeley: University of California Press.



bibliography

Vilà, C., J. A. Leonard, A. Götherström, S. Marklund, K. Sandberg, K. Lidén, R. K. Wayne, and H. Ellergren. 2001. Widespread Origins of Domestic Horse Lineages. Science 291:474–77. Vilà, C., J. E. Maldonado, and R. K. Wayne. 1999. Phylogenetic Relationships, Evolution and Genetic Diversity of the Domestic Dog. Journal of Heredity 90:71–77. Vilà, C., P. Savolainen, J. E. Maldonado, I. R. Amorim, J. E. Rice, R. L. Honeycutt, K. A. Crandall, J. Lundeberg, and R. K. Wayne. 1997. Multiple and Ancient Origins of the Domestic Dog. Science 276:1,687–89. Vinogradov, A. V. 1968. Neoliticheskie Pamyatniki Khorezma [Neolithic sites of Khorezm]. Materialy Khorezmskoi Ekspeditsii, Vol. 8, Moscow: Nauka. ———. 1979. Issledovaniya pamyatnikov kamennovo veka v severnom Afganistane’ [Studies of sites of the Stone Age in northern Afghanistan]. Drevnyaya Baktriya 2:7–62. ———. 1981. Drevnei Okhotniki i Rybolovy Sredneaziatskogo Mezhdurechiya [Ancient Hunters and Fishermen of the Central Asian ‘Mesopotamia’]. Trudy Khorezmskoi Arkheologo–Etnograficheskoi Ekspeditsii, Vol. 13, Moscow: Nauka. Vinogradov, A. V., and E. D. Mamedov. 1975. Pervobytnyi Lyavlyakan [Primitive Lyavlyakan]. Materialy Khorezmskoi Ekspeditsii, Vol. 10. Moscow: Nauka. Vinogradov, A. V., and V. A. Ranov. 1985. Les cultures de la fin de l’Age de la Pierre entre l’Hindukush et le Pamir-Alaï. In L’Archéologie de la Bactriane ancienne: Actes du Colloque franco-soviétique, Dushanbe (U.R.S.S.), 27 octobre–3 novembre 1982, pp. 63–75. Paris: CRNS. Vishnyatsky, L. B. 1989. The Discovery of Palaeolithic Handaxes in Western Turkmenia: A Preliminary Report. Paléorient 15(2): 95–98. ———. 1996. Paleolit Srednei Azii i Kazakhstana [The Palaeolithic of Central Asia and Kazakhstan]. Russian Academy of Sciences, Institute for the History of Material Culture. St. Petersburg: Evropeiskii Dom. ———. 1999. The Paleolithic of Central Asia. Journal of World Prehistory 13:69–122. von Bothmer, R., N. Jacobsen, C. Baden, R. B. Jørgensen, and I. Linde-Laursen. 1995. An Ecogeographical Study of the Genus Hordeum. 2nd ed. Rome: International Plant Genetics Resources Institute. von den Driesch, A. 1976. A Guide to the Measurement of Animal Bones from Archaeological Sites. Peabody

293

Museum Bulletin 1. Cambridge, MA: Peabody Museum of Archaeology and Ethnology, Harvard University. Waines, J. G., and D. Barnhart. 1992. Biosystematic Research in Aegilops and Triticum. Hereditas 116:207–12. Walden, J. 1999. Sample Collection and Preparation. In Environmental Magnetism: A Practical Guide, ed. J. Walden, F. Oldfield, and J. Smith, pp. 26–34. Quaternary Research Association Technical Guide 6. London. Walter, H., and E. O. Box. 1983. The Karakum Desert, an Example of a Well-studied Eu-biome. In Temperate Deserts and Semi-Deserts, ed. N. E. West, pp. 105–59. Ecosystems of the World, Vol. 5. Amsterdam: Elsevier. Ward, G. K., and S. R. Wilson. 1978. Procedures for Comparing and Combining Radiocarbon Age Determinations: A Critique. Archaeometry 20(1): 19–31. Wayne, R. K, J. A. Leonard, and C. Vilà. 2006. Genetic Analysis of Dog Domestication. In Documenting Domestication: New Genetic and Archaeological Paradigms, ed. M. A. Zeder, D. G. Bradley, E. Emshwiller, and B. D. Smith, pp. 279–93. Berkeley: University of California Press. White, K. H., and J. Walden. 1994. Mineral Magnetic Analysis of Iron Oxides in Arid Zone Soils, Tunisian Southern Atlas. In Environmental Change in Drylands: Biogeographical and Geomorphological Perspectives, ed. A. C. Millington and K. Pye, pp. 44–65. Chichester: John Wiley. Willcox, G. 2003. Chalcolithic Carbonised Cereals from Ubaid Burnt Storage Structures at Kosak Shamali. In Tell Kosak Shamali, Vol. II, ed. Y. Nishiaki and T. Matsutani, pp. 267–70. Tokyo: University of Tokyo Museum. ———. 2004. Measuring Grain Size and Identifying Near Eastern Cereal Domestication: Evidence from the Euphrates Valley. Journal of Archaeological Science 31:145–50. ———. 2005. The Distribution, Natural Habitats and Availability of Wild Cereals in Relation to Their Domestication in the Near East: Multiple Events, Multiple Centres. Vegetation History and Archaeobotany 14:534–41. ———. 2007. The Adoption of Farming and the Beginnings of the Neolithic in the Euphrates Valley: Cereal Exploitation between the 12th and 8th Millennia cal BC. In The Origins and Spread

294

bibliography

of Domestic Plants in Southwest Asia and Europe, ed. S. Colledge and J. Conolly, pp. 21–36. Walnut Creek, CA: Left Coast Press. Willcox, G., S. Fornite, and L. Herveux. 2008. Early Holocene Cultivation before Domestication in Northern Syria. Vegetation History and Archaeobotany 17:313–25. Williams, T. 2007. The City of Sultan Kala, Merv, Turkmenistan: Communities, Neighbourhoods and Urban Planning from the Eighth to the Thirteenth Century. In Cities in the Pre-Modern Islamic World: The Urban Impact of Religion, State and Society, ed. A. K. Bennison and A. Gascoigne, pp. 42–62. London: Routledge. ———. 2008. The Landscapes of Islamic Merv, Turkmenistan: Where Do We Draw the Line? Internet Archaeology 25: http://intarch.ac.uk/journal/issue25/merv_index.html. Williams, T., K. K. Kurbansakhatov, M. Ziebart, D. Gilbert, J. Hill, D. Hopkinson, F. Vardy, G. Puschnigg, P. Brun, A. Annaev, A. Upson, C. Cavanagh, K. May, G. Watkins, B. Alvey, A. Kurbanov, J. Keily, L. Cooke, and G. Palumbo 2002. The Ancient Merv Project, Turkmenistan: Preliminary Report on the First Season (2001). Iran 40:15–41. Williams, T., K. K. Kurbansakhatov, V. Zavyalov, G. Puschnigg, J. Keily, D. Hopkinson, K. May, F. Vardy, C. Cavanagh, P. Brun, A. Annaev, A. Feuerbach, D. Gilbert, A. Vince, S. Moriset, M. Bendakir, and L. Cooke. 2003. The Ancient Merv Project, Turkmenistan: Preliminary Report on the Second Season (2002). Iran 41:139–70. Wilson, D. E., and D. M. Reeder, eds. 2005. Mammal Species of the World: A Taxonomic and Geographic Reference. 3rd ed. 2 vols. Baltimore: Johns Hopkins Press. Witcombe, J. R. 1978. Two Rowed and Six Rowed Wild Barley from the Western Himalaya. Euphytica 24:431–34. Wright, H. E., Jr. 1977. Environmental Change and the Origin of Agriculture in the Old and New Worlds. In Origins of Agriculture, ed. C. A. Reed, pp. 281–318. The Hague: Mouton. Wulsin, F. R. 1932. Excavations at Tureng Tepe, near Asterabad. Supplement to American Institute for Persian Art and Archaeology Bulletin No. 2(1). New York. Wünnemann, B., K. Hartmann, M. Janssen, and H. C. Zhang. 2007. Responses of Chinese Desert Lakes

to Climate Instability during the Past 45,000 Years. In Late Quaternary Climate Change and Human Adaptation in Arid China, ed. D. B. Madsen, F-H. Chen, and X. Gao, pp. 11–24. Amsterdam: Elsevier. Yongi, W., and L. Houyuan. 1992. The Study of Phytoliths and Its Application. Beijing: China Ocean Press [in Chinese]. Yuan, J., and R. Flad. 2002. Pig Domestication in Ancient China. Antiquity 76:724–32. ———. 2006. Research on Early Horse Domestication in China. In Equids in Time and Space, ed. M. Mashkour, pp. 124–31. Oxford: Oxbow. Zabakov, V. A. 1986. Global’nye sobytiya pleistyotsena [The main events of the Pleistocene]. Leningrad: Gidrometeoizat. Zeder, M. A. 2001. A Metrical Analysis of a Collection of Modern Goats (Capra hircus aegagrus and C. h. hircus) from Iran and Iraq: Implications for the Study of Caprine Domestication. Journal of Archaeological Science 28:61–79. Zeder, M. A., D. G. Bradley, E. Emshwiller, and B. D. Smith, eds. 2006. Documenting Domestication: New Genetic and Archaeological Paradigms. Berkeley: University of California Press. Zeder, M. A., and B. Hesse. 2000. The Initial Domestication of Goats (Capra hircus) in the Zagros Mountains 10,000 Years Ago. Science 287:2,254–57. Zeuner, F. E. 1963. A History of Domesticated Animals. London: Hutchinson. Zhou, L. P., A. E. Dodonov, and N. J. Shackleton. 1995. Thermoluminescence Dating of the Orkutsay Loess Section in Tashkent Region, Uzbekistan, Central Asia. Quaternary Science Reviews 14:721–30. Zohary, D. 1989. Domestication of the Southwest Asian Neolithic Crop Assemblage of Cereals, Pulses and Flax: The Evidence from the Living Plants. In Foraging and Farming: The Evolution of Plant Exploitation, ed. D. R. Harris and G. C. Hillman, pp. 355–75. London: Unwin Hyman. ———. 1999. Monophyletic vs. Polyphyletic Origin of the Crops on which Agriculture Was Founded in the Near East. Genetic Resources and Crop Evolution 46:133–42. Zohary, D., and M. Hopf. 2000. Domestication of Plants in the Old World: The Origin and Spread of Cultivated Plants in West Asia, Europe and the Nile Valley. 3rd ed. Oxford: Oxford University Press. Zvelebil, M. 1986. Mesolithic Societies and the Tran-



bibliography

sition to Farming: Problems of Time, Scale and Organisation. In Hunters in Transition: Mesolithic Societies of Temperate Eurasia and Their Transition to Farming, ed. M. Zvelebil, pp. 167–88. Cambridge: Cambridge University Press. Zvelebil, M. 2002. Demography and Dispersal of Early

295

Farming Populations at the Mesolithic–Neolithic Transition: Linguistic and Genetic Implications. In Examining the Farming/Language Dispersal Hypothesis, ed. P. Bellwood and C. Renfrew, pp. 379–94. Cambridge: McDonald Institute for Archaeological Research.

Author Note D

avid R. Harris is Emeritus Professor of Human Environment at the Institute of Archaeology, University College  London (of which he was formerly Director), and a Fellow of the British Academy. His research has focused on plant and animal domestication and the emergence and spread of agriculture, for which he has undertaken fieldwork in Europe, Asia, Australia, and the Americas. His many publications include The Origins and Spread of Agriculture and Pastoralism in Eurasia, 1996 (Washington DC: Smithsonian Institution Press and London: UCL Press) and (with Gordon Hillman) Foraging and Farming: the Evolution of Plant Exploitation, 1989 (London: Unwin Hyman).

Index Note: The following are not included in the Index: names of authors and other individuals mentioned; most place names other than archaeological sites (which are indicated in bold); frequently used names of archaeological and geological periods; and the contents of the Appendices. Animals and plants are indexed by their common English names.

Achaemenian civilization/period 26, 50 Adjikuli, Turkmenistan 36, 116 agro-pastoralism 26, 52, 107, 218, 227, 230–33, 234–36 Akchadarya, ancient delta 22, 45, 59, 64, 66, 228 Ak-depe, Turkmenistan 45 Akhal region 60–61, 68, 218 alder (Alnus sp.) 32, 66, 167–69, 192 Ali Kosh, Iran 85, 162, 236 Ali Tappeh, Iran 49, 55–57, 63, 78–79, 81, 87, 90 Allerød interstadial 24, 55 almond (Amygdalus sp.) 9–10, 33, 74, 216, 225 Altai mountains 54, 87, 89 Altyn-depe, Turkmenistan 14, 45, 46, 79, 81, 83, 89, 213 Amorgos, Greece 151, 154, 156–57 Anatolia 67, 77, 78, 86, 170, 180, 226 Anau, Turkmenistan 28, 44–47, 49, 62, 68–69, 73, 78–81, 84, 87, 89, 194, 227, 236 Anau IA phase 60, 68–69, 218 ancient DNA 73, 79, 86, 226 apple (Malus spp.) 9–10, 32, 216, 225 apricot (Armeniaca vulgaris) 9 Aq Kupruk sites, Afghanistan 51, 55, 58–59, 82, 84, 87, 224, 232, 234 Aq Tepe, Iran 50 Aral Sea basin 7, 19, 222 fluctuations 22 Arkash region 60–61, 217, 218 Armadlu Tepe, Iran 50 ash, Syrian (Fraxinus syriaca) 9, 32, 166 ass/donkey (Equus asinus) 18, 68, 82 Auchin-depe, Turkmenistan 46 aurochs (Bos primigenius) 54–55, 66–67, 79–80, 91, 227. See also cattle, wild Ayakagytma, Uzbekistan 22, 48–49, 66–68, 78–79, 81, 84, 88–89, 91, 221 animal domestication 67–68, 79, 81, 91, 221

Bacha Well, Turkmenistan 61, 218 badgers 11, 17, 34, 66 Badghyz plateau 7, 10, 31, 33, 166, 213 Natural Reserve 10, 11, 14, 16, 74, 177, 213 Bami, Turkmenistan 46, 60–63, 162, 186, 217–19 barley (Hordeum spp.) 74–76, 162, 188 cultivation 59, 62, 119, 215–16, 225, 231 domestic (H. vulgare) 62, 68–69, 73–74, 153–54, 192, 198, 215, 232, 236 domestication 75–76, 91, 226, 231–32; genetics 74–75 wild (H. spontaneum) 10, 33, 74–75, 91, 163, 231 Bashkovdan, Turkmenistan 36, 116, 186, 207 beads 58, 61, 97, 179, 192, 212, 214, 217 bear, brown (Ursus arctos) 11, 15, 33, 34 cave 54 Belt Cave, Iran. See Ghar-i Kamarband Beshbulak, Uzbekistan 65 birds, domestic 18 remains 58, 66–67, 176, 179, 193, 201–2, 214 waterfowl 18, 66, 228 bison 67 Bølling interstadial 24, 55 Bolshoi Balkhan massif 5, 9–10, 12, 35–40, 57, 64, 84, 95, 113, 188, 234 climate 39 mammals 40 piedmont 35–37, 40, 57, 113, 234 sites 35, 113–17, 186, 197–207, 228, 234 vegetation 37, 39–40 Borj Tepe, Iran 51 Botai, Kazakhstan 82, 89–90 buckthorn (Rhamnus coriacea) 199 burial, dog 89–90, 177 goat 86 human 58, 86, 89–90, 195, 217 buried soil. See palaeosol

300 Burzahom, Kashmir 89 camel 66–67, 80–83, 229 Bactrian (Camelus bactrianus) 18, 81–83, 91, 227 domestic 62, 83, 227 domestication 67, 81–83, 91, 221 dromedary (Camelus dromedarius) 18, 82–83, 91 wild 80–83, 91 caper (Capparis sp.) 153, 164, 193, 214, 216 caracal/sand lynx (Caracal caracal) 16, 33, 58 Caspian Mesolithic 35, 45, 49, 55–58, 63, 90 Caspian Sea, regressions 20–22, 24, 38, 230 transgressions 20–22, 24, 35, 38–39, 230 cat 178–179 jungle/swamp (Felis chaus) 17, 33 sand (Felis margarita) 16, 17, 34, 58, 213 steppe/manul (Felis manul) 16, 17, 33–34, 40, 58, 98, 176, 213 wild/yellow (Felis silvestris) 16–17, 34, 40, 176, 177, 193, 213 Çatalhöyük, Turkey 67, 123, 170, 180, 182, 184, 194–95 cattle/oxen, domestic 18, 56, 62, 67–69, 79–80, 201–2, 219, 226–27, 232–33, 236 domestication 67–68, 79–80, 91, 221, 227, 232 genetics 80 taurine 79–80 wild 54, 58, 69, 79–80, 227. See also aurochs zebu/indicine 79–80 Caucasus mountains 76–77, 88 cereals 62, 73–74, 95, 232–33. See also barley, wheat cultivation 10, 24, 33, 44, 58, 60, 67–69, 108, 158, 229, 231–36 wild-grain collection/harvesting 33, 58, 67, 216, 224, 229 Chagylly, Turkmenistan 46, 60–62, 79, 81, 89, 162, 186, 218– 19, 221, 236 Chakmakli, Turkmenistan 46, 60–61, 68–69, 186, 218–19 Chandyr rockshelter, Turkmenistan 109, 111–12 charcoal analysis 166–70, 171–73, 199–200 Charla’uk, Turkmenistan 36, 116–17 Chashmanigar, Tajikistan 23, 53 cheetah (Acinonyx jubatus) 11, 16, 33 cherry (Cerasus spp.) 9–10, 32–33, 39, 216 chert 181–83, 194, 203, 213 Cheshmeh-Ali, Iran 50 pottery 50 China 24–25, 75–76, 79, 82, 89–90, 235 Chopan, Turkmenistan 46, 60–63, 79, 119, 162, 186, 217–19, 221, 228, 236 Climatic Optimum/Altithermal 25–26, 31, 40, 66, 74, 81, 108, 172, 212, 221, 227, 230–31, 233–34, 236 Dam Dam Cheshme rockshelters, Turkmenistan 35–36, 45, 55, 59, 64, 81, 84, 87, 90, 113–16, 186, 197–207, 217–18, 226, 228, 230, 232

index animal remains 115, 201–2 charred seeds 115, 197–98 pottery 113–15, 204–5, 206–7 stone tools 54, 113–15, 203–5 wood charcoal 115, 199–200 Darai Kalon, Tajikistan 23 Dara-i Kalon, Afghanistan 51, 55 Darbazakir, Uzbekistan 55 Dargaz/Darreh Gaz plain/zone 50–51, 60, 64, 218, 222–23, 235 Dar-i Kur, Afghanistan 51, 53 Dashlidji, Turkmenistan 69, 78, 79, 236 deer 11, 54, 56, 66–67 Bokhara (Cervus elephas subsp. bactrianus) 15, 16, 33–34 Deh Keir Tepe, Iran 50–51 Deh Morasi Ghundai, Afghanistan) 51 Dereivka, Ukraine 82 desert kite 213 desert vegetation 10–11, 14, 31, 231 desiccation/oasis theory 44, 230 diffusion, cultural 225, 231, 233 demic 185, 231, 233 dog 58, 67–68, 88–90 domestic 18, 56, 62, 88–90, 91, 130, 174–77, 193 domestication 67–68, 89–90, 91 genetics 89 hunting 89–90 Dorian, Turkmenistan 109, 110 Elburz/Alborz mountains 5, 9, 49, 55, 62, 87, 222–23, 226, 234–35 elm (Ulmus carpinifolia) 9, 32, 166 equids 67, 80–82, 177, 201–2, 221, 227. See also ass, horse, onager erosion, soil 26, 64, 108, 112 wind 32, 65, 192, 228 Etek region 60–61, 218 faecal spherulites 129, 143–44, 147–49, 193, 214 Fergana depression/oasis/valley 53, 55, 59, 222 Fertile Crescent, Southwest Asian 74–78, 80, 85, 87–88, 90–91, 162, 215, 217, 229, 231–32, 235–36 fig (Ficus carica) 9–10, 32–33, 39–40, 199–200, 216 figurines 192, 212, 234 animal 63, 83, 89, 97, 102 human 97 firewood 9, 11, 26, 31, 136, 148, 168–70, 189, 213 fish 18, 58, 66–67, 176, 179, 193, 201–2, 214, 228 fishing 58–59, 67–68, 205, 229, 235 forests broadleaf deciduous 9, 23, 32, 55, 166, 216, 231 evergreen coniferous 9 riparian. See tugai fox 11, 30, 54, 66, 98, 177–179



index corsac (Vulpes corsac) 15, 17, 33–34, 176, 193, 213 red (Vulpes vulpes) 15, 17, 33–34, 40, 213

Gadymi, Turkmenistan 60–61, 218–19, 220 Ganj Dareh, Iran 85–86, 232, 236 gathering 32–33, 58–59, 67–68, 185, 216. See also wild plant foods gazelle 11, 16, 24, 55, 58, 66–67, 174, 176–77, 179, 193, 201–2, 228 goitered (Gazella subgutturosa) 13–14, 17, 33–34, 58, 213–14 Geoksyur oasis 46, 69, 79, 82, 236 Geok-tepe region/zone 60–61 Ghabristan, Iran 50, 82 Ghar-i Kamarband, Iran 49, 55–57, 63, 78–79, 81, 84, 87, 90, 232 Gievdzhik, Turkmenistan 60–61, 219 goat 51, 67–68, 83–86, 175 bezoar (Capra aegagrus) 12–13, 16, 17, 33–34, 40, 83–86, 91, 175–76, 193, 213, 226, 231 domestic 17–18, 56, 58–59, 62, 68–69, 84–86, 98, 175–77, 179, 193, 201–2, 214–15, 223, 233–34, 236 domestication 51, 58, 67–68, 83–86, 91, 218, 221, 226, 228, 232 genetics 85–86, 91, 226 ibex (Capra ibex) 13, 83 markhor (Capra falconeri) 13, 83, 85 wild 11, 53–54, 83–85 Gobi desert 14, 82 Gohar Tepe, Iran 50 Gonur 1, Turkmenistan 46, 47 grapevine (Vitis sylvestris) 9, 32, 216, 225 cultivation 27, 32, 131, 171 grindstones 58, 60, 67, 205, 224, 229 hare 11, 66 desert (Lepus tolai) 16, 17, 33–34, 176, 177, 178–79, 193, 213 hawthorn (Crataegus spp.) 9, 32 hedgehog, long-eared (Hemiechinus auritus) 17, 34, 176–79, 193, 213 European (Erinaceus europaeus) 178 Himalaya mountains 74, 87 Hindu Kush mountains 5, 9, 22, 51, 55, 58–59, 225 Hissar Culture 59, 224, 229, 234 Hodja-su, Turkmenistan 55, 58 horse 67 domestic 18, 62, 81–82, 227 domestication 81–82, 91 genetics 82 wild (Equus ferus) 54, 80–82, 227 Hotu, Iran 49, 55–57, 63, 232 hunting 11, 33–34, 58–59, 66–69, 89–90, 185, 205, 213–14, 217, 220, 229, 235 hyaena, cave 54

301 striped (Hyaena hyaena) 11, 15, 33–34, 40

ice cores, Greenland 24 Ilgynly-depe, Turkmenistan 14, 16, 46, 79, 81, 89, 213, 236 Io Sea 22, 191, 219 Iranian plateau 5, 47, 49, 50, 60, 61–63, 81–82, 86, 194, 219, 226 irrigation agriculture 7, 18, 26, 28–29, 44, 62, 69, 79, 219, 227, 233, 235–36. See also Jeitun, irrigation Islamic civilization/period 26, 51, 59 jackal, golden (Canis aureus) 11, 15, 30, 40, 89, 177 Janbas, Uzbekistan 45, 64–66, 228 Jarmo, Iraq 90, 96, 119, 182, 220, 236 Jebel, Turkmenistan 35–36, 45, 55, 57–59, 64, 81, 84, 87, 90, 113, 186, 188, 207, 217–18, 224, 226, 228, 230, 232, 234 Jeitun, Turkmenistan 11, 14, 17, 22, 27–34, 46, 48, 60–63, 84, 69 (n.), 95–107, 207, 211–18 animal remains 78, 84, 87–89, 174–79, 193, 213–14 bone working 178–79 building materials and yard deposits analysis 125–30, 146, 148 built structures 96–106, 123, 192, 194–95 charred plant macro-remains 73–74, 95, 98, 150–65, 166–70, 191, 215 climate 27 crop composition and processing 154–57, 162–65 cultivated plants 73–74, 147, 151–53, 164 cultivation 119, 124, 136, 140, 147, 164, 214–17 Culture 22, 46, 59, 60–62, 68, 95, 107, 185, 211, 218–24, 227–31, 233–36 ditch-like features 119, 123–24, 131, 136, 137–40, 191, 216 excavations 96–107, 137, 191–92, 211–12 herding 164, 214–16 hunting 179, 185, 213–14, 216–17 irrigation 124, 135, 137, 147, 149, 163–64, 188, 191, 215–16 phytoliths 126–28, 136, 142–49, 191, 193 pollen 171–73 pottery 49–51, 57, 61–64, 97, 102, 180–89, 193–94, 207, 212–13, 222–24 radiocarbon dating 119–24, 136 rainfed agriculture/cultivation 124, 147, 163–64, 215 site occupation 106–7, 121–23, 164, 177, 190–94, 211–12 soils 29–31, 123–24, 131–41, 146, 216 stone tools 97, 180–85, 194, 213, 217, 224, 233 vegetation 31–33, 191 Joyruk, Turkmenistan 36, 64, 117, 186, 207 juniper, Turkmen (Juniperus turcomanica) 9–10, 13, 39–40, 166, 199–200 Kailyu, Turkmenistan 55, 58, 64, 217, 228, 232 Kalate Khan, Iran 51

302 Kantar, Turkmenistan 61 Karabil plateau 7 Kara-Bogaz Gol bay 7, 15, 58 Kara-depe, Turkmenistan 46, 89 Kara Kamar, Afghanistan 51, 54–55, 58, 82, 87 Karakesy rockshelters, Turkmenistan 109–11 Karakum, desert 7, 11, 22, 96, 190, 213–14, 221–22, 233–34; Canal 5, 14, 18, 26, 27–28, 96 former rivers and lakes 25–26, 221–22, 231 Kara Su river 5, 28–29, 32, 96, 124, 168–71, 179, 190, 214–16 Karatau, Tajikistan 23, 53 Kavat, Turkmenistan 65, 66, 228 Kelleli sites, Turkmenistan 46 Keltiminar Culture 45, 48, 59, 64–68, 220–22; groups 68, 188, 194, 221 Keltiminar-related sites 59, 64–66, 186, 189, 217, 228, 234 pottery 186, 188, 207, 218, 234 sites 22, 45, 49, 64–67, 186, 189, 221–22, 228–29, 233–34 Kelyata, Turkmenistan 61 Kepele, Turkmenistan 61 Khapuz-depe, Turkmenistan 46 Kharkush, Tajikistan 54 Khoresmia 45, 59 Khorramabad, Iran 50 Kok Jar, Afghanistan 51 Kopetdag mountains 5, 9–10, 27, 32–33, 35, 54, 61, 64, 77, 84, 107–8, 213–15, 222, 233, 235 piedmont 5, 10, 28–29, 32–33, 44–46, 59, 60–62, 68–69, 190, 213–14, 218–23, 226–29, 231–36 Krasnovodsk plateau 7, 12, 45, 54, 58, 64, 213 Kugitang mountains 7, 10, 13 Kuldara, Tajikistan 23, 53 Kyzylkum desert 7, 22, 48–49, 59, 64, 78, 217, 221, 233 former rivers and lakes 25, 66, 221–22, 231 Lakhuti, Tajikistan 23, 53 Last Glacial Maximum 20–21, 23–24, 54, 230 Lateglacial 20–21, 24, 49, 55, 229–30 legume crops/pulses 74, 162, 215, 217, 225, 236 leopard 11, 40 Persian (Panthera pardus) 16, 33 Levant 55, 74–78, 80, 87, 90, 119 lion, Asiatic (Panthera leo) 16, 33 cave 54 liquidambar (Liquidambar sp.) 19 livestock, domestic 11, 26, 29, 33, 67, 108, 164 herding 68–69, 85, 89, 214–15, 229, 231–35. See also ass, camel, cattle, goat, horse, pastoralism, stockbreeding, sheep, transhumance lizard 176–177 gray monitor (Varanus griseus) 18, 30, 34, 193, 213 loess 22–23, 28, 53–54, 112 Lyavlyakan, Uzbekistan 65 ancient lake 7, 25, 64, 66, 228

index lynx, Eurasian (Lynx lynx) 11, 16, 33–34, 40 magnetic susceptibility 124, 137–41, 91 Maly Balkhan massif 5, 35, 36 manul, see cat, steppe maple, Turkmen (Acer turcomanicum) 9–10, 32, 166 Margiana 46, 83, 222 marten, rock/stone (Martes foina) 11, 17, 33–34 Meana-Chaacha district/zone 46, 47, 60–61 Mediterranean-type climate 10, 33, 107, 235 vegetation 10, 108 Mehrgarh, Pakistan 75, 78, 80, 86, 88 Merv, Turkmenistan 7, 44, 47–48 Mesolithic–Neolithic transition 113, 116, 207, 217–18, 225, 228, 232–33 microliths 97, 180, 194, 218, 224 geometric 54, 57, 61, 180, 185, 203–5, 212–13, 217, 224, 233 micromorphological analysis 126–28, 132–35, 191 milk 82, 176–77, 193, 214 mollusks 55, 66, 228 Mongolia 14, 87 Monjukli, Turkmenistan 46, 60–61, 68–69, 186, 219 monsoon 8, 25 Mousterian period 53–54 Naiza, Turkmenistan 61 Namazga-depe, Turkmenistan 45 Neanderthal 53–54 New Nisa, Turkmenistan 45–46, 60–61, 219 Novkhandan Tepe, Iran 51 Novroz Tepe, Iran 51 oak (Quercus spp.) 9, 19, 77 Old Nisa, Turkmenistan 45 Older Dryas stadial 24, 55 oleaster/Russian olive (Elaeagnus orientalis) 9, 166 onager/hemione (Equus hemionus) 11, 14, 16, 24, 33–34, 55, 58, 62, 66–67, 80–82, 177, 213, 227 organic residue 68 analysis 57, 82, 214 oriental plane (Platanus orientalis) 9, 166 otter, Eurasian (Lutra lutra) 11, 17, 34 Oyukli, Turkmenistan 36, 64–65, 117, 186, 207 palaeosol 19, 23, 25, 29–30, 53, 98, 119, 123–24, 131–41, 146, 190–91, 216 Pamir mountains 5, 7, 22, 55, 74 Parkhai cemetery/settlement, Turkmenistan 46, 64, 110 Parkhai rockshelter, Turkmenistan 109–10 Parthian civilization/period 26, 109, 111–12 particle-size analysis 126–27, 129–30, 132–35, 140 pastoralism 26, 58, 221. See also livestock herding, stockbreeding, transhumance



index

caprine 60, 179, 221, 229, 233 pear (Pyrus spp.) 9–10, 32, 108, 216, 225 Pessedjik , Turkmenistan 46, 60–61, 186, 219 phytolith analysis 126, 143–49 pig, domestic 56, 62, 69, 78–79, 227 domestication 67–68, 78–79, 91, 221 genetics 78–79 wild (Sus scrofa) 11, 15, 16, 17, 33–34, 40, 55, 66, 78–79, 91, 177, 179, 193, 213, 227 pistachio (Pistacia vera) 10, 33, 74, 108, 166, 216 semisavanna 10, 33 plum (Prunus spp.) 9–10, 32, 216 polecat, marbled (Vormela peregusna) 11, 17, 33–34 pollen analysis/palynology 23, 166, 171–73 pomegranate (Punica granatum) 10, 33, 108, 216, 225 poplar (Populus spp.) 9, 23, 32, 40, 166–69, 177, 189 population growth/increase 26, 99, 194, 212, 228, 233, 235–36 Postglacial 20–21, 24–26, 49, 55, 230–31 8.2 ka climatic event 24–25, 230–31, 233, 235–36 Pottery Neolithic period 78, 85, 88, 91 (n.), 185, 215, 236 Pre-Pottery Neolithic period 75, 77–78, 80, 85, 88, 91 (n.), 185, 226, 229, 232 pulses. See legume crops Qaleh Khan, Iran 51, 222 Qar-i Komishan, Iran 50, 55 radiocarbon dating 26 (n.), 50, 55–57, 63, 69 (n.), 119–24, 198, 201, 217–18 rainfed agriculture/cultivation 27, 64, 107, 147, 163–64, 215, 223, 231, 233, 235–36 reed (Phragmites australis) 9, 32, 40, 143, 147–48, 167–69, 171, 177, 192–93 Remisowka, Kazakhstan 23 Repetek Sand Desert Reserve 11, 31, 166 Sadeghabadi, Iran 50 sagebrush (Artemisia spp.) 10, 23–24, 31–32, 40, 166 Sagzabad, Iran 50, 82 saiga (Saiga tatarica) 11, 14–15, 17, 24, 33–34, 66, 213–14 saksaul 57, 170 black (Haloxylon persicum) 11, 31, 40, 146–147, 166, 213 white (H. aphyllum) 11, 31, 40, 146–47, 166 salinization 26 Samarkand site, Uzbekistan 54 Sang-i Čakmaq, Iran 49, 51, 63–64, 69 (n.), 222–23 Sarab, Iran 220 Sarazm, Tajikistan 47 Sarykamysh depression/lake 7, 15, 18, 19, 25–26, 59, 64, 66, 228 seal, Caspian (Phoca caspica) 55 Sel-Ungur, Kirgizstan 53–54 Shar-i Sokhta, Iran 47, 83 Shah Tepe, Iran 49, 223

303

sheep 67–68, 86–88, 175 argali (Ovis ammon) 12, 86–88 Asiatic mouflon (O. orientalis) 12, 86–88, 91, 226, 232 domestic 17–18, 56, 58–59, 62, 68–69, 86–88, 98, 175–77, 179, 193, 201–2, 214–15, 224, 233–34, 236 domestication 51, 58, 67–68, 88, 91, 221, 226, 228, 232 genetics 87–88, 226 urial (O. vignei) 12, 17, 33–34, 40, 86–88, 175–76, 193, 213, 226 wild 11, 54–55, 66, 86–88 shiblyak vegetation 10, 33, 39, 216, 235 shifting/swidden cultivation 214, 235–36 Shir-i Shayn, Iran 62, 223 Shugnou, Tajikistan 54, 82 Siabcha, Uzbekistan 54 Sialk, Iran 61, 83, 220, 223 sickle, blades 50, 60–61, 67, 95, 180, 183–85, 194, 212, 224, 229, 234 gloss 68, 183 handles 63, 179, 212 solonchak soil 7, 29, 30, 32, 35, 40 solonetz soil 29, 191 squirrel, long-clawed ground (Spermophilopsis leptodactylus) 17, 98, 34, 176–78 steppe vegetation 10, 23, 33, 54–55, 77, 216 stockbreeding 59, 67–68, 221–22, 224, 229, 234 Surabe Tepe, Iran 223, 227, 234 Tajik–Afghan basin 19, 22 Takhirbai-depe, Turkmenistan 46 takyr 6–7, 10, 29, 30, 59, 190, 215, 224 tamarisk (Tamarix spp.) 9, 23, 32, 40, 146–47, 166–69, 171–72, 177, 189, 192, 199 Taurus mountains 85, 226, 232 tectonic activity 19, 35, 64, 108, 112 Tepe Guran, Iran 96, 119, 220 Tepe Hissar, Iran 49, 62 Tepe Pardis, Iran 50 Tepe Yahya, Iran 47 terrace, river 25, 29, 38–39, 59, 108, 219, 224 marine/shoreline 20, 25, 35, 57, 58 Teshik Tash, Uzbekistan 53 Tibetan plateau 25, 87 Tien Shan mountains 5, 7, 9, 22, 54, 87 tiger, Caspian/Turanian (Panthera tigris) 11, 16, 33–34 Togolok, Turkmenistan, Kopetdag piedmont 46, 60–61, 119, 186, 217–19, 228, 231, 236 Togolok, Turkmenistan, Margiana 46, 83 Tolstov, Uzbekistan 66 tortoise 67 steppe (Agrionemys horsfieldii) 18, 33–34, 106, 176–77, 179, 193, 213–14 transhumance 17, 215, 234 tugai 9, 23, 32–33, 40, 66, 166–68, 177, 213, 214

304 Tugh Tepe, Iran 50, 63, 223 Tureng Tepe, Iran 49, 62, 223 turquoise 61, 217 Tutkaul, Tajikistan 59, 224 Uchashchi, Uzbekistan 22, 228 Ulug-depe, Turkmenistan 46 urban settlements 26, 43, 69 Urkutsay, Uzbekistan 23 Ustyurt plateau 7, 14, 64, 228 Uzboi channel/river 7, 21, 25, 35–40, 57, 59, 64, 66, 117, 207, 228, 230, 234 former delta 35, 40, 228 river terraces 38–39, 219 walnut (Juglans regia) 9, 32, 166, 225 water buffalo 67 wheat (Triticum spp.) 62, 162, 188, 198 bread (T. aestivum) 62, 69, 236 cultivation 59, 62, 119, 215–216, 231–32 domestication 76–78, 232 einkorn, domestic (T. monococcum) 62, 68, 74, 76–78, 90, 152–53, 162, 192, 215, 225–26, 232, 236 einkorn, wild (T. m. boeoticum) 76–77, 163, 232 emmer, domestic (T. dicoccum) 74, 76, 152, 162, 192, 215, 232, 236

index emmer, wild (T. dicoccoides) 76, 163 free-threshing (T. aestivum/durum type) 62, 74, 77, 91, 153–54, 162, 192, 215, 232 genetics 76–77 glume 62, 74, 76–78, 91, 151–155, 162, 192, 215, 217, 226, 232, 236 Timopheev’s, domestic (T. timopheevi) 76, wild (T. araraticum) 76 T. urartu 76–77 wild plant foods 9–10, 32–33, 67, 193, 216, 220, 229. See also entries for species bearing edible fruits and nuts willow (Salix spp.) 9, 32, 166–69, 177, 189 wolf, gray (Canis lupus) 11, 15, 17, 33–34, 40, 54, 88–90, 91, 177 woodland 9–10, 23, 33, 39, 74, 77, 166, 231, 235 Yam, Iran 50, 68 Yarti-Gumbez, Turkmenistan 61 Yarim Tepe, Iran, Khorassan 51 Yarim Tepe, Iran, Mazandaran 49, 62–63, 220, 222–23 Yaz I, Turkmenistan 46, 47 Younger Dryas stadial 24, 55, 230 Zaghe, Iran 50, 82 Zagros mountains 54, 74–76, 85–87, 180, 226, 232, 234–35 Zaman-baba, Uzbekistan 68