Southern Turkmenistan in the Neolithic: A petrographic case study 9781841718620, 9781407328652

The Neolithic period of southern Turkmenistan, Central Asia is the primary focus of this study. During the Neolithic, so

160 70 72MB

English Pages [206] Year 2005

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Southern Turkmenistan in the Neolithic: A petrographic case study
 9781841718620, 9781407328652

Table of contents :
Front Cover
Title Page
Copyright
Dedication
Abstract
Acknowledgements
Table of Contents
TABLE OF FIGURES
INTRODUCTION
CHAPTER 1 PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA
CHAPTER 2 HISTORY OF ARCHAEOLOGICAL INVESTIGATION
CHAPTER 3 EXCAVATION METHODOLOGY AND RESULTS
CHAPTER 4 CERAMIC TYPOLOGY
CHAPTER 5 INTRODUCTION TO PETROGRAPHY
CHAPTER 6 THE KOPET DAG CASE STUDY
CHAPTER 7 SYNTHESIS AND INTERPRETATION
CHAPTER 8 NEOLITHIC CONTRASTS AND COMPARISONS
APPENDIX 1 GLOSSARY
APPENDIX 2 CERAMIC TYPOLOGY
APPENDIX 3 FABRIC CLASSIFICATION TEXTUAL DESCRIPTIONS
APPENDIX 4 CHAFF OBSERVATIONS
APPENDIX 5 GENERAL DATABASE
BIBLIOGRAPHY

Citation preview

BAR S1423 2005

Southern Turkmenistan in the Neolithic

COOLIDGE

A petrographic case study

Jennifer Coolidge SOUTHERN TURKMENISTAN IN THE NEOLITHIC

B A R

BAR International Series 1423 2005

Southern Turkmenistan in the Neolithic A petrographic case study

Jennifer Coolidge

BAR International Series 1423 2005

Published in 2016 by BAR Publishing, Oxford BAR International Series 1423 Southern Turkmenistan in the Neolithic © J Coolidge and the Publisher 2005 The author's moral rights under the 1988 UK Copyright, Designs and Patents Act are hereby expressly asserted. All rights reserved. No part of this work may be copied, reproduced, stored, sold, distributed, scanned, saved in any form of digital format or transmitted in any form digitally, without the written permission of the Publisher.

ISBN 9781841718620 paperback ISBN 9781407328652 e-format DOI https://doi.org/10.30861/9781841718620 A catalogue record for this book is available from the British Library BAR Publishing is the trading name of British Archaeological Reports (Oxford) Ltd. British Archaeological Reports was first incorporated in 1974 to publish the BAR Series, International and British. In 1992 Hadrian Books Ltd became part of the BAR group. This volume was originally published by Archaeopress in conjunction with British Archaeological Reports (Oxford) Ltd / Hadrian Books Ltd, the Series principal publisher, in 2005. This present volume is published by BAR Publishing, 2016.

BAR

PUBLISHING BAR titles are available from:

E MAIL P HONE F AX

BAR Publishing 122 Banbury Rd, Oxford, OX2 7BP, UK [email protected] +44 (0)1865 310431 +44 (0)1865 316916 www.barpublishing.com

…IN MEMORY OF JEREMY THOMAS LARGE 1968-1993

ABSTRACT The Neolithic period of southern Turkmenistan, Central Asia, is the primary focus of this thesis. A comparison of the settled agro-pastoral Jeitun Culture and the nomadic hunting/gathering/stockbreeding Keltiminar Culture illuminates the existence of a significant archaeological borderzone in prehistoric Turkmenia. This borderzone is representative of a “hard frontier” (Sherratt pers. comm. 2001) characterised by dryness, and a prolonged situation of widely differing subsistence systems existing contemporaneously in close proximity to one another. The notion of southern Turkmenia as a borderzone holds archaeological implications for the migration to, and the peopling of, southwestern Central Asia, the subsistence regime of the region, its origins and adaptations, as well as contact and communication networks between Turkmenia, via the Iranian plateau, to Mesopotamia and the Indus valley. Southern Turkmenia also functioned as a corridor to the steppe zone of inner Eurasia to the north. In this light, it is crucial to look at the Jeitun and Keltiminar Cultures as individual entities, and to understand them as such, but also to look at the dynamics at work between the two cultures, their respective geographical interaction spheres, their differing trajectories of economic and social development, and the borderzone between them as a locus for cultural exchange and the dispersal of traditions and knowledge. A petrographic case study of Jeitun Culture ceramic assemblages is used to explore issues of mode of production, distribution and exchange, and economy in a Neolithic society. The petrographic case study uses techniques of optical and polarising microscopy of ceramic samples and thin-sections, and entails the following three-fold analysis: 1) The creation of a fabric classification system which allows for the comparison of regional and temporal variation within the general assemblage, and also between the Jeitun Culture assemblages and those from adjacent regions; 2) A description of the pottery production technology utilised by the Jeitun peoples, allowing for a more comprehensive understanding of the Neolithic mode of production and economy; 3) A provenancing study which addresses the issues of population movement, distribution, trade, and exchange. The provenancing study uses techniques of geomorphological analysis, geological ground survey, and the geoarchaeological analysis and interpretation of remotely-sensed images (various spectral bands in the visible-infrared-radar range, different geometrical resolution, and different time of acquisition), appropriately elaborated (geometric and radiometric corrections, thematic classification, principal components analysis). The Jeitun peoples of the Kopet Dag piedmont zone were the first agro-pastoralists of southwestern Central Asia, and as such, they determined the path of future economic intensification, population expansion, and the subsequent origins of urbanisation. A comparison of the Jeitun and Keltiminar serves to more accurately place the prehistoric cultures of southern Turkmenia within the broader trends towards settled agriculture, nomadic pastoralism, and subsequent urbanism in southwestern Central Asia and inner Eurasia.

i

ACKNOWLEDGEMENTS Thanks go first and foremost to my supervisor, Chris Gosden, whose invitation to join the British/Soviet archaeological research team prompted this thesis. He has provided countless hours of discussion and encouragement, with consistent invaluable insight and level-headed advice. A wealth of scholarship precedes this research, most notably that undertaken in the former Soviet Union, by Soviet academicians. I am greatly indebted to both our Turkmenian and Russian colleagues, most importantly Kakamurad Kurbansakhatov and Sonya Lollekova for providing logistical support for archaeological fieldwork and survey, gracious hospitality, scholarly advice, and access to archived collections and bibliographic resources at the Institute of Archaeology, Turkmen Academy of Sciences, Ashkhabad. Heartfelt thanks also go to Vladimir Zavyalov and Vadim Masson, who organised archival and bibliographic study in St. Petersburg at both the Hermitage Museum and the Institute for the History of Material Culture, Russian Academy of Sciences. Bruno Marcolongo, Director of the Institute for Applied Geology, National Research Council of Italy, Padova, provided resources and expertise in the analysis of satellite and other remotely sensed images for geoarchaeological interpretation. Anatoly Bushmakin, Director of the Geological Museum, Turkmen Geological Expedition, Ashkhabad, served as expert geological consultant for the petrographic study. I would like to express my extreme gratitude to the Research Laboratory for Archaeology and the History of Art, University of Oxford. I am especially indebted to Director Mike Tite and my second supervisor Chris Doherty. The research laboratory provided me with a desk at which to write this thesis (in the infamous office also occupied by Alistair Pike and Karen Privat), expert teaching in laboratory techniques, consistent support, and a keen interest in my research. Thanks also go to the other members of the British research team for their camaraderie, advice, and sense of adventure and humour in the most drastic and dirty situations. In particular, project co-directors David Harris (University College London) and Chris Gosden, Mike Charles (University of Sheffield), Keith Dobney (University of Durham), Patrick Blackman and Greger Larson (both University of Oxford). Mike Charles and Jo Bending (University of Sheffield) carried out an eleventh-hour pilot study and analysis of archaeobotanical remains from the ceramic samples in this study. Gratitude is due to a host of other researchers with whom I consulted, including Fred Hiebert, Philip Kohl, C. C. Lamberg-Karlovsky, Pavel Dolukhanov, Gennagi Markov, Yuri Berezhkin, and Andrew Sherratt. Undertaking this thesis, including the fieldwork, academic interviews, and bibliographic research that were integral to the project, would have been entirely impossible without the expert Russian language teaching I received from Natasha Walker for four years. Chief computing gurus, Peter Bentley and Tamsin O’Connell, dedicated editor Barbara Large MBE, and talented illustrators Kathy Allinson and James Reeve were absolutely essential to the final production stage of this thesis. Thanks are also due to Valentina Kostina and Faina Vasilieva, for gracious hospitality and companionship in Moscow and St. Petersburg. University College, the British Schools and Universities Foundation, the British Academy, the Gilchrist Trust, and the Meyerstein Fund have all provided generous financial support. Finally and most importantly I come to my parents, John and Helen Coolidge, who have been a constant source of support and encouragement in this and every other endeavour I have ever undertaken. Thank you could never be enough.

iii

TABLE OF CONTENTS Abstract.............................................................................................................................................................. i Acknowledgements.......................................................................................................................................... iii Table of Contents...............................................................................................................................................v Table of Figures ............................................................................................................................................... ix Introduction ........................................................................................................................................................ xiii CHAPTER 1: PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA ....................................................................1 Hydrology ..........................................................................................................................................................1 Mountains ..........................................................................................................................................................1 Piedmont Zone...................................................................................................................................................3 Alluvial Floodplain ............................................................................................................................................3 Desert.................................................................................................................................................................3 Palaeoclimate and Palaeohydrography ..............................................................................................................4 CHAPTER 2:HISTORY OF ARCHAEOLOGICAL INVESTIGATION ..................................................................................7 Chronology ...........................................................................................................................................................10 The Mesolithic .................................................................................................................................................10 Transition from the Caspian Mesolithic to the Jeitun Neolithic ......................................................................12 Neolithic Subsistence Adaptations ..................................................................................................................13 The Neolithic ...................................................................................................................................................14 The Jeitun Culture of the southern super-zone............................................................................................14 Iranian Plateau: Sang-e Caxamaq..............................................................................................................14 Gorgan Valley: Tureng Tepe and Yarim Tepe ............................................................................................14 Southern Turkmenistan: Jeitun and the sites of the Kopet Dag piedmont...................................................16 Jeitun Culture Phases..................................................................................................................................17 The Keltiminar Culture of the northern super-zone ....................................................................................18 Keltiminar Culture transition from the late Neolithic to the early Aeneolithic...........................................19 The Keltiminar Culture and northward.......................................................................................................19 Transition from the Jeitun Culture to the Anau 1A period..........................................................................20 The pre-Namazga Aeneolithic Anau 1A Period (Kopet Dag) .........................................................................20 CHAPTER 3: EXCAVATION METHODOLOGY AND RESULTS ....................................................................................23 Iranian Plateau: Sang-e Caxamaq ....................................................................................................................23 Jeitun Culture sites: Kopet Dag central zone ...................................................................................................23 Jeitun: Soviet excavations ...........................................................................................................................23 Jeitun: British excavations ..........................................................................................................................28 Chopan ........................................................................................................................................................30 Togolok........................................................................................................................................................31 Pessedjik......................................................................................................................................................32 New Nisa .....................................................................................................................................................33 Gievdzhik.....................................................................................................................................................33 Jeitun Culture sites: Meana-Chaacha district (Kopet Dag eastern zone) .........................................................34 Chagylly ......................................................................................................................................................34 Chakmakli ...................................................................................................................................................36 Mondjukli ....................................................................................................................................................37 Gademi ........................................................................................................................................................38 Jeitun Culture sites: Kopet Dag western zone..................................................................................................39 Bami ............................................................................................................................................................39 Keltiminar and Keltiminar-related sites...........................................................................................................40 The lower Amu Darya and Akcha Darya delta: Djanbas............................................................................40 The Zeravshan, Ayakagitma, Daryasai, and Makhandarya ........................................................................43 Lake Lyavlyakan and the inner Kyzyl Kum .................................................................................................43 Uzboi river, north Pribalkhan, south Pribalkhan........................................................................................43 The Bolshoi Balkhan mountains (Figure 3.24) ...........................................................................................48 v

The Bolshoi Balkhan mountains: Dam Dam Cheshme I .............................................................................53 The Bolshoi Balkhan mountains: Dam Dam Cheshme II............................................................................55 CHAPTER 4: CERAMIC TYPOLOGY .........................................................................................................................59 Form and Decoration: Jeitun Culture phases ...................................................................................................59 Jeitun Culture Phase 1A..............................................................................................................................59 Jeitun Culture Phase 1B..............................................................................................................................59 Jeitun Culture Phase 2 ................................................................................................................................64 Jeitun Culture Phase 3 ................................................................................................................................65 Ceramic Assemblages by site ..........................................................................................................................66 Jeitun...........................................................................................................................................................66 Chopan ........................................................................................................................................................66 Togolok........................................................................................................................................................66 Pessedjik......................................................................................................................................................66 New Nisa .....................................................................................................................................................66 Kelyata, Kepele, Kantar, Naiza, and Chakmak Dash Beik .........................................................................67 Chagylly ......................................................................................................................................................67 Chakmakli ...................................................................................................................................................68 Mondjukli ....................................................................................................................................................68 Bami ............................................................................................................................................................68 Djebel ..........................................................................................................................................................69 Dam Dam Cheshme I ..................................................................................................................................72 Dam Dam Cheshme II .................................................................................................................................72 Oyukli ..........................................................................................................................................................72 Form and Function...........................................................................................................................................72 Statistical Analyses ..........................................................................................................................................76 Identification of Research Questions ...............................................................................................................83 Fabric Classification...................................................................................................................................83 Production Technology ...............................................................................................................................84 Provenancing ..............................................................................................................................................84 CHAPTER 5: INTRODUCTION TO PETROGRAPHY.....................................................................................................85 General Petrographic Methodologies...............................................................................................................86 Fabric Classification and Research Questions ...........................................................................................86 Study of Production Technology .................................................................................................................87 Provenancing Study.....................................................................................................................................87 Specific Research Methodology ......................................................................................................................87 CHAPTER 6: THE KOPET DAG CASE STUDY ............................................................................................................89 Research Aims and Questions .........................................................................................................................89 Fabric Classification...................................................................................................................................89 Technology Study ........................................................................................................................................90 Provenancing ..............................................................................................................................................90 Geological Predictions................................................................................................................................90 Western Zone .........................................................................................................................................93 Central Zone...........................................................................................................................................93 Eastern Zone...........................................................................................................................................99 Geological Overview...................................................................................................................................94 Methodology....................................................................................................................................................96 Sample Selection .........................................................................................................................................96 Cataloguing.................................................................................................................................................96 Thin-Section Analysis..................................................................................................................................98 Results .............................................................................................................................................................99 Fabric Classification...................................................................................................................................99 Technology Study ......................................................................................................................................100 Collection of Raw Materials.................................................................................................................101 Preparation of Raw Materials...............................................................................................................101 Preparation of the Ceramic Fabric........................................................................................................101 vi

Vessel Forming ....................................................................................................................................102 Drying of the Vessel.............................................................................................................................102 Surface Treatment ................................................................................................................................102 Firing ....................................................................................................................................................102 Temper .................................................................................................................................................104 Vessel Repair........................................................................................................................................105 Provenancing ............................................................................................................................................105 CHAPTER 7: SYNTHESIS AND INTERPRETATION ...................................................................................................111 Statistical Analyses: Correlation of Form and Fabric ....................................................................................111 The ceramic assemblages as evidence for Neolithic development ................................................................112 Mode of Production .......................................................................................................................................113 Distribution and Exchange.............................................................................................................................114 Creating a Cohesive Picture of the Jeitun Neolithic ......................................................................................115 CHAPTER 8: NEOLITHIC CONTRASTS AND COMPARISONS ...................................................................................117 Southern Turkmenia as an Archaeological Border Zone ...............................................................................118 The Jeitun Culture as a semi-sedentary adaptation ........................................................................................119 Relationships with Iran and Afghanistan .......................................................................................................122 Jeitun Culture relations with the Keltiminar ..................................................................................................123 Toward the Aeneolithic .................................................................................................................................123 Future Research .............................................................................................................................................124 APPENDIX 1: APPENDIX 2: APPENDIX 3: APPENDIX 4: APPENDIX 5:

GLOSSARY ..........................................................................................................................125 CERAMIC TYPOLOGY ..........................................................................................................129 FABRIC CLASSIFICATION TEXTURAL DESCRIPTIONS ..........................................................149 CHAFF OBSERVATIONS .......................................................................................................153 GENERAL DATABASE .........................................................................................................155

BIBLIOGRAPHY ....................................................................................................................................................181

vii

TABLE OF FIGURES CHAPTER 1: PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA Figure 1.1 Southern Central Asia .................................................................................................................2 Figure 1.2 Caspian Sea Transgression Regression Events ...........................................................................4 CHAPTER 2:HISTORY OF ARCHAEOLOGICAL INVESTIGATION AND CHRONOLOGY Figure 2.1 Archaeological Chronology ......................................................................................................11 Figure 2.2 Distribution of Jeitun Culture sites ...........................................................................................15 Figure 2.3 Keltiminar Culture area ............................................................................................................16 Figure 2.4 Keltiminar sites by phase...........................................................................................................19 Figure 2.5 Distribution of Anau 1A sites.....................................................................................................21 CHAPTER 3: EXCAVATION METHODOLOGY AND RESULTS Figure 3.1 Jeitun Culture architecture........................................................................................................24 Figure 3.2 Jeitun Culture phases and corresponding artefactual assemblage ...........................................25 Figure 3.3 Jeitun architecture.....................................................................................................................26 Figure 3.4a Jeitun original site photos .......................................................................................................26 Figure 3.4b Jeitun original site photos .......................................................................................................27 Figure 3.5 Chopan architecture and stratigraphy ......................................................................................30 Figure 3.6 Chopan ......................................................................................................................................30 Figure 3.7 Togolok site stratigraphy...........................................................................................................31 Figure 3.8 Togolok......................................................................................................................................31 Figure 3.9 Pessedjik architecture ...............................................................................................................32 Figure 3.10 Pessedjik..................................................................................................................................32 Figure 3.11 Chagylly architecture ..............................................................................................................34 Figure 3.12 Chagylly...................................................................................................................................34 Figure 3.13 Chakmakli................................................................................................................................36 Figure 3.14 Mondjukli.................................................................................................................................37 Figure 3.15 Gademi architecture................................................................................................................38 Figure 3.16 Bami site stratigraphy .............................................................................................................39 Figure 3.17 Bami ........................................................................................................................................40 Figure 3.18 Keltiminar Culture architecture ..............................................................................................41 Figure 3.19 Keltiminar Culture artefactual assemblage.............................................................................42 Figure 3.20 Bolshoi Balkhan, lower Uzboi, north and south Pribalkhan ...................................................44 Figure 3.21 Bolshoi Balkhan and Uzboi river ............................................................................................45 Figure 3.22 Upper and Lower Uzboi river artefactual assemblage............................................................46 Figure 3.23 Oyukli artefactual assemblage ................................................................................................47 Figure 3.24 Bolshoi Balkhan.......................................................................................................................48 Figure 3.26 Djebel artefactual assemblage ................................................................................................49 Figure 3.27 Djebel site stratigraphy ...........................................................................................................50 Figure 3.28 Djebel cave profile ..................................................................................................................50 Figure 3.29 Djebel cave plan view..............................................................................................................50 Figure 3.30 Djebel lithic assemblage..........................................................................................................51 Figure 3.31 Djebel ......................................................................................................................................52 Figure 3.32 Dam Dam Cheshme I cave plan view and site stratigraphy ....................................................54 Figure 3.33 Dam Dam Cheshme I...............................................................................................................54 Figure 3.35 Dam Dam Cheshme II cave plan view.....................................................................................55 Figure 3.36 Dam Dam Cheshme II .............................................................................................................56 CHAPTER 4: CERAMIC TYPOLOGY Figure 4.1 Jeitun Culture ceramic motifs by zone and cultural phase........................................................60 Figure 4.2 Jeitun Culture ceramic vessel forms..........................................................................................60 Figure 4.3 Jeitun Culture ceramic assemblages by phase and site.............................................................61 Figure 4.4 Jeitun Culture ceramic motifs....................................................................................................62 Figure 4.5 Jeitun Culture ceramic motifs by phase.....................................................................................63 Figure 4.6 Neolithic artefactual assemblages from Turkmenia and the Near East ....................................63 ix

Figure 4.7 Comparison of Jeitun Culture and Sialk 1 ceramic motifs ........................................................64 Figure 4.8 Jeitun ceramic form statistical analyses....................................................................................74 Figure 4.9 Jeitun ceramic ware and form histograms ................................................................................75 Figure 4.10 Jeitun Culture comparative histograms of ceramic ware and form ........................................77 Figure 4.11 Jeitun ware distribution...........................................................................................................77 Figure 4.12 Chopan ware distribution........................................................................................................78 Figure 4.13 Togolok ware distribution .......................................................................................................78 Figure 4.14 Pessedjik ware distribution .....................................................................................................79 Figure 4.15 Chagylly ware distribution ......................................................................................................79 Figure 4.16 Chakmakli ware distribution ...................................................................................................80 Figure 4.17 Mondjukli ware distribution ....................................................................................................80 Figure 4.18 Jeitun vessel form distribution.................................................................................................81 Figure 4.19 Jeitun vessel form distribution by ware ...................................................................................81 Figure 4.20 Distribution of ware by site assemblage..................................................................................82 Figure 4.21 Distribution of consistency by site assemblage .......................................................................82

CHAPTER 6: THE KOPET DAG CASE STUDY Figure 6.1 Geological Basemap Figure 6.2 Solid geology of Kopet Dag study area .....................................................................................91 Figure 6.3 Geomorphological Map.............................................................................................................95 Figure 6.4 Kopet Dag satellite image example ...........................................................................................95 Figure 6.5 Overview of mineralogy ..........................................................................................................106 Figure 6.6 Overview of mineralogy ..........................................................................................................106 Figure 6.7 Geologically diagnostic mineralogy........................................................................................106 Figure 6.8 Geologically diagnostic mineralogy........................................................................................106 Figure 6.9a Thin-section photomicrographs.............................................................................................107 Figure 6.9b Thin-section photomicrographs.............................................................................................108 Figure 6.10 Thin-section photomicrographs.............................................................................................109 CHAPTER 8: NEOLITHIC CONTRASTS AND COMPARISONS Figure 8.1 Settlement distribution on the Darreh Gaz plain.....................................................................120 APPENDIX 2: CERAMIC TYPOLOGY Figure A2.1 Jeitun ceramic vessel photographs .......................................................................................129 Figure A2.2 Jeitun ceramic vessel photographs .......................................................................................130 Figure A2.3 Jeitun ceramic forms and decoration....................................................................................131 Figure A2.4 Jeitun vessel forms ................................................................................................................131 Figure A2.5 Jeitun ceramic forms, decoration and artefactual assemblage.............................................131 Figure A2.6 Jeitun ceramic forms and decoration....................................................................................131 Figure A2.7 Jeitun ceramic forms and decoration....................................................................................132 Figure A2.8 Jeitun ceramic forms and decoration....................................................................................132 Figure A2.9 Jeitun ceramic forms and decoration, grinding stones .........................................................132 Figure A2.10 Jeitun ceramic motifs ..........................................................................................................132 Figure A2.11 Jeitun ceramic motifs ..........................................................................................................133 Figure A2.12 Jeitun ceramic motifs ..........................................................................................................133 Figure A2.13 Jeitun ceramic motifs ..........................................................................................................133 Figure A2.14 Jeitun ceramic motifs ..........................................................................................................133 Figure A2.15 Jeitun ceramic motifs ..........................................................................................................134 Figure A2.16 Chopan ceramic forms and decoration ...............................................................................134 Figure A2.17 Chopan ceramic forms and decoration ...............................................................................134 Figure A2.18 Chopan ceramic forms and decoration ...............................................................................134 Figure A2.19 Chopan ceramic forms and decoration ...............................................................................135 Figure A2.20 Chopan ceramic motifs .......................................................................................................135 Figure A2.21 Togolok ceramic forms and decoration ..............................................................................135 Figure A2.22 Togolok ceramic forms and decoration ..............................................................................135 Figure A2.23 Togolok ceramic motifs .......................................................................................................136 Figure A2.24 Togolok, Pessedjik, and Chakmakli ceramic motifs............................................................136 x

Figure A2.25 Togolok, Pessedjik, and Chakmakli ceramic motifs............................................................136 Figure A2.26 Pessedjik ceramic forms and decoration.............................................................................136 Figure A2.27 Pessedjik ceramic forms and decoration.............................................................................137 Figure A2.28 Pessedjik ceramic forms and decoration.............................................................................137 Figure A2.29 Pessedjik ceramic forms and decoration.............................................................................137 Figure A2.30 Pessedjik ceramic forms and decoration.............................................................................137 Figure A2.31 Pessedjik ceramic forms and decoration.............................................................................138 Figure A2.32 Pessedjik ceramic motifs .....................................................................................................138 Figure A2.33 Pessedjik ceramic motifs .....................................................................................................138 Figure A2.34 Pessedjik ceramic motifs .....................................................................................................138 Figure A2.35 Pessedjik ceramic motifs .....................................................................................................139 Figure A2.36 Pessedjik ceramic motifs .....................................................................................................139 Figure A2.37 Pessedjik ceramic motifs .....................................................................................................139 Figure A2.38 New Nisa ceramic forms and decoration ............................................................................139 Figure A2.39 Kelyata, Kantar, Kepele, Naiza ceramic forms and decoration..........................................140 Figure A2.40 Chagylly ceramic forms and decoration .............................................................................140 Figure A2.41 Chagylly ceramic forms and decoration .............................................................................140 Figure A2.42 Chakmakli ceramic forms and decoration ..........................................................................140 Figure A2.43 Chakmakli ceramic forms and decoration ..........................................................................141 Figure A2.44 Chakmakli ceramic forms, decoration, and artefactual assemblage ..................................141 Figure A2.45 Chakmakli ceramic motifs...................................................................................................141 Figure A2.46 Chakmakli ceramic motifs...................................................................................................141 Figure A2.47 Chakmakli ceramic motifs...................................................................................................142 Figure A2.48 Chakmakli ceramic motifs...................................................................................................142 Figure A2.49 Chakmakli ceramic motifs...................................................................................................142 Figure A2.50 Mondjukli ceramic forms and decoration ...........................................................................142 Figure A2.51 Bami ceramic forms and decoration ...................................................................................143 Figure A2.52 Bami ceramic forms and decoration ...................................................................................143 Figure A2.53 Bami ceramic forms and decoration ...................................................................................143 Figure A2.54 Djebel ceramic typology by occupational phase.................................................................143 Figure A2.55 Djebel ceramic vessel with decoration................................................................................144 Figure A2.56 Djebel ceramic forms and artefactual assemblage .............................................................144 Figure A2.57 Djebel ceramic forms ..........................................................................................................144 Figure A2.58 Djebel ceramic vessel with decoration................................................................................144 Figure A2.59 Djebel ceramic vessel with decoration................................................................................144 Figure A2.60 Djebel ceramic forms and decoration.................................................................................145 Figure A2.61 Djebel ceramic forms and decoration.................................................................................145 Figure A2.62 Djebel ceramic forms ..........................................................................................................145 Figure A2.63 Djebel ceramic forms and decoration.................................................................................145 Figure A2.64 Djebel ceramic forms ..........................................................................................................146 Figure A2.65 Djebel ceramic forms ..........................................................................................................146 Figure A2.66 Djebel ceramic forms and decoration.................................................................................146 Figure A2.67 Djebel ceramic forms ..........................................................................................................146 Figure A2.68 Dam Dam Cheshme I ceramic profile and motifs ...............................................................147 Figure A2.69 Oyukli ceramic profiles and motifs .....................................................................................147 Figure A2.70 Oyukli ceramic profiles and motifs .....................................................................................147 Figure A2.71 Oyukli ceramic profiles and motifs .....................................................................................147

xi

INTRODUCTION The Neolithic period of southern Turkmenistan (referred to as Turkmenia in the prehistoric context), Central Asia is the primary focus of this thesis. During the Neolithic, southern Turkmenia was inhabited by two main groups living in two discrete ecological-environmental zones: the Jeitun Culture of the southern super-zone and the Keltiminar Culture of the northern super-zone. The Jeitun peoples practised an agro-pastoral settled (or semi-settled) lifestyle in the upland intermontane valleys, the fertile piedmont zone, and the alluvial floodplain of the Kopet Dag mountains. The Keltiminar peoples practised a mobile hunting, gathering, fishing, and stockbreeding seasonal-round subsistence system while inhabiting the semi-desert, desert, and deltaic areas of the Kara and Kyzyl Kum deserts, and the lower Amu Darya and Zeravshan rivers.

workings of the Jeitun Culture. This framework is then used for comparison with the Keltiminar Culture (based on a literature survey rather than a ceramic case study), and more broadly with the prehistoric cultures of the ancient Near East, Mesopotamia, the Indus valley, and the steppelands of inner Eurasia. Had I not encountered problems in obtaining samples from Keltiminar Culture ceramic assemblages housed at Moscow State University, the Moscow branch of the Russian Academy of Sciences, and the Bukhara and Samarkand branches of the Uzbek Academy of Sciences, a similar petrographic analysis of Keltiminar Culture ceramics would have been undertaken. Unfortunately, academicians in the former Soviet Union are still enmeshed in the politics that dictate certain rules of collaboration. One often finds that collaboration with scholars from certain institutes automatically precludes collaboration with investigators from competing institutes. This is (and historically was) the case for the Moscow and Leningrad (St. Petersburg) branches of the Soviet Academy of Sciences (now the Russian Academy of Sciences) Undoubtedly, this phenomenon makes carrying out comprehensive and comparative studies difficult.

Initially, this study was intended primarily as a petrographic analysis of Jeitun Culture ceramic assemblages excavated by both Soviet and British archaeologists since the 1950s at Neolithic sites across the Kopet Dag piedmont zone (Figure 1.1). Beyond a straightforward analysis of Jeitun Culture ceramic assemblages, I originally hoped to obtain a significant sample of Keltiminar Culture ceramics for petrographic comparison. Unfortunately, this proved impossible due to logistics and the difficulties often encountered in obtaining archaeological samples in the former Soviet Union. However, in translating and synthesising the pertinent Soviet archaeological literature (published almost exclusively in Russian) for the Jeitun and Keltiminar Cultures, it became apparent that a comparison of the Jeitun and Keltiminar Cultures pointed to the existence of a significant archaeological borderzone in prehistoric Turkmenia. This borderzone is representative of a “hard frontier” (Sherratt pers. comm. 2001) characterised by dryness and a prolonged situation of widely differing subsistence systems existing contemporaneously in close proximity to one another. The development of my interest in the notion of southern Turkmenia as an archaeological borderzone led to an expansion of comparisons with areas both further south (the mountain and desert areas of northeastern Iran and northwestern Afghanistan), and further north (the desert, semi-desert, semi-steppe and steppelands of Uzbekistan, Kazakhstan, western Siberia and the Urals) based on a survey of the existing archaeological literature.

It is with the above outline of focus that I present the structure of this thesis. Chapters 1-3 provide the background critical to an accurate understanding of the typological and petrographic case studies, the insight those studies can provide to our knowledge of the structure of the Jeitun and Keltiminar Neolithic adaptations, and the notion of prehistoric Turkmenia as an archaeological borderzone. Initially, I present an overview of the palaeogeography, palaeoenvironment and palaeoclimate (Chapter 1), followed by the history of archaeological excavation in Turkmenistan and an archaeological chronology for the region (Chapter 2). Finally, narrowing focus from the broad-based to the specific, I present the excavation results for the sites involved in the typological and petrographic case studies (Chapter 3). The ceramic assemblages for the petrographic case study are initially introduced in the context of ceramic typology (Chapter 4), and subsequently in terms of general petrography (Chapter 5) and the Kopet Dag case study (Chapter 6). Finally, Chapter 7 represents a synthesis and interpretation of the data and results from Chapters 4 and 6. This synthesis and interpretation serves as a precursor to a final discussion of contrasts, comparisons, and possibilities for future research in the region (Chapter 8).

Although this thesis still retains its main attributes as a ceramic petrographic case study, the contrast between the Jeitun and Keltiminar Cultures and the existence of this archaeological borderzone has become a much larger part of the thesis than initially planned. The petrographic case study provides insight into the structure of the Jeitun Neolithic, specifically the mode of production, and serves to provide a framework for the understanding of the inner

The conventions used in this thesis for transliterating from Russian to English are based on the official romanization tables of the United States Library of Congress. Every effort has been made to preserve original place names, site names, and archaeological terminology in the most accurate manner possible. xiii

CHAPTER 1 PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA Southwestern Central Asia (Figure 1.1), as defined by the Caspian Sea to the west, the Kopet Dag mountains and the Iranian plateau to the south, the Pamir, Hissar, and Tien Shan mountains to the east, and the Kazakh steppe to the north, is an area marked by significant physiographic boundaries and vast differentiation in environment and physical setting. The convergence of the Kopet Dag mountains and the Kara Kum desert in southern Turkmenistan may be seen as “the northeastern frontier of the ancient Near East” (Tosi 1973-74), representative of a “hard frontier” (Sherratt pers. comm. 2001) characterised by dryness. The arid nature of this borderzone area served to promote the shift from a Mesolithic hunter/gatherer lifestyle to the Neolithic agropastoral adaptation. There was little potential for variation in subsistence regime to support the growing population of the area. Southern Turkmenia also functioned as a corridor to the steppe zone further north. The idea of southern Turkmenia as a borderzone and “hard frontier” holds archaeological implications for the migration to, and the peopling of, southwestern Central Asia, the subsistence regime of the region, its origins and adaptations, as well as the contact and communication networks between Turkmenia, via the Iranian plateau, to Mesopotamia in the west and the Indus valley in the east. The environmental zones of southern Turkmenia, which encompassed distinct physical settings and supported characteristic flora and fauna include: the mountains and upland intermontane valleys, the piedmont or foothills, the alluvial plain, and the desert. It is crucial to lay out a broad environmental framework in order to facilitate a comprehensive understanding of the archaeological cultures of southern Turkmenia. The subsistence systems, mode of production, size, and mobility of population were, in large part, determined by environmental factors. The geography, hydrography, and climate, all in the broader confines of the palaeoenvironment, differed significantly from those in place today. With this rationale, I will discuss each environmental zone individually after a brief overview of the hydrological forces that shaped the region. Finally, and perhaps most importantly, the continuing controversy surrounding the palaeo-climate will be addressed.

aerial photos reveal a multitude of dry river courses...” in southern Turkmenistan (Larson 2000, 1), evidencing the wide variation in the region’s hydrology over time. “Though our maps depict them as such, rivers are by no means static entities. Fluctuations in the courses of rivers...and even radical shifts in the location of deltas...on geological timescales, tend to be the rule, not the exception. These course shifts...have subsequently had a resounding impact on the populations dependent on the rivers for survival” (Larson 2000, 5). In the upland intermontane valleys of the Kopet Dag, there was sufficient rain water to support cultivation without irrigation. However, the lack of archaeological data from this zone makes it impossible to conclude whether or not there was prehistoric cultivation of domestic species occurring, or if the land was otherwise used to pasture domestic herds of sheep or goats (Kohl 1984, 33-34). Wild plant species growing in the mountain areas included oak, juniper, maple, hackberry, fruit trees, deciduous forest, and ephedra (Kohl 1984, 32; Atamuradov 1994, 63). At certain times of the year, purely rain-fed agriculture was possible in the intermontane valleys, but in the piedmont, lowland alluvial plains, and marginal desert areas, the necessity of irrigation canals, damming of rivers, and exploitation of available groundwater shaped the agricultural regime (Kohl 1984; Harris and Gosden 1996). Mountains The mountains of Central Asia are focal points of tectonic activity. In particular, southern Turkmenistan is subject to frequent seismic activity caused by the subsidence of the Iranian plateau (Persian plate) beneath the Asian continent (Turanian plate) at the Kopet Dag fault front, resulting in the uplift which created the mountains (Kohl 1984, 26; Babaev 1994, 20). The Kopet Dag form part of the larger Turkmeno-Khorassan plicative system, which is itself a segment of the Alpine-Himalayan orogenic belt. It is an anticlinal range characterised by the faulting and warping of Mesozoic, Palaeogenic and Neogenic rocks. Late Quaternary tectonic activity caused a general tilting and sinking toward the northwest of the Turanian plate, thus causing the rivers of the Turcoman plain, including the Sumbar, Chandyr, Atrek, Tedjen and Murghab, to flow in a west-northwesterly direction toward the Caspian Sea (Marcolongo and Mozzi 1992). The Kopet Dag range is marked by extremely sharp and linear fault-line scarps, consistent with continuing post-Neogenic uplift and neotectonic activity. Fauna of the mountain zone were bezoar goat, wild sheep, boar, otter, snow leopard, lynx and lion (Kohl 1984, 32).

Hydrology In order to understand these environmental zones in the correct context, an understanding of the relatively unstable hydrological regime, and its changes through time, is essential. Throughout history, water has remained the primary force affecting the climate, the vegetation, and the suitability of particular areas for human and animal habitation in southwestern Central Asia. “Satellite and

The Maly and Bolshoi Balkhan systems, located northwest of the Kopet Dag range, are considerably lower and

1

PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA

Figure 1.1 Southern Central Asia 2

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY orogenically distinct. They separate the Kara Kum desert from the Ust-Yurt plateau to the north. The Maly Balkhan is a small anticlinal range stretching from west-southwest to east-northeast and formed by Cretaceous and Tertiary sediments. It is demarcated on both north and south by depressions that connect the Caspian lowlands with the Kara Kum desert (Babaev 1994, 8). The Bolshoi Balkhan is a structurally more complicated mountain system. “Geologically, it is a diffuse anticline with its core comprised of Jurassic sediments, and its limbs, of Cretaceous ones...it is stretched latitudinally and is bordered in the north by a rocky, sometimes vertical cliff, while in the south it forms a steep slope, and in the west gives two offshoots. The southern-inclined surface of the Bolshoi Balkhan is dissected by numerous ravines. The eastern part of the range is lower and its northern slope, dissected by short ravines, is less steep than the southern one” (Babaev 1994, 8). There is archaeological evidence for the prehistoric occupation of cave sites in the Bolshoi Balkhan. Djebel, Dam Dam Cheshme I and II were focal points of excavation for Soviet researchers and will be discussed in detail in chapters 2 and 3.

untempered reality the further west one proceeded along the piedmont strip”. The Darreh Gaz Atak (central piedmont), which is a large upland plain, stretches from Anau to Kaakha and the watershed of the Ab-i Archingar (Kohl 1984, 31). The Atak-i Kelat (eastern piedmont) encompasses the area from Dushak to the district of Chaacha, and was once an alluvial plain, previously watered by a now dry extension of the Tedjen river (Kohl 1984, 31). Alluvial Floodplain From west to east across the region, the riverine flow from the Kopet Dag piedmont zone, the western Hindu Kush, and the Hissar toward the Kara and Kyzyl Kum deserts and the Bactrian plain created the lowland alluvial plains of the Meshed and the Gorgan, the Tedjen and Murghab, the Zeravshan, the Kashka Darya, and the Amu Darya (ancient Oxus) and Syr Darya (ancient Jaxartes) rivers. This alluvial plain is “the morphological counterpart of a subsiding basin (named the Turcoman trough), mainly filled with the alluvial deposits of major rivers...” (Marcolongo and Mozzi 1998, 9). Three concentrated areas of archaeological significance in the alluvial plains of Turkmenistan were identified. The first, and least well-documented, area is the lowland plain of the Meshed-Misrian and the Gorgan, situated near the Caspian Sea. Further east is the oasis that existed between the Tedjen and Murghab rivers, termed the “little Central Asian Mesopotamia” (Kohl 1981, xi). Easternmost is the area encompassed by the Amu Darya and Syr Darya rivers, or the “large Central Asian Mesopotamia”. Both were extremely fertile river oases in antiquity and contain numerous archaeological sites of various time periods. The lowland plains and river deltas were fertile with the adaptation of irrigation agriculture, despite having a relatively low amount of rainfall in comparison to the piedmont and mountainous zones. The climate was extremely hot in summer and allowed for cultivation throughout the mild winter. The vegetation was made up primarily of tugai thicket, grassland, marsh, and tree species of poplar, maple, ash, and elm (Kohl 1984, 32). The dense tugai thickets or forests formed at the edges of reedy marshland, along riverbeds and in delta areas. The tugai vegetation incorporated a dense woody canopy of Euphrates poplar and wild olive with a lower level and extremely dense thicket of willow, tamarisk, and smaller shrubs (Hiebert 1994, 7; Walter and Box 1983, 95-97). The fauna of the alluvial plains and the Caspian lowlands included deer, tiger, boar, lynx, hyena, cheetah, duck, rabbit, rat, mice, reptiles, and amphibians along with a wide variety of bird species, pheasant, flamingo, and swan (Hiebert 1994, 7; Kohl 1984, 32).

Piedmont Zone The piedmont zone is characterised by a series of coalescent alluvial fans created by the streams that run out of the Kopet Dag. These streams have an entirely seasonal nature, with peak flow during the spring and autumn (Marcolongo and Mozzi 1992), rendering the piedmont cultivable with the employment of irrigation agriculture. The possibilities for irrigation agriculture, crop-growing winters, use as pastoral land, and easy access to both the upland and lowland areas for exploitation of the various wild plant and animal resources made the piedmont an ideal environmental niche to inhabit. The piedmont is characterised by fertile loessic soil which supports semi-savannah vegetation, meadow grass, wormwood, juniper, oak and pistachio; animal species included gazelle, saiga antelope, onager, roe deer, snake and steppe tortoise (Kohl 1984, 32). In order to maintain consistency with the existing literature regarding geographical areas of the piedmont, Kohl’s terminology (after Berdiev) will be employed. It should be noted, however, that in later sections of this thesis, southern Turkmenistan will be separated into three geological zones which do not correspond to Berdiev’s or Kohl’s piedmont watershed areas, although the terms western, central and eastern will be used with geological definitions provided. Berdiev separated the piedmont into three sections based on the respective amounts of watershed. The Akhal Atak (western piedmont) ranges from Kyzyl Arvat to Anau. It is relatively dry in comparison with the central and eastern piedmonts today, although modern satellite imagery shows evidence of heavy riverine sedimentation, which suggests not only that the area was once well-watered, but also that there is a high probability that archaeological sites in this area could be buried under alluvial deposits (Kohl 1984, 3031). There are fewer streams that feed this section of the piedmont, and as Kohl (1981, xi) comments, “The desert to the north always was a much more immediate and

Desert Beyond the river deltas and alluvial plains lie the Kara Kum and Kyzyl Kum deserts. The sands of these deserts were created by the deposition and subsequent windblown erosion of alluvial sediments from the palaeoAmu Darya, palaeo-Murghab, and palaeo-Tedjen rivers

3

PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA

(After Kosarev & Yablonskaya 1994, Atamuradov 1994, Larson 2000)

Figure 1.2 Caspian Sea Transgression Regression Events (Kohl 1984, 26). The topography of the area is a combination of sand dunes and ridges, as well as intermittent takyr zones and salt flats. There are two types of sand dunes: moving barchan dunes and stable dunes with seasonal vegetation (Hiebert 1994, 8). The takyr are clay flats created due to the accumulation of dry, elutriated alluvium in depressions. They form smooth, thin, hard layers which are cemented on the surface by calcium carbonate and are the result of rapid drying (Gerasimov 1978, 320, Fig. 1; Kohl 1984, 26). These alkaline soil formations contain algae and lichen, and because of their impermeability, they collect water in small depressions forming periodic oases. They become vegetated after runoff from spring rains and can be exploited by digging temporary wells. Hiebert hypothesises that the takyrs held previous economic importance in antiquity as sources of clay and gypsum (1994, 7). Kohl suggests (1984, 26) that these takyr, which formed along the courses of ancient rivers and their alluvial fans, can act as possible indicators of archaeological sites. This idea stems primarily from the assumption that settlement occurred around these river delta areas, but also that the population gradually shifted southward toward the Kopet Dag as the alluvial fans were overtaken by desert due to increasing aridification. Gerasimov (1978, 322) supports the idea of the takyr as hydrological indicators and states that they provide evidence of alluvial drainage in what are now entirely arid expanses.

Regarding the archaeological sites of the Kara Kum desert, Kohl interestingly suggests that “...several of the Early Neolithic Djeitun sites, including Djeitun itself, were situated not exclusively along seasonal streams flowing down from the Kopet Dag, as has been assumed, but on the edge of terminal swamps created by the Tedjen” now lying in the Kara Kum desert. An unknown number of archaeological sites in the Kara Kum, which existed along the ancient palaeo-river courses, are presumably obscured by the shifting barchan dunes. Those that have been excavated, either by Soviet or western research teams, will be discussed more fully in Chapters 2 and 3. Due to the lack of significant vegetation to retain solar heat, the desert climate, although extremely hot in summer, can also become quite cold especially at night, in winter (Orlovsky 1994, 39-40). The desert vegetation is very sparse, but does include shrubs (saksaul, tamarisk, calligonum, and ephedra), semi-shrubs, herbaceous plants, sedges, and desert acacia. Palaeoclimate and Palaeohydrography A discussion of the palaeoclimate and palaeohydrography of Central Asia necessitates a preface that there has historically been marked debate in these two research areas, and more specifically, controversy regarding the relative aridity of the region during the Holocene. Larson (2000, 14) aptly observes that there is a rough academic 4

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY consensus surrounding climate fluctuation, changing sea levels, and tectonic activity, but that there is heated debate as to the underlying causes of the various events and cycles. The following is an attempt to summarise, and to an extent to synthesise, the range of interpretations.

desert” (1978, 323), rather than akin to the truly humid conditions prevalent in the forests of the Eurasian steppe. Gerasimov supports his argument with evidence for a complicated network of Bronze Age irrigation canals in the present day hyper-arid desert areas by noting the hydrological instability of the region and the multiple course shifts of the major Central Asian rivers over time. In the opposite camp, among those who favour the idea of more pluvial periods during the Holocene, Mamedov presents a convincing argument supported by geomorphological and archaeological evidence from the inner Kyzyl Kum desert. He stipulates that Central Asia, in the early and middle Holocene, was in fact much more moist than it is today (Mamedov 1980, 170-171; Kohl 1984, 25-26). Mamedov gives the existence of the following factors as evidence: proluvial and lake sediments in large depressions of the Kyzyl Kum desert, carbonate formations in the soil and bedrock indicating the previous existence of larger amounts of fresh ground and surface water, the presence of freshwater organisms in Holocene deposits, the increased flow of the Zeravshan river during the prehistoric period indicated by its now altered meanders, the formation of soils recognised in late Holocene deposits due to their thickness, high density, and high carbon dioxide carbonate content, the archaeological excavation of numerous prehistoric settlements in the Kara Kum and Kyzyl Kum deserts as well as on the now arid Ust-Yurt plateau, and finally, zooarchaeological and archaeobotanical findings indicating a prehistoric forest or forest-steppe environment during the early and middle Holocene (Mamedov 1980, 170-171; Kohl 1984, 25-26).

It is generally agreed that as the Oligocene period ended approximately 23 million years ago, the Tethys Sea covered only the lowest lying portions of western Central Asia, or much of what is now modern day Turkmenistan. During the Miocene and Pliocene, intense tectonic activity caused the remnants of the Tethys Sea to recede into the proto-Caspian basin. The first time that this area of western central Asia “was completely liberated from water” (Atamuradov 1994, 55) came during the early Miocene (approximately 7 million years ago). From that point onward, the area has been subject to a series of at least 11 major transgression and corresponding regression events (Figure 1.2), including the Sarmatian, Akchagyl, Apsheron, Baku, Khazar, Khvalyn, Mangyshlak and NeoCaspian, over the past 7 million years (Kosarev and Yablonskaya 1994; Atamuradov 1994; Larson 2000, 14). During the mid-Pliocene (approximately 3 million years ago) the Caspian and Black Seas separated for the first time, only later to reconnect along with the Aral Sea, during the Akchagyl transgression (Pliocene/Pleistocene boundary, approximately 2 million years ago). By approximately 95,000 years ago, at the end of the Pleistocene, Central Asia had taken shape roughly as it remains today. It is generally accepted that there were more pluvial periods during the Pleistocene (Kohl 1984, 25), but beyond that assertion, there are a variety of postulated theories concerning the Holocene. These arguments fall into two basic camps: the first proposing an intensely arid Holocene, and the second arguing for pluvial periods interspersed with xerophytic periods.

Favouring a pluvial Holocene becoming gradually more arid from approximately 4,000 BP, Dolukhanov argues that during the entire period between 7,000-2,000 BC, the region was much more moist than it is today (Dolukhanov 1981b, 359-360; Kohl 1984, 25). He substantiates this claim by explaining that the palaeoclimate was closely linked to the Lower Khvalyn (40,000-70,000 BP) and Upper Khvalyn (10,000-20,000 BP) transgressions of the proto-Caspian Sea, the causes for which are most likely a combination of three possible factors (1986, 122-123). Firstly, Dolukhanov links the transgressions and regressions to tectonic activity that characterised the area. He also cites Zubakov and Borzenkova’s work (1983, 118-119) which focused on changes in temperature and precipitation as causal factors. Finally, Dolukhanov cites Fedorov (1978) and Kvasov (1975) in postulating that the Lower and Upper Khvalyn transgressions were most likely also due to shifts in the river catchment areas caused by changes in polar ice sheet glaciation. Marcolongo and Mozzi (1998, 3; Ehlers 1971; Grosswald 1980; Varushchenko, 1980) agree and state that during the Lower Khvalyn, “...a continuous polar ice sheet did not allow the flow of fluvial waters to the north. This situation led to the creation of a series of pro-glacial lakes in the sub-Arctic and Siberian regions, whose waters ran ‘in cascade’ to the Aral Sea through the Turgai channel; from there they

These theories are vital for the archaeological interpretation of the subsistence regime of the region. In particular, the assumption of a wetter Neolithic period is consistent with the evidence for early plant domestication. Rain-fed agriculture practised at Jeitun during the final 7th and 6th millennia BC, supplemented by ancient irrigation canals were together able to support settled populations in an area which is today climatically stressed. The following are variations on the theories of climatic change (or lack thereof) during the Holocene. In the first theoretical camp, Lisitsina argues that the climate of Central Asia during the early and middle Holocene was similar to that of the present day, characterised by intense aridity (Lisitsina 1978, 189-193; Kohl 1984, 25). Gerasimov (1978, 324) asserts that pluvial periods were linked solely to the glacial epochs, and that increased wetness during the more recent Holocene could not have occurred due to the lack of glaciation during that period. He defines pluvial periods as “...epochs when the surface water was more abundant than at present in regions which were still nevertheless 5

PHYSICAL ENVIRONMENT OF SOUTHERN CENTRAL ASIA flowed to the Caspian Sea through the Uzboi channel and finally, to the Black Sea”. The hypothesis that the transgressions were caused by shifts in river courses due to polar glaciation is supported by the dating of marine sediments from the Caspian Sea (Kaplin et al. 1977; Dolukhanov 1986, 123). The Lower Khvalyn deposits date to between 65,000-35,000 BP, corresponding to the early Valdai or Wurm and Weichselian glaciations. During the Lower Khvalyn, the level of the proto-Caspian was 78 metres above its present-day level of 28-29 metres below sea level, creating a lagoon or estuary environment covering most of western Turkmenia (Dolukhanov 1986, 123). The Mangyshlak regression, which occurred between the Lower and Upper Khvalyn transgressions, was a period of extreme aridity. The proto-Caspian Sea level dropped to 58-50 metres below present day sea level and palynological data shows evidence of desert and semi-desert vegetation overtaking the area (Maev and Maeva 1977; Fedorov 1978, 120-121). The Upper Khvalyn transgression deposits date to between 20,00010,000 BP, corresponding to the late Valdai glaciation (Dolukhanov 1986, 123). The later transgressions of the Neo-Caspian (also referred to as the Atlantic period or Hypsithermal) again caused the Caspian Sea level to rise between 8,000-7,000 BP, with subsequent high levels between 6,700-6,400 BP and 3,400-3,000 BP (Maev and Maeva 1977; Kaplin et al. 1977; Dolukhanov 1986, 123).

During the above periods (as well as at other times), the increased flow of the Amu Darya and Syr Darya rivers caused the transferral of water from the Aral to the Caspian Sea via the Uzboi channel. There is, unsurprisingly, also substantial controversy surrounding the formation and activity of the Uzboi river. The various points of view are best summarised in Larson’s (2000) recent work on the subject. To summarise, it appears that Dolukhanov convincingly demonstrates that Central Asia was more moist, and therefore less continental, in climate during prehistoric times than at present, despite the continued controversy surrounding the causal factors. These indications combine to create a picture of the present day desert zones of Central Asia as having once been completely different ecological and environmental zones which would have been extremely suitable for early plant and animal domestication, including both rain-fed and irrigation agriculture as well as stockbreeding. The above arguments supporting the theory of a wetter Neolithic period are congruous with the archaeological evidence for such a settled subsistence regime.

6

CHAPTER 2 HISTORY OF ARCHAEOLOGICAL INVESTIGATION Building on the broad environmental framework for the archaeological study of prehistoric Turkmenia, it is necessary to outline the history of archaeological excavation in the region. In this way, I will illuminate the continuity and discontinuity of Soviet and foreign research and the effects of archaeologically artificial modern-day political boundaries on the discipline. The main theoretical premises will be incorporated into the discussion, especially the shift to a predominantly Marxist viewpoint which occurred in tandem with the creation of the Soviet Union subsequent to the Russian Revolution. The current Soviet and foreign research agendas will be reviewed, as well as future research opportunities in the region. In the second section of this chapter, I will sketch a broad overview of the archaeological chronology of the region, the sequence from Mesolithic to early Aeneolithic times.

Archaeological investigation in the modern day territory of Turkmenistan began during the 1880s. The initial stage of activity was marked by rampant plundering of sites for the international art market as well and the use of ad hoc or haphazard excavation techniques. The sites of Merv and Anau were subsequently investigated more scientifically by V. A. Zhukovsky at Merv in 1890, and General Komarov at Anau north mound shortly thereafter (Kohl 1984, 17). Komarov excavated a substantial trench, slightly offset from the centre of the mound but bisecting it, while looking for the remains of Alexander the Great. Instead, he wound up excavating the older of the two prehistoric kurgans at Anau. This substantial trench was later cleaned out by the geologist Raphael Pumpelly and his team of American and German explorers during the extremely well-documented “Explorations in Turkestan” of 1903-1904 (Pumpelly 1908a, 1908b). In 1904, Pumpelly conducted extensive excavations of both the north (Aeneolithic) and south (Bronze and Iron Age) mounds as well as Anau city (Antiquity and Medieval periods) using a cultural-ecological methodology (Kohl 1984, 18). Pumpelly's exhaustive treatment of the various categories of artefactual assemblages and the continual emphasis on the natural causes for cultural phenomena put his research well ahead of its time. Pumpelly (1908a, 1908b) thus constructed the first prehistoric cultural sequence for the region, sparking the interest of, and providing future direction for, later Soviet and Western archaeological research in Central Asia. Further investigation of any significance did not take place until after the Russian Revolution, and was subsequently highly structured, wellfunded, state controlled, and oriented toward revealing the origins of communism. The Soviet approach deserves praise for providing an extremely organised formula for archaeological excavation, cataloguing and analysis of artefactual assemblages, and the reporting and dissemination of results.

The overwhelming majority of references I will refer to are original Russian texts which I translated. I cannot stress strongly enough the monumental nature of the task of translation, compilation, and synthesis from Russian to English, especially because I was not a Russian speaker when I embarked on this thesis. The vast majority of these Russian articles, monographs, and textbooks have never been read or synthesised by a western scholar. Interestingly, a large proportion of the Russian literature has not been synthesised in any detail by Soviet scholars in Russian or English either. Certainly there are basic overview texts in Russian, but often these omit much of the detail of the original research reports or publications. I believe this is due, in part, to the intense rivalry and lack of collaboration between academicians from the respective branches of the Academies of Science in Moscow and Leningrad, and the unofficial governance by the Moscow and Leningrad branches over those of the former republics. I see a secondary reason for the historical lack of Soviet synthesis (although this situation is changing with the dissolution of the former Soviet Union) as prescribed and predetermined by the Soviet formulaic system of conducting archaeological research and interpreting archaeological data. This system required investigators to “plug” an archaeological site into a three-tiered hierarchy of local variant, archaeological sub-culture, and ultimately a designated “group” of archaeological cultures. This practice often encouraged comparison with “like”, but in the same light, discouraged (or more passively, did not foster) comparison with “unlike”. I see this as a potentially causal factor in the study of archaeological sites or prehistoric cultures as isolated insular entities.

Archaeology in the Turkmen Soviet Socialist Republic during the Soviet period began soon after the demarcation of the political boundaries of the republics in 1924 with the discovery and excavation of Namazga Depe by D. D. Bukinich (Kohl 1984, 18). Bukinich later went on to find cultural layers 7 metres beneath the current floodplain surface level near the site of Anau, suggesting that heavy sedimentation had potentially masked archaeological sites in the vicinity of the alluvial fans flowing northward across the piedmont from the Kopet Dag mountains. In 1925, an ethnographic-archaeological association was created in Ashkhabad, but was soon made redundant through the creation of an archaeological branch of Turkmen'kult, the Institute of Turkmen Culture.

7

HISTORY OF ARCHAEOLOGICAL INVESTIGATION Turkmen'kult together with Turkomstaris, the Turkestan Committee for the Preservation of Monuments of Antiquity and Art, organised expeditions to archaeological sites in the piedmont zone of the Kopet Dag, the Murghab and Tedjen oases, and along the Amu Darya river during the 1930s and 1940s until the end of World War II (Kohl 1984, 19). The well-renowned Soviet archaeologists, A. A. Marushchenko and A. F. Ganyalin, conducted excavations at Old and New Nisa, Altyn Depe, Ak Depe, and Anau during this period (Kohl 1984, 19). The final major pre-war push in Central Asian archaeology was the formation of the Khoresmian Archaeological Ethnographic Expedition by S. P. Tolstov under the jurisdiction of the Soviet Academy of Sciences in 1937. The project was later headed by M. A. Itina and is currently under the supervision of L. M. Levena of the Russian Academy of Sciences in Moscow. As Kohl recounts, “this project pioneered numerous field techniques in Central Asian archaeology, including settlement pattern studies, aerial reconnaissance and photography, and the use of mechanised digging equipment” and was extremely successful at training local archaeologists and producing doctoral dissertations (1984, 20).

cemetery of Parkhai II in 1977 (Kohl 1984, 20). V. I. Sarianidi went on to excavate at the Bronze Age sites of Ulug Depe and Khapuz Depe in the late 1960s. The Institute of Archaeology of the Soviet Academy of Sciences of Moscow and the Institute of History of the Turkmen Academy of Sciences created the Margiana Archaeological Expedition (MAE), with V. I. Sarianidi as director, in 1972. Sarianidi's team carried out an extensive survey in Margiana, revisiting the previously excavated sites of Auchin Depe and Takhirbai and also discovering new sites including Gonur 1 in 1974. In 1978, the decision was made to split the Margiana Archaeological Expedition into two parts, one led by Sarianidi operating in the southern and eastern area of Margiana at Togolok and Gonur, and the other led by I. S. Masimov of the Turkmen Academy of Sciences working to the north of Sarianidi's project at Kelleli (Hiebert 1994, 17). O. K. Berdiev, who died in a car accident at the age of 38 in 1973, deserves special recognition for his intensive and broad-ranging research on the Jeitun Culture and the agricultural Neolithic of the Kopet Dag piedmont. Still unsurpassed in his efforts to carry out a comprehensive survey of the Neolithic piedmont sites from Bami in the west to Meana-Chaacha in the east, Berdiev began his fieldwork with the doctoral dissertation entitled “The Neolithic period of southern Turkmenistan” (Berdiev 1963b). Between 1963 and 1973, Berdiev excavated and published the results of work at Bami (1963a), Togolok (1964c), Chagylly (1964a, 1966), New Nisa (1965), Anau (1971b), Chakmakli (1968a, 1968b, 1968c, 1969), Pessedjik (1968c, 1973), Chopan (1968b, 1972a), Mondjukli (1972c, 1972d), and a variety of general works on the Neolithic of southern Turkmenistan (1963b, Berdiev and Attagariev 1967, 1968b, 1970, 1971a, 1971c, 1972b, 1976). During excavations at Pessedjik in 1967, Berdiev discovered a large fresco, similar to those found at Çatal Höyük in central Anatolia. Berdiev published a monograph in Russian entitled “Ancient agriculturalists of southern Turkmenistan” (1969) synthesising his work up to that point. Descriptions of Berdiev's excavations and methodology can be found in Chapter 3.

Post-war, IuTAKE (Southern Turkmenistan Complex Archaeological Expedition) was organised by M. E. Masson in 1946. The IXth Brigade of IuTAKE, led by A. P. Okladnikov, worked extensively on Palaeolithic and Mesolithic sites on the Krasnovodsk peninsula and in the Bolshoi Balkhan range between 1947-1952. Okladnikov excavated the cave site of Djebel, which was later reinvestigated along with the other Bolshoi Balkhan cave sites of Dam Dam Cheshme I and II by G. E. Markov of Moscow State University in the 1970s (Okladnikov 1956; Markov 1966a, 1966b). IuTAKE sponsored numerous research projects between 1952-1965. The most notable of these were B. A. Kuftin's excavations at Namazga Depe in 1952, published posthumously in 1956, which established a comprehensive Aeneolithic to late Bronze age chronological sequence (Kuftin 1956; Kohl 1984, 20). In 1947, the VIIth Brigade of IuTAKE, directed by M. E. Masson and G. A. Pugachenkova conducted fieldwork at Altyn Depe, which was later continued intermittently by S. A. Ershov and A. F. Ganyalin from 1951-1961 (Kohl 1984, 20). From 1954-1956, M. E. Masson's son, V. M. Masson, discovered and excavated the Iron Age Yaz Culture complex in Margiana. Continuing in his father's footsteps, the younger Masson went on to direct the XIVth Brigade of IuTAKE with V. I. Sarianidi and I. N. Khlopin, excavating nine sites in the Geoksyur oasis of the Tedjen, at Kara Tepe, and at the Neolithic site of Jeitun near Ashkhabad (Kohl 1984, 20). Further work was also undertaken by the XIVth Brigade of IuTAKE at Altyn Depe from 1965-1967, after which time Masson became the sole project director. Masson continued actively at Altyn Depe until 1978, and has returned intermittently through the present (Kohl 1984, 20; Masson 1988). I. N. Khlopin, together with his wife L. I. Khlopina, turned his efforts to fieldwork in the Sumbar valley, discovering the early Bronze Age

In the early 1970s, foreign academics again began to take an archaeological interest in Soviet Central Asia. C. C. Lamberg-Karlovsky of Harvard University and Maurizio Tosi of the Italian Institute of Middle and Far East Studies (IsMEO) in Rome, began speculating about the interaction between the Iranian sites of their then current research, and those to the north of the Iranian plateau in Turkmenistan. Lamberg-Karlovsky had worked at Tepe Yahya in southeastern Iran while Tosi had carried out extensive excavations at Shar-i Sokhta in Iranian Seistan during the early 1970s. Lamberg-Karlovsky and Tosi described parallels between the sites circumscribing the Iranian plateau (Lamberg-Karlovsky and Tosi 1973; Lamberg-Karlovsky 1975), lending strength to Masson and Sarianidi's prior suppositions of the possible 8

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

existence of trade networks connecting the two areas (Masson and Sarianidi 1972). This initial research provided the background for both Lamberg-Karlovsky and Tosi to later organise academic collaboration with Soviet archaeologists and to conduct fieldwork in the Murghab delta.

delta was carried out by Tosi and his team of archaeologists and geologists including Bruno Marcolongo of the Institute for Applied Geology at Padova, along with Gubaev and Koshalenko from 19901995. This study, groundbreaking in its use of satellite and other remotely-sensed images to detect ancient irrigation features and locate new archaeological sites, resulted in the publication of “The Archaeological Map of the Murghab Delta: Preliminary Reports 1990-1995” (Gubaev, Koshalenko and Tosi 1998). Tosi and Marcolongo have also conducted smaller-scale geoarchaeological studies in the Sumbar valley of southwestern Turkmenistan and the Meana-Chaacha district of southeastern Turkmenistan (Marcolongo pers. comm. 2000).

After a hiatus of more than half a century since Pumpelly's initial research in Turkestan, Soviet collaboration with foreign research institutes again began to play a role in Central Asian archaeology. After conducting archaeological reconnaissance and excavation in northern Afghanistan prior to the Soviet invasion, and in Iranian Khorassan prior to the Iranian revolution, Philip Kohl of Wellesley College conducted a study tour of Soviet Central Asian archaeological sites in 1979 and 1980 (Lamberg-Karlovsky in Hiebert 1994, xx). The invitation for this collaboration was sparked by the meeting of R. Munchaev, N. Merpert, M. Tosi, and C. C. Lamberg-Karlovsky at a conference on the origins of early agriculture in Denmark in 1977 (LambergKarlovsky in Hiebert 1994, xxvii). Excavation was impossible during these early exchanges given the political climate of the Cold War, but Kohl was able to study museum collections from previous excavations and make site visits which resulted in two of the most significant volumes written to date by a western archaeologist in English, providing a comprehensive overview of Central Asian archaeology from the Palaeolithic through the Iron Age (Kohl 1981, 1984). Following Kohl's visits, the first joint USA-USSR archaeological symposium was organised by Kohl and Lamberg-Karlovsky for the following year at the Peabody Museum of Harvard University (Lamberg-Karlovsky in Hiebert 1994, XXI). This joint USA-USSR collaboration continued under the sponsorship of the International Research and Exchanges Board (IREX), alternating between the USA and the USSR in Samarkand in 1983, Washington DC in 1986, and concluding at Tbilisi in 1988.

K. Kurbansakhatov, a prominent figure in Turkmenian archaeology today, re-opened research at Anau in the late 1970s for fieldwork related to his doctoral dissertation. In 1987, he started research at Jeitun, opening large tracts of the .3 hectare site on the edge of the Kara Kum desert 15 km north of Ashkhabad under the direction of V. M. Masson. Kurbansakhatov has since become central in organising and undertaking foreign collaboration in Turkmenistan. He is currently the director of the Department of Archaeology and Foreign Expeditions of the Institute of History of the Turkmen Council of Ministers and is responsible for one American and two British collaborations in the region. These are led respectively by Fred Hiebert of the University of Pennsylvania continuing research at the Aeneolithic to Iron Age north mound of Anau; Georgina Herrmann of University College London conducting research at medieval Merv; and the consortium of David Harris of UCL, Chris Gosden of the University of Oxford, and Mike Charles of the University of Sheffield, researching the origins of agriculture at Jeitun in the Kopet Dag piedmont. The last group has also undertaken fieldwork in the Sumbar valley and the Bolshoi Balkhan range of southwestern Turkmenistan.

Maurizio Tosi began a collaboration in 1989 in the Murghab delta to study the ancient Islamic city of Merv and its environs with A. G. Gubaev of the Turkmen State University and G. A. Koshalenko of the Institute of Archaeology of the Soviet Academy of Sciences. In 1981, Koshalenko had begun creating an archaeological map of the ancient city of Merv, a pivotal trading stop on the Silk Route. The site has sections that date from the Iron Age Yaz Culture discovered by V. M. Masson in the 1950s. Substantial excavations and mapping at Merv were also carried out by Georgina Herrmann of the Institute of Archaeology, University College London in conjunction with St. John Simpson and K. Kurbansakhatov from 1992-2000. These excavations focused primarily on the later Medieval portions of the city and yearly published field reports may be found in the journal Iran from 1993-2001. Further archaeological fieldwork and geological reconnaissance in the Murghab

The research consortium originally led by David Harris and later co-organised by Harris and Gosden became involved with V. M. Masson and K. Kurbansakhatov at Jeitun after the World Archaeological Congress in 1986 held in Southampton, United Kingdom and organised by Peter Ucko, then archaeological department head of the University of Southampton (Harris pers. comm. 2000). Masson subsequently invited Peter Ucko and Tim Champion, also of Southampton, to visit the Soviet Union in order to develop a UK-USSR collaboration in 1987. During that initial visit, the UK delegates visited the early agricultural site of Jeitun where trenches had been reopened by Kurbansakhatov under the supervision of Masson. Masson expressed his interest in expanding the project to encompass archaeometric research methods then common in the western academic community, but still unused in Soviet archaeology. In particular, the Soviets were interested in employing fine-grained

9

HISTORY OF ARCHAEOLOGICAL INVESTIGATION archaeobotanical retrieval methods such as flotation in order to substantiate their claims of evidence for early crop domestication at the site. With the initial help of Ucko, a collaboration was organised. In 1989, David Harris and Gordon Hillman, also of UCL, made an initial study trip to Jeitun, which at that point was under the Soviet supervision of Yuri Berezhkin, still directed by Masson and assisted by Kurbansakhatov. There, Harris and Hillman performed bucket flotation, obtaining the sought after archaeobotanical remains and recognising the potential for archaeobotanical research at the site. Arrangements were made to continue the collaboration, and in 1990 Harris and Hillman returned with a larger research team including archaeobotanist Mike Charles (still involved with the present project), a zooarchaeologist and a soil micromorphologist. The UCL team worked in conjunction with the Soviets led by Berezhkin and assisted by Kurbansakhatov and N. Salavyova, but carried out their own excavations in a separate pit. In 1991 Chris Gosden, present project organiser, joined the research team at the suggestion of Peter Ucko. In 1991 and 1992, Masson, Berezhkin, Kurbansakhatov, Salavyova and the Russian team continued their previous excavations while Gosden, Charles, and the British team dug a series of test pits using standard British context excavation techniques. With the break-up of the former Soviet Union and the gradual phasing out of Russian archaeologists, the British team was given increasing autonomy between 1991-1994, which eventually led to the British team, in collaboration with Kurbansakhatov, being entirely in charge of excavations on the site in 1993 and 1994. Results of the collaborative research at Jeitun were published in Harris et al. 1993, Harris and Gosden 1996, and Harris et al. 1996. An account of both the British and Soviet excavation methodologies and results from Jeitun is detailed in Chapter 3. The British team, under the direction of Gosden and Harris, in conjunction with Kurbansakhatov have since carried out one field season in the Sumbar valley in the spring of 1996, and one field season in the Bolshoi Balkhan at the cave sites of Dam Dam Cheshme I and II during the spring of 1997.

which is the Neolithic period. The Mesolithic period variants of northwestern Afghanistan and the Caspian Sea area will be reviewed because of their geographical pertinence to this study. The transition from the Mesolithic to the Neolithic will be outlined. In the analysis of the Neolithic, the focus will be the two Neolithic adaptations which create a significant Near Eastern/Eurasian archaeological borderzone in Turkmenistan: the semi-sedentary agro-pastoral subsistence system of the Jeitun Culture and the stockbreeding-hunting-fishing subsistence system of the Keltiminar and Keltiminar-related groups. Finally, the transition from the Jeitun Neolithic to the pre-Namazga Anau 1A period of the Kopet Dag piedmont will be outlined and the basic parameters of this economic and cultural adaptation covered. The Mesolithic In categorising the Mesolithic of southern Central Asia, Masson and Sarianidi (1972, 26) describe this period as a link between “the two great epochs in the history of man: the period of a food-gathering economy and the period of a production economy”. They define a series of regional cultural variants for the Mesolithic dated to between 10,000-5,000 BC for southern Central Asia. Differentiation in terms of cultural variation and geography is observed in the Ferghana valley, the Pamirs, southwestern Tadjikistan, northwestern Afghanistan, and the Caspian Sea area of Turkmenistan. A brief overview of the latter two areas will be presented. The Mesolithic of northwestern Afghanistan was wellsurveyed and documented by Vinogradov (1979; Kohl 1984, 40-41). During three field seasons, Vinogradov undertook surveys in the area to the south of the Amu Darya in which he systematically checked the border area of the takyr zone and the desert where the terminal runoff for smaller streams flowing north out of the Hindu Kush existed. In so doing, Vinogradov located over 150 areas of archaeological material which he categorised as “points” or isolated surface scatters, “stations” or what he considered to be poor sites, and “sites” defined as having a diameter of 10-30 metres. These sites were typologically dated by similarities in the artefactual assemblages with those at Tutkaul in Tadjikistan and the Caspian caves of Djebel and Dam Dam Cheshme I and II (Kohl 1984, 41). Vinogradov felt that later Bronze Age and Achaemenid sites which he discovered further to the south, testify to the gradual drying of former Bactrian streams, and thus to the necessity for the population to follow the retreating water resources to the south (Vinogradov 1979; Kohl, 1984, 41). Kohl believes, however, that it would seem strange that these northern Afghan sites would only cluster around the terminal runoff of such streams, and further suggests that other earlier sites to the south have now been covered by alluvial deposits (Kohl 1984, 41). Masson and Sarianidi reference Gari-Mar cave in Afghanistan as showing

Chronology The current study represents the first attempt by any scholar, Soviet or western, to summarise in the most comprehensive manner possible, the entirety of the archaeological literature for the Neolithic period of Turkmenia written in Russian and published in the former Soviet Union. In translating and synthesising the relevant archaeological literature, I found myself making repeated mental notes of the distinct lack of Soviet archaeological comparison for the Jeitun and Keltiminar (or Keltiminarrelated) Cultures and their respective subsistence adaptations during the Neolithic period. It is in this light that the following chronology (Figure 2.1) for the Mesolithic, Neolithic and early Aeneolithic is presented. It is my intention to provide a temporal setting and frame for the interpretative sections of this thesis, the focus of 10

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Near East and Mesopotamia

Turkmenia: Southern Super-Zone

Turkmenia: Northern SuperZone

Iran and Afghanistan

Late Bronze Age: Middle Babylonian and Assyrian, Hittite Middle Bronze Age: Hittite, Old Babylonian, Old Assyrian Early Bronze Age: Akkadian, 3rd Dynasty Ur

Iron Age: Yaz I Complex (Margiana and Bactria), Anau IV

Steppe Iron Age: Chust, Zamanbaba and Amirabad Cultures

Middle Elamite (Iran); Mundigak (Afghanistan)

Late Bronze Age: Namazga VI, Anau III Takhirbai, Auchin, Tekkem, Gonur I

Steppe Bronze Age: Tazabagyab Culture, Timber Grave Culture, Andronovo Culture Late Keltiminar Phase: Djanbas, Djebel layers 1-3, pre-Chust Bronze Age

Old Elamite (Iran), Dashli, Shortugai, Bactrian Bronze Age (Afghanistan)

Early Bronze Age: Nineveh, Early Dynastic Mesopotamia Early Bronze Age: Uruk, Jemdet Nasr

Early Bronze Age: Namazga IV, Anau III, Altyn Depe, Khapuz Depe (Geoksyur oasis) Late Aeneolithic: Namazga III, Anau II, Kara Depe, Altyn Depe, Ilginli Depe, Ulug Depe, Geoksyur Depe, Chong Depe (Geoksyur oasis) Middle Aeneolithic: Namazga II, Anau II, Kara Depe, Altyn Depe, Yalangach Depe, Mullali Depe (Geoksyur oasis) Early Aeneolithic: Namazga I, Anau 1B, Mondjukli, Kara Depe, Altyn Depe, Dashlidji Depe (Geoksyur oasis) Pre Namazga Aeneolithic: Anau 1A: Chakmakli, Mondjukli (MeanaChaacha) Late Neolithic: Jeitun Culture III: Bami, Pessedjik, Chagylly, Chakmakli, Mondjukli Middle Neolithic: Jeitun Culture II: Bami, Naiza, Chopan, Pessedjik, New Nisa, Togolok,Gievdzhik Early Neolithic: Jeitun Culture phase I: Jeitun, Chopan, Togolok

Time Period Late 2nd Millennium BC Early 2nd Millennium BC Late 3rd Millennium BC Early 3rd Millennium BC Late 4th Millennium BC

Early 4th Millennium BC

Early Bronze Age: Ubaid Period, Uruk, Choga-Mami

5th Millennium BC

Neolithic: Ubaid Period

Final 6th and Early 5th Millennium BC Late 6th Millennium BC

Neolithic: Ubaid Period

Mid 6th Millennium BC Final 7thEarly 6th Millennium BC pre-8,000 BP

Neolithic: Catal Hoyuk, Hacilar, Jarmo, Hassuna, Halaf, Samarra Neolithic: Catal Hoyuk, Hacilar, Jarmo, Hassuna, Halaf, Samarra Neolithic: Catal Hoyuk, Hacilar, Jarmo, Hassuna, Halaf, Samarra Neolithic: Catal Hoyuk, Jarmo, Hassuna, Halaf, Samarra

Middle Bronze Age: Namazga V, Anau III, Altyn Depe

Aceramic Neolithic? Mesolithic?

Late Keltiminar Phase: Djanbas, Djebel layers 1-3 Late Keltiminar Phase: Djanbas (Akcha Darya), Djebel layers 1-3, Oyukli, turquoise working begins in Kyzyl Kum and lower Zeravshan Transitional Keltiminar Phase: Djanbas (Akcha Darya), Djebel layer 4, Dam Dam Cheshme I and II layer 3, Oyukli Transitional Keltiminar Phase: Djanbas (Akcha Darya), Djebel layer 4, Dam Dam Cheshme I and II layer 3, Oyukli Early Keltiminar Phase: Djebel layer 4, Dam Dam Cheshme I and II layer 3, Oyukli Early Keltiminar Phase: Djebel layer 5, Oyukli

Tureng Tepe II, Sialk II (Iran)

Tureng Tepe II, Sialk II (Iran)

Tureng Tepe I, Sialk I (Iran)

Early Keltiminar Phase: Djebel layer 5, Oyukli

Sang-e Caxamaq, Tureng Tepe I, Yarim Tepe, Hotu, Sialk I (Iran)

Early Keltiminar Phase: Djebel layers 5a, 5-6, 6 (Mesolithic-Neolithic transition) Caspian Mesolithic: Djebel layers 7-8, Dam Dam Cheshme I and II, Kailiu, Hodja-Su

Sang-e Caxamaq, Tureng Tepe I, Yarim Tepe, Hotu, Sialk I (Iran)

Figure 2.1 Archaeological Chronology

11

Old Elamite, Hissar III, Tureng Tepe III, Shah Tepe II (Iran), Bactrian Bronze Age(Afghanistan) Proto-Elamite Shahr-i Sokhta II/III, Tepe Yahya IVB, Hissar IIIB, Tureng Tepe III, Shah Tepe II Susa, Uruk, Shahr-i Sokhta I, Tepe Yahya IVC, Sialk III, Hissar IIA, Shah Tepe III, Tureng Tepe IIA Hilmand valley sites (Iran), Mundigak and Quetta (Afghanistan) Shah Tepe III, Tureng Tepe IIA, Sialk II, Hissar IIA (Iran)

Gar-i Kamarband, Hotu, Ali Tappeh, Sang-e Caxamaq (Iran), Gar-i Mar cave (Afghanistan)

HISTORY OF ARCHAEOLOGICAL INVESTIGATION evidence for the early domestication of sheep and goats (Dupree 1967, 24; Masson and Sarianidi 1972, 32). Without absolute dating, the evidence for domesticated species from Gari-Mar may be explained as a technologically underdeveloped later Neolithic variant.

via convenient geographical corridors from the Iranian Khorassan and/or northwestern Afghanistan, or (3) a combination of the two, entailing Neolithic influence through diffusion over the local Mesolithic populations in terms of economy and culture as well as the possible movement of people onto the Kopet Dag piedmont from either northeastern Iran or northwestern Afghanistan.

The Caspian Mesolithic, most pertinent to the present study, is represented at cave sites located near the southern and eastern shores of the Caspian Sea, the Krasnovodsk peninsula, and the Bolshoi Balkhan range. Among these sites are Gar-i Kamarband (Belt), Hotu, and Ali Tappeh caves in northern Iran, Djebel and Dam Dam Cheshme I and II on the southern face of the Bolshoi Balkhan mountains, Kailiu cave on the Krasnovodsk peninsula, and Hodja-Su on the eastern coast of the Kara Bogaz Gol (Masson and Sarianidi 1972, 26). The subsistence regime for this region included hunting (onager, gazelle, goat and sheep), fishing (in the Caspian and the Uzboi river), gathering, and caprine domestication.

In addressing the first possibility, that of a direct transition from the Caspian Mesolithic to the Jeitun Neolithic, supported by Tosi (1973-1974), I feel it is important to outline the parameters of the Caspian Neolithic, especially in terms of its differentiation from the Jeitun agricultural Neolithic. The two subsistence adaptations exhibit characteristic differences in mode of production, intensity of labour, population mobility, and perhaps most importantly, facility of exit and entry. In order to be forced to leave or “exit” a Mesolithic way of life, a Mesolithic population which exploits abundant natural resources in fertile oases or forested areas requiring a relatively low intensity of labour and supporting a relatively small and mobile hunter/gatherer population, either must find itself hard-pressed to meet the demands of a growing population, or may find itself surrounded by non-indigenous Neolithic farmers practising an economically and culturally distinct and more productive way of life, or a combination, with subtle internal variations, of the two. In terms of this definition of the forces which ultimately cause “exit” from the Mesolithic way of life and subsequent “entry” into a Neolithic agro-pastoral way of life, the transition from Mesolithic to Neolithic in fertile oases and forested areas was slowed by the potential for continued exploitation of the environment without intensification of labour to support a specific population. Most fundamentally, maintenance of the status quo by Mesolithic hunter/gatherer bands was the path of least resistance and therefore continued in environmental areas suited to such a subsistence system. However, in less fertile steppe, semi-steppe, and semi-desert areas (the so-called “hard frontier” zones), the transition from Mesolithic to Neolithic was often forced or necessitated much earlier than in the naturally fertile areas, specifically due to the marginality of the environment. The Caspian region of Turkmenia was likely a fertile oasis, especially along the southern Caspian shore in what is modern-day Iran (Sherratt pers. comm. 2001). Therefore, indigenous populations may have been reticent to adopt a more labour intensive way of life if population pressure was not imminent. If the Jeitun agricultural Neolithic did evolve directly from the Caspian Mesolithic, one might expect a greater degree of similarity between the Caspian Neolithic and the Jeitun Neolithic than is evidenced in the archaeological record. Harris (1997, 9) remarks that “It took some 2,000 years for agro-pastoralism to spread throughout South-west Asia. Probably this process depended more on colonisation by farmers [the Jeitun people] than on the adoption of agriculture by foragers [the indigenous groups] because the settled agricultural populations

Djebel, excavated by Okladnikov (1956, 11-219) is considered the type-site for the Caspian Mesolithic. The lithic assemblages show a shift from the predominant pebble-tool tradition to the incorporation of geometric microliths. This development is thought to represent diffusion from more developed cultures to the west. The lithic assemblages found in the Caspian region are more advanced than contemporaneous collections from northern Afghanistan. Based on the characteristics of the stone tool assemblage at Djebel (despite the lack of a sound chronological division), Okladnikov (1966, 59-63) provides a typological separation for the early and late Mesolithic periods. Dolukhanov (1986, 124) interprets the Central Asian Mesolithic sites as hunting camps utilised for the seasonal exploitation of resources as part of a mobile hunter/gatherer lifestyle and considers it improbable that long distance migration from the Zagros range occurred during the Mesolithic (1986, 124). The site of Gar-i Kamarband (Belt cave) in northern Iran evidences late Mesolithic as well as subsequent Neolithic and Bronze Age levels. The site contains 11 occupational layers, the earliest of which dates to 6,000 BC. The Neolithic layers contained shell-tempered ceramics and the lithic industry was analogous to that found in Djebel layer 5 (Okladnikov 1956, 210). Transition from the Caspian Mesolithic to the Jeitun Neolithic There has been considerable speculation and controversy regarding the links between the Caspian Mesolithic and the Jeitun Neolithic and the peopling the Kopet Dag piedmont. Three main possibilities present themselves: (1) a direct transition, in other words, a contiguous indigenous shift from the Caspian Mesolithic to the Jeitun Neolithic, (2) the migration of Jeitun Neolithic peoples 12

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

would have tended to expand both numerically and territorially at the expense of the hunter-gatherers”. This distinction still remains unclear and certainly necessitates further research. Of the three options, (1) a direct indigenous transition from the Caspian Mesolithic to the Jeitun Neolithic, (2) the colonisation of the Kopet Dag piedmont by Jeitun people from northern Iran, or (3) a combination, with inherent and subtle variations of the two, the final possibility seems most likely. The existence of a separate and very distinct Caspian Neolithic tradition, which clearly stems from the Caspian Mesolithic, but is quite different from the Kopet Dag agricultural Neolithic, supports this hypothesis. It appears that there may have been remnants of earlier Mesolithic adaptations, including aspects of the Mesolithic lithic industry, incorporated into the new Neolithic agricultural adaptation which appeared on the Kopet Dag piedmont. This is suggestive of the peopling of the area by newcomers moving through convenient corridors either from northeastern Iran or northwestern Afghanistan, together with some assimilation of local Mesolithic population. Harris states (1997, 9) “That we are observing the diffusion of agro-pastoralism by farmers rather than its adoption by foragers is suggested by the comprehensive nature of the changes in material culture that accompany the appearance of the crops and domestic animals at Neolithic sites”. I see this certainly as part, but possibly not all, of the transitional puzzle.

assemblages from different levels at Dam Dam Cheshme I and II evidence two distinct lithic traditions, one similar to the assemblages found at Gar-i Kamarband (Belt) and Hotu caves in northern Iran, and the other similar to the Iranian Zarzian complex. Later publications (Masson 1996) fail to recognise this earlier distinction, hence the significance of such a distinction is dubious. Using the most current literature as a guide (Masson 1996), however, the Balkhan Keltiminarrelated lithic assemblages consisted of geometric microliths, large trapezes, symmetrical and asymmetrical triangles, elongated triangles, elongated lunates, and Keltiminar-type arrowpoints made on shouldered blades. The Jeitun Culture stone tool assemblages were dominated in the early Neolithic by geometrical microliths and trapezes, but in the middle and late Neolithic there was a gradual decrease in microliths and an increase in denticulate blades and sickles used for harvesting grains and cereals. Masson and Sarianidi disagree with Markov (Masson and Sarianidi 1972, 170) in his assumption that the Jeitun Culture was contemporaneous with layer 4 at Djebel and that the Kopet Dag piedmont was peopled by nomadic settlers from the Balkhan foothills (Markov 1971, 61). Masson and Sarianidi do, however, believe that there was a chronological correlation between the Jeitun Culture and levels 5 and 5a of Djebel and layer 3 at Dam Dam Cheshme I and II. More research, especially the absolute dating of secure archaeological contexts, will be necessary in order to substantiate these hypotheses. It would seem that although the Jeitun and Keltiminar Cultures overlapped chronologically, they could not have grown from a single Mesolithic origin based on the differentiation in material cultural assemblages.

Neolithic Subsistence Adaptations The subsistence systems of the Jeitun and Keltiminar Cultures during the Neolithic exhibited marked differentiation. The Jeitun peoples were settled or semisettled agro-pastoralists while the Keltiminar and Keltiminar-related groups were semi-nomadic/nomadic hunter/gatherers, fishers, and stockbreeders practising a seasonal-round subsistence system. These differences were undoubtedly due to environmental restrictions on the productive economy of each group. However, such striking differentiation in material culture exists, that it seems difficult to accept the theory that the Jeitun Neolithic and the Caspian Neolithic grew from one mutual origin, the Caspian Mesolithic. In terms of material culture, the ceramic assemblages of the Caspian cave sites, the Keltiminar, and other Keltiminar-related groups are different from the Jeitun Culture ceramics of the Kopet Dag area to the south in form, ware, and fabric. The Keltiminar ceramic assemblages consist primarily of hand-made, calcitetempered, incised, pointed-bottomed pottery whereas the Jeitun ceramic assemblages are chaff-tempered, painted, and flat-bottomed. Korobkova (Korobkova and Masson 1978) characterised the lithic industries of the Keltiminar-related Bolshoi Balkhan sites as a separate Balkhan (Keltiminar-related) group rather than strictly Keltiminar, although parallels do exist. Masson and Sarianidi (1972, 26-28) report that the lithic

In attempts to prove the geographical origins of the Jeitun Culture Kopet Dag settlers, Mellaart (1975, 194) and Kohl (1984, 45) suggest that the sites of the Iranian Khorassan, in particular Sang-e Caxamaq which was excavated by Seichi Masuda of Tsukuba University in 1971 and 1973, could provide insight into this transitional time period. The hypothesis that the Mesolithic/Neolithic transitional sites lie further to the south and west of the Kopet Dag piedmont zone prompted recent research by the British team in the Sumbar and Chandyr river valleys in 1996 as well as the Bolshoi Balkhan mountains at Dam Dam Cheshme I and II in 1997. Most researchers agree that the earliest transition to the Neolithic has not yet been found (Dolukhanov 1981b, 371). It is also theoretically possible that there is no Mesolithic/Neolithic transition evidenced in the Kopet Dag, and that the Jeitun peoples migrated from Iran, bringing with them an already established culture, basic architectural complex and subsistence regime that was further developed in the Kopet Dag during the Neolithic and subsequent periods.

13

HISTORY OF ARCHAEOLOGICAL INVESTIGATION agriculture, architectural traditions and ceramics (Dolukhanov 1986, 129). Whether communities migrated from as far afield as the Zagros is a matter of contention. More likely, taking into account the convenient sub-tropical corridor leading into southwestern Turkmenistan from northeastern Iran, was the movement of people from the Iranian Khorassan.

The Neolithic The Neolithic of southern Central Asia is more clearly defined than the Mesolithic in terms of two major geographical areas and their respective subsistence regimes. These areas were termed the southern and northern super-zones by Masson (1996, 92) and Dolukhanov (1986, 124-128). The Jeitun agropastoralists lived in the piedmont of the Kopet Dag mountains of southern Turkmenistan, the southern superzone. The Keltiminar Culture and Keltiminar-related groups practised hunting, fishing, gathering, and stockbreeding from the Neolithic to the Bronze Age in the desert, semi-desert and steppelands of the northern super-zone. Local variants existed in the Kara and Kyzyl Kum deserts, the deltaic regions of the Amu Darya and Zeravshan and the steppe and semi-steppe regions further north. Environmental characteristics and constrictions proved to be the determining factors in the existence of the Keltiminar variant, and for the continuation of the Keltiminar Culture long after agriculture and urbanisation took hold in the Kopet Dag piedmont. Dolukhanov insists that “the Neolithic population of the northern zone maintained extensive economic and cultural contacts with its agricultural neighbours in the south” (1986, 128) but that “there were no major population displacements from the south to the north or vice versa in the Neolithic” (1986, 131). Unfortunately, we are left to wonder on what these claims are based because Dolukhanov fails to substantiate either of them with archaeological evidence. The two main Neolithic variants undoubtedly had some chronological overlap, but due to the frequently deflated nature of Keltiminar sites, and thus the lack of secure absolute dating, the chronological relationship between the Jeitun agricultural Neolithic and the Keltiminar Culture remain an interesting and essential point for future research. The geographical areas inhabited by the Jeitun and Keltiminar Cultures are outlined in Figures 2.2 and 2.3.

Sarianidi separates Iran into three distinct zones in his interpretation of the origins and interconnection of the Jeitun agriculturalists of the Kopet Dag piedmont. These are (1) the north, including the Elburz and Turkmeno-Khorassan mountain ranges, (2) the west or Zagros mountain range, and (3) the south or Persian Gulf area (Sarianidi 1970, 19). The northern zone was the major focus for the comparison of the Jeitun peoples in northern Iran and southern Turkmenistan. The settlement distribution of the Jeitun Culture ranged from the Iranian plateau and the Gorgan river valley to the Kopet Dag piedmont zone, encompassing temporal differentiation between the early-, middle-, and lateJeitun Culture phases. An overview of the sites in northern Iran will explore possible similarities with the Jeitun Culture. Iranian Plateau: Sang-e Caxamaq Sang-e Caxamaq near Shahrud on the southern slope of the Elburz range was excavated by Seichi Masuda in 1971 and 1973 (Masuda 1974, 1976). The site has east and west mounds about 150 metres apart. Masuda reports that the western mound, the earlier of the two, rises 3 metres from the surrounding alluvial plain and contains five occupational horizons of mud-brick structures with red-painted lime-plaster floors, hearths, and sleeping platforms that appear to have been repaired on numerous occasions (Masuda 1974, 222). He gives the example of a house structure in occupational horizon 2 containing twenty floor layers and estimates that these floor layers may have been repaired at least once per year, thus accounting for twenty years of occupation of the house. Gupta suggests that because the western mound is almost entirely aceramic, it may represent the transition from the Caspian Mesolithic to the Jeitun Neolithic (Gupta 1979 vol. II, 49-52; Kohl 1984, 46).

The Jeitun Culture of the southern super-zone The Jeitun Culture of northern Iran and southern Turkmenistan spans the whole of the Neolithic dating from the final seventh to the sixth millennia BC. This complex shows the earliest known evidence for settled agriculture, crop domestication and animal husbandry in southern Central Asia. The shift from the Mesolithic to the Neolithic is marked by a more sophisticated lithic industry, the advent of ceramics, mudbrick villages, and a subsistence regime of semi-sedentary agropastoralism. In studying the origins of agriculture in southern Central Asia, Dolukhanov theorises that one must take into account the previous Mesolithic foraging economy, the ecological prerequisites for farming, and the demographic instability of farming which resulted in frequent migration (1981b, 141-148). He suggests that a small community from the Zagros dispersed among the local hunter/gatherer population and introduced

The eastern mound at Sang-e Caxamaq rises some 6 metres above the alluvial plain and was estimated to contain six occupational levels. Masuda reports similarities between finds from the east mound and artefactual assemblages of the Jeitun Culture in southern Turkmenistan. He also suggests contemporaneity with Yarim-Tepe, excavated by David Stronach in 1960, and with the Sialk II complex (Masuda 1974, 223). Gorgan Valley: Tureng Tepe and Yarim Tepe Evidence for Jeitun-like materials on the Gorgan plain in

14

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 2.2 Distribution of Jeitun Culture sites (after Kohl 1984)

15

HISTORY OF ARCHAEOLOGICAL INVESTIGATION

Figure 2.3 Keltiminar Culture area (after Kohl 1984) the Iranian Khorassan is reported at Tureng-Tepe (Deshayes 1967; Kohl 1984, 46) and Yarim-Tepe (Crawford 1963; Kohl 1984, 46). Deshayes (1967) provides lengthy descriptions of the ceramic assemblages from Tureng Tepe, relating them to Jeitun (but unfortunately lacking secure radiocarbon dates), Sialk I, Shah Tepe III, and Hissar Tepe III. Further details of the relationships between Jeitun Culture sites of northern Iran and southern Turkmenistan are covered in Chapter 8. After excavating the site of Yarim Tepe in the Gorgan valley, Kohl (1984,46) hypothesised that there are most likely other related sites in the region which were buried by alluvial deposits and are therefore difficult or impossible to locate.

Southern Turkmenistan: Jeitun and the sites of the Kopet Dag piedmont Jeitun, the type-site for the Jeitun Culture is situated 25 kilometres north-northwest of the modern-day capital of Ashkhabad. There are similar Jeitun Culture sites in the central and western Kopet Dag piedmont zones and in the Meana-Chaacha district (as defined by Berdiev 1969). I visited the majority of these sites either for fieldwork and excavation, or reconnaissance survey in 1997 and 1998. Berdiev separated the Jeitun Culture sites into three distinct phases in three geographical areas, inevitably with some overlap. The areas he identifies are the central

16

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

zone (early- and middle-Jeitun phases), the western zone (late-Jeitun phase), and the Meana-Chaacha district (lateJeitun phase and subsequent pre-Namazga Aeneolithic Anau 1A variants). Berdiev suggests that the Jeitun Culture started in the central piedmont, later spreading to both the western piedmont and the Meana-Chaacha district in slightly different but contemporaneous cultural variations (1964c, 276). Each of these late-Jeitun phase variations formed the basis for subsequent pre-Namazga Aeneolithic Anau 1A Cultures. The potential archaeological significance of the eastern Kopet Dag piedmont, between Anau and Tedjen, remains badly understood due to the absence of any known Neolithic sites in that area. Berdiev surmised that Neolithic sites do exist, but are covered by alluvial sedimentation and are as yet undiscovered. Kohl (1984) adopted Berdiev's terminology in referring to the three sections of the Kopet Dag piedmont as well as the Meana-Chaacha district for consistency.

archaeological stratigraphy and potential links with the related sites further west in northern Iran and the Bolshoi Balkhan and Caspian Sea areas. The sites near Kyzyl Arvat do not form a spatial cluster like those in the GeokTepe and Meana-Chaacha regions. The Meana-Chaacha region contains four sites at Gademi (not visited during the 1998 reconnaissance survey because of its proximity to the Iranian border and the necessity for special permits and armed Iranian escort guards to access the area), Mondjukli, Chagylly and Chakmakli. Berdiev suggests that Meana-Chaacha sites were settled by people from the Geok-Tepe region. In particular, the most striking difference in the artefactual assemblages between the central piedmont and MeanaChaacha district Jeitun variants is the significant decrease in the percentage of decorated ceramics in the MeanaChaacha district compared to assemblages from the Geok-Tepe region (Berdiev 1964c, 276).

The central piedmont includes the sites of Jeitun itself, Chopan, Togolok, Pessedjik, Gievdzhik (extensively documented by Korobkova 1975b; Lollekova 1982; Kohl 1984, 46, but not visited during our reconnaissance survey in 1998), Kelyata, Kantar and Kepele (impossible to locate but briefly documented by Berdiev 1971c), New Nisa, and Yarti-Gumbez (impossible to locate and undocumented). These sites are representative of the early- and middle-Jeitun phases. Jeitun is entirely attributed to the early-Jeitun phase while Chopan and Togolok are interpreted as having lower early-Jeitun phase levels covered by subsequent middle-Jeitun phase levels. Pessedjik, Gievdzhik, Kelyata, Kantar, Kepele, New Nisa, and Yarti-Gumbez are considered representative of the middle-Jeitun phase and possibly later.

Jeitun Culture Phases The early-Jeitun phase found at Jeitun, lower Chopan and lower Togolok is marked by the predominance of painted and unpainted hand-made ceramic bowls, a lithic assemblage including sickle blades and occasional geometric microliths, and other material remains including terra-cotta counters, discs (possibly spindle whorls), bone scrapers, bone piercers, turquoise, and shell beads (Kohl 1984, 48). The middle-Jeitun phase is represented in the upper levels of Chopan, Togolok, Bami 1, 2 and possibly 3, New Nisa, Pessedjik, and the lower levels of Mondjukli and Chagylly (Kohl 1984, 49). The middle-Jeitun phase is considered comparable to the Sialk I complex of Iran and is characterised by fewer painted vessels than found at sites of the early-Jeitun phase, a larger variety of types of ceramic vessels, a decrease in scraping tools and an increase in denticulate blades (Kohl 1984, 49).

The sites of the western piedmont between Kyzyl Arvat and Bakharden belong to the late-Jeitun phase (Berdiev 1964c, 276). There are two sites at Naiza and Bami and a surface scatter at the Bacha well. Naiza (which proved impossible to locate during our reconnaissance survey in 1998) and the surface scatter at the Bacha well are reported by Berdiev and Kohl to be quite small. Bami, near the modern-day town of Bami, is a large circular mound rising some 6-7 metres above the surrounding floodplain. Of all the Jeitun Culture sites in the piedmont, Bami is closest to the Kopet Dag mountains. We descended into the deep (approximately 5 metres) excavation trench originally dug by O. K. Berdiev during our reconnaissance survey in 1998. It appeared that approximately 2 metres of sediment at the bottom of the trench had caved in, obscuring the lowest Neolithic levels of the site. Time did not permit the re-opening and shoring up of the caved-in trench, hence the absence in this thesis of potentially interesting ceramic samples from the Neolithic levels at Bami. Because of its westernmost position in terms of known Jeitun Culture sites on the Kopet Dag piedmont, Bami would be an ideal location for future excavation documenting a significant

The late-Jeitun phase is found at Bami 4 and 5, the upper levels of Chagylly and the lower levels of Mondjukli. This period is marked by fewer geometric microliths, the absence of bone scrapers, increasingly naturalistic motifs on the ceramics, mudbricks and proto-blocks instead of the previous cylindrical or rounded bricks, small pieces of copper, stone door sockets, an increase in the number of denticulate blades, and the evidence of bread wheat at Chagylly (Kohl 1984, 49). The Jeitun Culture of the Neolithic period can be seen as a regional adaptation to favourable environmental conditions allowing for the early transition to a settled agricultural and a caprine pastoral subsistence regime. This complex will be further examined in comparison to the nomadic hunting and stockbreeding Keltiminar Neolithic of the northern super-zone. Dolukhanov states that the hunter/gatherers of the north maintained contacts 17

HISTORY OF ARCHAEOLOGICAL INVESTIGATION with the southern agricultural communities (1986, 121), but could not adopt similar subsistence strategies due to the more arid environmental conditions of the northern zone. This seems entirely likely. Nevertheless, the relationship between the two zones is complex and interesting.

the degree of difference and thus the potential artificial nature of these inter and intra-cultural divisions”. Discrepancies in the interpretation of the Keltiminar variants aside, there is the ever-present dilemma which seems to have become a recurring mantra in this thesis, of the lack of a secure absolute dating chronology. The absence of such a chronology has caused still-unresolved controversy concerning the contemporaneity of the Keltiminar with the Neolithic Jeitun Culture, and subsequent Aeneolithic and Bronze Age developments on the Kopet Dag piedmont. Kohl devotes an entire section in his chapter on the Keltiminar (1984, 61-64) to the problem of chronology. The generally accepted view is that the Keltiminar variants represent a period stretching from the Neolithic to the late Bronze Age and dating to the 6th-3rd millennia BC (Korobkova 1975a, 48). Kohl rightly states that the “dating of the Keltiminar culture is particularly problematic given the astratigraphic or deflated nature of most of the sites and the dearth of convincing parallels with better dated sites to the south”. He also reiterates that “the beginning of the Keltiminar culture is unclear and depends upon whether one accepts ceramic or flint comparisons as primary evidence for dating” (1984, 61).

The Keltiminar Culture of the northern super-zone “At the beginning of the 6th millennium BC when agricultural communities developed in southern Turkmenistan, Central Asia was divided into two major cultural and economic zones: the farmers and pastoralists of the south (Jeitun) and the hunter-gatherer-fishers in the north (Keltiminar, Hissar, Fergana)” (Masson 1996, 91). Tolstov identified the Keltiminar Culture in 1939 after his excavations at Djanbas 4 in the Akcha Darya delta of the lower Amu Darya river (Tolstov 1948). As the primary focus of the Khoresmian expedition, more than sixty Keltiminar sites were identified through aerial reconnaissance and/or excavated after the initial discovery of Djanbas. Markov interprets the Keltiminar Culture as an archaeological front, marginal to the “Forest Cultures” of Europe and Asia, which spread from the southern steppeland of Eastern Europe, through the Pri-Black Sea, the Caucasus and the Urals, western and southern Siberia, into Kazakhstan and more broadly into Central Asia (Markov 1971, 66; Korobkova 1981, 31).

The following overview is presented to synthesise the interpretations of the Keltiminar Culture structure. Keltiminar or Keltiminar-related sites exist in five main geographical zones: 1) the lower Amu Darya and Akcha Darya delta; 2) the Zeravshan including the Ayakagitma, Daryasai, and Makhandarya; 3) the ancient Lake Lyavlyakan area of the inner Kyzyl Kum desert; 4) the upper and lower Uzboi; and 5) the Bolshoi Balkhan mountains and the Krasnovodsk peninsula.

Let me preface the discussion of the Keltiminar with the note that I have found considerable variation in what various Soviet researchers consider to be “Keltiminar”, and contradictions concerning how many geographical variants exist and how they are differentiated. I have been told by Soviet academics on more than one occasion that in order to secure funding for “original research”, more differentiation than was actually realistic may have been stressed. Thus, “splitting” as opposed to “lumping” became the norm in terms of the vast expanse between the Kara Kum, the Kyzyl Kum and the Eurasian steppe. Only recently has Masson published an article (1996) which attempts to draw generalist conclusions, to reunite the many variants into some sort of cohesive whole. In this article, he provides a brief description of the Soviet archaeological methodology, thus shedding some light on why so many variants were identified. He states that a “three components system is usually employed for the classification and grouping of sites: the local variant, the archaeological culture, and the community of cultures”. Similarly, Markov (1971) criticises the imposed archaeological policy of defining archaeological cultures, sub-cultures and variants. Markov advocates linking archaeological cultures or sub-groups when it seems appropriate, but veers away from a “forced” representation of cultural phenomena. It seems reasonable that some researchers found it impossible to “plug” a site into the already existing framework, thus creating the necessity to identify a new variant or culture. I agree with Larson (2000, 29) in his comment that “based strictly on the literature, it is difficult to ascertain

Korobkova (1975a, 49-51) outlines three main phases of development for the Keltiminar Culture: early, transitional, and late (Figure 2.4). The early Keltiminar-phase is characterised by leafshaped arrowheads, microblades, and pointed-bottomed ceramic vessels with zigzag or waving lines ornamentation. The transitional Keltiminar-phase is marked by a stone tool industry containing microblades, scrapers, and scythes characteristic of the early Keltiminar-phase with the addition of double-sided arrowheads and the disappearance of geometric microliths. The ceramic assemblage contains thickbottomed vessels ornamented only on the upper part of the vessel walls (Korobkova 1975a, 50). The late Keltiminarphase is characterised by large double-sided arrowheads and knives. There is almost a complete absence of the typical early Keltiminar-phase leaf-shaped arrowheads and microscrapers. The ceramic vessels change in size and proportions becoming lower and wider, flat-bottomed and undecorated (Korobkova 1975a, 51). Masson (1996, 91) clarifies the existence of two main types of settlements and economies in the Keltiminar zone based on differing subsistence models and 18

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Keltiminar sites by phase Early

Transitional

Late western Kazakhstan Saksaulska 2

Akcha Darya delta

Akcha Darya delta

Djanbas 4, 5, and 12 Kunyak 1 Kavat 5 Tadjikazgan

Kavat 7 Dingilje 6 inner Kyzyl Kum lake Lyavlyakan sites

inner Kyzyl Kum Beshbulak

upper Uzboi Kyuganok 22 Balaisham 9

upper Uzboi site 178 Ortaguyu

Topograficheskaia

western Kazakhstan Agispe Saksaulska 1

lower Zeravshan Kyzylkir Bolshoi Tuzkan 11, 14, 28 Maly Tuzkan 1, 2, 3

lower Zeravshan Darbazakir 1, 2 Bolshoi Tuzkan 3, 7, 17-22 lower Kashka Darya Penkendji

upper Uzboi

Zhelgi-Zagan 3 Balaish 8 Kygunok 14-16 Akcha Darya delta Tadjikazgan 2, 3, 8 Kamishli 1 Uchtagan 2

Figure 2.4 Keltiminar sites by phase zone based on differing subsistence models and environmental constrictions. Large wooden dwellings in sedentary settlements were constructed by the huntergatherer-fishers of the lower Amu Darya and the Zeravshan deltas who had access to relatively stable food and water resources. The desert, semi-desert and steppic regions were populated by nomadic Keltiminar hunters and stockbreeders who used seasonal campsites. Evidence of this adaptation is found along the course of the Uzboi river, the Bolshoi Balkhan mountains, and in the Kara and Kyzyl Kum deserts.

Jeitun people influenced the Keltiminar and vice versa. These theories supports the growing image of Turkmenia as a true archaeological borderzone in prehistory. The Keltiminar Culture and northward Although I will focus my concluding comparisons primarily on the relations between the Jeitun Culture and the Keltiminar Culture, and the relations between the Jeitun Culture and northeastern Iran/ northwestern Afghanistan, I think it is ultimately crucial to lay a broader framework to convey just how significant Turkmenia was as an archaeological borderzone in prehistory. Turkmenia lies at a major juncture, a significant cultural boundary where traditions to the north, beginning with the Keltiminar, stretch thousands of miles to the north, northwest and northeast. Similarly, contemporaneous cultural traditions of the Jeitun Culture evidenced in the Kopet Dag piedmont, were broadly analogous in structure and economy to subsistence adaptations in Iran, Mesopotamia, the Near East, Afghanistan, Pakistani Baluchistan and the Indus valley. I am attempting to paint a picture of Eurasia tinted with two broad swathes of colour, one representing the settled agriculturalists of the south and the other representing the nomadic stockbreeders of the north. Because so much of the archaeological research during the Soviet period was limited to the investigation of individual sites as insular entities, only more recently have Soviet scholars (Masson and Dolukhanov in particular) joined western scholars (Tosi, Lamberg-Karlovsky and others), in their attempts to create broader generalisations covering large

Keltiminar Culture transition from the late Neolithic to the early Aeneolithic The Neolithic/Aeneolithic transition among the Keltiminar and Keltiminar-related variants is evidenced in layer 3 of Dam Dam Cheshme I and II, Djebel layer 4, Djanbas layer 4, and the sites of western Siberia. These are considered contemporaneous with similar developments at Shah Tepe in Iran and Ak Tepe in southern Turkmenistan (Markov 1971, 62). Markov further suggests that the Pribalkhan was a crossroads during the late Neolithic, an area of exchange in both knowledge and cultural traditions. He hypothesises that the movement of people would have occurred from the north along the Uzboi corridor, from the south along the coast of the Caspian, the Aktam and Kelkyor rivers, and from the east along the Kopet Dag piedmont (Markov 1971, 62). In this way, Turkmenia was a focal point of interaction with influence coming from three different directions. Korobkova (1975a, 48) also suggests that the 19

HISTORY OF ARCHAEOLOGICAL INVESTIGATION geographical areas during prehistory. Although I have come across many smaller-scale comparisons between areas immediately adjacent to one another, I prefer to address this issue of cultural similarity and differentiation as broadly as possible. In accordance with Christian (1998), I believe that this broad differentiation in subsistence system is caused in large part by environmental and geographical restrictions.

of an eventual further shift in economy and way of life of the northern tribes during the Bronze Age. He remarks on the eventual shift to farming and stockbreeding in the oasis zones of Central Asia, including the riverine cultures of the Uzboi and Amu Darya, the Urals and the Altai; and the steppe adaptation in Kazakhstan, and western Siberia during the 3rd millennium BC (Okladnikov 1956, 216). After this point, the “Forest Cultures” of Eurasia were increasingly contrasted with the above-mentioned Bronze Age variants of Central Asia. There was, however, no overall Central Asian synthesis in economy and culture until the Islamic period.

Interestingly, Okladnikov noted that the Pricaspian peoples were more broadly related to “tribes of the north” and suggested that the microlithic culture found in the Pricaspian area actually spread from south to north (Okladnikov 1956, 213). He later suggests that cultural innovation and technology could have spread in both directions, from the south and from the north, with the Pricaspian area as a focal crossroads. Okladnikov reports that Tolstov noted (Okladnikov 1956, 213) that the ceramics from Djanbas, the type-site for the deltaic Keltiminar variant, were similar to the assemblages of the “Forest Cultures” of Priobya (the region of the river Ob) in central Siberia. Tolstov reports on the work of Chernetsov, pertaining to the similarities between the technique of production, form and fabric of Keltiminar ceramic assemblages and those from Priobya (Okladnikov 1956, 213). Okladnikov also makes comparisons between the ceramic forms and motifs from Djanbas and the assemblages of the Andreev Lake region, the Urals, and western Siberia (Okladnikov 1956, 214); and between the vessels from Djebel with ceramic forms from Crimea, the Ukraine, the Pri-Black Sea region “Ditch Culture” to the west, and the Afanav, Minuzin, and Altai Cultures to the east (Okladnikov 1956, 215). Similarities between the Keltiminar lithic assemblages, Keltiminar arrowheads in particular, and those from the Andreev Lake region, Priaralia, the Urals, Priobya, and the Volga river region also exist (Okladnikov 1956, 215). In attempts to substantiate his theory of predominant cultural influence from south to north, Okladnikov suggests that the Pricaspian tribes were the source of innovation for the above-mentioned groups to the north. He attempts to prove this idea of influence from south to north with three major points. The first is the existence of a “solid chronology” from the late Mesolithic to the Neolithic (Okladnikov 1956, 215). This “solid chronology”, which is in fact slightly dubious, is seen as the basis for an understanding of the transition from the Mesolithic to the Neolithic and possible contact between the Jeitun farmers and the Pricaspian hunter/fisher/stockbreeders, which led to an exchange of technology and ideas. The unstated assumption in all of this is that the southern agriculturalists were more developed, and therefore it is logical that influence would have spread from south to north. However, this Marxist model of economic progression is at times too narrow to promote an accurate understanding of the causal factors for influence and dispersal of technology and ideology. The final point that Okladnikov makes, which stems from his discussion of chronology and transition, is that

Transition from the Jeitun Culture to the Anau 1A period Back on the Kopet Dag piedmont in the realm of the Jeitun Culture, Mondjukli is the only site in which Jeitun late Neolithic layers are directly overlain by Anau 1A layers. Based on the similarities of the stone and ceramic assemblages between Chagylly and Mondjukli, Berdiev suggests that the radiocarbon date obtained for Chagylly (which in 1969 was the only radiocarbon date available for all of southern Turkmenistan) is thus also relevant for Mondjukli. In this way, Masson and Berdiev related the late-Jeitun and Anau 1A phases to chronologically contemporaneous cultures in Iran and Iraq, establishing that southern Turkmenia was similarly one of the earliest areas of the development of settled farming and pastoralism (Berdiev 1969, 55). Masson and Berdiev further related the cultural adaptations of the inhabitants of Jeitun and Chopan to the inhabitants of Jarmo, and the settlement of Bami to the complexes of Hassuna and Sialk. They therefore attributed the Jeitun Culture as a whole to the 6th millennium BC and possibly the final 7th millennium BC (Berdiev 1969, 55). These dates have now been substantiated by a radiocarbon dating programme carried out by the British/Soviet collaboration and dated at the Research Laboratory for Archaeology and the History of Art, Oxford University. The pre-Namazga Aeneolithic Anau 1A Period (Kopet Dag) Kohl (1984, 65) accurately and succinctly describes the Anau 1A (or pre-Namazga Aeneolithic) complex in saying that “perhaps no prehistoric period within the Western Turkestan sequence is more enigmatic than the Anau 1A period (end of sixth to beginning of fifth millennia BC)”. The type-site for the Anau 1A complex is the basal level of Anau north kurgan originally excavated by Pumpelly. A. A. Marushchenko collected further data documenting the Anau 1A complex in the 1930s. This information unfortunately remained unpublished, and in so doing contributed to the appraisal of the Anau 1A levels at Anau as “an accidental phenomenon” (Khlopin 1963, 21; Kohl 1984, 65).

20

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 2.5 Distribution of Anau 1A sites (after Kohl 1984) 21

HISTORY OF ARCHAEOLOGICAL INVESTIGATION Putting the initial ignorance of the Anau 1A Culture aside, there has also been considerable controversy surrounding the definition, extent and sub-periods of the archaeological complex. The first contributing factor to this problem is the discrepancy between the degree of excavation of western and eastern piedmont zone Anau 1A sites. Ironically, the greatest concentration of Anau 1A sites exists in the western piedmont, but the majority of the archaeological investigation of the complex has taken place at the eastern piedmont sites. Masson and Sarianidi (1972, 50-51) acknowledge this fact and stipulate that more research is necessary to fully understand the complicated nature of the transition from the Jeitun Culture to the significantly more developed Anau 1A Culture. “The period Anau 1A represents the link between the Dzeitun horizon and the ordered chronological-stratigraphic sequence established on the basis of a model settlement, the largest of the preprotohistoric era: Namazga depe. The whole period Anau 1 has been linked with Hassuna IV-V (5,500-5,000 B. C.) [Iran], as well as with Siyalk I-II [Iran], although no radiocarbon determinations are available to confirm the association fully...It is indeed in the second half of the 6th millennium that southern Turkmenia and most of eastern Iran are definitively included in the Near East zone of interchange” (Tosi 1973-1974, 31).

this definition, Berdiev identified chronological overlap between the Chakmakli sub-period and the late-Jeitun period. The distribution of Anau 1A sites ranges from the Iranian Khorassan across the western, central, and eastern Kopet Dag piedmont zones (Figure 2.5). The most well documented examples of the Anau 1A complex are at Anau north mound in the central piedmont and Chakmakli and Mondjukli in the eastern piedmont. However, Roberta Vienco Ricciardi (1980) details the excavation of Anau 1A assemblages at Tepe Yam and a site called XA6, both in the upper Atrek river valley of the Iranian Khorassan. In the western Kopet Dag piedmont, there are also similarly lesser known sites at Beurme near the modern-day town of Bami, Ekin, Chuli, Ovadan, and Gavoutch. Kurbansakhatov excavated two large trenches criss-crossing the site of Tilkin Depe in the late 1970s, but the results of this study unfortunately remain unpublished. Kurbansakhatov and I visited Tilkin during the 1998 reconnaissance survey and collected ceramic samples for the current petrological study. In the central piedmont, there is a site containing Anau 1A levels at Koushut near the modern-day town of Kaakha (Kohl 1984, 67). Finally, in the eastern piedmont, there is one settlement in the Serakhs river delta in addition to the Anau 1A sub-period typesites of Chakmakli and Mondjukli.

There were initial attempts to divide the Anau 1A complex into regional variants, with Sarianidi (1969) defining a “Mondjukli Culture” as representative of the transition from the Jeitun Culture to the Namazga Complex. Later, Berdiev suggested (1976) that a more reasonable format for understanding the Anau 1A Culture was in terms of the central and eastern piedmont sites considered as one cultural complex consisting of earlier Chakmakli and later Mondjukli sub-periods. In creating

The current chapter provides a general overview of the history of archaeological excavation and the archaeological chronology for southern Turkmenia. Chapter 3 contains detailed archaeological descriptions for the northern and southern super-zone sites, focussing specifically on those relevant to the typological and petrographic studies which follow in Chapters 4, 5, and 6.

22

CHAPTER 3 EXCAVATION METHODOLOGY AND RESULTS Chapter 2 provided a broad-based background in the archaeology and chronological sequence of southern Turkmenia. I will use the current chapter to increase focus and illuminate the specific excavation methodologies and results from the British and Soviet excavations of Jeitun Culture sites in this study (Jeitun, Chopan, Togolok, Pessedjik, New Nisa, Gievdzhik, Chagylly, Chakmakli, Mondjukli, Gademi, and Bami), and contrast those with Keltiminar Culture sites (Djanbas, Uchashi, Ayakagitma, Kavat, and Darbazakir) and Keltiminar-related sites (Kizpara, Chaloi, Sengradji, Djoiruk, Oyukli, Mollakara, Djebel, Dam Dam Cheshme I, and Dam Dam Cheshme II) in southern Turkmenia. This discussion will also include a brief overview of the results of Japanese excavations at Sang-e Caxamaq in the Gorgan valley of northeastern Iran. The excavation descriptions are given in chronological order with respect to geographical area.

absolute dating is needed to establish where the site fits within a regional chronology (1974, 223). Masuda also recounts the existence of two special types of structures in the western mound which he interprets as “not an ordinary hearth but a holy place to keep the fire in the house” and the second as a “holy place for offering” based on the retrieval of numerous mother-goddess and zoomorphic clay figurines found in the structure (1974, 222). These types of structures are found predominantly in the top three occupational horizons. The east mound contained Jeituntype pottery “decorated in red-brown paint on cream or reddish ground with horizontal and vertical parallel lines and other geometric patterns. The body is tempered with chaff and made by the ring building technique”. Masuda remarks that the pottery found in the upper horizons is decorated with zoomorphic designs, whereas the pottery of the lower horizons is not. This is entirely consistent with the progression of decoration found in the Jeitun Culture assemblages found on the Kopet Dag piedmont. The east mound also contained complex architecture including house structures with kilns and courtyards, human remains buried with painted pottery bowls covering the skulls, zoomorphic figurines, bone implements, polished stone adzes, flint blades, geometric microliths, stone and clay cosmetic implements, conical clay objects, and square and round spindle-whorls. The west mound at Sang-e Caxamaq probably represents a small, potentially semi-sedentary farming village, at least during the Neolithic. The nature of the site is somewhat ambiguous, perhaps evidencing an aceramic Neolithic or late Mesolithic variant. It remains unclear whether there is a true Mesolithic component to the site, and thus a continuous chronological sequence spanning the Mesolithic-Neolithic transition. Only future excavation will illuminate these issues.

Samples for the petrographic study detailed in Chapter 6 were collected at Jeitun, Chopan, Togolok, Pessedjik, Chagylly, Chakmakli, Mondjukli, Djebel, and Tilkin. Jeitun is the only site that was excavated using both western and Soviet archaeological methodology, therefore, attempts will be made to describe the two methodologies in some detail and to provide a basis for their comparison using the two sets of data from Jeitun as a baseline. The descriptions of the Soviet excavation methodology and results were translated from the original Russian texts. In some cases, excavations were carried out by different investigators over extended periods of time, so varying strategies within the overall Soviet methodology will also be addressed. The most significant and potentially confusing discrepancy I encountered was the variation in the numeration of occupational horizons (also called phases, building horizons or occupational levels) from the bottom up (occasional pre-war period investigators like A. A. Marushchenko) versus from the top down (post-war period standardisation). In these cases, I make every attempt to provide a clear explanation of what is meant by the numbering system used.

Jeitun Culture sites: Kopet Dag central zone Jeitun: Soviet excavations Jeitun is considered the type-site for the Jeitun Culture (Figures 3.1 and 3.2). The site is located approximately 25 km north-northwest of Ashkhabad, the capital of Turkmenistan, and is situated adjacent to the margin of the alluvial fan of the Kara Su, a stream originating in the Kopet Dag which cuts through the southernmost dune ridge of the Kara Kum and enters the desert near the site (Harris et al. 1996, 423). It is a small mound of 0.7 hectares rising 5.5 metres above the surrounding plain with 3-4 metres of cultural material representing an as yet unclear number of occupational phases. Prior to the British collaboration at Jeitun, Soviet archaeologists employed large-scale excavation methods. The site was categorised chronologically by five occupational horizons.

Iranian Plateau: Sang-e Caxamaq Sang-e Caxamaq consists of two tell mounds, an earlier western one and a later eastern one. The artefactual assemblage from the west mound contained bone implements, flint cores, blades and microliths, obsidian blades, and interestingly, almost no ceramic material. After the first field season, Masuda reports that initially he was inclined to attribute the west mound to the aceramic Neolithic, but following the discovery of 3 ceramic sherds at the site during the second field season, he refutes the earlier hypothesis and states that more research and 23

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.1 Jeitun Culture architecture and site layout (Pessedjik 1-5) anhjd Jeitun (6) (after Ribakov 1996)

24

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.2 Jeitun Culture phases and corresponding artefactual assemblage (after Ribakov 1996)

Figure 3.3 Jeitun site layout (after Masson 1971)

25

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.4a Jeitun original site photos (after Masson 1971)

26

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.4b Jeitun original site photos (after Masson 1971)

27

EXCAVATION METHODOLOGY AND RESULTS

Under the direction of V. M. Masson, occupational horizon 1 (uppermost) was determined to be deflated and was stripped to uncover horizon 2 (Figures 3.3, 3.4a, and 3.4b). This second level revealed a settlement of approximately 30 mudbrick houses as previously described. Almost all architectural features and artefactual remains from horizon 2 were excavated and removed. Most intact ceramic vessels and other artefactual remains were subsequently taken to the Institute for the History of Material Culture of the Russian Academy of Sciences or to the Hermitage Museum, both in St. Petersburg, Russia. The Soviet excavations employed specialists in many aspects of material culture. Notably, G. H. Lisitsina (1965), a soil micromorphologist, conducted a study of the ancient environment at Jeitun. During prehistory, Jeitun was located near the banks of the Kara-Su, a stream which experienced seasonal flooding. Lisitsina notes that the Jeitun agriculturalists built dikes to keep the flood waters from dissipating into the desert, thus creating an estuary environment suitable for the intensification of farming (1965, 24). Lisitsina supervised the excavation of a trench opened to expose a soil profile in the nearby agricultural fields. Five soil layers were found including (1) cracked takyr made of slatey, compact, light grey, porous sub-clay; from 0-5 cm below the surface, (2) pale grey sub-clay of average porosity with compacted clay pellets; from 5-15 cm below the surface, (3) muddy brown compacted granular sub-clay with salt; from 1535 cm below the surface, (4) pale-grey humic granular sub-clay with charcoal and salts; from 35-120 cm below the surface, and (5) ancient alluvial sand and clay layers with lithic content; from 120-160 cm below the surface (Lisitsina 1965, 25). Lisitsina concludes, based on the results of her study, that farming at Jeitun appeared in a relatively advanced form. Because of this, she hypothesises that agriculture did not develop at the site, but that traditions were brought with the settlers who populated the Kopet Dag foothills (1965, 26). Jeitun is representative of the archaeological borderzone in which it exists, between the hunting-fishing-stockbreeding cultures of the semi-desert and steppe regions to the north, and the settled agricultural-pastoral communities to the south.

Jeitun: British excavations The original Russian determination of five occupational phases was made based on sets of floor levels encountered in an initial deep sounding, but is now under consideration as a result of a recent suite of radiocarbon dates obtained by the British research team. The site is made up of approximately 30 rectangular mudbrick houses constructed from oval proto-bricks measuring between 2025 cm in width by 60-70 cm in length (Kohl 1984, 49). The house structures, typically between 4-6 metres on a side, are of similar construction with hearths, raised areas interpreted as sleeping platforms, lime plaster floors with holes for storage vessels, painted walls, and communal yard areas possibly used for the processing of cereals, the production of household goods, and the night-time penning of animals (Masson and Sarianidi 1972; Kohl 1984, 49; Harris et al. 1996, 425). The artefactual assemblage recovered from the site includes ceramic vessels, flint blades and microliths, mortars and pestles, polished stone axes, bone points and needles, and clay counters and figurines (Masson and Sarianidi 1972; Harris et al. 1996, 425). Jeitun was located in an environmentally strategic area for the cultivation of cereal and other grain crops during prehistory. The Kara-Su, which passed by the site and disappeared in the sands of the Kara Kum, was a stream affected by the shifting barchan dunes close to the site. The movement of these dunes temporarily caused the damming of the Kara-Su and the creation of a well-watered boggy estuary environment ideal for farming (Berdiev 1965, 241; Lisitsina 1965, 24). The theory that the site was much more well-watered in prehistory is supported by Lisitsina’s archaeobotanical findings of certain tree species, including maple, poplar and saksaul (Lisitsina 1965, 24). The methodology used by Russian investigators throughout the history of archaeological research at the site differs from the stratigraphic context-based excavation techniques typically employed in British archaeology. Initially, the British research team was invited to collaborate at Jeitun with the Southern Turkmenistan Multi-Disciplinary Archaeological Expedition of the Turkmen Academy of Sciences, Ashkhabad (IuTAKE), in order to obtain archaeobotanical samples by means of flotation for AMS (accelerator mass-spectrometric) radiocarbon dating, to analyse human and animal bone samples, and ceramic and stone tool assemblages. During the final four of six field-seasons (1989-1994), however, the British investigators conducted separate but simultaneous excavations alongside their Russian colleagues at Jeitun. Due to the use of differing methodologies, I will attempt to reconcile and link both the Russian and British data.

The Soviet archaeological method used at Jeitun provides a broad framework for two occupational phases across the entire site, whereas the British methodology provides detailed insight into two chronologically distinct house structures and the spatial variation within those structures. In linking the results relevant to these two approaches, a more comprehensive picture of Jeitun as a Neolithic agricultural and pastoral settlement is constructed. The two sets of data from Jeitun also help to better correlate the general British methodology and findings with those from Soviet excavated sites in other areas of the Kopet Dag piedmont, the Meana-Chaacha district, the Kara and Kyzyl Kum deserts, the Bolshoi Balkhan mountains, and the ancient course of the Uzboi river.

The excavation methodology employed by the British research team in directing their own excavations at Jeitun was “to use fine-grained stratigraphic excavation techniques to distinguish individual features and fine layering on the site, and to tie this to a programme of 28

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

fine sieving and systematic sampling for soils, charred plant remains, bones, and dating materials” (Harris et al. 1996, 426). In order to maintain some continuity with the previous Russian excavations, the five phase terminology was retained as a reference point.

al. 1996, 429). Test pit 8 revealed no evidence of architectural features and very few material remains. The main purpose of the excavation of houses “A” and “B” was to study the lowest known architectural features on the site. After the removal of the superposed, slightly larger and offset house “A” in 1993, house “B” was dug during the subsequent 1994 field season. House “A” was found to have a basal mudbrick floor underlying numerous layers of lime plaster interspersed with sand and interpreted as successive floor coverings (Harris et al. 1996, 431). The oven and a sleeping platform were constructed against the northeastern wall of the house, and there was also a series of small post holes cut into the floor, the nature of which still remains unclear. It is possible that these were made during a later period of non-permanent residence in the disused house (Harris et al. 1996, 431). After the excavation of house “A”, the west wall was removed and a slot trench dug in order to uncover the smaller house “B”.

In 1991 the British archaeologists opened a 1.5 x 3 metre excavation area in the southeast corner of the site with the aims of compiling detailed stratigraphic information pertaining to site formational processes and to recover samples from these known contexts for botanical and petrological analyses as well as radiocarbon dating (Harris et al. 1996, 426). This trench revealed a series of lined hearth features against two superimposed mudbrick structures, from horizons 2 and 3 respectively (Harris et al. 1996, 428). A small sondage of 1 x 1.5 metres was extended from the central excavation in order to study the relatively unknown lower levels of the site. These lower layers were distinguished by house structures located farther apart and containing less artefactual material (Harris et al. 1996, 428). Simultaneously on the southern edge of the site, the second group of British excavators were re-excavating and deepening a test pit originally opened by Yuri Berezhkin in 1987. The main purpose of re-excavating the test pit (later labelled test pit 7) was to study the sand layers underlying the human occupational layers at the site (Harris et al. 1996, 428). This was, however, not possible due to the increasingly dangerous nature of the 1.5 x 1.5 metre test pit which was dug to a depth of 4.7 metres without reaching undisturbed sand. Test pit 7 was abandoned after obtaining radiocarbon dating samples from one of the lowest levels reached at the site.

During the 1994 field season, house “B” was excavated to reveal a more simple construction than that of house “A”. The existence of sterile sand underneath house “B” indicates that this was the earliest usage of the northern part of the site (Harris et al. 1996, 435). House “B” had a solid mudbrick floor overlain by successive layers of lime plaster and sand from a period of disuse. On top of this sand layer was a perhaps “temporary” floor constructed of clay with gypsum patches (Harris et al. 1996, 436). It is unclear whether this floor was laid while the house was still standing or after it had collapsed. On top of this clay and gypsum flooring were two contexts which apparently functioned as a foundation for house “A”, consisting of mudbrick, charcoal, and sand (Harris et al. 1996, 436). On top of these foundation contexts lay the actual mudbrick flooring of house “A”.

During the 1992 field season, the British excavation team dug a series of six test pits measuring 1.5 x 1.5 metres across the western and central areas of the site. These again were intended to investigate the deeper site stratigraphy in areas thought to contain occupational horizon 3 and 4 deposits (Harris et al. 1996, 428). Test pits 5 and 6 revealed structures from horizons 2 and 3. Digging was then halted in order to preserve these features in situ (Harris et al. 1996, 428). Test pits 1 and 4 uncovered and excavated features from horizons 2 and 3, but showed no evidence of features below those levels (Harris et al. 1996, 429). Test pit 4 was dug in the bottom of the already existing central excavation area from 1989 and 1990. Test pit 2 uncovered no features whatsoever and is thought to represent an open area of the site (Harris et al. 1996, 429). Test pit 7, which had been re-opened in 1991 was again extended to reveal three yard layers overlying 0.4 metres of sterile sand and even earlier floor levels separated by layers of ash and sand (Harris et al. 1996, 429).

Harris et al. comment that “The site of Jeitun has had a long and varied history of excavation by different teams employing varying techniques. Nevertheless, we believe it is possible to come to some conclusions through a synthesis of all the results obtained so far from the onsite investigations” (1996, 439). Harris et al. thus conclude that, the British excavations at Jeitun served to distinguish two or three, rather than five, main occupational phases at the site. The results of the radiocarbon dating programme show that these phases span a relatively short amount of (radiocarbon) time. The basis for this identification of only two or possibly three phases is based on the main structural and architectural levels, and the fundamental differentiation between them. That is, the hearth pits found in test pits 4 and 7 and attributed to phase 1 (if differentiation is indeed warranted) may represent “a phase of ephemeral occupation prior to the first houses being built on the site as a whole” (Harris et al. 1996, 439). Phase 2 is evidenced by the two house structures excavated by the British team. These differ from the houses of phase 1, which were excavated by the Soviets in open-plan style,

In 1993 and 1994, the British archaeologists at Jeitun conducted grid-unit excavation by stratigraphic context of two house structures (houses “A” and “B”) in the northern end of the site and dug test pit 8 in an attempt to locate the limits of the western end of the site (Harris et 29

EXCAVATION METHODOLOGY AND RESULTS

contained numerous destruction layers intermixed with rebuilt plaster floor layers. This destruction and replastering phenomenon is similar to that found at Jeitun and Chagylly (Berdiev 1968b, 9). Berdiev noted that the structures on the northern side of the mound had thicker walls, most likely for protection from the northern wind and sand blowing in from the desert. Berdiev (1968b, 10-11) provided an extremely detailed account of the orientation of the house structures, their corresponding yard areas, and the general site layout, which was slightly more complex than at Jeitun due to the occurrence of house structures with shared walls.

mainly in their density and the density of their yard layers (and debris or rubbish) across the site. As time went on, the occupation and usage of the site gradually became more intensified, with a greater number of house structures with increased activity associated with each house. Thus, the overall architectural plan became more compact. The site of Jeitun probably represents a small sedentary or semi-sedentary settlement of agro-pastoralists who cultivated cereal crops including wheat (Triticum monococcum and Triticum dicoccum) and barley (Hordeum sativum), practised hunting, and engaged in caprine pastoralism. As the typesite for the Jeitun Culture, the subsistence regime found at Jeitun is broadly representative of the subsistence regimes found at the related Jeitun Culture sites (Chopan, Togolok, Pessedjik, New Nisa, Gievdzhik, Chagylly, Chakmakli, Mondjukli, Gademi, and Bami) of the Kopet Dag piedmont. Chopan Discovered by A. A. Marushchenko in 1952 and excavated by S. A. Ershov in 1953 (Ershov 1956), and later by D. Durdiev from 1957-1959 (Durdiev 1959), A. A. Marushchenko in 1959 (Durdiev 1959) and O. K. Berdiev in 1967 (Berdiev 1968b, 1972a), Chopan (Figures 3.5 and 3.6) is the largest of all known Jeitun Culture sites, representing the early and middle Jeitun phases. The site is located 7.5 km east of the modernday city of Geok-Tepe and rises 7.5 metres above the surrounding floodplain with a total size of approximately 2 hectares. Chopan is well preserved and appears not to suffer from erosion or any undo trampling due to its location in an agricultural field. Modern-day access to the site is treacherous and requires tightrope-walking across a narrow irrigation pipe high above a fast-running stream and subsequently climbing through a field of nettles and snakes. During prehistoric times, Chopan, Togolok and Pessedjik all lay on the banks of the ancient river Sekiz-Yab, with Chopan situated southernmost (upstream), followed by Togolok (5-6 kilometres to the north downstream) and finally Pessedjik (2 kilometres northwest of Togolok further downstream).

Figure 3.5 Chopan site layout (after Ershov 1952)

Figure 3.5 shows the three occupational horizons with plaster floor layers and intermittent cultural material and debris in profile. Ershov reported that the uppermost occupation level of the site was badly deflated and therefore most excavation concentrated on occupational horizons 2 and 3. Ershov began by excavating a 4 x 4 metre trench in the topographically most prominent southwest end of the mound. Ershov reduced this trench to 2 x 4 metres at 2 metres depth and subsequently 2 x 2 metres at 3 metres depth (Ershov 1956, 13) through what he interpreted as three occupational horizons to a total depth of 6.4 metres. Berdiev opened a trench measuring 5 x 10 metres between the areas previously excavated by Ershov and Marushchenko, excavating a total of 10 oneroomed proto mud-brick house structures in occupational horizons 2-5 (numbered from the top down). The houses

Figure 3.6 Chopan Ceramics, clay figurines, clay beads, stone and bone pendants, mortars and pestles, stone tools, stone 30

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

weights, shell fragments and an abundance of animal bone made up the artefactual assemblage. The bone assemblage contained the remains of goat, sheep, bull, dog, gazelle, and fox (Berdiev 1972a, 78). Berdiev reports that the stone used in the on-site flintknapping was provenanced to the nearby Kopet Dag (1968b, 11). Chopan is especially notable for the burials of at least three adults, two of whom were buried extended on their backs while the other was buried in a contracted position on its side. There were also five child burials found, some with traces of red ochre on the skull (Berdiev 1972a, 73-74). Kohl (1984, 49) notes that this trait was also recorded at the Caspian cave site of Kailiu. On-site burials, however, are atypical for the Jeitun Culture as a whole. Chopan is also unique among the Jeitun Culture Kopet Dag piedmont sites in terms of the existence of a supposed “pottery production location”. Berdiev suggests that pottery was manufactured in a workshop on-site in the southern area of the mound. This pottery production area consists of 3 large hearths associated with 5 small structures, presumably used for the preparation of clay and organic temper, and the drying of unfired vessels (Berdiev 1972a, 79). Interestingly, and in accordance with this hypothesis, is the homogenous nature of the ceramic assemblage from Chopan. The assemblage is entirely redware, with a much more consistent fabric, clay matrix, and decorative standard than the pottery from other Jeitun Culture sites. Therefore, Chopan probably represents the earliest shift (middle Neolithic, Jeitun phase 2) to a “workshop” mode of production for the Jeitun Culture. A discussion of this shift to workshop production follows below in more detail. The shift in mode of production was perhaps necessitated by the large site size and the needs of the relatively large population that the site supported. In general, the inhabitants of Chopan practised a sedentary or perhaps semi-sedentary agro-pastoral lifestyle supplemented by hunting.

1 = stone, 2 = ash, 3 = construction layers, 4 = wall, 5 = plaster floor, 6 = rubble, 7=– hearth, 8 = sterile sub soil

Figure 3.7 Togolok site stratigraphy (after Berdiev)

Togolok Togolok (Figures 3.7 and 3.8) is located 5-6 km north of Chopan and 2 km to the south of the first dune ridges of the Kara Kum desert in the vicinity of the modernday city of Geok-Tepe. The Neolithic levels at the site are at least 3 metres thick containing four occupational horizons, but are entirely buried underneath the remains of an early Medieval fort. The site has a diameter of approximately 50 metres and rises more than 10 metres above the surrounding floodplain (Berdiev 1964c, 271).

Figure 3.8 Togolok

Kurbansakhatov and I visited the site during the 1998 reconnaissance survey to find the eastern face of the site being actively bulldozed as a source of fertiliser for the surrounding agricultural fields. Active bulldozing encouraged large birds to dig nests on the open eastern face of the mound, contributing to further erosion and destruction. Damage since that time is unknown.

Togolok was discovered by A. A. Marushchenko in 1939 and the Medieval sections were first excavated by Ershov in 1950. The lower Neolithic levels on the southern side of the mound were excavated by Berdiev in the spring of 1961 (1964c, 271) and in the late 1970s by K. Kurbansakhatov. Berdiev reports the excavation 31

EXCAVATION METHODOLOGY AND RESULTS

of four occupational horizons of the Neolithic levels at Togolok in one deep sounding on the south side of the mound (1964c, 272). Togolok was not excavated in the typical strip-removal open-plan style until 1967 because of the Medieval fortress which stands atop the prehistoric remains of the mound. During the 1967 excavations, Berdiev opened a 10 x 10 metre square of the site and dug through Antiquity period and Sassanian ruins in order to reach the lower Neolithic remains. Below the lowest of the four Neolithic occupational levels lies sterile soil. Berdiev compares the mud-brick walls of the house structures encountered to those at Jeitun, Chopan, Chagylly and the Neolithic levels at Bami. He remarks that at Togolok, there is also a predominance of house structures with shared walls (Berdiev 1968b, 13). On the eastern side of the excavation, a series of low parallel walls, often interpreted as an area for the processing and drying of grain, was uncovered. These typical sets of low parallel walls are found at early agricultural sites throughout Turkmenistan, including Jeitun, Chagylly, Yalangach, and Kara Depe (Berdiev 1968b, 13).

Figure 3.6 shows house structures from the same main occupational horizon excavated during different fieldseasons. Pessedjik follows the same building tradition of Jeitun, Chopan and Togolok. The occupational horizons at Pessedjik form a total cultural deposit of 3.5 metres with sterile soil beneath the lowest of the four levels.

Ceramics, stone implements including scrapers, geometric microliths, trapezes, grinding stones, mortars and pestles, shells from the Caspian Sea, anthropomorphic clay figurines, and a variety of animal bones made up the remainder of the artefactual assemblage at the site (Berdiev 1968b, 14-15). The Neolithic levels at Togolok represent a small sedentary or perhaps semi-sedentary agricultural-pastoral village whose inhabitants supplemented their subsistence regime through hunting.

Figure 3.9 Pessedjik architecture (after Berdiev 1970)

Pessedjik Pessedjik Depe (Figures 3.9 and 3.10) lies 1.5 kilometres northeast of Togolok on the banks of the ancient SekizYab, on the modern-day Maksima-Gorkogo GeokTepinsky collective farm. During our reconnaissance survey in 1998, Kurbansakhatov and I noted that more than half of the mound is currently under crop and the remainder is extremely trampled and suffering from severe erosion. We witnessed three herds of livestock and numerous villagers walk across a path on the southeast uncultivated part of the site during our two hour visit. Pessedjik was discovered and first excavated by V. N. Pilipko and G. Gamayunovo in 1967. The site was further excavated by O. K. Berdiev in 1968 and 1969 and O. Lollekova in 1976 and 1977. Pessedjik is representative of the middle- and late-Jeitun phases, and Berdiev (1970, 14) suggested that the site was potentially inhabited through the early pre-Namazga Aeneolithic or Anau 1A period. An initial trench of 3 x 2 metres indicated that there were four occupational horizons, the lower three of which were well-preserved while the uppermost was found entirely deflated but evidencing the remains of a plaster floor layer (Berdiev 1968c, 15). The small scale nature of the initial excavations prevented the investigators from understanding the intricacies of the site lay-out and architectural structure.

Figure 3.10 Pessedjik Similar agricultural and pastoral traditions were practised by the Jeitun Culture people at Pessedjik. Lisitsina provided a similar analysis of the agricultural environment at Pessedjik. Most likely, water was diverted from the Sekiz-Yab to irrigate agricultural fields near the settlement, thus creating a similar boggy estuary environment (Lollekova 1978b, 189). The existence of desert saksaul points to the aridification of the area, even as early as the 6th millennium BC. An area of 1500 square metres of the uppermost occupational horizon at Pessedjik was stripped due to its deflated nature. Horizon 1 contained ceramics, stone tools such as sickle blades, one trapeze, knife blades, a chisel and flint debitage (Berdiev 1968c, 16). 32

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Subsequently, a 10 x 10 metre trench was opened in horizon 2, revealing 6 typical Jeitun-type mud-brick structures with courtyards and outbuildings, hearthpits, storage niches, lime plaster floors and reed matting. Perhaps most notable was the unearthing of a larger (64 square metres) structure that was interpreted as a communal ritual sanctuary (Lollekova 1978b, 177). This building originates in horizon 3 and is distinguished from the other house structures by its size and the relatively finer quality “alabaster-like” floor. The ritual sanctuary contained no utilitarian goods whatsoever. Most striking was the discovery of a polychrome fresco reminiscent of those found at Çatal Höyük in central Anatolia. The fresco contained both geometric (black and red on a white background) and anthropomorphic, feline, and representations of herbivores. In addition to the Jeitun-style house structures and the ritual sanctuary, several series of low parallel walls, generally interpreted as granaries, were also found in the courtyards separating the house structures. These low parallel walls measured 40 cm long by 1 cm wide by 25 cm tall and ceramics and numerous fire-cracked stones were found in the space between them (Lollekova 1978b, 178). Several courtyard areas were excavated revealing alternating layers of sand and ash to a depth of 40-45 cm. Similar excavations of plaster floor layers within the house structures revealed multiple floor layers, each 5-6 cm thick. The remaining artefactual assemblage included storage vessels dug into the floors of house structures, Jeitun-style redware ceramics, sickle blades, knife blades, and a piercer (Lollekova 1978b, 178-180).

prehistoric mound of New Nisa lies to the north of the Parthian fortress at Old Nisa. The site would appear to present an excellent possibility for future research. However, substantial excavation of overlying cultural deposits would be required in order to successfully uncover the Neolithic extent of the site. Discovered in 1896 by V. A. Zhukovsky and surveyed by M. E. Masson in 1929, initial excavations were undertaken from 1935-1936 during an expedition of Turkmen’kult led by A. A. Marushchenko. Post-war, excavations at New Nisa were re-opened by a IuTAKE expedition led by A. P. Okladnikov (Berdiev 1965, 237) which uncovered 39 house structures. In his attempts to incorporate New Nisa into the overview of Neolithic sites of the Kopet Dag piedmont, Berdiev provides a detailed account of the types of archaeological materials recovered, including ceramics and stone tools. Unfortunately, it would seem that Berdiev did not have access to Okladnikov’s site maps or notes, most likely because Okladnikov was based in Novosibirsk. Hence, we do not have an idea of the overall layout of the 39 house structures, or exactly how the site was excavated (i.e. by occupational horizon). The ceramic assemblage at New Nisa included coarsely made thick-walled storage vessels (mainly cylindrical vases) and well-made thin-walled domestic vessels (mainly bowls). These vessels were chaff-tempered, low-fired, slipped, and painted, typical for the Jeitun Culture (Berdiev 1965, 238). Berdiev also recovered clay weft weights, ceramic discs, knife blades, sickle blades, and microscrapers with retouch (Berdiev 1965, 239-240).

Over the course of four field-seasons between 1976 and 1977, O. Lollekova continued Berdiev’s excavations of a 10 x 10 metre area in horizon 3. She encountered further building levels of the mud-brick house structures and the so-called ritual sanctuary. Lollekova continued excavations of the sanctuary complex to uncover more granary walls in the adjacent courtyard, fragments of wall paintings, ceramics, stone tools and interestingly, large numbers of cracked and burnt stones between the granary walls. Pessedjik represents a small agriculturalpastoral sedentary or semi-sedentary village. Lollekova (1978b, 191) interpreted the ritual sanctuary and fresco as indicative of the site’s regional status as the capital and religious centre of the surrounding oasis during the middle-Jeitun phase. She also noted parallels with similar ritual centres of the Near East and Central Asia, namely Çatal Höyük in Anatolia and Zaraut-Kamara in southern Uzbekistan (Lollekova 1978b, 191). This existence of such a ritual sanctuary points to the increasing socio-cultural complexity of the Jeitun Culture during the middle Neolithic.

Berdiev suggests that New Nisa was populated by the former inhabitants of Jeitun who abandoned their original settlement due to an environmental trend towards desertification and the drying of streams in the Kopet Dag floodplain. These Jeitun people, supposed in this hypothesis to have been entirely sedentary, left in search of better-watered and cultivable lands. This theoretical abandonment led to the population of New Nisa and the other middle-Jeitun phase sites. Berdiev concludes that, because of the ever-narrowing nature of the piedmont strip due to the increasing size of the Kara Kum, the Jeitun Culture populations were forced to build new settlements nearer the Kopet Dag range, such as New Nisa, in order to continue their agro-pastoral lifestyle (Berdiev 1965, 242). Gievdzhik The site of Gievdzhik lies in the foothills of the Kopet Dag in the Geok-Tepe region. Kurbansakhatov and I were unable to visit the site during our 1998 reconnaissance survey because of its proximity to Soviet tank bunkers currently manned by Russian border guards on the Iranian frontier.

New Nisa The fortress and mound complex of Old and New Nisa is located 18 km west of Ashkhabad, surrounded by grape vineyards near the modern-day settlement of Bagir. The 33

EXCAVATION METHODOLOGY AND RESULTS

Jeitun Culture sites: Meana-Chaacha district (Kopet Dag eastern zone)

horizon 2 included grinding stones, beads, clay figurines and stone tools (Berdiev 1966, 5).

Chagylly Kurbansakhatov and I visited Chagylly in 1998 and remarked that the site is suffering from severe erosion and the so-called Soviet donut-ring excavation malaise. I mention this pathology because it is actually named in Soviet excavation manuals which insist that it is forbidden to undertake excavations leaving the unexcavated part of the site in a “donut-ring” shape, thus promoting rapid deterioration of the remaining cultural deposits through erosion. Similar to what we found at Chakmakli, any future excavation at the site should be undertaken in a timely fashion due to the current state of degradation. Chagylly (Figures 3.11 and 3.12) is located 8-9 km east of Meana in the Meana-Chaacha district of southeastern Turkmenistan. The site is between the Meana and Chaacha streams, which at the time of occupation would have created a well-watered estuary environment suitable for the domestication and hybridisation of cereal and grain species (Berdiev 1966, 26). Discovered by A. F. Ganyalin in 1961, Chagylly was first excavated by Ganyalin in 1962-1963 and later by Berdiev in 1966. The site contains a cultural deposit 6.5 metres thick spanning 12 building (or occupational) horizons. The main subsistence adaptation was farming by a method of collective labour, collective ownership of land, and a share-out of the harvest (Berdiev 1966, 26). Agriculture was supplemented by stockbreeding. Berdiev suggests that domesticated animals were also collective property (Berdiev 1966, 27).

Figure 3.11 Chagylly architecture (after Berdiev 1969)

Occupational horizon 1 was categorised by surface finds and was otherwise completely eroded (Berdiev 1966, 3). Horizon 2 was completely excavated and contained architecture of one-roomed dwellings with adjacent courtyards and outbuildings constructed with rounded proto-mudbricks, plaster floors painted red or black, and flat wooden roofs. The architectural plan is slightly more densely packed than at the site of Jeitun (Berdiev 1966). There is evidence that the house floors were pounded in order to make them harder and more compact. In some cases, the hearths were made of stone with storage niches located nearby and vessels for collecting charcoal cleaned from the hearth (Berdiev 1966, 4). Reed mats, characteristic of Near Eastern sites like Jarmo, were found covering the floors (Berdiev 1966, 4). There was also an adult burial found underneath the floor of house 18, lying on its right side with the head facing southwest, arms elongated, legs bent, and a stone vessel placed next to the head. Another burial, this time of a baby, was found under the floor of one of the outbuildings of house 18. The house was determined to be an outbuilding because of the lack of a hearth. The baby was buried 50 cm under the floor facing north (Berdiev 1966, 5). Other artefacts from

Figure 3.12 Chagylly Occupational horizon 3 was widely excavated, covering approximately two thirds of the area of the entire site. Horizon 3 contained standard one-roomed houses, courtyards and outbuildings with plaster floors painted red, or floors and walls painted black, oval-shaped hearths, and abundant fire-cracked stones near the hearths (Berdiev 1966, 5-6). Interestingly, the hearth in house 9 contained a slit or flue in the brick construction to allow smoke to escape outside the house (Berdiev 1966, 7). Other typical Jeitun-type architectural features include several series of low parallel walls, either used as a granary or for drying meat. There were also hearths outside the dwellings interpreted as communal features 34

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

(Berdiev 1966, 8). A second adult burial in horizon 3, this time in house 6, was lying on the right side with legs bent, facing northeast. The left arm was straight while the right arm was bent so that the right hand was resting under the head (Berdiev 1966, 7). Other artefactual remains included animal figurines, ceramics and stone tools.

rather than the rounded Jeitun-type proto-blocks. The Aeneolithic levels of the site contained a large house, which Masson interprets as a “clanhouse” or home of the chief (1971, 52). Although Chagylly is attributed to the late Neolithic Jeitun phase 3, the site is considered transitional to the subsequent Anau 1A period (Berdiev 1969, 32).

Horizons 4-12 were excavated in a single deep trench rather than as an open area. Horizon 4 was a yellowgreenish clay layer containing ceramics, stone and bone tools characteristic of Chopan (Berdiev 1966, 10). Horizon 5 was a greenish soil layer containing intermittent black charcoal layers with destruction materials, stone tools and animal bones (Berdiev 1966, 10). Horizon 6 contained further floor levels, a hearth and walls made of plaster approximately 5 cm thick with chaff admixture (Berdiev 1966, 10). Horizon 7 contained a floor made of bricks measuring 60 x 20 x 10 cm covered in plaster with a chaff admixture. On this floor were abundant fire-cracked stones (Berdiev 1966, 10). Horizon 8 was a red layer of burned soil mixed with ash, ceramic fragments and bones (Berdiev 1966, 10). Horizon 9 contained walls made of mudbricks measuring 50-55 cm long, 20-25 cm thick and 8-10 cm high. The floor was 5-6 cm thick made of plaster with a chaff admixture (Berdiev 1966, 10). Horizon 10 contained a plaster floor, 4-5 cm thick, and the remains of a campfire which had been lit on top of it (Berdiev 1966, 10). Horizon 11 was a plaster floor, 5 cm thick (Berdiev 1966, 10). Horizon 12, the lowest occupational level at Chagylly, contained mudbrick walls 10-12 cm thick. Notably, these walls were thinner than the house walls of the upper levels. Beneath horizon 12 was pure sand from some sort of “mountain torrent” overlying red alluvial clay (Berdiev 1966, 10).

There is evidence of a shift towards increasingly technological stone tool industries and a complex economy (Berdiev 1969, 23). A large amount of flint debitage and cores excavated at Chagylly point to stone tool making on-site. The types of tools found included sickle blades, one-sided and double-sided blades, geometric tools, scrapers, microscrapers, arrows, hammers, sharpening stones, and axes (Berdiev 1969, 25-29). Berdiev notes that the microscrapers, typical for the earlier phases of the Jeitun Culture, were very infrequent. He interprets their existence as traditional rather than actually useful (Berdiev 1969, 27). In general, the stone tool assemblage from Chagylly is more refined in comparison to assemblages from sites of earlier periods. An abundance of mortars and pestles were found in the Neolithic levels of the site, evidencing the grinding of grain and mineral dyes for paint (Berdiev 1969, 28). The upper Aeneolithic levels at Chagylly contained sickles blades attached to handles, which points to an increase in technology warranted by an intensification in agriculture (Berdiev 1969, 26). There were also bone tools, needles, awls and piercers used for pottery production (vessel smoothing or burnishing), weaving reed mats, and tanning hides (Berdiev 1969, 31). Berdiev synthesised the transition of the stone and bone tool industries and increasing technology in two main ways. The first is the transition from stone to bone tools in the late Neolithic evidenced by the multitude of efficient and refined bone tools at Chagylly. Second, Berdiev notes that not only does the quality of stone tools decline, but the variety of stone tool types is reduced (Berdiev 1969, 31). Eventually, stone tools are completely replaced by bone tools, which are in turn replaced by metal implements (Berdiev 1969, 32).

Based on Berdiev’s above description for the occupational horizons at Chagylly, it would appear that not all Soviet researchers interpreted the words “occupational horizon” in the same manner. Some of the lower layers, notably horizons 4, 5 and 8, seem to refer to archaeological contexts rather than occupational levels. This is the misuse of archaeological terminology in its most exaggerated form. I am not attributing this exaggeration only to Berdiev and his work, but to the Soviet archaeological formula for documenting excavations. It would seem that “occupational horizon” (a direct translation from the original Russian) is not the appropriate wording. Denoting each floor layer, which could have been laid down on a seasonal basis, as a separate occupational horizon unfortunately gives the reader an impression of a significant time period and transition from one horizon to another. It would seem that “occupational horizon” at times refers to what it signifies, but at other times refers to what is considered a “context” in British archaeological terminology.

In the upper early Aeneolithic levels of Chagylly, fragments of copper and the Near Eastern tradition of placing burials in the floors of houses or outbuildings were found. There was also evidence of an increasingly diversified production economy in terms of tanning and leatherworking, woodworking, boneworking, and the use of mineral dyes for painting floors and walls (Berdiev 1969, 32). Berdiev suggests that along with increasing artistic and aesthetic appreciation came spiritual development, evidenced by the production of anthropomorphic figurines (Berdiev 1969, 32). Berdiev synthesises the results from the excavations at Chagylly by suggesting that the site is typologically contemporaneous with the Sialk I complex. He notes exchange and connections with Iran evidenced by a turquoise bead found at the site. The nearest known

The upper Aeneolithic levels at Chagylly were composed of house structures made of “Aeneolithic-type bricks” 35

EXCAVATION METHODOLOGY AND RESULTS

sources of turquoise are at Nishapur in Iran or in the mountains near Samarkand in modern-day Uzbekistan (Berdiev 1966, 27). He also suggests that some of the shell jewellery found at Chagylly points to contacts with the Caspian Sea area and the Indian Ocean. Volcanic glass excavated at the site is attributed to a source area on the Krasnovodsk peninsula and evidences contacts with the Keltiminar or Keltiminar-related groups. Berdiev suggests that these contacts started during the Neolithic period and expanded during subsequent periods to include contacts with Pri-Aralia, Khoresmia, the Iranian Khorassan, Djanbas, the upper and lower Uzboi sites, and the Murghab and Tedjen oases (Berdiev 1966, 28).

48) that Chakmakli was situated in the floodplain of a large river delta channel and suffered from seasonal inundation evidenced by a layer of alluvial sand 50 cm. thick found between occupational horizons 4 and 5. As previously mentioned, the generally accepted Soviet interpretation of the site is that it was occupied by a technologically more advanced intrusive population from Iran who coexisted with the Neolithic Jeitun Culture peoples of nearby Chagylly. Kohl (1984, 69) does not accept this interpretation and suggests that the Anau 1A Culture may have developed from the late Neolithic Jeitun Culture at the site. Both interpretations seem questionable, and as is characteristic of most archaeological research in former Soviet Central Asia, the situation is exacerbated by the lack of a secure radiocarbon dating chronology.

Chakmakli The site of Chakmakli (Figure 3.13) is located 1 kilometre from Chagylly, near the Chaacha stream. During our reconnaissance survey in 1998, K. Kurbansakhatov and I visited the site after considerable difficulty in locating it. Having not visited Chakmakli previously, Kurbansakhatov questioned local informants from Chaacha as to the exact location of both Chakmakli and Chagylly. Unfortunately, the only surviving man in Chaacha old enough to remember Berdiev’s excavations could not take us to the site, but drew a rough map which eventually led us to it. Normally, in situations such as this, a GPS would prove essential. The political climate of Turkmenistan, marred by former Soviet paranoia, makes the use of such equipment at present difficult or impossible. Moreover, it proved impossible to locate many of the Kopet Dag Neolithic sites Kurbansakhatov and I looked for during our two week survey.

Figure 3.13 Chakmakli Marushchenko’s deep soundings (approximately 5 metres) provide evidence for buried cultural layers, to a depth of approximately 2.85 metres from the top level of the site (Kohl 1984, 67). Marushchenko identified five building or occupational horizons, some of them separated by sterile layers. Interestingly, the lowest level of the site, called level 5 (separated from level 4 by 50 centimetres of sterile alluvial sand), contains no ceramic remains. Masson and Sarianidi state that “this absence of pottery makes the whole picture very confusing” (1972, 48). Kohl (1984, 68) suggests, contrary to the significance given by Masson and Sarianidi (1972, 48) and Gupta (1979, 58-64) to the lack of ceramics in level 5, that actually the area of level 5 excavated was too small (1.5 by 2 metres) to draw overarching conclusions relative to the overall nature of the occupational level.

At present, Chakmakli rises less than one metre above the surrounding floodplain. Formerly oval in shape, it is now a heavily eroded narrow ring-shape. Erosion has been promoted by extensive excavations of the middle of the site by Marushchenko and Berdiev. Because of the degree of degradation, a crucial factor in our ability to locate the site was the contrast between what remains of the bare sandy mound and the surrounding grassy plain. In spring, the floodplain is overgrown with desert grass watered by the spring rains and run-off from the Kopet Dag for a period of one to two months. It was necessary to periodically climb on top of our vehicle in order to successfully spot the site. It is clear that if there is to be further archaeological excavation to establish an absolute dating chronology, it must be done swiftly prior to the imminent disappearance of the above-ground portion of the site.

Berdiev (1968a) carried out excavations at Chakmakli in 1965, excavating deflated level 1, which he interpreted as a transitional stage between Anau 1A and Anau 1B (also referred to as Namazga 1 or the early Aeneolithic), and subsequently exposed a large section of level 2. Level 2 contained a main street bisecting the site and one smaller alleyway on either side of this street. Berdiev identified four self-contained occupational areas with shared walls and courtyards, separated by the main street and the two alleyways. There was a set of low parallel

Chakmakli, originally discovered by Ganyalin in 1962 and partially excavated by Marushchenko in 1963, previously rose 1.5 metres above the surrounding floodplain and was an oval shape of approximately 80 by 60 metres. Based on an examination of the underlying sedimentary rock, Masson and Sarianidi report (1972,

36

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Discovered by Marushchenko in 1935, and not excavated until 1959 and 1960, Mondjukli lies 2 kilometres southeast of the modern-day town of Meana. Marushchenko identified nine building levels and a total cultural deposit of 5.6 metres (Berdiev 1972c, 1972d). Mondjukli had a more complicated architectural structure than the other Neolithic Jeitun Culture sites of the Kopet Dag piedmont, but a less complicated and less well-planned structure than that found at Chakmakli. The site is bisected by a main street and contains three types of houses as described by Berdiev (1972c, 1972d). These include: small rooms divided into several smaller sections by projecting walls, rooms with hearths, and long narrow spaces interpreted as entrance hallways. In the uppermost layers, 40 rooms with plastered floors and walls were exposed through excavation.

walls constructed in one of the courtyards similar to features interpreted as granaries or areas for drying meat at Jeitun Culture sites. The house structures were made of uniformly sized mud-bricks with hearths and painted plaster floors. The site layout at Chakmakli was significantly more complicated than that of other Neolithic Jeitun Culture sites, including nearby Chagylly. The ceramics from Chakmakli are of the Anau 1A type, well made, thin walled, sand rather than chafftempered, and slipped or burnished. The decoration includes geometric designs, especially triangles. The lithic remains at Chakmakli also differed from those at Chagylly in the use of white flint for making stone tools and the existence of a stone hoe in the uppermost level of the site. Berdiev interpreted this stone hoe as further evidence for technology intrusive from Iran and related to the Sialk I complex. Other material remains excavated at Chakmakli include spindle whorls for weaving and copper pins, awls, piercers, and chisels. The nature of the site remains slightly ambiguous due to the controversy surrounding the origins of its inhabitants. The Chakmakli people were either immigrants from northern Iran or an indigenous group practising an increasingly complex and technologically advanced way of life. Only future excavation and radiocarbon dating will provide answers to these questions and more clearly place Chakmakli within the overall chronological and socio-cultural framework of prehistoric Turkmenia.

The Anau 1A levels at Mondjukli existed in the top three metres of the cultural deposits, underneath were found late Jeitun Culture levels. Mondjukli is the only site representative of the shift from the Jeitun Neolithic to the pre-Namazga Anau 1A Aeneolithic period. Berdiev (1969, 29-30, 40, 45) divided the cultural levels into three distinct sub-phases. These are: Mondjukli 1 (middle-Jeitun phase), Mondjukli 2 (late-Jeitun phase) and Mondjukli 3 (Anau 1A phase). Korobkova disputes Berdiev’s claim of a continuous archaeological sequence with lithic evidence linking Mondjukli to the middleJeitun phase but lacking evidence for the late-Jeitun phase. Korobkova suggests that the Anau 1A levels at Mondjukli were contemporaneous with the late-Jeitun levels at Chagylly, a phase which is missing from the Mondjukli stratigraphic sequence. Whatever the case, it seems clear that until there are secure radiocarbon dates for the Neolithic Jeitun phases and the Anau 1A and 1B periods, there will be unresolved controversy caused by typological analysis.

Mondjukli Kurbansakhatov and I also visited Mondjukli (Figure 3.14) during our reconnaissance survey. It lies between Altyn Depe and the Meana-Chaacha road and is easily visible rising 2 to 3 metres above the surrounding floodplain, with a diameter of 20 metres. It is currently seasonally inhabited by several shepherds who have erected small lean-to structures, dug cooking hearths, and who use the mound as a base from which to watch their flocks of sheep. Because of this modern-day usage of the mound as a convenient observation spot, it has become extremely trampled by humans and sheep.

The 3 lowest building levels at Mondjukli contained Jeitun-type proto-bricks, Jeitun-type door sockets, and Jeitun Culture ceramics. The fourth level (numbered from the bottom up), represents a transition, with Jeitun Culture assemblages and house structures found together with more well-made ceramics. The sixth level contained Anau 1A ceramics, spindle whorls and sling balls made of stone. The seventh level contained more Anau 1A ceramics found together with other vessels bearing decoration reminiscent of the parallel vertical line decoration typically found at Jeitun. The houses of level 8 had plastered walls, uniform mud-bricks and a child burial similar to those found at Anau itself. As mentioned, the final 9th level is interpreted as representative of the transition to Anau 1B and contained some Namazga-type artefactual remains (Berdiev 1972c, 1972d). In the Anau 1A levels at Mondjukli, the ceramic assemblages were yellow and brown slipped, containing typical Anau 1A decoration of hatched and solid triangles, spindle whorls, animal figurines, polished

Figure 3.14 Mondjukli. 37

EXCAVATION METHODOLOGY AND RESULTS

stone chisels, handled weights, grinding implements, digging stick weights, metal items including punches, awls, and needles, a lump of copper ore suggesting final stage metallurgical production on site, and luxury items such as an alabaster pendant, a lapis lazuli plate, and a shell ornament (Berdiev 1972c, 1972d).

excavation will be precluded by the political climate of the region for some time.

To reiterate, the only known stratigraphic evidence for Jeitun levels overlain by Anau 1A levels exists at Mondjukli. Kohl recounts (1984, 65) that the transition from the Jeitun Culture to the Anau 1A Culture represents the “classic problem in archaeological interpretation: continuity or discontinuity with the preceding period”. Again we find various possible explanations for the archaeologically undocumented “leap” in technological innovation, material culture and complicated architectural complex. The historically unanimous Soviet explanation of the Anau 1A complex is that of an intrusive cultural phenomena. The idea that Anau 1A peoples migrated from Iran, bringing with them a more complicated subsistence and societal regime, is unquestioned within the Soviet academic community. Berdiev (1972c, 30) hypothesised that the more highly developed Anau 1A people of Chakmakli coexisted with the Jeitun Culture people who had migrated from the Geok-Tepe region and settled at Chagylly and Mondjukli in order to exploit the relatively empty fertile plains of the eastern piedmont. There is no absolute dating to substantiate any of the chronological claims and it is obvious that an intensive radiocarbon dating program is necessary to clear up many of the unanswered chronological queries. Kohl (1984, 70) accurately summarises the problem saying that “it is not easy to interpret the Anau 1A period because it is not clear whether one is dealing with a separate chronological period; a distinctive culture that is possibly intrusive from Iran; functionally distinct communities from contemporary Late Jeitun sites; or some combination of two or three of these possibilities”.

Figure 3.15 Gademi architecture (after Lollekova 1980) The houses at Gademi were of typical Jeitun-type mudbrick with plaster floors, hearthpits, and adjacent sleeping platforms. Lollekova reports the size of the larger of two house structures that she excavated as 6 x 4.5 metres with a shared courtyard and a series of parallel low “granary” walls (Lollekova 1980, 18). The inhabitants of Gademi relied not only on hunting upland animal populations of gazelle and goat, but also on some form of agro-pastoralism as evidenced by domesticated animal bones, ancient irrigation ditches and trace-wear analysis on the stone tool assemblage (Lollekova 1980, 19). The ceramic assemblages at Gademi are Jeituntype, entirely undecorated, and analogous to the ceramics from Chagylly, Chakmakli and the Neolithic levels of Mondjukli. Korobkova carried out a study of the stone tool assemblage from Gademi and reports that it is typologically similar to the assemblages from the Jeitun Culture sites of Chopan and Pessedjik including sickle blades, knife blades, scrapers, microblades, microscrapers, symmetrical and asymmetrical trapezes (Korobkova and Volovik 1972, 42). Geometric microliths are absent from the stone tool industry. Korobkova attributes Gademi to the middle-Jeitun period based on the analysis of stone tools and mud-brick architecture (Korobkova and Volovik 1972, 42; Korobkova 1969b, 58-59). Pendants made of Didacna shell were recovered, evidencing some form of longdistance exchange with the Caspian region (Korobkova and Volovik 1972, 42). Korobkova notes that the inhabitants of Gademi had a slightly different

Gademi Gademi (Figure 3.15) is located 7 km south of the southern border of the Chaacha-Kaakhinskogo region of the Marinskoi Oblast of southeastern Turkmenistan. The Neolithic mound, which rises 3 metres above the surrounding terrace on the right bank of the Chaacha river, lies in the demilitarised buffer-zone between the former Soviet Union and Iran. The site was discovered in 1971 by a Soviet geological survey team led by V. T. Volovik (Korobkova and Volovik 1972, 41-42) and was excavated by Korobkova in 1972 and Lollekova some years later. Because of the site’s sensitive location, only limited excavations were permitted under the constant supervision of armed Iranian soldiers (Lollekova pers. comm. 1998). It thus proved impossible to visit the site during the reconnaissance survey in 1998. Despite the potentially interesting nature of the site due to its location actually in the Kopet Dag foothills rather than on the broader piedmont strip, it would seem that further 38

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

subsistence regime to their Jeitun Culture counterparts who lived further down the piedmont slopes (Korobkova and Volovik 1972, 42). She suggests a predominance of stockbreeding and hunting with relatively little agriculture being practised because of the environmental conditions and location of the site. Korobkova concludes by suggesting future research into the varying economies and adaptations of Jeitun people based on environmental and geographical restrictions (Korobkova and Volovik 1972, 42).

southeast (northern Afghanistan), arriving on the Kopet Dag piedmont fully equipped with agricultural and pastoral traditions. This new subsistence system would likely have overridden the nomadic hunting/gathering adaptation of the local groups. Such a theory, although controversial, is also viable, and will require further investigation at the site for verification. Unfortunately, the artefactual assemblages excavated by Marushchenko and Berdiev at Bami were destroyed. Between the prehistoric and later Antiquity levels, there is a 1 metre thick deposit of gravel, interpreted by Berdiev as debris from some natural event which he terms a “mountain torrent” (Berdiev 1963a). Bami is located near an ancient now-dry riverbed.

Jeitun Culture sites: Kopet Dag western zone Bami Bami is an interesting site because of its long occupational sequence, albeit with several chronological gaps. The site (Figures 3.16 and 3.17) was first documented and excavated by A. A. Marushchenko in 1960 as the leader of an expedition of the Institute of History, Archaeology, and Ethnography of the Turkmen Soviet Socialist Republic. A. F. Ganyalin and O. K. Berdiev were members of Marushchenko’s archaeological research team. Marushchenko is unfortunately notorious in Soviet archaeology for his lack of research publications and cryptic field notes (Belova pers. comm. 1998). It would appear that this task was deferred to Berdiev (1963a, 188-194). Kurbansakhatov and I visited Bami during our reconnaissance survey in 1998. The prehistoric kurgan is a large well-preserved mound near the modern-day town of Bami. There is a second Antiquity and Medieval period mound nearby. Marushchenko’s team excavated a 5 metre deep sounding of 3 x 2 metres (decreasing to 2 x 2 metres at 2.5 metres depth) in the centre of the mound. This trench is intact except for the bottom 1-2 metres which have caved in, currently obscuring the lowest Neolithic levels of the site which lie partially below the current surface level of the surrounding floodplain. Bami is the westernmost of all Jeitun Culture sites and contains a long stratigraphic sequence. The main occupation was during the middle and late-Jeitun phases followed by subsequent occupation in the Parthian period. Berdiev (1964c, 277) also suggests, based on the artefactual assemblages of the lowest occupational horizon at Bami, that the site could have been occupied by semi-nomadic hunter/gatherers during the early Neolithic before the Jeitun people arrived during the middle Neolithic. He cites similarities in the artefactual assemblage of this lowest horizon with those found at nearby Kelyata, a nomadic hunting camp. In attempts to reconcile the controversy surrounding the lack of evidence for the transition from the Caspian Mesolithic to the Jeitun Neolithic, Berdiev takes this evidence as pointing to the hypothesis that Jeitun peoples migrated from both the west (the Iranian plateau) and the

1= stone wall, 2 = pebble wall, 3= ash, 4= building layers, 5 = hearth, 6 – mudbrick wall, 7= mudbrick rubble, 8= red ochre, 9 = clay, 10= sterile sub-clay

Figure 3.16 Bami site stratigraphy (after Berdiev 1963) 39

EXCAVATION METHODOLOGY AND RESULTS

gravel layer contained Madaun-type ceramics dated to the 1st millennium BC. Above the gravel layer there were Medieval layers and walls (Berdiev 1963a, 192). Bami may be representative of a small nomadic hunter/gatherer camp during the early Neolithic (or even late Mesolithic?). Later (during the middle and late Neolithic), the site represents a small sedentary or semisedentary agricultural-pastoral village also evidencing hunting. A future programme of excavation and radiocarbon dating at this well-preserved site will elucidate these ideas. The site has a particularly long occupational sequence, probably due to its location along an ancient riverbed as well as its proximity to the Kopet Dag and diverse environmental zones suitable for the exploitation of various natural resources.

Figure 3.17 Bami Marushchenko’s excavations unearthed mud-brick walls and multiple plaster floor layers in the lower 2 metres of the site which are attributed to the Neolithic. Interestingly, some of the plaster floor layers were underlain by stone foundations, the first evidence of such complicated construction for the Jeitun Culture (Berdiev 1963a, 194). Marushchenko identified 5 building horizons within the Neolithic deposit (Berdiev 1963a, 188). These building horizons are labelled Bami 1, 2, 3, 4, and 5. Contrary to the standard methodology of numbering from the top down, Marushchenko chose to number from the bottom up. Hence, Bami 1 is the lowest building level at the site, overlying sterile soil.

Keltiminar and Keltiminar-related sites (Figures 3.18 and 3.19) The lower Amu Darya and Akcha Darya delta: Djanbas Djanbas, located in the southern Akcha Darya delta, is the type-site for the Keltiminar deltaic variant and is contemporaneous with the late Neolithic/early Aeneolithic adaptations on the Kopet Dag piedmont (Vinogradov 1981). Djanbas contains a large (400 square metre) house structure of wood and reed post-hole construction. Tolstov hypothesised (Vinogradov 1968, 35) that because this structure is more than ten times the size of the largest house at Jeitun, the Keltiminar most likely had a non-nuclear extended-family societal structure.

In occupational horizon 1, the architecture was constructed of rounded proto-mudbricks, ceramics included vases and bowls, knife blades with retouch, and sickle blades (Berdiev 1963a, 188-189). Bami horizon 2 contained further plaster floor layers 5-7 centimetres in thickness painted with red ochre, hearthpits, ceramic vases and bowls, flint blades with retouch, and sheep and goat bones (Berdiev 1963a, 191). Occupational horizon 3 contained architecture consisting of plaster floors with an admixture of straw and proto-mudbrick walls with bricks of dimensions 40 x 17 cm. In the southern part of the trench, an unspecified number of alternating layers of charcoal and ash were found, while in the northern part of the trench there was a pile of mudbricks (Berdiev 1963a, 191). The ceramics were predominantly “reddish-yellowish” with painted decoration (Berdiev 1963a, 192). Occupational horizon 4 contained further architecture with mudbricks of dimensions 20 x 30 cm and plaster floors with intermittent ashy layers. The ceramics again included both vases and bowls, stone and bone tools (Berdiev 1963a, 192). Occupational level 5 had a floor made of coarse yellow plaster 5-7 cm thick covered with ash and showing evidence of hearthpits. Other artefacts included ceramics, blades with retouch and trapezes (Berdiev 1963a, 193). Berdiev suggests that the triangle ceramic motifs were prototypes for the later Anau 1A motifs (1963, 193). Above occupational level 5, there was a gravel layer, the result of the abovementioned “mountain torrent”, which was 120 cm thick. The lower part of the gravel layer contained ceramics with Anau 1A decoration, while the upper part of the

The Keltiminar of the deltaic regions subsisted by hunting deer and wild boar in the tugai forest near the rivers and their distributaries, goitre gazelle, camel and wild ass in the adjacent desert areas, and fishing for pike using both nets and spears (Masson 1996, 96). The artefactual assemblage is characterised by calcitetempered stamped or incised pottery as opposed to the chaff-tempered painted pottery of the Jeitun Culture. The Keltiminar lithic assemblages, studied extensively by Korobkova, are dominated by arrowpoints made on shouldered blades, the so-called Keltiminar arrowheads. Notched blades with retouch, backed microblades, elongated asymmetrical triangles, end scrapers, leafshaped arrowpoints with bi-facial retouch, and geometric microliths, notably horned trapezes, were also found (Masson 1996, 99). Tolstov describes Keltiminar life in the deltaic areas as being hampered by swamp-like conditions, thus preventing the Keltiminar peoples from adopting the same developments of the Jeitun culture to the south (1962, 31). This Marxist viewpoint is perhaps overly-simplified due to the inherent assumption that increased labour and output is necessarily a societal, economic, and cultural step forward. Djanbas represents a small semi-nomadic community who practised a seasonal round of resource exploitation. Djanbas itself 40

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.18 Keltiminar Culture architecture (Kavat 2-3 and Djanbas 4-8) (after Ribakov 1996)

41

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.19 Keltiminar Culture artefactual assemblage (Tolstov and Djanbas) (after Ribakov 1996)

42

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

acted as the main base for this seasonal round, whereas the satellite hunting and fishing camps were much more seasonal in nature and occupied for short intermittent periods of time.

complex is generally accepted as contemporaneous with the Jeitun Neolithic adaptations of the Kopet Dag piedmont (Vinogradov 1981, 89). Uzboi river, north Pribalkhan, south Pribalkhan

The Zeravshan, Makhandarya

Ayakagitma,

Daryasai,

and Markov splits the territory surrounding the Bolshoi Balkhan mountains into three distinct regions (Figure 3.20) (Markov 1971). The sites in the mountains themselves are categorised separately and can be found below. The three sub-areas include: the east Pribalkhan, otherwise known as the upper and lower Uzboi; the north Pribalkhan (territory of Oyukli); and the south Pribalkhan (territory of the Aktam and Kelkyor rivers). A brief overview of each of these areas follows.

The Neolithic site of Uchashi 131, a Keltiminar campsite located on the multi-layered sandy peat Holocene alluvial terrace of the Daryasai, contains six cultural horizons and shows evidence for the exploitation of gazelle, pig, deer, wild boar, goat and various species of fish (Korobkova 1975a, 31). Uchashi is considered contemporaneous with layers 5 and 5a at Djebel and the upper 2nd horizon at Tutkaul in Tadjikistan (Vinogradov 1981, 92). The site is therefore dated, in relative terms only, to the 6th millennium BC. Vinogradov (1981, 92) reports that, based on other surveys he carried out, the sites of the Daryasai are similar to sites in southeast UstYurt, northeast Priaral, northern Afghanistan and eastern Iran. Vinogradov notes that the (unnamed) sites of the Makhandarya are broadly similar to other sites in greater Khoresmia. Korobkova suggests that the stone tool assemblages found in the Zeravshan valley are analogous to those from Mondjukli and Gademi, thus evidencing cultural contact between the Jeitun and Keltiminar Cultures (Korobkova 1981, 31). Although I have no relevant literature regarding the Neolithic site of Ayakagitma (on the bank of the Ayakagitma river), the site was recently re-opened by a Polish team from Warsaw University in conjunction with the Institute of Archaeology in Bukhara, a branch of the Academy of Sciences of Uzbekistan. Vinogradov reported the existence of Keltiminar arrowheads at Ayakagitma (Vinogradov 1981, 96).

As Larson (2000) reports, there are a multitude of archaeological sites along the course of the now-dry Uzboi river (Figures 3.21 and 3.22) with potential archaeological dates ranging from the Neolithic to the Medieval periods. The Keltiminar-related Uzboi variants would thus be attributed to the period between the 6th and 2nd millennia BC. There is archaeobotanical evidence pointing to the Uzboi river valley in prehistory as a green oasis with pines, fir, birch, oak, beech, alder, and various species of nut trees. The riverine environment would have provided an excellent source of fish and waterfowl to complement the wild boar, deer and goat living in this once-fertile valley (Korobkova 1981, 29). Archaeological sites have been found on both the right and left banks of the ancient Uzboi. There are two sites on the left bank of the lower Uzboi river, Kizpara and Chaloi. Kizpara, near Chaloi, is located on a terrace of the lower Uzboi. The artefactual assemblage at Kizpara included stone tools, ceramics, charcoal and the remains of campfires (Markov 1971, 29). The site of Chaloi is attributed to the late Neolithic and considered contemporaneous with layer 4 at Djebel. Chaloi had a significantly larger artefactual assemblage including 200 fragments of undecorated coil ceramics (either jars or vases) with a grey or grey-brown paste containing quartz or sand temper, and stone tools (Markov 1971, 29). There are three main sites located on the right bank of the lower Uzboi river: Sengradji and Djoiruk I and II. Sengradji, considerably smaller than either Djoiruk I or II, yielded three fragments of coil vessels and one fragment of a wheel-turned vessel. Based on the stone tool assemblage, which is unfortunately not welldocumented, Markov considers Sengradji to be more archaic than the other sites of the lower Uzboi (Markov 1971, 36). Markov does not, however, specify how “archaic” he means. The sites of Djoiruk I and II are situated near one another on the right bank of the lower Uzboi and are attributed to the Bronze Age. Markov encountered some difficulty in excavating these sites because they were partially covered with deep drifts of sand from nearby barchan dunes (Markov 1971, 43-44).

Lake Lyavlyakan and the inner Kyzyl Kum Vinogradov surveyed the inner Kyzyl Kum and conducted excavations at the Neolithic Keltiminar sites of Kavat and Darbazakir in the area of ancient Lake Lyavlyakan. These sites were dated in relative terms, through a comparison of stone tool and ceramic assemblages, to the early 5th millennium BC and are considered contemporaneous with layers 5-6 at Djebel (Vinogradov 1981, 88). Six radiocarbon dates were obtained from sediments of Lake Lyavlyakan, four of which suited the framework of the 4th millennium BC, one of which referred to the third or fourth quarter of the 5th millennium BC, and the last of which was dated to the first half of the 3rd millennium BC (Vinogradov 1981, 94). The sites of the inner Kyzyl Kum were previously considered earlier in terms of chronology than many of the other Keltiminar or Keltiminar-related variants because of the dating of these ancient lake sediments. Further research by Masson and Korobkova, however, pushed the dating of the entire Keltiminarrelated complex back so that the Neolithic Keltiminar 43

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.20 Bolshoi Balkhan, lower Uzboi, north and south Pribalkhan (after Markov 1971) The pottery at Djoiruk was stamped rather than incised, which appears to be a Bronze Age variation in the assemblages of the Keltiminar-related groups. At Djoiruk I, Markov recovered 31 fragments of coil vessels (as well as later Medieval ceramics), which he remarks were similar to part of the assemblage at Oyukli (Markov 1971, 43). The archaeological chronology at Djoiruk II, however, was more difficult to determine due to severe alluvial erosion and the movement of the nearby barchan dunes. Markov recovered three fragments of coil pots as well as stone tools close in form and shape to vessels from other lower Uzboi sites (Markov 1971, 44). In providing a synthesis regarding the sites of the Uzboi, Markov (1971, 57) rejects the previously accepted theory that there were separate cultural variants in the upper and lower Uzboi regions during the Neolithic. He does, however, mention that during the Bronze Age, the upper Uzboi is characterised by sites of the Tazabagyab Culture, which become more infrequent as one travels down the course of the Uzboi toward the Caspian Sea. Markov also notes the existence of a separate (but unnamed) lower Uzboi Bronze Age Culture characterised by flat-bottomed vessels with stamped decoration (Markov 1971, 44).

by Markov in 1957 (Markov 1961, 67). The main site in the north Pribalkhan, Oyukli is interpreted by Markov and Khamrakuliev as a separate archaeological culture although it fits into the scheme of Keltiminar-related variants based on its artefactual assemblage. The British team visited the site, which was entirely deflated, in the spring of 1997. Markov’s original excavations at Oyukli unearthed prehistoric ceramics, metal objects, and later Medieval remains pointing to the long time-span of the site’s occasional use as a campsite or stopover. Markov suggests that no long-term settlement took place at Oyukli as evidenced by the lack of either mudbrick structures or post-holes indicating previous wooden wattle and daub houses like those found in the deltaic regions (Markov and Khamrakuliev 1980, 68). Oyukli was a nomadic camp, used periodically as part of a seasonal-round subsistence system. Interestingly, the prehistoric assemblage from Oyukli (i.e. possibly Mesolithic, but generally Neolithic through Iron Age), appears to be more advanced and unrelated to the artefactual assemblages of nearby contemporaneous prehistoric sites in the Akcha Darya delta, the Bolshoi Balkhan, and along the Uzboi river (Markov 1961, 80; Markov and Khamrakuliev 1980, 67).

The site of Oyukli (Figure 3.23), located in the north Pribalkhan or north Balkhan steppe, is located just south of the Chil Mamed Kum desert and northwest of the course of the ancient Uzboi. The site is set on sandy takyr solonchak proluvium 200-250 metres south of the first ridge of barchan dunes and was originally excavated

Markov reports that it was impossible to determine any sort of stratigraphy for Oyukli due to its deflated nature (Markov 1961, 68). The archaeological finds were very mixed and included charcoal and evidence of hearths, prehistoric flint and stone tools, ceramics, metal objects, bone, and later Medieval remains. 44

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Due to the complete deflation of this seemingly important site, Markov decided to carry out site deflation experiments in order to attempt to better understand the artefactual assemblage at Oyukli. These site deflation experiments entailed reburying archaeological finds at a prescribed depth and subsequently returning (he does not detail the length of time) to record the extent of exposure and disappearance of the overlying sand. Site deflation appeared to be quite rapid based on these experiments (which were presumably conducted between one fieldseason and the next) (Markov 1961, 68). Markov and his team located excavation trenches at Oyukli based on the highest density of surface finds. There is no stratigraphy reported whatsoever, hence Markov categorises the entire artefactual assemblage as a single entity, despite the obvious chronological differentiation within it. Interestingly, the part of the assemblage attributed to prehistoric times (i.e. Neolithic through Iron Age), appears to be more advanced and unrelated to the artefactual assemblages of nearby prehistoric sites in the Bolshoi Balkhan and along the Uzboi river (Markov and Khamrakuliev 1980, 67). Markov reports that although some of the ceramic motifs at Oyukli are similar to those found at Djanbas, the stone tool assemblages from the two sites are discrete. Markov and Khamrakuliev compare the Oyukli ceramics to vessels from Eastern Europe, the Dnieper river region, the Afanav and Tazabagyab Cultures, and the Ukraine (Markov 1961, 80; Markov and Khamrakuliev 1980, 68). However, Markov does consider prehistoric Oyukli to be contemporaneous with Djebel layers 4 and 5 (1961, 82). Unfortunately, Markov gives no backing whatsoever for this statement. He concludes by suggesting a seasonalround subsistence system where Oyukli was used as a springtime stockbreeding/hunting camp because of the lack of nearby available water at other times of the year (Markov 1961, 82). Markov suggests that other Oyukli Culture camps were located on the banks of the Uzboi river and synthesises his findings in attributing the material remains found at Oyukli to a large nonindigenous semi-nomadic Neolithic hunting, fishing and stockbreeding tribe (1961, 82). It is interesting that Markov insists that the Oyukli Culture is more advanced than its contemporaneous variants and that it could therefore not have developed in situ. It appears that this hypothesis will be difficult to verify or refute based on the deflated nature of the site. The area of the south Pribalkhan contains a high density of nomadic campsites and stretches to the southwest of the Bolshoi Balkhan mountains, including the mouth of the Uzboi, and the courses of the Aktam and Kelkyor rivers. Markov reports on Fedorov’s previous research in this region. Fedorov hypothesised (Markov 1971, 46) that a large estuary environment was formed by the confluence of the waters of the Uzboi and the Kelkyor rivers which created an ideal location for the easy exploitation of flora and fauna. Fedorov concludes that the inhabitants of the south Pribalkhan practised hunting, gathering, fishing, and stockbreeding.

Figure 3.21 Bolshoi Balkhan & Uzboi river (after Tolstov 1960)

45

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.22 Upper and Lower Uzboi river artefactual assemblagem (after Ribakov 1996)

46

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.23 Oyukli artefactual assemblage (after Ribakov 1996).

47

EXCAVATION METHODOLOGY AND RESULTS

The primary site in the south Pribalkhan is Mollakara, located on the western slope of the Kelkyor river. Markov attributes Mollakara to the Neolithic and considers the site to be contemporaneous with Sengradji layer 6, Djebel layer 5 and Oyukli, based on the stone tool and ceramic (both wheel-turned and coilconstruction) assemblages (Markov 1971, 48). Markov also documents three unnamed hunting and fishing camps near Mollakara, all of which contained bell beaker-shaped pottery with an unusual flat-bottom, sand-temper, red slip and grey/black paste (Markov 1971, 48). The campsites of the Aktam and Kelkyor riverine area represent small nomadic hunting and fishing campsites used as part of a seasonal-round mobile Keltiminar-related subsistence adaptation.

shore of the ancient Caspian Sea. This is evidenced by the ancient terraces and beachfront water-rounded stones on the slopes to the south of the cave (Okladnikov, 1956, 14). Djebel is presently situated between 64-85 metres above sea-level. Okladnikov creates quite a complex and occasionally confusing picture of the chronology and stratigraphic sequence at Djebel. He makes many useful cultural analogies but unfortunately contradicts himself on quite significant issues like economy or degree of sedentism. In his synthesis, there is a resounding undertone that this important type-site lacks a secure radiocarbon chronology, and thus a multitude of relative chronologies are proposed. I visited the site with the British team in the spring of 1997, but no prehistoric soil or cultural material whatsoever remains in the cave. There is a significant amount of modern-day debris. Unfortunately, we were unable to find any remaining archaeological surfaces or profiles that could yield material for radiocarbon dating.

The Bolshoi Balkhan mountains (Figure 3.24)

Although Okladnikov’s published field notes are extremely lengthy and detailed, I will attempt to summarise the main points. Okladnikov provides a recount of the habitation sequence at the cave, suggesting that the site was inhabited on a “permanent basis” from the Mesolithic to the early Bronze Age (1956, 196). This conclusion is based on the progression of the stone tool industry through the cultural layers at the site and the cultural traditions which supposedly developed at the site and which form the basis of the Caspian variant for the southern Turkmenian prehistoric sequence. Okladnikov suggests a kind of “ethnic unity” among the inhabitants of the Bolshoi Balkhan cavesites and the wider Keltiminar-related groups from the Krasnovodsk peninsula, the Bolshoi Balkhan, and the upper and lower Uzboi river area (1956, 196). Okladnikov interprets the site as a seasonal hunting camp used over thousands of years because of the natural convenient shelter it provided, but not inhabited on a constant basis due to the environmental restrictions which required a seasonal-round type of subsistence system (Okladnikov 1956, 11). The cavesite contains ten main cultural layers, 1-8, 5a, and 5-6, numbered from the top down. Okladnikov interpreted layers 7 and 8 as representative of the aceramic Mesolithic and compared them with the European Tardenoise Culture (northern France) and other unnamed Mesolithic variants in Jutland, Denmark (presumably Ertebølle). The Mesolithic economy was based on hunting and gathering, with the possible domestication of the dog (Okladnikov 1956, 198). Okladnikov stresses that the transitional layers 5a, 5-6, and 6 are crucial to understanding the MesolithicNeolithic transition and chronological sequence at the site, and to how the Djebel sequence correlates with the Kopet Dag sequence. Layers 5a, 5-6 and 6 represent the transitional period and the shift from the Mesolithic to

Figure 3.24 Bolshoi Balkhan Djebel (Figures 3.26, 3.27. 3.28, 3.29, 3.30, and 3.31) is the type-site for the Caspian area Keltiminar-related groups. Djebel is located on the southern escarpment of the Bolshoi Balkhan mountains near the modern day settlement of Djebel, west of Nebit Dag on the Ashkhabad-Krasnovodsk railway line. The site was excavated from 1947-1950 by a team directed by A. P. Okladnikov of Soviet Academy of Sciences, Novosibirsk, in conjunction with IuTAKE. The site is unique in its long stratigraphic sequence, spanning from the Mesolithic to the early Bronze Age. During the time of its prehistoric inhabitation, Djebel was located on the 48

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.26 Djebel artefactual assemblage (after Ribakov 1996)

49

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.27 Djebel site stratigraphy (after Okladnikov 1956)

Figure 3.29 Djebel cave plan view (after Okladnikov 1956)

Figure 3.28 Djebel cave profile (after Okladnikov 1956)

50

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 3.30 Djebel lithic assemblage (after Ribakov 1996)

51

EXCAVATION METHODOLOGY AND RESULTS

Caspian and the Uzboi. The suggestion that wild grain was collected is evidenced by grinding stones found in layers 1 and 2 (Okladnikov 1956, 201). Okladnikov alludes to contacts with the Kopet Dag agriculturalists, the exchange of ideas and material items, but does not say explicitly whether he thinks that domestication of wild grain was attempted in the Bolshoi Balkhan. He provides a detailed explanation of the prehistoric environmental conditions, including the nearby water resources of the Uzboi and the Caspian Sea in transgression. Okladnikov suggests that the Caspian gulf near Djebel was a large freshwater environment constantly fed by the water flowing from the delta of the ancient Uzboi into the Caspian near the site, creating an ideal area for collecting water, fishing and hunting waterfowl (1956, 203). Layers 1 and 2, attributed to the late Neolithic/early Bronze Age, were very dry, non-compacted, ashy layers differentiated by colour rather than texture or contents. Layer 1 was blue and extremely loose while layer 2 was darker in colour and relatively more compacted. The uppermost cultural layers were basically level with a slight inclination to the north. Both layers contained hearth stones, grinding stones, white, red, and black charcoal, and abundant modern-day sheep dung (Okladnikov 1956, 19).

Figure 3.31 Djebel

Layer 3, also attributed to the late Neolithic/early Bronze Age, was identified as “detritus”, and Okladnikov noted that the soil was typical of archaeological sites, as opposed to the abundant sheep dung ashy layers above. Layer 3 was muddy yellow with streaks of brown and contained a turquoise bead, ceramics, stone tools, shells, and the remains of burned saksaul in the hearths (Okladnikov 1956, 20). Layers 1-3, however, are quite distinct, not only because of the artefactual assemblage within them, but because of their ashy dung consistency. Okladnikov suggests that the differentiation of the type of deposit, namely massive amounts of ashy animal dung, points to the advent of stockbreeding, and places this shift chronologically somewhere between the late Neolithic and early Bronze Age. He provides analogies between the polished, wheel-turned, greyware found in layers 13 with assemblages from Shah Tepe in Iran, Tell Hassuna in Iraq, and the Eastern European Tripolian Culture (Okladnikov 1956, 201). Okladnikov gives a relative date for layers 1-3, based on radiocarbon dating carried out at Shah Tepe by T. Arne, for the end of the 4th millennium BC (3200 BC) (Okladnikov 1956, 204205). Okladnikov also provides analogies for the stone tool assemblage, mainly arrowheads, from layers 1-3 with Tell Halaf VII-VIII, Tepe Hissar (3500 BC), the lower levels at Anau north mound, Ust-Narim, the Pribaikal region, and the broader Neolithic of northern Eurasia (Okladnikov 1956, 205-206).

the early Neolithic with the characteristic advent of pottery production, and the use of geometrical microliths and asymmetrical triangles (Okladnikov 1956, 197). Layers 4 and 5 represent the early Neolithic, characterised by symmetrical tools and rounded- or pointed-bottomed vessels analogous to assemblages found in both the northern and southern European Neolithic (Okladnikov 1956, 200). The Neolithic economy at Djebel retained some of the earlier Mesolithic stone tool making traditions, and included hunting, gathering and fishing as the main subsistence systems. Layers 1-3, however, are quite distinct, not only because of the artefactual assemblage within them, but because of their ashy dung consistency. Okladnikov suggests that the differentiation of the type of deposit, namely massive amounts of ashy animal dung, points to the advent of stockbreeding, and places this shift chronologically somewhere between the late Neolithic and early Bronze Age. He provides analogies with Shah Tepe in Iran and the Eastern European Tripolian Culture, citing similarities in stone tool industries as well as ceramic forms (Okladnikov 1956, 201). Thus, Okladnikov interprets the economy of the late Neolithic and early Bronze Age as characterised by caprine pastoralism, hunting, fishing, the collection of wild grain, and a primitive network of trade and exchange. The species exploited included goitre gazelle, urial sheep, bezoar goat, and wild ass. The remains of fish bones found Djebel, Dam Dam Cheshme I and II attest to the supplementation of hunted game with fish from the

Layers 4 and 5 were thick, pure, clean ashy layers with hearths. The colour of layers 4 and 5 varied from light 52

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

grey to dark brown. The hearths in these layers contained hearthstones and charcoal, other artefactual remains included stone tools (geometric microliths), ceramics, and a shell bead (Okladnikov 1956, 20). Layers 4 and 5 represent the early Neolithic, characterised by symmetrical tools and rounded- or pointed-bottomed ceramic vessels analogous to assemblages found in both the northern and southern European Neolithic (Okladnikov 1956, 200). The Neolithic economy at Djebel retained some of the earlier Mesolithic stone tool making traditions, and included hunting, gathering and fishing as the main subsistence systems. Okladnikov makes various chronological analogies to sites in the Pribalkhan region, including the upper and lower Uzboi sites as well as Djanbas in the Akcha Darya delta. Okladnikov also reports that Formozov suggested similarities between the lithic assemblages from the Neolithic layers at Djebel with sites in western Kazakhstan (Okladnikov 1956, 208). Finally, Okladnikov states that layer 5 at Djebel was contemporaneous with the early farming (Jeitun) culture of the Kopet Dag piedmont and the Neolithic inhabitation of Gar-i Kamarband (Belt Cave) in northern Iran. Gar-i Kamarband evidences the same Mesolithic/Neolithic transition which proved so vital to the understanding of the Djebel sequence, and so elusive elsewhere on the piedmont (Okladnikov 1956, 210).

economy was initially a simple hunter/gatherer subsistence regime, supplemented in the early Neolithic by fishing, and subsequently in the late Neolithic/early Bronze Age by caprine pastoralism and the collection of wild grain (Okladnikov 1956). In summarising his field notes for the excavations at Djebel, Okladnikov contradicts an initial statement that the cave was inhabited “permanently” (1956, 196) in saying “the inhabitants of Djebel did not live permanently in the cave, but stayed from time to time and led a semi-nomadic way of life” (1956, 202). This discrepancy, based on direct translations of the original Russian text, is difficult to reconcile. It seems likely that Djebel was one of several campsites used in a seasonalround subsistence system spanning the chronological range represented at the site, and that his first reference pertains to the overall chronological sequence, spanning several thousand years, as relatively continuous. The Bolshoi Balkhan mountains: Dam Dam Cheshme I The sites of Dam Dam Cheshme I and II have long stratigraphic occupational sequences containing levels contemporaneous with the Jeitun Culture Neolithic sequence. Dam Dam Cheshme I and II were excavated by Okladnikov during the 1950s, Markov in the 1960s, Khamrakuliev in the 1970s and the British team in 1997. These cave sites are critical to the understanding of the Caspian area Keltiminar-related groups. The British team led by Harris and Gosden (including myself) conducted test excavations in small trenches in both Dam Dam Cheshme I and II together with Kurbansakhatov, Khamrakuliev and Zavyalov. The main aim of the British excavations was to locate and excavate any remaining undisturbed areas in the caves (which contained evidence of the long stratigraphic sequence previously reported by Soviet investigators) in attempts to obtain samples for radiocarbon dating to determine a secure absolute chronology. Markov (1966a, 1966b) recounted a stratigraphic sequence from Mesolithic to early Bronze Age at Dam Dam Cheshme II, which if found undisturbed could provide necessary insight into the understanding of the Mesolithic/Neolithic transition of southern Central Asia. Unfortunately, it became evident during our fieldwork that the Soviets had excavated the caves sites almost in their entirety. We were able to locate one small section on the northwest side of Dam Dam Cheshme II with a partly unexcavated stratigraphic profile. Samples for radiocarbon dating were collected and the results of the field-season in the Bolshoi Balkhan are currently in preparation by Harris et al..

Layers 5a and 5-6 were so-called “transitional” layers, representing the late Mesolithic and early Neolithic, with the same general texture and consistency as layer 5, but more limestone detritus. These layers contained geometrical microliths, asymmetrical triangles, ceramics, hearths, and a shell bead (Okladnikov 1956, 21). Okladnikov stresses that the transitional layers 5a, 5-6, and 6 are crucial to understanding the Mesolithic-Neolithic transition and chronological sequence at the site, and to how the Djebel sequence correlates with the Kopet Dag sequence as well as contemporaneous sites in the Akcha Darya delta (Djanbas). Layers 5a, 5-6 and 6 represent the transitional period and the shift from the Mesolithic to the early Neolithic with the characteristic advent of pottery production, and the use of geometrical microliths and asymmetrical triangles (Okladnikov 1956, 197). Okladnikov makes further analogies with lithic assemblages of the Ural and Siberian Neolithic (Okladnikov 1956, 209). Layer 6, attributed to the late Mesolithic/early Neolithic, was another ashy layer with ceramics and trapezes. Layers 7 and 8, attributed to the aceramic Mesolithic, were composed of “fine detritus” in various areas of the cave, but not covering the entire cave. In layers 7 and 8 there were stone tools and one fragment of ceramic interpreted as intrusive (Okladnikov 1956, 21, 198).

The cavesites of Dam Dam Cheshme I and II are located near Djebel, on the southern face of the Bolshoi Balkhan mountains to the east of the modern-day town of Nebit Dag. Dam Dam Cheshme I (Figures 3.32 and 3.34) was excavated by G. E. Markov of Moscow State University in collaboration with the Turkmen branch of the Soviet Academy of Sciences from 1963-1964. The site contains

Okladnikov outlines three main phases of occupation at Djebel, the Mesolithic, the early Neolithic, and the undistinguished late Neolithic/early Bronze Age. The 53

EXCAVATION METHODOLOGY AND RESULTS

a stratigraphic sequence from the upper Palaeolithic to the early Bronze Age (Markov 1981, 41). Markov identified five main cultural layers or occupational horizons at Dam Dam Cheshme I.

assemblages and because of the lack of a secure radiocarbon dating sequence. However, based on the stone tool industry and characteristics of ceramic production, Markov links Dam Dam Cheshme I to the Pribalkhan and lower Uzboi sites, but not to Dam Dam Cheshme II and Djebel, in terms of stone tools and ceramics (Markov, 1981, 51). Markov suggests that based on provenancing studies carried out on the lithic assemblage from Dam Dam Cheshme I, the tools were not made on site but at a workshop/hunting campsite in the Uzboi valley. Markov links layer 5 at Dam Dam Cheshme II to the late Mesolithic. He creates a further analogy between layer 4 at Dam Dam Cheshme I and layer 4 at Dam Dam Cheshme II, stating their synchronicity. The same applies for layer 3 at both sites, these are interpreted as dating to the Neolithic and early Aeneolithic (Markov, 1981, 51). Layer 2 at both Dam Dam Cheshme I and II are attributed to the late Aeneolithic and early Bronze Age. Finally, Markov provides comparisons between layers 2 and 3 at Dam Dam Cheshme I and II and the sites of Shah-Tepe in Iran and Ak-Tepe in southern Turkmenistan.

1= spring, 2 = runoff, 3 = cultural layers, 4 = extension of cave roof, 5 = rear overhang of cave, 6=limestone rockfall

Figure 3.32 Dam Dam Cheshme I cave plan view and site stratigraphy (after Markov 1981)

1= horizon 1, 2= horizon 2, 3= horizon 3, 4 = horizon 4, 5= horizon 5, 6= sterile clay, 7= bedrock

Figure 3.33 Dam Dam Cheshme I (after Markov 1981) Horizon 1 was 20 cm thick and made up entirely of sheep dung, however, Markov does not specify whether this was ancient sheep dung (as Okladnikov notes at Djebel) or modern sheep dung. Horizon 2 was approximately 80 cm thick and consisted of “terra-cotta” coloured clay with evidence of hearth fires, stone and flint tools, ceramics and metal remains (Markov 1981, 43). Horizon 3, 110 cm thick, consisted of “lightcoloured” sub-clay, limestone rockfall from the cave roof, flint tools and ceramics. Horizon 4, 160 cm thick, existed at the same level as sterile blue clay in other parts of the cave and was a “notably” compact, solid brown sub-clay with limestone fragments, evidence of campfires, hearthpits, ceramics and stone tools (Markov 1981, 43). Horizon 5 was identified in several areas of the cave, very sparse, and contained stone tools.

Figure 3.34 - Dam Dam Cheshme I Both of these sites are attributed to the 3rd millennium BC. Markov provides even broader generalisations in linking the Bolshoi Balkhan cavesites to contemporaneous developments in northern Iraq, southern Pricaspia, southeastern Iran, and with the “related cultures to the east” (Markov 1981, 53). Presumably in this statement, Markov is referring to the Keltiminar and Keltiminar-related groups of the Priaral,

Markov notes that constructing a relative chronology for Dam Dam Cheshme I and II was difficult. This was due in part to the bad preservation of the stone tool 54

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Amu Darya and Zeravshan deltas. Markov suggests that cultural contact and influence stretched readily eastward due to the lack of geographical boundaries that would otherwise have inhibited the movement of people and ideas (Markov 1981, 53).

discovered, makes excavation during the rainy season (winter and early spring) exceedingly difficult due to the periodic inundation of the cave which creates a deep swirling pool of sheep dung and dung beetles. The British team conducted surface collection on the slope immediately below the cave entrance and recovered a significant amount of archaeological remains from the cave.

The Bolshoi Balkhan mountains: Dam Dam Cheshme II The cavesite of Dam Dam Cheshme II (Figures 3.35 and 3.36) is located in the valley immediately to the east of the valley in which Dam Dam Cheshme is situated. Dam Dam Cheshme II was excavated in 1949 and 1950 by A. P. Okladnikov and subsequently by G. E. Markov and a team from the Institute of Archaeology of the Turkmen Academy of Sciences in 1963. The site is quite large, measuring 50 x 10 metres. A total of 23 sub-layers of 20 cm excavation spits were excavated (Markov 1966b, 85). Markov reports that “The overhanging rock-shelter above the middle part of the cave is washed by a waterfall which on rainy days flows from the roof (30 metres above) and washes out the cave surface and the slope which leads to the valley below” (1966a). This phenomenon is indeed true, and as the British team

Markov’s excavations (250 square metres) established that the cultural layers in the cave are approximately 3.2 metres thick. The deepest cultural and natural clay deposits in the cave were determined to be 5.2 metres thick on top of the cave bedrock (Markov 1966a). The archaeological stratigraphy consisted of nine cultural layers. Layer 1 was a dust-like mixture of ash and rotten sheep dung, 25 cm thick. Layer 2 was primarily clay sub-soil and mineral salts, measuring 20 cm in thickness, and indicative of climatic change (Markov 1966b, 85). Layers 1 and 2 contained numerous fragments of flatbottomed Bronze Age-type ceramics and stone tools, therefore Markov dates the two uppermost layers to the Bronze Age (1966b, 86).

Figure 3.35 Dam Dam Cheshme II cave plan view (after Markov 1966)

55

EXCAVATION METHODOLOGY AND RESULTS

Figure 3.36 Dam Dam Cheshme II

56

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Layer 3 is 20-40 cm in thickness and separated into two distinct sub-layers. The upper part of layer 3 is interpreted as dating to the Aeneolithic while the lower part of layer 3 is interpreted as late Neolithic. These conclusions are based entirely on typological dating of the geometrical microlithic industry in comparison with the stone tool assemblages recovered from Gar-i Kamarband (Belt Cave) in northern Iran and Djebel layer 4 (Markov 1966b, 88). The only absolute dating that enters into this picture is from Gar-i Kamarband. Markov relies entirely on typological comparison of assemblages to conclude contemporaneity and to create a chronological sequence at Dam Dam Cheshme II.

interpreted as belonging to the “same group of hunters” (Markov 1966b). Layer 8 was similar in composition to layer 7, but with finer-grained clay particles and an overall depth of 40 cm (Markov 1966b). Layer 9 is representative of the lowest cave level containing cultural remains. It was deposited on a thin sterile sublayer of bright yellow sand mixed with grey-blue sand probably weathered from the cave bedrock. The economy of the inhabitants of Dam Dam Cheshme II changed from the initial hunting, gathering, fishing and possible domestication of animal species of layers 69 (late Mesolithic-early Neolithic), to the hunting and productive economy of layers 4-5 (Neolithic), and concluded in the Bronze Age (layers 1-3), with hunting, fishing, stockbreeding and possible evidence of domestication (grinding stones from layer 2 pointing at least to collection) of wild grain species (Markov 1966b, 91). Markov concludes that the cultural layers at Dam Dam Cheshme II show evidence for both the transitions to a stockbreeding economy and later to the steppe nomadic economy evidenced throughout the regions to the north as far as the Urals and the Siberian steppe (Markov 1966b, 91).

Markov reports that dating layers 4-8 was extremely difficult and concludes that the only remaining link he can make with great confidence is that between the ceramics of layer 5 at Dam Dam Cheshme II with similar finds at Gar-i Kamarband layer 12 and Djebel layer 7 (Markov 1966b, 88). Layer 4 was separated from layer 3 by a compact mineralised sub-layer. Layer 4 was separated into five sub-layers which differed from one another in the colour and consistency of the constituent red or brown clays (Markov 1966b). The average depth of layer 4 was 60 cm, but in places it reaches up to 75-80 cm overall thickness. Layer 4 contained stone tools, ceramics, bone needles, and numerous perforated shells. Layer 5, 10 cm thick, was also separated from layer 4 by a mineralised sub-layer. It consisted of large-sized granular reddish sandy clay. Layer 5 contained numerous stone tools, the details of which are presented with unremitting accuracy by Markov (1966a). Layer 6, 15 cm thick, was a grey-green clay layer containing a significant amount of organic remains and stone tools. Layer 7 was composed of reddish sandy clay and 25 cm in thickness and evidenced a significant decrease in the size and variation of the stone tool assemblage. Layers 8 and 9 contained very few stone tools and based on the “kind and colour of the flint, and the product” are

The above site descriptions for the Jeitun, Keltiminar, and Keltiminar-related sites of prehistoric Turkmenia serve to place the ceramic case study which follows within a defined archaeological context. Chapters 4 (typology) and 6 (petrography) will provide an analysis of the ceramic assemblages from the Jeitun Culture sites outlined above. The results presented in Chapters 4 and 6 will be used to outline the structure of the Neolithic, specifically the mode of production, including aspects of regional and temporal variation. This Jeitun Culture structure will then be compared with what is known of the Keltiminar Culture structure, based on the pertinent archaeological literature and the site descriptions outlined above.

57

CHAPTER 4 CERAMIC TYPOLOGY The first three chapters of this thesis provide background information pertinent to the study of the Neolithic period of southern Turkmenia, and more specifically for the ceramic case study. Although the main focus of this ceramic analysis is ceramic petrography, it is ultimately crucial to provide a well-rounded study of the ceramic assemblages in question, incorporating previous Soviet ceramic typological data. Thus, the incorporation of the current chapter, which overviews form, function, and decoration. In this chapter, the Soviet ceramic typologies for ceramic ware and vessel form will be outlined. Initially, I will review form and decoration by cultural phase. This will include detailed overviews for the ceramic assemblages of the sites represented in the petrographic case study as well as those documented in the archaeological literature for which no samples were available. A discussion of form with relation to function follows. This illuminates issues of technological complexity and societal preference. Next, statistical analyses based on form, consistency, and ware distributions for Jeitun, Chopan, Togolok, Pessedjik. Chagylly, Chakmakli, and Mondjukli will be presented. Finally, I will identify the research questions which point to the necessity to conduct a petrographic case study.

and Bronze Age (Anau 1A, Namazga I-IV) assemblages from Mondjukli, Obadan, Kashut, Dashlidji, Geoksyur, Altyn, Khapuz, and Ulug will also be reviewed. Form and Decoration: Jeitun Culture phases Berdiev (1969) separated the Jeitun Culture Neolithic period into three main phases based on the development of the stone tool industries and ceramic assemblages: phase 1 (the early Neolithic), phase 2 (the middle Neolithic), and phase 3 (the late Neolithic). He further distinguished two sub-divisions of phase 1, sub-phases 1A and 1B, and acknowledged the possibility for future modification of the scheme following further archaeological investigation (Berdiev 1969, 37). It is necessary to clarify that not only was the “Jeitun Culture” given a series of phases (i.e. Jeitun Culture phase 1, 2, and 3), but each site was also categorised by Soviet investigators in terms of occupational horizon (numbered from the bottom up) and phase (i.e. Chopan phases 1 and 2, Bami phases 1 and 2, etc.). The phases for each site do not correspond to the phases for the Jeitun Culture as a whole, adding to the potential confusion. It is potentially easy to mistake Bami occupational horizon 1 for Bami phase 1, unless this differentiation is spelled out explicitly. I will therefore endeavour to provide as much detail as possible in keeping with the original Soviet categorisations, when referencing levels, occupational horizons, and phases.

The form and ware categorisation scheme combines previous typologies, regarded as a baseline, with additional examples from the current study. O. K. Berdiev (1969) and V. M. Masson (1971) established the following basic typologies for form and decoration for the Neolithic Jeitun Culture of the Kopet Dag piedmont (Figures 4.1, 4.2, 4.3, 4.4, and 4.5). Figures 4.6 and 4.7 represent ceramic form and decoration comparisons with assemblages from the Near East and the Sialk 1 complex of Iran. All other illustrations of ceramic form and decoration from specific site assemblages may be found in appendix 2.

Jeitun Culture Phase 1A The Jeitun 1A phase, represented in occupational horizons 1-3 at Jeitun and Chopan and occupational horizon 1 at Togolok, contained ceramics of three main vessel forms. These are: 1) cylindrical vases, 2) bowls, and 3) four-cornered vessels of the so-called “saladnitsa” type. The term “saladnitsa” was coined by Soviet investigators due to the similarity of this type of vessel to a modern-day salad bowl or salad server. For the sake of continuity, terminology consistent with the direct English translations of the original Russian text will be used. The main ceramic ware reported by Berdiev (1969, 38) during the Jeitun 1A phase was buffware, either decorated or undecorated. If decorated, the vessels had simple motifs of 1) frequent horizontal wavy lines, 2) frequent horizontal straight lines, or the so-called 3) “bracket” motif, all painted in dark red.

It is important to note that, in certain instances, I will not include Masson’s interpretations of form and decoration because I find that they do not add scientifically based information, and in some cases actually detract from Berdiev’s interpretations by creating confusion. Masson’s use of unconventional and non-scientific colour names for the description of ceramics with terms like dark chestnut, raspberry, and yellow-ginger, or the renaming of already complicated but descriptively accurate names for design motifs using terminology like “slanting stripe thunderbolt variant” seems unnecessary and impertinent.

Jeitun Culture Phase 1B Y. E. Berezhkin’s statistical attempts to classify the Jeitun Culture ceramic forms and F. L. Hiong’s technological and mineralogical analyses of Aeneolithic

The Jeitun Culture 1B phase is represented in the uppermost occupational horizon at Jeitun, occupational 59

CERAMIC TYPOLOGY

Figure 4.1 Jeitun Culture ceramic motifs by zone and cultural phase (after Berdiev 1969)

Figure 4.2 Jeitun Culture ceramic vessel forms (after Berdiev 1969) 60

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 4.3 Jeitun Culture ceramic assemblages by phase and site (Ribakov 1996) Late Jeitun phase (1-25); middle Jeitun phase (26-40); early Jeitun phase (41-54); Bami (1-16); Chagylly (17-25); Togolok (26, 28, 31, 32); Pessedjik (27, 29, 30, 33); Chopan (30-40); Jeitun (41-54)

61

CERAMIC TYPOLOGY

Figure 4.4 Jeitun Culture ceramic motifs (after Masson 1971) horizons 4-5 at Chopan, and occupational horizon 2 at Togolok. Berdiev (1969, 38) distinguishes the Jeitun Culture 1B phase from the Jeitun Culture 1A phase due to the appearance of new vessel types in addition to types 1, 2, and 3 mentioned above, including 4) jars and 5) dishes. Ceramic ware and decoration similarly exemplifies increased variety with the addition of redware vessels. Berdiev also suggests that from the Jeitun Culture 1B phase through the end of the Jeitun Culture 2 phase, redware dominated the assemblage. This trend was then followed during the Jeitun Culture phase 3 by a resurgence of buffware. There is, however, some inherent ambiguity in Berdiev’s categorisation. It is unclear in most cases whether his

descriptions of “buff” and “red” refer to the actual ceramic fabric, or to the slip applied to the surface of the vessel. The slip colour of a vessel often differs from the fabric colour itself. At present, there is no way to reconcile this ambiguity and the related discrepancies in ceramic data. It proved impossible to find Berdiev’s original field notes. Most of the ceramic assemblages originally housed at the Institute of Archaeology in Ashkhabad were destroyed due to a lack of storage space, therefore it is also impossible to check the original assemblages. The assemblage was almost exclusively redware with red/dark brown painted decoration. The motifs used included those mentioned above typical of Jeitun Culture phase 1A with the 62

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 4.5 Jeitun Culture ceramic motifs by phase (after Berdiev 1969)

Figure 4.6 Neolithic artefactual assemblages from Turkmenia and the Near East (after Berdiev 1969) 63

CERAMIC TYPOLOGY

Figure 4.7 Comparison of Jeitun Culture and Sialk 1 ceramic motifs (after Masson 1971) addition of three new types: 4) frequent horizontal wavy lines with occasional crossing vertical lines, 5) frequent horizontal straight lines with occasional crossing vertical lines, and 6) a dot motif. Berdiev (1969, 38) interprets the motifs of Jeitun Culture phase 1B as a transitional stage from the more “realistic” patterns of the early Neolithic to the increasingly “geometric” patterns of the middle and late Neolithic.

Jeitun Culture Phase 2 Jeitun Culture Phase 2 (middle Neolithic) is represented by Chopan phase 2, Togolok phase 2, Bami phase 1, Chagylly phase 1, Mondjukli phase 1, New Nisa, Kelyata, Kepele, Kantar, and Naiza (Berdiev 1969, 40). Berdiev also notes that if Jeitun was inhabited during the middle Neolithic, evidence of this occupational continuity

64

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

was destroyed by site deflation and was therefore unrecognisable when excavation commenced (1969, 41). Berdiev provides an analysis of the new aspects of form, ware, and decoration on a site by site basis. Unfortunately, he often focuses solely on decorative motifs and omits any reference to form or ware (fabric). Presumably, this omission implies that the vessels were redware based on his general statement of the increase in redware and decrease in buffware during the middle Neolithic.

Jeitun Culture phase 2) and buffware (characteristic of Jeitun Culture phase 1) were found. Painted decoration remained consistently red or dark brown and decorative motifs included: 1) the triangle motif in “chess order”, 2) frequent horizontal straight lines with occasional crossing vertical lines, 3) frequent vertical straight lines with occasional crossing horizontal lines, 4) the socalled “inclined steps” motif, and 5) the triangle motif with broken lines (Berdiev 1969, 46). In particular, Berdiev notes the continuity of some motifs from Jeitun Culture phase 2, with the addition of several new motifs.

Large rounded redware storage jars appear in abundance at Chopan. It is likely that these were made in the ceramic workshop or “production location” identified at the site. The painted decoration is dark brown with both new and old motifs including 1) frequent vertical straight lines with occasional crossing horizontal lines, 2) frequent horizontal straight lines with occasional crossing vertical lines, 3) the dot motif, 4) the so-called diagonally inclined “grid pattern”, and 5) the triangle motif (Berdiev 1969, 40). Berdiev reports that the upper two occupational horizons at Togolok contained similar assemblages and were probably contemporary with Chopan.

In the upper occupational horizons at Chagylly (Chagylly phase 2), Berdiev unearthed ceramics with Jeitun Culture phase 1 and 2 style motifs. He attributes this evidence not to site disturbance and inherent stratigraphic mixing, but to a revival of ancient motifs among the Jeitun Culture phase 3 inhabitants (1969, 46). This conclusion strikes me as unlikely, but without the use of context excavation or the ability to verify exact provenance, it is of course impossible to refute or substantiate this claim. Berdiev identifies the following motifs: 1) the dot motif, 2) frequent horizontal wavy lines with intermittent sets of crossing straight vertical lines, 3) a horizontal band motif around the vessel rim (which continues into later time periods), and 4) the socalled “tree” motif, also found in Mondjukli phase 2 and continuing into the Anau 1A phase (1969, 47-48). Berdiev delineates Jeitun Culture phase 3 (late Neolithic) as a time of major transition. Architecturally, the proto-blocks consistently used for construction at Jeitun and elsewhere become “proper” rectangular mudbricks during Jeitun Culture phase 3. This transformation in building technology was also found at early Bami and later Chopan. Berdiev hypothesises a direct link, previously refuted by Marushchenko but supported by Masson and Kuftin, between the Jeitun Culture and the subsequent Anau 1A Culture (1969, 5153).

At New Nisa, motifs included: 1) frequent vertical straight lines with occasional crossing horizontal lines and 2) the diagonally inclined “grid pattern”. The vessels of Bami phase 1 (occupational horizons 1-2) are similar to the vases of Chopan phase 2 (occupational horizons 45). Decoration motifs from Bami occupational horizons 1-2 are limited to 1) frequent vertical lines with occasional crossing horizontal lines, 2) the triangle motif similar to that of Chopan phase 2, 3) frequent horizontal and vertical straight crossing lines, 4) frequent horizontal wavy lines with crossing straight vertical lines, and 5) the dot motif (Berdiev 1969, 41). Berdiev relates Chagylly phase 1 and Mondjukli phase 1 to Jeitun Culture phase 2 (middle Neolithic) and reports the following decorative ceramic motifs: 1) frequent vertical straight lines with occasional crossing horizontal lines, 2) a motif consisting of two crossed lines found at Chagylly, 3) a vase with horizontal arched lines (unique specimen among Jeitun Culture assemblages), and 4) the triangle motif, which was found only at Mondjukli (1969, 43). Berdiev attributes the lower levels of Mondjukli to Jeitun Culture phase 2, due to the existence of this triangular decoration motif (1969, 43).

Masson (1971), however, provides a more simplified typology in comparison to Berdiev’s categorisation above. Masson identifies early, middle and late Jeitun Culture phases and divides the ceramic assemblages into undecorated and decorated categories of vessel types. In the undecorated category there were bowls and jars, while in the decorated category there were bowls, small jars, and large jars (or vases). He likens these major vessel forms to those found in the Near East, especially at the sites of Hassuna and Jarmo (1971, 39). Masson outlines seven major categories of design motifs for decorated vessels, including most already proposed by Berdiev. As mentioned at the beginning of this chapter, Masson’s motif categories will not be detailed because they contribute confusion to Berdiev’s previous work. In fact, Masson (1971, 36-37) chose to omit several motif categories altogether in his outline of the related assemblages of other piedmont sites.

Jeitun Culture Phase 3 Jeitun Culture phase 3 encompasses Bami phase 2 (occupational horizons 3-5), Chagylly phase 2 (upper 3 occupational horizons), and Mondjukli phase 2 (middle occupational horizon) (Berdiev 1969, 45). Again, Berdiev recounts the trends for Jeitun Culture phase 3 site by site. At Bami, both redware (characteristic of

65

CERAMIC TYPOLOGY Ceramic Assemblages by site

Togolok

Jeitun

Berdiev presents a detailed description of the ceramics found at Togolok (appendix 2, 326-330), comparing them typologically to the assemblages found at Jeitun and in the lower levels of Chopan (1964c, 273). He notes characteristic similarities in form and decoration and a generally high percentage (more than 50%) of decorated vessels and vessel fragments in the upper three Neolithic occupational horizons. The ceramics from the lowest of the four occupational levels (called horizon 1) were undecorated, a fact which led Berdiev to attribute the Neolithic levels of Togolok to Jeitun Culture phases 1 and 2 (1964c, 275). Berdiev describes the ceramics as consisting of a redware and a buff (or yellow) ware, in some cases with a light or dark red slip and dark reddishbrown painted decoration (1968b, 13). Variations in decoration motifs found in horizon 2 included frequent wavy vertical lines and frequent straight vertical lines with occasional crossing straight horizontal lines. In occupational horizon 3, this vertical with intersecting horizontal variant persisted while the wavy arching line variant diminished. In the uppermost horizon 4, the wavy line variant completely disappeared and the assemblage was dominated by the frequent vertical with crossing horizontal straight lines variant (Berdiev 1964c, 275). The forms of undecorated and decorated vessels included flat-bottomed cups with narrowing bases, flat-bottomed thin-walled plates, and large storage vessels (Berdiev 1968b, 14; Berdiev 1964c, 273-274).

Masson constructed the following outline of form and decoration for the ceramics at Jeitun (1971). The main forms (appendix 2, 306-320), as mentioned above, were decorated or undecorated bowls, dishes, and jars of varying size. The undecorated vessels from Jeitun were a “muddy red” or “brown muddy” colour, while the decorated vessels either had a “light terracotta” or “yellow ginger” background with “dark brown”, “chestnut”, “raspberry”, or “dark chestnut” decoration (Masson, 1971, 37). The decoration motifs found at Jeitun (appendix 2, 306-320) were predominantly variants of the 1) wavy line design, which was usually found around the vessel shoulder, rim, or base. The 2) “bracket ornament” varied from geometrically organised to free form and later morphed into the 3) “net design” and subsequent “cellular pattern” (Masson 1971, 36). Other versions of the “bracket ornament” consisted of vertical rows of brackets crossed by horizontal bands. Originating from the “net design” is the 4) “slanting net design” (Masson 1971, 37). The remaining rarer motifs include the 5) “horizontal lines in bands design” which was either standard or with crossing vertical lines, the 6) “diagonal lines and bands design” with diagonal lines crossed by vertical lines on the exterior of a vessel and horizontal lines on the interior of the vessel, and 7) the “silhouette triangle motif”, which was seen as the basis for the later widespread Anau 1B and Namazga decorative motifs (Masson, 1971, 37).

Pessedjik

Chopan

There are four occupational horizons at Pessedjik forming a total cultural deposit of 3.5 metres with sterile soil beneath. Vessel forms included large and small bowls and vases, all flat-bottomed. There were both undecorated and decorated vessels, either red or buffware, and the motifs were painted in dark red (Lollekova 1978b, 180) (appendix 2, 331-342). Occupational horizon 1, although deflated, revealed Jeitun-style ceramics with the 1) frequent vertical lines with occasional crossing horizontal lines motif, a new variant of the 2) frequent horizontal lines with occasional crossing vertical lines motif, 3) the bracket design, 4) frequent wavy lines, and 5) the net design (Lollekova 1978b, 177). Occupational horizon 2 contained decorated ceramics, mostly redware, with the 6) vertical wavy line motif, the 7) side by side bracket motif, and the 8) frequent vertical and occasional crossing horizontal lines motif (Berdiev 1968c, 16; Lollekova 1978b, 178-179). In occupational horizon 3, Berdiev reports the recovery of Jeitun-type ceramics with the frequent vertical and occasional crossing horizontal lines motif and the side by side bracket motif (Berdiev 1968c, 16).

The ceramic assemblage from Chopan (appendix 2, 321325) contains typical Jeitun Culture pottery which was manufactured on-site in the southern area of the mound in 3 large hearths (Berdiev 1972a, 79). These hearths were associated with 5 small structures, presumably used for the preparation of the clay and temper, and the drying of the unfired vessels. This complex, interpreted as a pottery production area, is the only one of its kind among the Jeitun Culture sites. Berdiev separates the ceramics into two main sub-phases: Chopan 1 and Chopan 2. Chopan 1, the earlier of the two variants, is itself divided into two sub-phases Chopan 1A and Chopan 1B. The ceramics from Chopan 1A were characterised by two forms of vessels and cups with two types of painted decoration which Berdiev (1972a, 67) labels as 1) “stream painting” and 2) “cone ornamentation”. Chopan 1B ceramics were subdivided into vessels, vases, bases and jars. The decoration of the Chopan 1B ceramics was characterised by 1) frequent vertical straight lines with occasional crossing horizontal lines (Berdiev 1972a, 68). The Chopan 2 sub-phase contained the previously mentioned vessel forms with the addition of pots. The predominant decoration style shifted from line decoration to triangular ornamentation (Berdiev 1972a, 70).

New Nisa The ceramic assemblage at New Nisa included coarsely made thick-walled storage vessels (mainly cylindrical 66

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

vases) and well-made thin-walled domestic vessels (mainly bowls) (appendix 2, 343). These vessels were chaff-tempered, low-fired, slipped and painted, and typical for the Jeitun Culture (Berdiev 1965, 238). The main decorative motif was frequent vertical lines with occasional crossing horizontal lines typical of the upper levels of Chopan and Togolok, the lower levels of Bami and Chagylly, and the uppermost occupational level at Jeitun (Berdiev 1965, 239). Berdiev also points out one example of a rare motif, the so-called “rhomboid grid”, and suggests that this design was a predecessor of the later Anau 1A geometrical motifs.

typically chaff-tempered and slipped, with a surface colour of red, light brown, buff and very occasionally black. The texture was porous and a discontinuity between the colour of the vessel core and margins points to a low firing temperature. An open-firing method was probably used (Berdiev 1966, 12). Berdiev separates the ceramics from Chagylly into two main groups: Chagylly 1 (horizons 4-12), and Chagylly 2 (horizons 1-3). The vessels in the sub-group Chagylly 1 included flat-bottomed vases and bowls. The vases were cylindrical in form, had walls of 7-25 mm thickness, a rim diameter of 15-35 cm, and a base diameter of 10-30 cm. The motifs for the decorated vases included 1) vertical lines crossed with horizontal bands and 2) the vertical staircase or “lightning bolt” design that was typical for Chopan and Bami (Berdiev 1966, 12).

Kelyata, Kepele, Kantar, Naiza and Chakmak Dash Beik The campsites of Kelyata, Kepele, Kantar, Naiza, and Chakmak Dash Beik are attributed to the middle phase of the Jeitun Culture (Jeitun Culture phase 2) and are thought to have been inhabited for a short period of time due to the unsuitable nature of their environs and lack of water. Berdiev suggests that these sites may have been used after the abandonment of Jeitun. The sites were deflated and contained very thin material cultural layers consisting of typical Jeitun Culture ceramics (appendix 2, 344), figurines, pendants, stone tools, and hearths, but no architecture.

The bowls attributed to the Chagylly 1 complex included both large storage bowls and small domestic bowls. The large storage vessels were made of a coarse fabric and had a rim diameter of 60 cm and a base diameter of 35-40 cm. The smaller, thin-walled bowls had walls from 5-15 mm thick, a rim diameter of 30 cm and a base diameter of 15-22 cm. The motifs on the smaller domestic bowls that were decorated included 1) vertical straight lines, 2) the horizontal staircase design, and 3) wide horizontal bands (Berdiev 1966, 14).

Kelyata, discovered in 1931 by hydrologists, contained thick-walled chaff-tempered ceramic sherds, presumably from cylindrical vessels. The decoration was dark red on a red background and the motifs included: 1) frequent vertical lines with rare crossing horizontal bands and 2) vertical lines with crossing horizontal lines (Berdiev 1971c, 76). Kepele, southeast of Ashkhabad near the Kepele well, was discovered in 1941 by A. A. Marushchenko. It contained typical Jeitun-type ceramics with the vertical lines motif (Berdiev 1971c, 77). Kantar, near the site of Chopan and the Kantar well, was discovered by S. A. Ershov during his excavations at Chopan in 1952. There were Jeitun-type ceramics and stone tools recovered from Kantar (Berdiev 1971c, 79). The site of Naiza was discovered by Marushchenko in 1946 and is located between Bami and Kodj in the Kopet Dag foothills. Marushchenko recovered Jeitun-type ceramics with chaff temper, including a square “saladnitsa” type vessel (Berdiev 1971c, 78). Chakmak Dash Beik was discovered by a geologist in 1937, and first surveyed in 1944. The site is located 30 km northwest of Ashkhabad, and although attributed to the Jeitun Culture due to the majority of ceramics found, also contained Mesolithic stone tools, Kelyata-type ceramics, and Anau III ceramics (Berdiev 1971c, 78).

The Chagylly 2 ceramic complex, representative of the top three occupational levels at the site, had vases of cylindrical form and both large and small bowls. The decorative motifs on the vases included 1) vertical lines, 2) wavy lines, 3) vertical wavy lines, 4) the bracket motif, 5) horizontal straight lines, 6) horizontal bands beneath rare vertical lines, 7) triangles in chess order, and 8) the dot ornamentation reminiscent of the ceramic assemblages from Chopan, Bami, Sialk, Nineveh, and Hassuna (Berdiev 1966, 14). The large bowls, called tagora, had three main decorative motifs: 1) vertical lines in groups inside vertical bands, 2) horizontal wavy lines divided by vertical bands, and 3) the staircase ornament (Berdiev 1966, 14). The smaller and finer bowls for domestic use had six different decorative motifs. These were: the 1) wavy lines on both vessel interior and exterior, 2) vertical lines on the vessel exterior with horizontal bands on the vessel interior, 3) frequent vertical lines on the vessel exterior with rare wavy lines on the vessel interior, 4) horizontal bands with parallel staircase design on the vessel exterior and horizontal parallel lines on the vessel interior, 5) the tree ornamentation, and 6) the net design (Berdiev 1966, 15). These last two decorative motifs were typical of later Anau 1B or Namazga I ceramics.

Chagylly The ceramics from Chagylly (appendix 2, 345-346) are similar to those from Jeitun in terms of paste, firing, decoration, and surface treatment. More than 5000 vessel fragments were analysed from the site, only 100 of which were decorated (Berdiev 1966, 10). The ceramics were

In addition to ceramic vessels, there were also five examples of stone vessels excavated at Chagylly. These were made of highly-polished crystalline limestone or 67

CERAMIC TYPOLOGY alabaster and had a square form and straight walls similar to the so-called “saladnitsa” ceramic vessels. Berdiev makes analogies to similar square vessels from Mersin, Tepe Sarab, and round stone vessels from Jarmo (1966, 20).

the upper levels of Jeitun and the site of Chopan (Berdiev 1963a, 188-189). Bami horizon 2 contained two types of ceramic vessels, vases and bowls. The decoration on vases included the 1) frequent vertical with occasional crossing horizontal lines motif and 2) the triangle motif similar to ceramic decoration found in the upper levels at Jeitun and the sites of Chopan and Togolok (Berdiev 1963a, 189). Other motifs were 3) the rhomb design with dots in the centre, and 4) the arch design found on the interior of the vase rims. The decoration on bowls included 5) straight vertical lines on the exterior of the vessel with straight horizontal lines on the interior of the vessel, and 6) horizontal lines on both the inside and outside of the vessel (Berdiev 1963a, 190).

Chakmakli The ceramics from occupational levels 2-4 at Chakmakli (appendix 2, 347-354) were of the Anau 1A type: wellmade, thin-walled, sand rather than chaff-tempered, and slipped or burnished. The decoration included geometric designs, especially triangles. In occupational level 4, there were also Chagylly-type ceramics found, which led Berdiev (1968a) to suggest the coexistence of the Chagylly and Chakmakli communities during the period that level 4 was occupied. Kohl (1984, 69) interprets the existence of Chagylly-type ceramics in level 4 of Chakmakli as representative of the development of the Anau 1A Culture (represented by the Mondjukli and Chakmakli sub-phases) from the late Neolithic Jeitun Culture at the site of Chakmakli, rather than accepting Berdiev’s interpretation of the migration of the non-indigenous Chakmakli people from northern Iran.

Occupational horizon 3 at Bami contained ceramics that were predominantly “reddish-yellowish”, with motifs including 1) triangles one above another, 2) triangles in “chess order”, 3) frequent vertical lines with rare crossing horizontal lines, 4) horizontal lines with crossing vertical lines, 5) the staircase or “lightning bolt” motif, 6) a bowl with horizontal lines on both inside and outside, and 7) rare vertical lines on the interior of the vessel with rows of arches on the exterior of the vessel (Berdiev 1963a, 191).

Mondjukli Occupational levels 1-3 at Mondjukli contain Jeitun Culture ceramics (appendix 2, 355). Two types of vessel forms were identified for these Neolithic levels: vases and bowls. The vases were cylindrical and decorated whereas the bowls were either thick-walled and coarse or thin-walled, fine, and undecorated. One decorated square vessel with the triangle motif was recovered in the Neolithic levels of Mondjukli (Berdiev 1971c, 79). Level 4 (numbered from bottom to top), represents a transition from the Neolithic Jeitun Culture to the Aeneolithic Anau 1A Culture, with Jeitun Culture assemblages and house structures found together with more well-made ceramics. Level 6 contained an Anau 1A artefactual assemblage, including Anau 1A ceramics which were yellow and brown slipped, decorated with typical Anau 1A motifs of hatched and solid triangles. Level 7 contained more Anau 1A ceramics found together with vessels bearing decoration reminiscent of the parallel vertical line decoration typically found at Jeitun. Level 9 is interpreted as representative of the transition from Anau 1A to Anau 1B and contained some Namazga-type artefactual remains including ceramics (Berdiev 1972c, 1972d).

Occupational horizon 4 at Bami contained both vases and bowls. The motifs for the vases included 1) triangles one above another, 2) frequent vertical lines with rare crossing horizontal lines, 3) vessels decorated on the exterior with vertical lines and on the interior with horizontal lines, 4) the staircase or “lightning bolt” motif, 5) the broken line motif, and 6) upside-down triangles connected to vertical lines (Berdiev 1963a, 192). Interestingly, there were also undecorated vessels tempered with chamotte, a previously unused ceramic temper among the Jeitun Culture. The vessel with chamotte temper was flat-bottomed with a protruding foot, similar to vessels found at Jeitun (Berdiev 1963a, 192). Occupational level 5 at Bami contained ceramic with motifs including 1) frequent vertical lines, 2) the staircase motif, 3) the triangle motif, 4) vertical wavy lines, and 5) two crossing bands. The vessels were often decorated on both the interior and exterior, and the rims were typically decorated with arches (Berdiev 1963a, 192). The ceramic production technology for the assemblage at Bami was coil-construction, typical of the Jeitun Culture. The surface of most vessels was a red, dark red, or brown background with brown or very dark red painted decoration. The vessel surface was often stained with grey or black patches due to poor firing (Berdiev 1963a, 192). The ceramic assemblage at Bami is striking in its differentiation from assemblages of other Neolithic Kopet Dag piedmont sites. In addition to the typical Jeitun-style chaff-tempered painted ceramics, the assemblage also

Bami The site of Bami was separated into five occupational horizons numbered from bottom to top. Note that this convention is opposite to the norm. Ceramics were recovered from all five occupational layers (appendix 2, 356-358). Occupational horizon 1 contained vases and bowls with the 1) frequent parallel lines motif in dark brown on a light background, similar to ceramics from 68

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

contained grey and black shell-tempered incised pottery (Berdiev 1964c, 276). Grey and black shell-tempered incised wares were typical of the Keltiminar and Keltiminar-related groups. This occurrence points to Bami, in its position western-most among Jeitun Culture sites of the Kopet Dag piedmont, as having extensive ties with the Keltiminar or Keltiminar-related groups further west.

the cave as a shepherd’s lookout or picnic spot evidenced by abundant modern debris on the present-day cave surface (1956, 35). Vessel 2 was wheel-turned, red in colour, and contained chamotte temper. Vessel 3 was wheel-turned, light yellow and good quality. Vessel 4 was wheel-turned, well-made, black in colour and wellpolished. The core appeared light grey at the sherd edge. The rim was decorated and the vessel was flat-bottomed. Vessel 5 was roughly made of coil-construction with an uneven, rough surface, flared rim and rounded bottom. Vessel 6 was similarly black in colour with incised decoration in the hatches and lines motif, shell-tempered, poorly-fired, round-bottomed and friable (Okladnikov 1956, 36). Vessel 7 was black in colour, roundbottomed, and contained crushed quartz temper (Okladnikov 1956, 39).

Djebel The ceramic assemblage from Djebel (appendix 2, 359372) was categorised in two main ways: 1) as part of the overall artefactual assemblage documented by Okladnikov in the main body of the published site report, and 2) as a fabric and technology study conducted by A. N. Avgustinik and V. I. Baranova, published as a separate article in the same volume as the main site report. Although it would seem that there should have been more cooperation and communication between Okladnikov and the pottery researchers in terms of coordinating categorisation schemes and presenting data, I will attempt to present both sets of data as clearly as possible. Unfortunately, there is no system for the direct correlation of vessel fragments cited by Okladnikov or Avgustinik and Baranova.

The ceramics from layer 2 were more abundant than in layer 1 and may be divided into four basic groups. Type 1 consists of thin-walled wheel-turned, black, polished vessels with a solid compact dark grey fabric lacking any admixture. Type 2 was unpolished, roughly-made, shell or feldspar-tempered black coil pottery of oval or semiegg form with some evidence of handles. Type 3 was a very badly preserved, wheel-turned, polished, thinwalled, very well-fired redware vessel. Type 4, a large pot with no further description, was considered intrusive from layer 1 (Okladnikov 1956, 39).

The ceramic assemblage from Djebel was characterised by Keltiminar-type sand-tempered, incised, pointedbottomed vessels. This assemblage was representative of a completely separate cultural tradition, evidenced by the striking differentiation in fabric, form, technology and decoration between the Djebel ceramics and the typical Jeitun-type chaff-tempered, painted, flat-bottomed ceramics. While Okladnikov documents the forms, decoration, and to an extent, the fabrics for the Djebel ceramic assemblage, Avgustinik and Baranova document the ceramics solely in terms of fabric types. As mentioned, Okladnikov and Avgustinik and Baranova published their results in a format that does not make it easy or practical to present the data contiguously. Okladnikov presents the main site report in the order of excavation; i.e. from top to bottom. Avgustinik and Baranova, however, present their findings in terms of archaeological chronology, i.e. from bottom to top. I will keep both categorisation schemes separate because Avgustinik and Baranova created their fabric classification in terms of development, making it possible to see certain fabric types develop and then phase out. Obviously, working through the same material in two different ways is not ideal, and potentially leads to confusion because the reader is never presented with a discussion linking the two sets of data. Okladnikov’s findings will be presented first.

The ceramic assemblage from layer 3 was much more complex than that of the two upper layers, with six main ceramic types. Type 1 was a wheel-turned, black, flatbottomed vessel with 3-4 mm thick walls, of which seven fragments were found and reconstructed. Type 2 was a very large “dirty-yellow”, shell-tempered, roughly-made, friable, coil pot with a rounded form, polished body and narrow neck, and incised decoration in the hatched lines motif. Type 3 was a dark red, smooth-rimmed, narrow, polished, chamotte and sand-tempered, pointed-bottomed, coil vessel with a compact paste (Okladnikov 1956, 76). Type 4 were sherds of several thick-walled, roughlymade, polished, round-bodied, coil-type vessels with a solid compact paste. The vessels were generally light yellow on the exterior and black on the interior, with incised decoration in the zigzag, crossing lines, or dots motifs, and were tempered with either crushed quartz, chamotte, or crushed shells. Type 5 was represented by a thin-walled, well-polished, straight-rimmed, roundbodied vessel of 4-5 mm wall thickness (Okladnikov 1956, 77). Type 6 was a coil pot incised on the entire exterior surface with stripes, interestingly similar to Neolithic vessels from Yakutia (Okladnikov 1956, 80). Layer 4 at Djebel contained ceramics of five main types. Fragments of type 1 were primarily from one large poorly-fired, dark-red coil vessel with smoothed walls, a solid dark reddish-brown fabric and vessel core, flared rim, narrow neck, round body, and pointed-bottom. The wall thickness was approximately 1 cm, the vessel height 28.5 cm, the rim diameter 18 cm, the neck diameter 17.5

Layer 1 (uppermost) contained fragments of seven distinct vessels. Vessel 1 was wheel-turned, red-orange in colour with slip. Okladnikov notes that this vessel was most likely modern and that its existence at Djebel would not be out of the ordinary, due to the modern-day use of 69

CERAMIC TYPOLOGY cm, and the widest part of the rounded body 29.5 cm (Okladnikov 1956, 105). Type 2 was represented by a large fragment of a well-fired, light-red, chamotte and shell-tempered, smoothed, rounded vessel with a narrow neck. The vessel had incised decoration made of parallel horizontal bands of raised bumps painted with red ochre. The form was semi-egg shape with a wall thickness of 56 mm (Okladnikov 1956, 106). Type 3 was a dark brown vessel with a smoothed, rounded body, 5-6 mm thick walls, and incised decoration in the rhomboid net and parallel lines and hatches motifs. Type 4 was represented by a rim fragment of a dark red, roughly-made, unsmoothed, large coil vessel with chamotte temper. Type 5 included four fragments of highly-polished black and reddish-orange wheel-made vessels which Okladnikov interpreted as intrusive, probably by means of an animal burrow (1956, 109).

fragment which Okladnikov insisted was intrusive. The vessel was a black roughly-made coil pot tempered with soft white crushed stone or shells. Layer 8 contained no ceramic material whatsoever (Okladnikov 1956, 188). The following is an outline of Avgustinik and Baranova’s fabric and technology study in the order of archaeological deposition. They interpret the overall assemblage in terms of fabric types, which they label by letter (i.e. A, B, C). There were no ceramics recovered in layers 7 and 8 that were deemed non-intrusive, hence Avgustinik and Baranova start their classification with layer 6. Layer 6 contained two main fabric types, type A and type B. Type A was a well-prepared paste with no voids and a 25-30% (determined by optical inspection) admixture of angular and rounded limestone temper (2-3 mm grain size). Avgustinik and Baranova hypothesise that it was necessary to add a significant amount of temper to the ceramic paste, due to the high shrinkage properties of the clay when fired (1956, 223). The vessel body was not friable, the surface was rough and the walls were approximately 4 mm thick. The fabric was fired, most likely in an open fire, at a low temperature for a short duration of time, as evidenced by the differentiation in oxidation of the vessel core and margins (Avgustinik and Baranova 1956, 222). Avgustinik and Baranova suggest that strength was added to the ceramic paste through the addition of a binding agent, probably albumen. They report that albumen aids the polymineralisation process of low-fired ceramic fabrics, but that further experimentation and chemical analyses will be necessary to substantiate this hypothesis (1956, 222). Type B was quite similar to type A, but contained a 25% admixture of fine limestone. Both types A and B had low porosity, low permeability, and showed little degradation of calcium carbonate in the fired state (Avgustinik and Baranova 1956, 223).

Layer 5 contained 45 small ceramic fragments, mainly from two or three coil vessels similar to the assemblage from layer 4. Okladnikov identified five basic groupings despite stating that the assemblage came from two or three main vessels. Type 1 were thin-walled, smoothed vessels of semi-egg shape with a flared rim and no neck. These were most likely intrusive from layer 4. Type 2 were vessels similar to those from layer 4 with thinned and smoothed rims, rounded bodies and pointed bottoms. Type 3 were thin-walled vessels with a rough surface, tempered with crushed white stone. Type 4 was the bottom fragment of a vessel found in layer 6. Type 5 was a very small “muddy yellow-grey” pointed-bottomed, round-bodied vessel with a dark red core evidenced at the sherd edge, a solid consistency, and tempered with roughly crushed grey limestone and white feldspar (Okladnikov 1956, 128). Layer 5a (transitional) contained six fragments of ceramic, five of which were from coil pots similar to those found in layer 5, and one of which was wheelturned. This wheel-turned vessel was black with incised decoration in the fir-tree motif. Okladnikov interpreted the vessel as intrusive from one of the upper layers, deposited in an animal burrow (1956, 153). Layer 5-6 (transitional) contained two main types of ceramics, wheel-turned and coil-construction. Four fragments from two roughly-made friable, coil vessels were recovered. Two of these fragments contained a nondescript crushed grey admixture, while the other two were shell-tempered and had incised decoration in the zigzag band motif. The wheel-turned fragment was from a red, thin-walled, wellpolished, small vessel, and like so many of the abovementioned samples, was interpreted as intrusive from upper layers (Okladnikov 1956, 161).

The ceramic fabrics present in layer 5 were predominantly type A and type C. Type B was otherwise absent. The type C fabric had a solid paste with an admixture of 27-28% very well-ground limestone (with grains between .1-.3 mm). The type C fabric was fired at a low temperature, had low porosity, was well-polished, and had a burnished appearance. The fabric exhibited no degradation of calcium carbonate in the fired state (Avgustinik and Baranova 1956, 223). Layer 4 contained fabrics of three main types: A, B, and D. Type C was otherwise absent. Types A and B, with some variability, were predominant in the assemblage. The type A fabric contained an admixture of 30-35% angular and rounded grains of limestone. The vessels were burnished and occasionally slipped, had low permeability, and a wall thickness of 3.5 mm. The firing was even and relatively low temperature (less than 600º Celsius). The fabric did not exhibit degradation of calcium carbonate, and polymineralisation during the firing process was probably aided through the use of a binding agent such as albumen (Avgustinik and Baranova

Layer 6 contained six small fragments of black or dark red coil-type vessels heavily tempered with rounded grains of crushed grey stone (possibly limestone?) (Okladnikov 1956, 178). Layers 7 and 8 were generally interpreted as pre-ceramic layers representative of the Mesolithic period. Layer 7 contained one ceramic 70

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

1956, 224). The type B fabrics had a solid paste, were low-fired, and contained an admixture of shell (ground to approximately .5 mm grain size). Evidence of food remains on the inside of the vessel points to its use as a cooking pot. There was no degradation of calcium carbonate, and a binding agent such as albumen caused polymineralisation of the ceramic matrix (Avgustinik and Baranova, 1956, 224). The first appearance of fabric type D was evidenced in layer 4. Type D was characterised by an iron-rich clay and a thick-walled poorly-fired vessel body with an admixture of finely ground limestone, sand, and gypsum, or an ordinary admixture of sand temper. The vessel core was not well-oxidised as evidenced by the differentiation in firing between the vessel core and margins. Fabric D exhibited a porosity of 12-13% (Avgustinik and Baranova 1956, 224). Avgustinik and Baranova conclude that fabric types A, B, and D were the main locally-produced types. They trace progress in production technology in terms of the use of slip on the interior and exterior of the vessel, the use of finely ground and possibly sieved limestone, sand, and shell tempers, and the introduction of wheel-turned pottery (1956, 225).

a well-prepared paste with low permeability that contained a staggering amount of very coarse sand and gravel temper, and some evidence of under-firing. The vessel was wheel-turned with a wall thickness of approximately 8-9 mm. Avgustinik and Baranova conclude that significant technological advances such as well-controlled firing and the grinding and sieving of tempering materials were accomplished by the end of the time period represented by layer 2 (1956, 225). Layer 1 contained fabric types A, C, E, F, G, and H. Fabric types B and D were otherwise absent. Fabric type A was very rare. Fabric type C was represented by one fragment of a thick-walled (6-7 mm) vessel with a solid paste and evidence of food remains on the vessel interior (Avgustinik and Baranova 1956, 226). Fabric type E was exemplified by a solid but unevenly mixed paste with a very fine admixture of 18.5% iron oxide. The vessel surface was “dark black coal colour” while the vessel core was grey. The firing temperature was estimated to be approximately 950º Celsius (Avgustinik and Baranova 1956, 226). Fabric type F was represented by a thick, solid vessel fragment with a significant iron oxide content, low permeability, and uneven firing. Type F was considered to be non-local in origin (Avgustinik and Baranova 1956, 226). Fabric types G and H appeared for the first time in layer 1. Fabric type G was represented by vessels fragments with a solid paste, an admixture of very fine sand with low permeability, prepared on a fast wheel and fired at a temperature between 950-980º Celsius. Fabric type G was considered to be non-local in origin (Avgustinik and Baranova 1956, 226). Fabric type H was produced with a mergilite clay, evenly-fired with no tempering material. Vessel fragments were “light terracotta” in colour, and had high porosity and high permeability. Chemical analyses revealed that the calcium carbonate content was 16% and the salt content was high. Alkali content was not analysed, but there was evidence of secondary iron oxide (Avgustinik and Baranova 1956, 226).

Layer 3 contained fabric types A, C, D, and E. Type B was otherwise absent. Type A was very well represented, while there was minimal representation of type C (Avgustinik and Baranova 1956, 225). Fabric type D showed a significant technological improvement from layer 4. It is clear from such a conclusion that the terminology used by Avgustinik and Baranova is not in accordance with currently accepted definitions for ceramic fabric. If indeed fabric type D showed significant technological improvement from layer 3 to layer 4, it would normally be considered as a new fabric type. Type D vessels were thinner-walled (5-6 mm in thickness) and contained an admixture of 30% limestone and sand. The vessels were higher and better-fired (at an approximate temperature of 950-980º Celsius) and had very low permeability (Avgustinik and Baranova 1956, 225). Fabric type E, interpreted as non-local in origin, appeared for the first time in layer 3. Type E contained a solid paste that was produced from a very well-prepared clay paste with no admixture. The vessels were primarily wheel-turned, exhibited high permeability, and were fired under well-controlled conditions evidenced by the uniform vessel colour throughout the core and margins (Avgustinik and Baranova 1956, 225).

Avgustinik and Baranova concluded that, in general, there were two or three types of clay evident in the Djebel assemblage. Fabric types A, B, and C were produced using a calcareous-rich clay, whereas fabric types F and H were produced using iron-rich clays. Types A, B, C, and D were considered to be of local origin due to the use of locally occurring tempering materials. Fabrics E, F, G, and H were interpreted as non-local due to the technology of production as well as the differentiation of clay types and tempering materials (Avgustinik and Baranova 1956, 227).

Layer 2 contained fabric types A, B, C, D, E, and F. Types A and B were very rare. Type C was represented by both under-fired and well-fired vessel fragments with low permeability. Type D was an evenly-fired clay paste containing an admixture of 20-25% sand with low permeability. The vessels were wheel-turned and assumed to be of non-local origin (Avgustinik and Baranova 1956, 225). Fabric type E was represented by a roughly-made solid paste with a black colour, low permeability, and an assumed non-local origin (Avgustinik and Baranova 1956, 225). Fabric type F had

In accordance with the findings outlined above, Okladnikov concludes that not all the ceramics found at Djebel were made in situ (Okladnikov 1956, 197). He notes the fact that the inhabitants of Djebel did not use chaff-tempered and painted ceramics characteristic of the assemblages found at sites of the Jeitun Culture (Okladnikov 1956, 211). 71

CERAMIC TYPOLOGY Dam Dam Cheshme I

these vessels, especially the flat-bottomed examples, to the Bronze Age.

The ceramic assemblage from Dam Dam Cheshme I (appendix 2, 373) contained small vessels of three basic forms from the top four horizons at the cave. The forms were 1) vessels with an outward curving rim and inward curving neck, rounded body, flat bottom and no ornamentation, 2) vessels with a straight neck without a protruding rim, ornamented with horizontal incised bands and an unknown bottom shape, and 3) small ornamented bowls (Markov 1981, 44).

There were 174 vessel fragments (12 rims, 118 body sherds, and 27 fragments of grey, black, or brown polished ware) recovered from layer 3. Most of these were wheel-turned (Markov 1966a). They were broadly representative of groups 4 and 5 from layer 2. Markov notes one coil vessel with a “comb” decoration in the form of triangles one inside another. The ceramics from layer 4 included 11 very small vessel fragments. Layer 5 contained one sherd from a coil pot (Markov 1966b, 86). When compared with Okladnikov’s detailed descriptions for the ceramics from Djebel cave, this sparse information is entirely unacceptable and exemplifies the recurring and problematic theme in Soviet archaeology. All too often, interpretative conclusions are published without sufficient data provided as basic background evidence.

The ceramic fragments from horizon 1 were crushed and unrecognisable. Horizon 2 contained both wheel-turned (126 fragments) and coil (214 fragments) vessels. The wheel-turned pottery was thin-walled, polished, high quality and interpreted as imported or intrusive. The coil vessels were roughly-made, red or yellow in colour, and constructed of a high shrinkage clay with an unspecified temper type (Markov 1981, 44). In the lower section of horizon 2, there were fragments of polished, thin-walled, greyware with horizontal and zigzag incised decoration classified separately from the above types (Markov 1981, 45). Horizon 3 contained 138 fragments of coil pottery consisting of roughlymade vessels with a curved neck and round or pointed bottom. In some cases the vessels were polished and ornamented with horizontal incised bands or horizontal incised zigzags. Horizon 4 contained very few unspecified ceramic fragments while horizon 5 contained none whatsoever (Markov 1981, 45).

Oyukli Markov and Bakhta (1959, 48) report on the ceramic assemblage from Oyukli (appendix 2, 374-376) as originating from a separate “workshop” despite the lack of evidence for pottery production at the site. There were 202 vessel fragments of two main types: coil-construction and wheel-turned. There were three main forms of coil pots: two variations of bell-beaker type and small dishes. The first bell-beaker variant was decorated, roundbottomed, with incised decoration in the fir-tree or halfmoon motifs. These vessels had a rim diameter of approximately 16 cm, wall thickness of 4-5 mm, a solid fabric tempered with sand and unevenly fired (Markov 1961, 68). The second type of bell-beaker was also round-bottomed with incised decoration in the horizontal or vertical waving line, broken line, zigzag line, or hash marks motifs. The major difference between type 1 and type 2 was the thick straight rim of the type 2 vessels. The rim diameter was approximately 21 cm and the shoulder diameter was approximately 17 cm (Markov 1961, 68). The third vessel type, the small decorated dish, had similar incised decoration in the fir tree, wavy line, or broken line motifs. The walls were typically 3-5 cm thick. The coil vessels were generally either grey with patches of pink or “rusty-coloured” with patches of dark grey and light grey (Bakhta and Markov 1959, 48). The vessels were constructed of a coarse paste with shell or quartz sand temper (Markov and Khamrakuliev 1980, 74). The wheel-turned vessels had a finer paste and consistency, sand temper, a “dark rusty”, “light rusty”, cream or grey colour, often polished with red slip on one or both sides. The wheel-turned vessels had incised decoration in the crosses or wide bands with crosses motifs (Bakhta and Markov 1959, 51).

Dam Dam Cheshme II The ceramic assemblage from Dam Dam Cheshme II provided significant comparative typological information for dating the chronological sequence of the cave. The ceramics from layer 1 were badly trampled and crushed. However, layer 2 yielded 458 ceramic fragments, 141 of which were wheel-turned, and 317 of which were fragments of coil vessels. There were 23 rim fragments, 8 bottom fragments, and 276 body sherds from the coil vessels (Markov 1966a). Based on this assemblage, Markov distinguished eight main ceramic types. Group 1 included rim fragments of a high quality, polished vessel with decoration in the form of uneven triangles filled with slanting lines (Markov 1966a). Group 2 included thinwalled vessels with incised rim, elongated narrow neck, and a rounded body. Group 3 were curved rim, flatbottomed bowls. Group 4 were kitchen vessels which evidenced sooting. Group 5 included vessels with a short neck and curved rim with a diameter between 8-20 cm (Markov 1966a). Group 6 were thick-walled vessels with a curved rim of diameter 10-12 cm. The vessel walls were between .5-.6 cm in thickness. Group 7 contained flat-bottomed, rounded vessels with a rim diameter of between 4-28 cm and walls between .5-.8 cm thick. Group 8 contained many varying fragments of polished, well-made, black, grey or brown vessels with geometrical decoration motifs (Markov 1966a). Markov attributed

Form and Function In terms of vessel function, three basic types were outlined: storage vessels, domestic ware, and cooking 72

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

pots. Storage vessels were typically poorly-fired, porous, thick-walled bowls and vases used for the storage of water and dry goods. Domestic vessels, used for serving and eating, were higher-fired, thin-walled, slipped and decorated dishes, small bowls and “saladnitsa-type” vessels. Cooking pots were, however, another source of controversy. There is considerable debate as to whether cooking pots were actually used on the fire. Masson (1971, 38-39) hypothesises that the Jeitun peoples did not put their cooking pots directly into the fire, but instead used hot stones to warm the contents of the pots (or otherwise grilled meat on a spit). He provides evidence for such a phenomenon by noting the abundance of fire-cracked stones found in and around the cooking hearths, and the absence of sooting on the outside of the ceramic vessels. Masson explains that cases of soot-like staining found on various vessels was not due to sooting, but rather to rotting foodstuffs being left in the vessel causing subsequent discoloration (1971, 38). Berdiev, however, points out the existence of sooting on cooking pots from both Jeitun and Chagylly, but implies the later absence of sooting during the middle and late Jeitun Culture periods. He explains these discrepancies as indicative of the use of cooking stands or the placement of cooking vessels in hot ashes rather than directly in the fire (1969, 78).

(d), Berezhkin obtained the relation of the height to the diameter of the vessel, and thus its proportions. Figure 4.8 shows the three tables he constructed, 1) the categorisation of vessels from Jeitun according to the height of the upper wall in relation to the rim diameter, 2) the categorisation of vessels from Jeitun according to the inclination of the upper wall with respect to vertical and the rim diameter, and 3) the categorisation of vessels from Jeitun according to the inclination of the lower wall with respect to vertical and the rim diameter. Based on the analysis of these measurements, Berezhkin concluded that the size of the base differs only slightly for any type of vessel (1970, 144). Berezhkin also classified the ceramics in terms of five main wares. These include A) red coarse undecorated vessels, B) decorated vessels with chaff temper and the wavy line, bracket and vertical bands motifs, C) decorated vessels with a coarse paste and the vertical, horizontal and inclined bands motifs, D) undecorated vessels with a coarse paste (more crudely made than group A), and E) very badly made undecorated vessels (more crudely made than group D) (Berezhkin 1970, 145). Using these five vessel categories, Berezhkin presents a series of histograms (Figure 4.9), noting that he determined a 5% margin of accuracy for the statistical analyses. The first histogram represents the quality of vessels from Jeitun in relation to their size. This shows that vessels from groups D and E are generally large, deep and wide pots with a rim diameter of 20-42 cm, or tagori (large vases with a wide bottom and oval form) with a rim diameter of 35-50 cm (Berezhkin 1970, 145). The next series of histograms is an analysis of vessels from Jeitun according to their size and proportions, with the following categories. Figure 4.10 shows 1) cylindrical vessels from Jeitun according to rim diameter, 2) red-walled undecorated vessels from Jeitun according to rim diameter, 3) bowls from Jeitun according to the relation of vessel height to rim diameter, 4) category 3 separated into two parts, 5) rim diameter of red-walled undecorated vessels from Chopan, 6) red-walled vessels from Chopan according to height divided by rim diameter, 7) red-walled coarse undecorated vessels from Jeitun, 8) decorated vessels from Jeitun with chaff temper, 9) decorated vessels from Jeitun with coarse paste, 10) undecorated vessels from Jeitun with very coarse paste, and 11) undecorated vessels from Jeitun with extremely coarse paste (Berezhkin 1970, 146).

Y. E. Berezhkin (1970) carried out a statistical study of the forms of Jeitun Culture vessels excavated at Jeitun, Chopan, Togolok, Pessedjik, Chagylly and Bami. This was the first scientifically objective study which categorised vessels according to mathematical formulas rather than using subjective categorisation that inevitably differed from researcher to researcher. Berezhkin (pers. comm. 1999) believes that this study was a necessary experiment, but that his attempts to categorise ceramic forms mathematically, albeit from an incomplete dataset, led to difficulties in the interpretation of an undoubtedly non-mathematical, subjective and artistic discipline. The following is a brief overview of Berezhkin’s methodology. The statistical analysis of ceramic forms was broken into two main classificatory parts: quantity features and quality features. A categorisation of the quantity features included the proportions and size of the vessels, while the categorisation of quantity features included the identification of fabrics, method of production, and type of decoration (Berezhkin 1970, 143). Berezhkin noted that, because of the predominance of small sherds in the ceramic assemblages in question, it was possible for him to analyse and classify the form of approximately 5% of all recovered material. The three main vessel types he studied were 1) vases, 2) bowls, and 3) large bowls. Because the vessel bottom was often absent, Berezhkin chose other structural features for analysis. These were the rim diameter (d), the height of the vessel (h), the inclination of the upper wall with respect to vertical (B or beta), and the inclination of the lower wall with respect to vertical (y) (Berezhkin 1970, 144). By dividing (h) by

Based on the analysis of these histograms, Berezhkin concludes that it is difficult to classify trends between form, function and decoration. He comments that the precision with which the painted vessels were made is worse than that of the more standardised storage bowls. Berezhkin concludes that the difference in this standardisation of vessel types is due to the differentiation in their usage: red-walled undecorated vessels were for domestic use while painted storage vessels were connected with cult rituals. In terms of 73

CERAMIC TYPOLOGY

Figure 4.8 Jeitun ceramic form statistical analyses (Berezhkin 1970)

74

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 4.9 Jeitun ceramic ware and form histograms (Berezhkin 1970)

75

CERAMIC TYPOLOGY these cult rituals, Berezhkin decided that he was able to demarcate a subconscious type of production standard, used by the potters, where ornamentation and decoration were more important than the form of the vessel. (Berezhkin 1970, 147-149). In reading the original paper, it is entirely unclear to me which sets of criteria led Berezhkin to this particular conclusion, especially given the generally scientific nature and careful detail with which the paper is otherwise presented. In his final remarks, Berezhkin comes to an even less supported conclusion that the forms found in the ceramic assemblages of Chopan and Togolok do not belong in the same general group as the forms found at Jeitun. In making this statement, he references the final series of histograms, comparing the forms from Chopan levels 1-2 and Togolok levels 1-2 (which are similar), the forms from Jeitun and Chopan levels 1-2 (relatively dissimilar), and the forms from Jeitun and Togolok levels 1-2 (relatively dissimilar). Thus, without providing reasons why certain forms should be considered earlier or later than others, Berezhkin concludes that the ceramics from Jeitun level 2 were younger than the ceramics from Chopan and Togolok levels 1-2 (1970, 149). Again, there is a massive gap in the data presented that led Berezhkin to this conclusion. There is also a complete lack of radiocarbon dating to support the relative ages of different layers from different sites. Berezhkin suggests that the further use of this statistical technique could be extremely useful in distinguishing subtle variations between site assemblages (1970, 149). This form of statistical analysis, which the author initially noted only applied to 5% of the overall excavated Jeitun Culture ceramic assemblage, is an entirely unsuitable method for reaching the conclusions presented in the paper.

rich), orangeware (naturally occurring amalgamation of calcareous and iron-rich), whiteware (highly calcareous), and greyware at the above-mentioned sites (Figures 4.11, 4.12, 4.13, 4.14, 4.15, 4.16, and 4.17). These charts show consistency on the x-axis and the percentage of the assemblage on the y-axis. The most striking comparison is evident in the assemblage from Chopan, which is entirely redware. As mentioned in Chapter 3, Chopan is the only site where a pottery production location or workshop was identified. Otherwise, the assemblages from Jeitun and Togolok are fairly analogous, consisting predominantly of buffware (38-47%) and redware (50-60%), with relatively small amounts of whiteware (2%) and greyware (1%) present. Whiteware occurs more rarely than buffware, probably due to the lesser availability of highly calcareous clay. The greyware in these assemblages is very infrequent and probably represents contact with and the dispersal of Keltiminar or Keltiminar-related ceramics. The assemblage from Pessedjik, however, consists to a large extent of orangeware. The clay used in the construction of the orangeware vessels was neither solely calcareous nor iron-rich, but an amalgamation of the two. This differentiation was most likely due to locally-sourced clay occurring naturally near the site. The assemblages from Chagylly, Chakmakli and Mondjukli (all late Neolithic) are considered separately due to their geographical and temporal distinction from the central zone sites of the early and middle Neolithic in the Geok-Tepe region. The assemblages from Chagylly and Mondjukli are roughly analogous, comprised in large part of buffware (91-94%) with the addition of some whiteware (6-9%). The assemblage from Chakmakli, however, consists predominantly of buffware (70%) and whiteware (19%), with the addition of a small amount of redware (3%) and greyware (2%), further diversifying the assemblage as a whole. This increased diversification may be interpreted as increased technological complexity progressing toward the Anau 1A Aeneolithic period, or possibly as indicative of outside influence, cultural contact, and dispersal of material culture. I am hesitant to suggest “trade” as an organised practice at this early stage. In any case, the size of the assemblages from the Meana-Chaacha district sites are smaller, and thus less representative of the excavated site assemblage as a whole, than those from the Geok-Tepe region.

Statistical Analyses The ceramic case study entailed a limited set of statistical analyses of the assemblages from Jeitun, Chopan, Togolok, Pessedjik, Chagylly, Chakmakli, and Mondjukli. The comparisons were derived from the database that was constructed. These charts show the distribution of ware by consistency, a basic breakdown which may be interpreted loosely in terms of vessel function, i.e. coarse wares are representative of large storage vessels whereas fine wares are representative of domestic vessels. The intermediary category of vessels of medium consistency is more ambiguous, but most likely also applies to the category of domestic wares. Attempts to separate cooking wares as a separate vessel category will not be made because of the lack of sufficient evidence to prove that one pot rather than another was used for cooking. This is due to both the absence of sooting and a lack in consistent use of tempertype for all vessels from a specific time period and a specific region.

Because there is significantly more data available for the site of Jeitun, based on both the Soviet and British excavation data, than for any of the other Jeitun Culture sites, charts of the distribution of forms and the form distribution by ware for Jeitun were compiled (Figures 4.18 and 4.19). The data for these charts is limited by the amount of partial or complete vessel profiles available from Jeitun. The charts show vessel type on the x-axis and either the percentage or number of vessels on the y-axis.

Initially, distribution of ware for each site was analysed. The following graphs show the breakdown by consistency for buffware (calcareous), redware (iron76

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 4.10 Jeitun Culture comparative histograms of ceramic ware and form (Berezhkin 1970)

100% 90%

Percentage of sherds

80% 70%

greyware

60%

redware

whiteware buffware

50% 40% 30% 20% 10% 0% coarse (n=184)

medium(n=374)

fine (n=34)

Consistency

Figure 4.11 Jeitun ware distribution

77

CERAMIC TYPOLOGY

100% 90%

Percentage of sherds

80% 70%

redware

60% 50% 40% 30% 20% 10% 0% coarse (n=8)

medium (n=18)

fine (n=1)

Consistency

Figure 4.12 Chopan ware distribution

100%

Percentage of sherds

80% whiteware redware

60%

buffware

40%

20%

0%

coarse (n=91)

medium (n=151)

fine (n=1)

Consistency

Figure 4.13 Togolok ware distribution

78

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

100%

Percentage of sherds

80% whiteware redware

60%

buffware

40%

20%

0%

coarse (n=91)

medium (n=151)

fine (n=1)

Consistency

Figure 4.14 Pessedjik ware distribution

100%

Percentage of sherds

80% whiteware redware

60%

buffware

40%

20%

0%

coarse (n=91)

medium (n=151)

fine (n=1)

Consistency

Figure 4.15 Chagylly ware distribution

79

CERAMIC TYPOLOGY

100% 90%

Percentage of sherds

80% greyware

70%

redware whiteware

60%

buffware

50% 40% 30% 20% 10% 0% coarse (n=8)

medium (n=9)

fine (n=14)

Consistency

Figure 4.16 Chakmakli ware distribution

100% 90%

Percentage of sherds

80% 70% 60%

whiteware buffware

50% 40% 30% 20% 10% 0% coarse (n=10)

medium (n=6)

fine (n=6)

Consistency

Figure 4.17 Mondjukli ware distribution

80

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

plates 12% small bowls 32%

storage jars 30% large bowls 26%

Figure 4.18 Jeitun vessel form distribution

100%

Percentage of examples

80%

60%

redware buffware

40%

20%

0% small bowls (n=24)

large bowls (n=19) storage jars (n=22) Vessel type

plates (n=9)

Figure 4.19 Jeitun vessel form distribution by ware

81

CERAMIC TYPOLOGY

100%

80% greyware whiteware

60%

orangeware redware buffware

40%

20%

0% Jeitun (592)

Chopan (27) Togolok (243)

Pessedjik (171)

Chagylly (101) Chakmakli (31)

Mondjukli (22)

Figure 4.20 Distribution of ware by site assemblage

100%

Percentage of sherds

80% whiteware redware

60%

buffware

40%

20%

0%

coarse (n=91)

medium (n=151)

fine (n=1)

Consistency

Figure 4.21 Distribution of consistency by site assemblage

82

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

At Jeitun, small bowls (32%) and storage vessels (30%) are predominant with a slightly smaller contribution of large bowls (26%) and a significantly small contribution of plates or platters (12%). In terms of the distribution of these four forms by ware, 80% of small bowls are redware and only 20% buffware, suggesting a preference for finer redware domestic vessels. The percentages of the other three categories, large bowls, storage jars, and plates or platters, are more or less similar, representing an approximate 50/50% split.

Identification of Research Questions Through the above analyses of form, ware and function, the following issues warrant more focused petrographic research. Although the Soviet categorisation of the Jeitun Culture ceramics is relatively comprehensive in terms of form, decoration and function, it lacks the scientific evidence and detail required to reach supported conclusions regarding mode of production, cultural influence and contact, and archaeological continuity. The three basic research aims to a petrographic study are: the creation of a fabric classification, the study of production technology, and the determination of origin (provenancing).

In terms of analysing the characteristics of the entire Jeitun Culture assemblage, two further charts were compiled (Figures 4.20 and 4.21). These represent the distribution of ware by site assemblage and the distribution of consistency (roughly analogous to vessel function) by site assemblage, where the assemblages are presented on the x-axis in chronological order and the percentage of the assemblage is presented on the y-axis.

The only previous study of ceramic classification and production technology (not including polarising microscopy or provenancing) carried out in Central Asia was undertaken by F. L. Hiong and published in 1982. Hiong analysed 500 ceramic samples from Aeneolithic and Bronze Age assemblages from Mondjukli, Obadan, Kashut, Dashlidji, Geoksyur, Altyn, Khapuz, and Ulug. Hiong, a Vietnamese student at Moscow State University, was one of the first proponents of mineralogical and technological ceramic studies in the former Soviet Union, and criticised the traditional ceramic studies of form and decoration as lacking scientific basis. Hiong analysed the clay fabric, inclusions and temper in order to classify the ceramic assemblages in groups, and studied production technology to supplement the mineralogical data. Hiong initially divided the ceramics into two groups: those with inorganic inclusions and those with both organic and/or inorganic inclusions (1982, 88). In the first group he identified five main variants including 1) sand, 2) sand and carbonate, 3) sand, carbonate and flint or granite, 4) sand and slate, and 5) sand and chamotte. In the second groups Hiong identified four main variants including 6) chaff temper and sand, 7) chaff temper, sand and carbonate, 8) chaff temper, sand, carbonate, and flint or granite, and 9) chaff temper, sand, flint or granite (1982, 88). He suggests the differentiated use of the nine main types depending on vessel function (Hiong 1982, 89). Hiong also notes the shift from the exclusive use of chaff temper, by the Jeitun peoples, to the use of sand temper during the Aeneolithic period (1982, 91). He does not, however, document a slightly earlier shift to the use of sand temper during the late Neolithic as will be outlined in Chapter 6.

The analysis of ware by site assemblage is no more informative than the description provided above for the same charts presented individually for each site. However, the trends presented in the chart of the distribution of consistency (roughly analogous to vessel function) by site assemblage provides an interesting look at change through time. As mentioned, the site assemblages are organised on the x-axis in chronological order. There is a distinct increase in coarse wares through time from the early Neolithic (represented at Jeitun, Chopan and Togolok), through the middle Neolithic (represented at Chopan, Togolok, and Pessedjik), to the late Neolithic (represented at Chagylly, Chakmakli, and Mondjukli). This trend remains on a steady increase until the incorporation of the assemblages from Chakmakli and Mondjukli, where there are significant increases in the amount of fineware evidenced. Again, this is most likely due to technological progression toward the Aeneolithic period combined with outside influence and the dispersal of material cultural remains. In summary, the findings of these statistical analyses are consistent with a progression from household to workshop pottery production which will be discussed in detail in Chapter 7. Initially, in the early and middle Neolithic assemblages from Jeitun, Togolok, and Pessedjik, there is a high level of variability in the assemblages. Chopan, however, represents a unique and relatively early shift to workshop production, most likely necessitated by the large size of the site and the demands of its population for domestic ceramic wares. By late Neolithic times, a further shift occurred, this time one of increased technology, a preference for and the ability to produce finer wares. This shift was most likely driven by outside influence, increased cultural contact, and the dispersal of material goods.

Fabric Classification The classificatory study of the Neolithic Jeitun Culture ceramics serves to document regional and temporal differentiation in ceramic fabric, and thus more broadly, in ceramic assemblages. The examination of regional differentiation will focus not only on an internal comparison of the western, central and eastern Kopet Dag piedmont zones of the Jeitun Culture, but also on an

83

CERAMIC TYPOLOGY external comparison between the Jeitun Culture and the Keltiminar Culture. In terms of temporal differentiation, I will address the issues of continuity and change from the early Neolithic to the Anau 1A Aeneolithic period. I will look at the pottery as evidence for outside influence and the dispersal of technology and culture. Finally, the fabric classification will be used to draw conclusions about household versus workshop-based pottery production.

Through the proposed ceramic petrographic study, I intend to tackle a consistently problematic theme in Soviet research methodology and interpretation. The tendency of Soviet scholars to reach unfounded or unsupported conclusions, and to present interpretations stemming from insufficient data is well known. Lamberg-Karlovsky remarked on the presentation of “results of their research [which] are unfortunately little known to western archaeologists and even more rarely incorporated into their discussions” (1973, 43). He also notes that data is provided in the original Soviet site reports, but not “laid out in such a manner as to allow one to test one’s own ideas as well as theirs” (LambergKarlovsky 1973, 45) and that “all too often the authors’ interpretations are subject to doubt and remain unconvincing” (1973, 46). I choose to call this phenomenon the “leap of faith”, and Soviet archaeological literature is rife with it. I do not necessarily imply that these “leaps of faith” are always incorrect, they are simply often unfounded in scientific fact or existing archaeological data. As Tosi noted, the tendency of the Soviet authors is “to give interpretations and speculate on the basis of finds made. Such a conjectural philosophy is detrimental to any systematic approach to the studies. This is no superficial criticism, but a more general controversy as to the value to give to certain systems for explaining the pure archaeological datum” (1973-1974, 22-23). Tosi makes these comments in the wider arena of the comparison of Soviet Marxist versus western structuralist and systems theory. I agree with Lamberg-Karlovsky in his conclusion that “these pioneer archaeologists have greatly advanced our knowledge of the substance of Turkmenian cultures, for which we shall be long indebted to them” (1973, 46). Entirely true, we are indebted to the Soviet investigators whose work and data forms the basis of our continued studies. I hope that this petrographic study, the first of its kind in the former Soviet Union, will function to foster future ceramic petrographic studies in Central Asia and the former Soviet Union as a whole, and that it will provide a reference point in terms of relating prior Soviet ceramic typologies, archaeological methodology and theory with current archaeometric method.

Production Technology In addition to a discussion of the basic production methodology used by the Jeitun Culture potters, two specific issues which serve to distinguish the Jeitun Culture will be detailed. The shift from chaff to sand temper during the late Neolithic in the Meana-Chaacha district will be addressed. The differentiation in the use of tempering material by distinct groups living in discrete geographical areas and/or during different time periods gives rise to technological and cultural issues. Secondly, the differences in bottom-shape for vessels of the Jeitun Culture and the Keltiminar Culture and Keltiminarrelated groups will be analysed. This study will focus on the technological and socio-cultural reasons for the existence of a stark contrast between the Jeitun flatbottomed vessels and the Keltiminar and Keltiminarrelated pointed-bottomed vessels. Provenancing The following points will be addressed specifically in light of provenancing: the correlation between the mineralogical characteristics of the inclusions or sand temper in the ceramics from specific site assemblages with the mineralogy of the local geology and locally available raw materials. Using these two sets of mineralogical data, I will draw conclusions regarding the local or non-local origin of the ceramic samples. This discussion will incorporate evidence for cultural contact and interaction through the occurrence of characteristic fabric types outside the expected area of exchange or dispersal.

84

CHAPTER 5 INTRODUCTION TO PETROGRAPHY Petrography is defined as “the study of mineralogical and textural relationships within rocks revealed by observation of thin-sections and hand specimens” (Stiegeler 1976, 208). For archaeological purposes, petrography is commonly used in the study of ceramic samples. Pottery is fundamentally a product of two component parts, clay and sand. Both clay and sand are largely composed of minerals, thus for archaeological applications, ceramic samples consisting of clay and sand can be studied mineralogically in the same way that rock samples are studied for geological purposes. Petrographic analysis is commonly used in the study of archaeological ceramics in three main ways: description or classification, the study of production technology, and provenancing. Naturally, there are limitations in ceramic petrography which I will discuss further in the following sections. I will give a brief overview of the origins of the discipline and the progression of the major contributions to archaeological ceramic studies in order to set the Kopet Dag case study in context among other petrographic studies and their respective research questions.

pottery production and usage for the differentiation and interpretation of various assemblages. The delineations he outlines include household production, household industry, individual workshops, nucleated workshops, the manufactory, the factory, estate production and military or other official types of production. Using these categories, Peacock studies mode of production, distribution and exchange, and provides detailed and previously unwieldly provenancing studies based on the collection and organisation of data that he constructs. Peacock follows this overview with an ethnoarchaeological attempt at creating analogies, generating models, and testing hypotheses (1982, 13). Owen Rye (1981) used ceramic petrographic techniques to research and write a volume covering the principles and reconstruction of ceramic production technology. Rye accurately covers the aspects of ceramic production using comprehensive ethnographic studies of modern-day communities in Israel, Palestine, Pakistan, Guyana, and Papua New Guinea. Rye details each step in the production sequence incorporating 1) the preparation of raw materials (clay and temper), 2) the preparation and mixing of the fabric, 3) vessel formation, 4) surface treatment, 5) decoration, 6) drying of the vessel, and finally 7) the firing of the vessel.

The earliest proponent of the use of ceramic petrography in archaeology was Anna Shepard (1956), who wrote the first significant volume on the subject, entitled “Ceramics for the Archaeologist”. Well ahead of her time, Shepard stresses the importance of collaborative analysis and interpretation by both the archaeologist and the ceramic technologist in their mutual understanding of “the fundamentals of ceramics, the principles and limitations of analytical methods, and the objectives of archaeological research” (1956, xxiii). Shepard provides a general overview of ceramic materials including clay, temper, pigment and glaze; ceramic production technology; and ceramic analysis studies. The ceramic analysis encompasses both form and fabric, including the archaeometric methods of binocular and polarising microscopy, chemical analysis, spectrographic analysis, differential thermal analysis, and x-ray diffraction. Finally, Shepard also addresses the limitations of the discipline and provides guidelines for the interpretation of data.

Prudence Rice (1984, 1987) is renowned for having written and edited overview texts covering a wide range of issues pertaining to archaeological ceramic analysis. In particular, Rice details the raw materials used in pottery production and their respective geological and chemical properties; techniques of pottery manufacture from preparation to forming and firing; issues of pottery production and distribution; an analysis of vessel form, function, decoration and use; the issues surrounding the archaeological, ethnoarchaeological and ethnographic studies of pottery; and a range of aspects of ceramic classification study (1987). Leading the way in the development of ceramic studies, Rice (1984) explored the aspects of “understanding...ancient potters within their society and their use of available resources; study of present-day potters and experimentation with potterymaking to gather observations that can be extended to the potters of prehistory; and, use of recent advances in physicochemical analytical techniques to sharpen the precision of ceramic technological analysis and characterization of ancient pottery” (1984, 245).

David Peacock (1977, 1982) conducted extensive studies of Roman pottery, effectively utilising aspects of ceramic petrography and ethnoarchaeology to research differing modes of production as well as distribution, trade and exchange across the Roman Empire. Peacock embraced ceramic petrography primarily in terms of its potential for the study of production technology. In his attempts to analyse and correlate enormous amounts of pottery found across the broad geographical area that was the Roman Empire, Peacock constructs a multi-tiered hypothesis of

Finally, Orton, Tyers and Vince (1993) provided a more recent volume in an attempt to synthesise the use of archaeometry in ceramic studies in light of new technological and theoretical advances. Quite

85

INTRODUCTION TO PETROGRAPHY importantly, they confront the reality of the need for standardisation in pottery studies and provide a series of guidelines for the discipline. These guidelines outline the history of archaeological ceramic studies and further elaborate on the topics of pottery as archaeological evidence, the logistical practicalities of pottery processing in conjunction with archaeological fieldwork, form and fabric analysis, and archaeological interpretation.

After a discussion of the geology of the study area, I will outline the specific research methodology for the Kopet Dag case study. This methodology consists of four basic parts: sample selection, cataloguing and databasing of site assemblages, and the analysis of thin-sections using techniques of geological polarising microscopy. Following the research methodology section, the results of the Kopet Dag case study will be presented in three parts. Initially, I will review the data and outline the fabric classification for the Neolithic Jeitun Culture ceramic assemblages. Subsequently, the parameters and results of the production technology study will be outlined, including relevant Soviet experimental data. Finally, the provenancing study will be detailed in relation to the above-mentioned section on the geology of the Kopet Dag study area.

In this chapter, I will cover the basic methodology for ceramic petrography and provide definitions of the terminology used to create a clear understanding of the discipline, its limitations and possibilities. I will explore the above-mentioned fundamental research aims of fabric classification, production technology study, and provenancing. After a discussion of potential research aims, general ceramic petrographic methodologies and the likely results of such a case study will be outlined. Chapter 6 focuses on the Kopet Dag case study in particular, following the structure laid out in the current chapter. The archaeological background for these research questions is provided in Chapters 2 and 3. The research aims which will be discussed are three-fold and include: a general fabric classification for the Neolithic Jeitun Culture ceramic assemblages of the Kopet Dag study area, an overview of production technology, and a more extensive overview of the provenancing study.

In Chapter 7, the data from both the form and ware study (Chapter 4) and the petrographic study (Chapter 6) will be synthesised, and interpretations will be made in terms of the original research questions. In this way, the petrographic data provided will act as a link between the study of the archaeological background, the initial research aims, and the final interpretations. Too often, archaeometric studies are carried out in a disjointed and dead-end fashion simply because the technology and capability to carry out archaeometric studies exists. Far more pertinent are archaeometric results when they form the basis of an interconnected and predetermined structure of comprehensive archaeological research directed at answering specific research questions.

Following this discussion of research aims and general petrographic methodology, the geology of the study area will be reviewed. This is necessitated by the research questions presented. A thorough study of the geology of southern Turkmenistan and northern Iran in the mountain zone of the Kopet Dag and its corresponding piedmont allows for the determination of the locally-occurring mineral constituents available for inclusion in the pottery clays or for use as sand temper. Our initial predictions and the delineation of the study area into three geological zones of potential differentiation will be outlined. To supplement this initial geological study, the elaboration and analysis of multi-spectral satellite and other remotely-sensed images was carried out in conjunction with Professor B. Marcolongo of the Institute for Applied Geology at Padova, National Research Council of Italy. The analysis of satellite and other remotely-sensed images was necessary because previous comprehensive geological mapping of the study area was poor. The data from both geological ground-survey and the study of satellite imagery is compared with the mineralogy of the thin-sections in order to make inferences regarding the possibilities of local or non-local production in the provenancing study. It is important to identify the limitations of such a geological study, to make the requisite observations, and to discuss the realistic potential for results. Too often, similar petrographic studies focus on attaining unrealistic results or gleaning extraordinary amounts of information from this form of analysis.

General Petrographic Methodologies There are three main research aims to a ceramic petrographic study: to create a general system for classifying fabric types, to study production technology, and to provenance ceramic assemblages based on the inclusion of diagnostic minerals in comparison to the local geology. Each of these will be addressed in turn. Fabric Classification and Research Questions A ceramic fabric is defined as the constituents of fired pottery, including the matrix, inclusions (solids and voids) and pores, but excluding any surface coatings. The term “fabric” may often be used synonymously with body, paste, or ware. The classification of fabric types proves useful in distinguishing temporal and regional variation within the assemblage of a distinct cultural group (i.e. the Jeitun Culture) as well as between cultural groups (i.e. between the Jeitun Culture and the Keltiminar Culture). Such a classificatory system provides a reference point from which ceramic assemblages from differing geographical areas or distinct archaeological time periods may be compared. The classification of fabric reflects not only the mineralogy of the assemblage,

86

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

but also grows from a typological understanding of vessel form or ware and function.

accurately suggest a unique source area because, as the size of the study area becomes larger, the chances that the geology will repeat also increases. Similarly, if the region is geologically homogenous, the petrographer will be unable to distinguish mineralogical differentiation which could lead to a determination of source area. The interpretation of the results of a provenancing study may include inferences not only of vessel origin but also broader frameworks surrounding trade and exchange as well as population movements and cultural contact. The geological and geographical suitability of the study area is paramount to the feasibility and success of a provenancing study and should be carefully considered from the outset of any programme of provenancing through petrographic analysis.

Once a fabric classificatory system is established, it is possible to determine whether stylistic variations in vessel form reflect the use of different raw materials. Various inferences arise including, most obviously, the comparisons of fabric and form as well as fabric and function. Additionally, the analysis of the constituent parts of the fabric, clay and inclusions, in relation to form and function may be used to distinguish separate archaeological traditions. It is possible to create general data spreadsheets and a variety of comparisons of different variables in order to answer specific questions surrounding the following issues: • determination of household, local, or regional production • changes through time and variability across a particular region • trends in the usage of certain tempers • the existence of a connection between amount of temper (coarseness) or temper type and vessel function • the existence of a connection between clay type and vessel function • ceramic typologies in relation to cultural phases

Specific Research Methodology The main analytical methods used in undertaking the above-mentioned studies of fabric classification, pottery production technology, and provenancing are optical and polarising microscopy. A polarising microscope incorporates filters to modify the vibration direction of light as it passes through a thinly-ground section of a mineral and permits the characterisation of that mineral by its optical properties. The polarising microscope employs transmitted light, but its distinguishing features are a rotating stage and a pair of polarising filters. These filters limit the vibrations of light to one plane and are mounted in the microscope at right angles to each other. When both are in position, referred to as crossed polars or XPL, the field of view is black until a thin-section is placed on the microscope stage. Many of the minerals present will have the ability to rotate the light so that it can pass through the top polariser, referred to as the analyser, and thus these minerals will appear coloured. When the stage is rotated, minerals or clay matrices that are anisotropic (optically active) will change colour, going black (extinction) at intervals of 90º, while isotropic (optically inactive) minerals and clay matrices will always appear dark. The properties of isotropism and anisotropism may help in the identification of the minerals present. When the lower polariser only is in position, referred to as plane polarised light or PPL, the field of view is white and many of the minerals present will be transparent. Others, however, may show colour and will exhibit pleochroism. Pleochroism is the property of a mineral of differentially absorbing light that vibrates in different directions as it passes through a crystal. More fundamentally, pleochroism is the change in colour (dependent on the rotation of the microscope stage) of a mineral viewed in plane polarised light (PPL) using a polarising microscope.

Study of Production Technology A study of pottery production technology includes the analysis of the raw materials used as well as the manner in which the vessel was formed. This covers all phases of production including the drying and crushing of the clay, the preparation of temper, the preparation or mixing of the fabric, formation of the vessel, drying of the vessel, surface treatment, firing, and post-firing treatment. It is also possible to study firing methods, the reaction of the clay body and sand inclusions to thermal shock, and to approximate kiln or firing temperatures based on chemical changes in the constituent materials of the pottery through firing. These sorts of analyses may provide insight into the archaeological economy and overall mode of production. Provenancing Study In a provenancing study, it may be possible to correlate diagnostic mineral and rock types to specific geological areas or specific outcrops in order to suggest a source for the clay and sand tempers used to construct the pottery. Although provenancing studies can be particularly useful and successful, evidenced by Yuval Goren’s (1995) extensive petrographic studies of Neolithic ceramics from Palestine, given a certain set of geological criteria and a manageable study area, there are, of course, potentially significant limitations to a provenancing study. An ideal provenancing study would take place in a compact region of marked differentiation in solid geological outcrop. If the region in question is too vast, it may prove difficult to

For thin-section analysis, samples once selected from the general catalogued database of ceramic samples, are viewed using the polarising microscope so that the observation of optical properties and mineralogical identification may be carried out. In plane polarised light

87

INTRODUCTION TO PETROGRAPHY (PPL), colour, relief, cleavage and pleochroism may be observed. In crossed polars (XPL), extinction, interference colours, zoning, twinning and exsolution may be observed. It is through a combination of these two sets of observations that mineralogical identification is determined.

seemingly differentiated case study scenario for the classification, study of technology, and provenancing of archaeological ceramic assemblages from the Neolithic Jeitun Culture of the Kopet Dag piedmont, the Bolshoi Balkhan mountains, and the Meana-Chaacha district of southern Turkmenistan. The Kopet Dag case study is the first study of its kind to be carried out on any Central Asian site assemblage or cultural group of assemblages. I hope that this study will therefore provide a useful reference point in terms of methodology and theory for future archaeometric ceramic studies in Central Asia and the former Soviet Union as a whole.

In summary, a programme of petrographic analysis may be used to determine a variety of the above-mentioned results pertaining to fabric classification, production technology, and provenancing. The study outlined in chapter 6 represents a geographically compact yet

88

CHAPTER 6 THE KOPET DAG CASE STUDY The Neolithic Jeitun Culture sites of the Kopet Dag piedmont and the Meana-Chaacha district present an interesting case study in terms of ceramic petrography. The data from this petrographic study will be used to complement the analysis of form, ware, and decoration (Chapter 4) in order to look at modes of production and exchange in a Neolithic society. Culturally, the area is one of development and change during the Neolithic. Changes within ceramic assemblages are typically interpreted as chronological markers, evidence of the mode of production, and indicators of outside influence, trade, and exchange. The shift to an early agricultural and pastoral adaptation in combination with a semisedentary lifestyle provides the first insight into such a subsistence system in southern Central Asia.

diagnostic factors in consideration are the use of temper and temper type, pottery form and fabric differentiation from one cultural area to another (i.e. between the Jeitun Culture and the Keltiminar Culture), and the pottery as evidence for population movements, outside influence, and cultural interaction. The fabric classification will also be used to support the typological evidence from Chapter 4 in order to draw conclusions regarding the likeliness of household versus workshop pottery production. The study of mode of production will be based on the mineralogical variability within the assemblage. A high level of variability represented in an assemblage points to production occurring in multiple households. A low level of variability, or a relatively homogeneous assemblage in terms of form, ware and mineralogy, points to a centralised workshop mode of production.

Research Aims and Questions The research aims for the Kopet Dag case study are threefold. 1. The creation of a fabric classification system will allow for the comparison of regional and temporal variation within the general assemblage and also between the Jeitun Culture assemblages and those from adjacent regions. 2. A description of the pottery production technology utilised by the Jeitun people will provide a more comprehensive understanding of the Neolithic mode of production and economy. 3. Finally, in light of the issues of population movement, trade and exchange, I will discuss the provenancing study and the potential that the Kopet Dag region provides for a relatively compact yet geologically differentiated case study area.

With regard to the incorporation of a study of form and function in this analysis, it is not my intention to draw more than basic conclusions regarding the relationship of form and function to any aspect of the fabric analysis. This, unfortunately, is due to the extreme discontinuity that exists between the current petrographic study and previous Soviet typological studies. As mentioned above, it is generally accepted that a petrographic study will contain comparisons of fabric and form as well as fabric and function, incorporating studies of both the clay and the tempering materials as separate factors linked to comparisons of form and function. The previous Soviet typologies, covered in Chapter 4, are comprehensive in their categorisation of form. However, it is rare in the Soviet literature to encounter even the most basic description of ware accompanying the description of form. On occasion, colour is mentioned, but it is often difficult to distinguish whether the colour noted refers to the ceramic fabric or to the vessel slip. Therefore, it is impossible to rely on the colour descriptions provided in the Soviet literature because, on a seemingly indiscriminate basis, vessel body colour may or may not differ from vessel slip colour. In the current study, colour is used as a descriptive tool for identification of waretypes only. It is well-known in the field of ceramic analysis that colour is a dubious characteristic on which to rely. Colour-based descriptions were used for naming the ware-types to provide some consistency with the only ceramic characteristic at times noted in the Soviet typologies.

Fabric Classification The fabric classification system, which is the first of its kind for the Neolithic of the region, was established to act as a baseline for further research and comparison. It is intended to address the following pertinent archaeological questions in light of Chapters 3 and 4. The most basic research aim of the classification is to document regional and temporal differentiation in ceramic fabric from west to east across the Kopet Dag study area, and from the early Neolithic through the early Aeneolithic. The examination of regional differentiation will focus on the Jeitun Culture of the Kopet Dag piedmont and MeanaChaacha district in comparison to the Keltiminar Culture and Keltiminar-related variants of the Kara Kum desert, the Bolshoi Balkhan, and the lower Uzboi river valley. In terms of temporal differentiation, the issues of continuity and change from the early Neolithic through the early Aeneolithic Anau 1A periods will be addressed. The

During my archival study at both the Hermitage Museum and the Institute for the History of Material Culture in St. Petersburg, Russia, I examined twelve complete pots from the Neolithic Jeitun Culture sites. These are documented in the literature as well as described in my 89

THE KOPET DAG CASE STUDY

field notes, but for obvious reasons, I was prohibited from taking any samples from these vessels for petrographic analysis. In general, the ceramic database for the current study is comprised almost entirely of vessel sherd fragments, in the majority of cases too small to provide insight into even partial vessel profiles. In most cases, it is possible to distinguish sherds of thickwalled storage vessels from those of finer bowls or plates. Therefore, it is difficult to relate the comprehensive Soviet typological dataset with the current petrological dataset with any consistency across the assemblages. Unfortunately, the discontinuity between the two bodies of information will allow me to draw only the most basic of conclusions comparing fabric with form and function. I interpret the difficulties I have faced in relating these two datasets as characteristic of the initial use of any new or previously unused archaeometric method. I hope that the current study acts to create a foundation for the use of ceramic petrography in Central Asia and more broadly in the former Soviet Union. In order to overcome the historically precedent language barrier that exists between the western academic community and the Soviet academic community, this thesis will ultimately be published in Russian.

because it was abundantly available, because the use of chaff temper followed a cultural tradition (perhaps imported from northern Iran), or because it was actually advantageous to use chaff temper rather than sand temper due to the properties of the pottery clay sourced in the Kopet Dag foothills? Similarly, was sand temper used by the nomadic stockbreeding Keltiminar and Keltiminarrelated Balkhan groups of the northern super-zone because it was the only material readily available, because of a cultural tradition, or because the addition of sand to the pottery clay of the Balkhan and Kara Kum regions was advantageous due to the physical properties of the pottery clay and sand temper when mixed? It seems reasonable, within the scope of the current study, to address the issues of the use of conveniently available tempering materials by these two distinct cultural groups. Provenancing In addition to the research questions concerning fabric classification and production technology, the case study will address the following issues specifically in light of provenancing. An investigation of the correlation between the mineralogical characteristics of the inclusions or sand temper in the ceramics from specific site assemblages with the mineralogy of the local geology and locally available raw materials is critical. Using these two sets of mineralogical data, conclusions will be drawn regarding probable local or non-local origin of the ceramic samples. Such a discussion will incorporate evidence for cultural contact and interaction through the occurrence of characteristic fabric types outside the expected area of dispersal. In order to answer research questions centred on mineralogical comparisons, it was necessary to conduct a geological survey of the study area. As mentioned, C. Doherty (Research Laboratory for Archaeology and the History of Art, University of Oxford) and I created a threezone geological hypothesis of the potential mineral constituents present in locally-sourced pottery clay and sand tempers. The initial geological hypotheses were verified and amended by A. G. Bushmakin of the Turkmen Geological Expedition, Ashkhabad, during the reconnaissance survey in the spring of 1998. The following section is an outline of the geological background data central to the provenancing study.

Technology Study The research aims of the current study entail a limited pottery production technology study. The ceramic assemblages incorporated in the study are hand-made vessels of the coil-construction type, tempered with either organic material (fine chaff, gramineae, cereal grains, or a combination thereof) or sand. I will discuss the fundamental aspects of pottery production, including the various preparation stages, drying and firing, and address the question of pottery production location. Only one suggested pottery production location, at the site of Chopan, is documented for the entire Kopet Dag study area (Berdiev 1972a). This negative evidence, of course, does not mean that others do not exist. It is my impression, however, that the majority of pottery production for the Neolithic Jeitun Culture was household production. This will be discussed as part of the results in further detail. In addition to the discussion of the basic production methodology used by the Jeitun Culture potters, I will look briefly at the two constituent parts of ceramic vessels, clay and temper. I did not conduct a physicochemical analysis of the pottery clays as part of this thesis due to time constraints, though such a study could be undertaken in the future. Therefore, the primary focus will be the chaff and sand tempers used in the production process. The differentiation in use of tempering material by distinct cultural groups living in discrete geographical areas presents the general technological question of why temper was used. Furthermore, the question of why certain types of temper were used by certain cultural groups will be addressed. Was chaff temper used by the Jeitun agriculturalists of the southern super-zone simply

Geological Predictions Geological ground survey combined with the elaboration and analysis of multi-spectral satellite images was carried out to facilitate the creation of petrographic hypotheses and the interpretation of results. Initially, Doherty and I constructed a three-zoned geological hypothesis based on the available literature concerning the Kopet Dag study area (Suslov 1961, Chistyakov 1974, Rozyyeva 1984, Ruttner 1984, Sholokhov et al. 1986, Yeziashvili 1986, Bazhenov 1987, Preobrazhensky 1987 and 1990, Moussavi and Brenner 1990, and Lasemi 1995). This geological basemap (Figure 6.1) separates the study area

90

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 6.1 – Geological Basemap into three geological regions: the western, central and eastern geological zones. We separated the study area into these three regions based on archaeological significance in combination with the most predominant geological differentiation that was observed. To reiterate, these areas do not correspond entirely to Kohl’s western, central, and eastern piedmont watershed zones of archaeological significance, but refer to zones of geological significance as outlined on the map. Kohl’s western, central, and eastern regions of archaeological significance correspond to our central and eastern geological zones. The western geological zone, which covers the Maly and Bolshoi Balkhan and Tuarkyr, is outside Kohl’s piedmont watershed archaeological area, but is included in the current study because of its pertinence as part of the northern super-zone inhabited by the Keltiminar Culture and Keltiminar-related groups. The map is intended to show possible petrological differences which might result in the sediments, and therefore the resulting pottery fabrics, being different in

these archaeologically defined regions. From the outset, the principal limitations of these general predictions were identified as: 1. a lack of detailed geological maps and reports 2. a lack of soil data 3. the possibility of lateral transport of sediments on the piedmont principally by aeolian (wind) processes 4. the dilution of regionally diagnostic minerals/rock fragments by much larger volumes of common mineral and rock types. The following is an overview of the premises on which we developed our hypotheses for the provenancing of archaeological ceramics based on the solid and drift geology of the region. As is the case for all provenancing studies, it is very difficult to accurately predict the degree to which bedrock geology correlates with the overlying drift topography without a site visit. Thus, the initial first step in creating a geological hypothesis for further verification is to construct a predictive basemap. 91

THE KOPET DAG CASE STUDY

Predicted contribution of solid geology to the clays and “sand” tempers available for pottery in the Turkmenistan-Kopet Dag study area. 1) This is a prediction based on solid geology. The unknown is how the drift geology correlates with the solid outcrop (eg. sand bodies may migrate some distance from the eroding outcrop). 2) Bed-rock solid geology suggests three broad zones which may have different/potentially diagnostic inputs into (pottery) clay. These zones are: West, Central, and East. 3) Western Zone: contains significant outcrops of terrigenous (ie. land-based) sedimentary rocks in addition to marine limestone. (weathering to clays) Rock Type conglomerates = (pebbly sandstone) sandstone sandy clay clay limestone Middle Jurassic sandstone

Possible grains quartz, quartzite, limestone quartz, glauconite fossil bivalves, ammonite fragments gypsum, clay pellets (intraclasts) calcite, dolomite, gypsum, oolitic limestone, micrite, rudistids and coral, fossil fragments ??

4) Central Zone: similar to west but with greater contributions from limestone and less from coarser terrigenous sediments ie. Limestone

calcite, dolomite, gypsum, fossil fragments, oolites quartz, quartzite, glauconite, fossil fragments

Sandstone

* This area does not contain outcrops of Middle Jurassic terrigenous deposits, but these are common in the West. Unfortunately we have no references giving a mineralogical description of these, therefore we do not know if these are useful discriminations. 5) East Kopet Dag: This zone is potentially the easiest to distinguish, that is if the drift geology (clays etc...) do reflect the solid geology. The East Kopet Dag contains the usual limestone and sandstone and therefore we can predict the same/similar suite of derived inclusions, which could find their way into the pottery clays. also contains: Permo-Triassic terrigenous deposits and Hercynian Basement These extend into northern Iran but erosion has probably transported these northward into the East Kopet Dag. (erosion/transport) Rock type Limestone Tuffaceous sandstone (tuff = volcanic ash) non-marine shales Basement (marble) Tectonised conglomerate Diabase

Possible Grains Calcite, limestone quartz, plagioclase feldspar, amphiboles, pyroxenes, iron-oxides (magnetite), volcanic ash/pumice. Iron-rich clay pellets coarse calcite +/- mica polycrystalline quartz and Jasper greenish igneous rocks, feldspar, etc...

Figure 6.2 Solid geology of Kopet Dag study area

92

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

However, this basemap is never more than a first approximation until the initial site visit is carried out. After the initial site visit, undertaken as part of the reconnaissance survey of 1998, our premises were verified and amended by A. G. Bushmakin of the Turkmen Geological Expedition, Ashkhabad (pers. comm. Bushmakin 1998). The provenancing study was carried out as a multi-step process beginning with the creation of the geological basemap, incorporating the initial site visit and extensive consultation with the local specialist (Bushmakin), carrying out a petrographic analysis of the samples collected, and finally comparing the petrographic data with the geological data.

Caspian shore which marks the tectonic path of the Turanian plate over a volcanic hotspot during the Palaeogene and Neogene which produced now extinct volcanoes at Boyadag and several other locations. In the Boyadag hills south of Nebit Dag there is abundant quartz, jasper and carnelian. Additionally, the Tuarkyr region to the north of the Bolshoi Balkhan differs significantly from the Kopet Dag in its orogenesis with the existence of kaolin clays and sandstones. Some 400 metres below the surface in Tuarkyr lay Palaeozoic deposits, coal and other organic deposits, diabase and schist. The kaolin deposits date to the middle Jurassic and were documented at Kyzylkaya. The middle Jurassic sandstone and limestone contains ammonite, flora and fauna, petrified wood, coral and argilite. The kaolin deposits of the western region are a structural anomaly representing the oldest surface sediments in Turkmenistan. Similar middle-Jurassic deposits lie 3-4 kilometres below the earth surface in the Kopet Dag region. The Tuarkyr region also contains jasper, quartz, calcite, kaolin, various clays, sandstone and limestone. Bushmakin remarks that the Bolshoi Balkhan contain bentonite clays, limestone, sandstone, a limestone and clay mixture, alevralite (clay and sandstone), argillaceous schist, calcite, quartz, serdolik, gypsum, carnelian, and limonite (iron oxide derived from clay). In addition to the volcanism and kaolin sediments of the western zone, Bushmakin reports glauconite found between Kyzyl Arvat and Kara Kala and in the Sumbar valley.

The predicted contribution of the local geology to the clays and sand tempers available for pottery in the Kopet Dag study area based on solid geology is critical to this study. This is essentially the correlation of the drift topography with the existing solid outcrop. Consideration and analysis of the drift topography was necessary due to the ability of sand bodies to migrate from the original eroding outcrop by alluvial or aeolian transport. Bed-rock solid geology could possibly have suggested three broad zones which have different and potentially diagnostic inputs into the pottery clay. These zones were termed the western zone (Maly and Bolshoi Balkhan and Tuarkyr), the central zone (west and central Kopet Dag), and the eastern zone (east Kopet Dag). Figure 6.2 outlines the rock types present in each zone as well as the constituent minerals expected in the sand of each zone following the weathering of the solid outcrop. Without much movement of surface deposits, one would expect the transition from one zone to the next to be gradational, exhibiting subtle differentiation.

Central Zone The central zone is similar to the western zone, but contains greater contributions from limestone and less from coarser terrigenous sediments. The limestone and sandstone weather to calcite, dolomite, gypsum, fossil fragments, oolites, quartz, quartzite, and glauconite. The central zone does not contain the middle-Jurassic terrigenous deposits common in the west. Unfortunately, it was impossible to obtain accurate mineralogical descriptions of these middle-Jurassic deposits. Therefore, it is unknown whether the presence or absence of middleJurassic sandstone is significant.

Western Zone The western zone contains significant outcrops of terrigenous (i.e. land-based) sedimentary rocks in addition to marine limestone. These include conglomerates (or pebbly sandstone), sandstone, sandy clay, clay and limestone. These outcrops weather to quartz, quartzite, limestone, ammonite, glauconite, fossil bivalves, clay pellets (or intraclasts), calcite, dolomite, gypsum, oolitic limestone, micrite, rudistids, coral, and other fossil fragments. There is also a possible contribution of Middle Jurassic sandstone from the Tuarkyr region to the north of the Bolshoi Balkhan mountain range.

Bushmakin (pers. comm. 1998) reports that the Kopet Dag mountains began forming during the Palaeogene in the area of the former Tethys Sea. Due to the subduction of the northward moving Iranian plateau and the folding tectonic activity occurring in the Kopet Dag, there is a significant fault line separating the Kopet Dag mountains and the Kara Kum desert. This fault line created thermal springs, underground caves, and water along its length near Ashkhabad (Bakharden), Kou Ata, Nebit Dag and Cheleken. The thermal springs contain high levels of sulphur, yodium and bromium. Bushmakin notes that there are limited Jurassic deposits exposed underground near the Bakharden subterranean lake. There is abundant barite which formed in the tectonic fissures of the Kopet Dag. Abundant calcite exists in the central Kopet Dag

Bushmakin (pers. comm. 1998) described the genesis of the western zone as different to that of the central and eastern zones. To our suggestions above, he added that there are deep clay deposits south of Cheleken and west of the Sumbar river valley, and that the area as a whole is dominated by a sandstone conglomerate. During the Palaeozoic, the Maly and Bolshoi Balkhan and the Kubadag were formed. To the southwest of these low mountain systems, there is an area of volcanism near the

93

THE KOPET DAG CASE STUDY

and quartz and gypsum in the Argman, Archmana and Bakharden regions. Boulder clay deposits in the Ashkhabad region were created by previous glaciation of the area. Alevralite deposits between the clay and sandstone layers in the central Kopet Dag contain quartz, glauconite, and the remains of ancient flora and fauna.

and interpretation of remotely-sensed images (various spectral bands in the visible-infrared-radar range, different geometrical resolution, and different time of acquisition) appropriately elaborated (geometric and radiometric corrections, thematic classification, principal components analysis, etc.). Marcolongo coordinated and collaborated on previous geoarchaeological projects throughout north Africa, the Middle East and Central Asia, including past fieldwork in the Sumbar valley, the Murghab oasis and the Meana-Chaacha district of Turkmenistan. His prior experience in the area was vital to our reassessment of the initial and amended geological predictions.

Eastern Zone The eastern zone is potentially the easiest of the three zones to distinguish. This differentiation, however, is dependent on the fact that the drift topography reflects the solid geology of the region. The east Kopet Dag contains the usual limestone and sandstone, therefore allowing for the prediction of a similar suite of derived inclusions which could appear in the pottery clays. More specifically, these include limestone, tuff, tuffaceous sandstone, non-marine shales, basement (marble), tectonised conglomerate, and diabase. These outcrops weather to calcite, limestone, quartz, plagioclase and other feldspars, amphiboles, pyroxenes, iron-oxides (magnetite), volcanic ash, pumice, iron-rich clay pellets, coarse calcite with or without mica, polycrystalline quartz, jasper, and various green igneous rocks. However, potentially diagnostic rock types, such as diabase, only occur in relatively small outcrops. Therefore, it is not certain that these would be major contributors to the general sediments developing from the east Kopet Dag pediment. The east Kopet Dag also contains Permo-Triassic terrigenous deposits and Hercynian Basement which extend into northern Iran. These sediments were transported northward into the east Kopet Dag study area by the Tedjen river and its distributaries.

The discovery of a problem with the three-zoned geological hypothesis subsequently explained the difficulty I was finding in detecting the above outlined geological differentiation manifested in the ceramic assemblages, given a presumption of local production. Figure 6.3 is an interpretative map, initially created by Marcolongo and P. Mozzi, based on Soyuz KFA 1000 satellite imagery (example, Figure 6.4) and ground survey, and later modified to address the Kopet Dag study area more directly. It is clear from this map that there is a general northwestward trend in alluvial deposition and drift geology in accordance with the direction of watershed toward the Caspian and the northwestward tilting of the Turanian plate. This includes riverine and aeolian transport of sediments along the Kopet Dag piedmont, and across the Kara Kum desert, and more markedly along the Tedjen and Murghab river courses. Also evident from the Soyuz KFA 1000 satellite imagery of the Kopet Dag mountains (including segments of northern Iran), is the existence of multiple river courses in the internal intermontane valleys of the mountain range itself. These rivers travel in a generally northwestward direction before their distributaries turn perpendicular to the Kopet Dag and flow out onto the piedmont in the form of streams or rivers which eventually disappear into the sands of the Kara Kum. The result of both alluvial and wind-blown transport of sediments in a generally northwestward direction caused the sediments of the eastern zone, originally potentially the easiest to distinguish of all three geological zones, to be carried westward and northward across the other two zones. Because of these geological, riverine, and aeolian processes, the Kopet Dag study area is in reality a much more homogenous environment than initially expected.

Bushmakin (pers. comm. 1998) verifies that the eastern zone is an area of metamorphism with outcrops of granite, jasper and carnelian. Barite in the Kopet Dag originates in fissures near the Iranian border which were created by hydrothermal genesis from tectonic movement rather than by riverine transport from elsewhere. Bushmakin notes that sediments from the east Kopet Dag region originate in Iran and Afghanistan and were transported by river courses such as the Tedjen from the Palaeogene to the present. There are significant loess deposits (approximately 20 metres thick) in the vicinity of the Murghab river. Geological Overview Following the initial creation and verification of these geological hypotheses in conjunction with the reconnaissance and ground survey carried out in 1998, we discovered a discrepancy in our theory of geological differentiation for the study area. This anomaly was illuminated by my collaboration with Professor B. Marcolongo (Institute for Applied Geology, Padova). Marcolongo specialises in geomorphological and geoarchaeological analyses, mainly based on the analysis

However, if northwestward sediment drift was solely responsible for the apparent mismatch between sediment and bedrock, this would result in all of the regions having similar “diagnostic” minerals as the east Kopet Dag. The minerals predicted for the east Kopet Dag, if present in sufficiently large amounts to survive into the sediments, would be feldspars and ferromagnesian minerals from the igneous diabase. The immediate problem is thus that ferromagnesian minerals rapidly degrade with increasing 94

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure 6.3 Geomorphological Map

Figure 6.4 Kopet Dag satellite image

95

THE KOPET DAG CASE STUDY

distance from the source area, and therefore should become noticeably rarer toward the west. This fact is inconsistent with the geological and petrological data which shows general homogeneity in mineral content. This geological homogeneity may also be explained by Palaeogene and Neogene volcanic activity on the eastern shore of the Caspian Sea at Boyadag. Our original map failed to predict the possibility of a recent (in geological terms) input of volcanic ash across the study area originating from the Boyadag volcano. On further consultation with Bushmakin, who referred to volcanic activity in the Caspian region, it became apparent that this eventuality was entirely possible. To summarise, there are two possible sources and methods of transport for the ferromagnesian minerals present in the sediments of the Kopet Dag study area. These are northwestward aeolian and riverine transport of volcanic minerals for a limited distance from the east Kopet Dag combined with more recent inputs of volcanic ash from the Boyadag volcano near the Caspian seashore.

each fabric type, as defined by visual inspection, was selected. Because the pottery was sampled by fabric type per excavation context and the fact that there is considerable overlap in fabric types from one context to another, this method of selection provided for considerable repetition in selection of fabric types. This proved useful in identifying the comprehensive range of characteristics attributable to each fabric category. Further differentiation and comparison, beyond visual inspection, of fabrics in group 1 could not be carried out in the field due to the difficulty in obtaining a suitable microscope. Sample group 2 was catalogued and collected at the Hermitage Museum and the Institute for the History of Material Culture of the Russian Academy of Sciences in St. Petersburg during a study trip in the winter of 1997. The ceramic assemblages that were catalogued included material previously excavated by Soviet archaeologists during the 1950s and 1960s primarily from Jeitun, but with smaller representative samples from Pessedjik and Chopan. Approximately 700 sherds were catalogued, one third of which (weighing two kilograms) was sampled and exported for further laboratory analysis at the Research Laboratory for Archaeology and the History of Art, University of Oxford.

Methodology “Archaeologists use a variety of methods to recover materials for analysis, including excavation, collection of materials from the surface of the ground, and the study of existing museum collections. The ways in which data are collected in each of these contexts has a clear impact on the content and representativeness of the ceramic assemblages that we study and on which we base our interpretations. It is therefore important to consider some of the general approaches to archaeological sampling that play a role in determining the structure and limitations of our ceramic datasets” (Sinopoli 1991, 47). In Chapter 5, I reviewed standard petrographic research methodologies. The following explanation of the methodology for the Kopet Dag case study includes an overview of the sample selection process, cataloguing and databasing, and thinsection analysis using standard optical and binocular microscopy of 1319 ceramic samples and 60 thin-sections from 16 sites across the Kopet Dag piedmont of southern Turkmenistan. The general ceramic database is presented in Appendix 5 and the thin-section database is presented in Appendix 3.

Sample group 3 was sampled during a second research trip to Turkmenistan in the spring of 1998. The ceramic assemblages catalogued included material from both the British and Soviet excavations at Jeitun, and the previous Soviet excavations at Chopan, Pessedjik, Togolok, Chagylly, Chakmakli, Mondjukli, Bami, and Djebel. These collections were housed at the Institute of Archaeology of the Turkmen Academy of Sciences in Ashkhabad. All material available was catalogued and databased. The assemblages from each of the above sites were labelled according to occupational horizon or excavation area. In similar fashion to the sampling of groups 1 and 2 above, a sample of each fabric, as distinguished by visual inspection, from each site assemblage was sampled. In some cases, for instance Chopan and Chakmakli, the assemblages available were quite small. Kurbansakhatov reported that this was most likely due to the lack of storage space at the Institute of Archaeology, Ashkhabad, necessitating selective conservation and storage of limited collections. To supplement this previously excavated material, field trips for surface collection and geological survey were made to all of the above-mentioned sites. This surface survey was undertaken to compile a wide-ranging sample in order to facilitate a comprehensive study of the temporal and regional variation found within the Jeitun Culture ceramic assemblages as a whole.

Sample Selection Three separate sherd sample groups were selected, which when combined form the ceramic database. These original sample groupings will be referred to as sample groups 1, 2, and 3. Sample group 1 was collected during the 1997 field season in Turkmenistan, at which time a representative sample of the range in pottery fabrics was chosen from material excavated during the 1993 and 1994 field seasons by the British team at Jeitun. Sample group 1 is representative of Houses “A” and “B” at Jeitun. This sample included approximately 200 sherds weighing eight kilograms. The pottery was sorted in the field laboratory by excavation context, and a sample of

Cataloguing The database was created using Microsoft Access and encompasses all of the sampled and unsampled collections. The following categories were used for 96

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

distinguishing each of the 1319 sherds/vessels databased. Note that only the most pertinent fields are presented in the database in Appendix 5. Records include: • lab number (arbitrarily assigned); • site; • date excavated; • field or excavation number; • museum number (if applicable, to facilitate comparison with the original Soviet recording system); • fabric code (which was initially used but later found to be unnecessary); • fabric consistency; • fabric colour; • types of inclusions identified; • box number (for the Soviet collections); • status in the analytical process; and • information pertaining to the location of the sample (i.e. sampled or left in the field).

The idea behind this was to avoid “forcing” samples into a previously constructed framework which, at a later stage of analysis, may not have been entirely representative of all aspects of the assemblage. After the initial inspection, the pottery was further treated in order to increase the efficiency of the binocular microscopic process. This entailed wet sawing a profile slice (running perpendicular to the lateral grain of the vessel) from each sherd in order to obtain a flat surface. This surface was then coated with lacquer to create a permanently-wetted surface. This coating helped to distinguish differentiation within the matrix, and provided increased clarity for the identification of the inclusions. Following procedures routinely used for fabric analysis at the Research Laboratory for Archaeology and the History of Art (Doherty pers. comm. 1997), the pottery samples were laid out by arbitrary series of context numbers in order to have a larger visual sample for comparison at one time. In this manner, I looked at all context numbers between 1 and 10, 11 and 20, 21 and 30, etc. simultaneously. The particular context numbers combined had no specific archaeological importance regarding their relationship to one another, except for their successive numeration. During this second process of inspection, the sherds were grouped cross-context by fabric type. In this way, it was possible to note the main characteristics of each fabric type. Subsequently, the characteristics of each fabric type were verified using the binocular microscope when visual inspection alone proved insufficient.

The further selection of samples for thin-section analysis was based on the range of fabrics differentiated by this initial databasing exercise. The thin-section database, which was compiled in Microsoft Excel, is documented below under the “Thin-Section” sub-heading. Sample group 1 was treated in as comprehensive and rigorous a manner as possible in order to familiarise with the southern Turkmenian Neolithic ceramic fabric types and to create a sound basis for further study before broadening the temporal and regional aspects of the study. First, an initial binocular microscopic inspection of all sherd samples was undertaken. An edge from each sherd was snipped which allowed for the examination of a fresh break, and notes were recorded based on five sets of criteria. These were: • body: colour, oxidation, and differentiation, uniformity, and offsetting of the vessel core and margins; • inclusion type: (mineralogical identification), occurrence (abundance), size, sorting, morphology and distribution; • porosity: low, medium or high (due to the burnout and subsequent voids left by organic inclusions); • post-burial alteration: existence of salt or gypsum derived from modern agricultural practices; and • surface treatment: polish, burnish, slip, or self-slip.

Once each fabric type was distinguished by arbitrary context series, the fabric types were combined and the range from each house structure was compared in order to compile an overall set of fabric types representative of the assemblage. The assemblages excavated during the two field seasons, and representing two chronologically distinct house structures, were treated separately. A comprehensive photographic record of all fabric types was compiled. Sample group 2 was catalogued during the 1997 study trip to St. Petersburg, but unfortunately it proved impossible to conduct comprehensive sampling due to restrictions on the exportation of archaeological artefacts from Russia. Using both visual and microscopic inspection in Russia, I created an extension of the basic categorisation for sample group 1 to encompass the further variation found within the Russian collections. Subsequent to the initial analysis of sample group 1 (which used the two-fold, multiple check description process), it no longer seemed necessary to carry out such an exhaustive double-checking procedure due to my everincreasing familiarity with the assemblages in this study. Sample group 2 received a thorough visual and binocular microscopic inspection in Russia. Illustrations of all profiles, rims, and sherds with visible form were also

After this exercise in familiarisation with the various fabric types, a list of characteristics particular to the Kopet Dag ceramic assemblages was compiled. This list, for eventual use in the creation of a spreadsheet to organise the data, was also used to check for accuracy on the second microscopic inspection of all samples in group 1. This two-fold process was used as a built-in check of all observations for sample group 1. I found it useful to create the parameters for the spreadsheet after an initial inspection, rather than attempting to compile a list of supposed characteristics before viewing the material. 97

THE KOPET DAG CASE STUDY

carried out. Each sherd in group 2 was snipped to create a fresh break for more efficient and correct identification. The Russian collections were recorded by structure or courtyard and occupational horizon only. Past Sovietsponsored excavations did not employ the context excavation technique used by British archaeologists, therefore there is less information available as to exact location of the Russian and Turkmenian collections excavated by Soviet archaeologists. I compiled an exhaustive catalogue of all Neolithic Jeitun Culture material housed in the Hermitage Museum and the Institute for the History of Material Culture. At both of these institutes, the ceramic assemblages are stored in boxes by structure or yard area. In many cases there is pottery from more than one structure in a given box, but all sherds are labelled and the immaculate and fully comprehensive Soviet recording system allows for the efficient interpretation of the labels. In similar fashion to the visual inspection carried out for the selection of samples from group 1, the Soviet boxing system of sample group 2 was used to form arbitrary groups for inspection and comparison. From any given box, whether it represented one structure or many, a sample of each fabric type present was selected. Therefore, as in sample group 1, there is also a high incidence of the same fabric types occurring in more than one structure, yard layer, or occupational horizon. There is a considerable amount of overlap in fabric types from one horizon or structure to another caused by this sampling procedure, thus again allowing for repeated comparison. Such repeated comparison serves to strengthen and broaden the observations made. The ceramic samples chosen for exportation and further analysis as well as those catalogued and described, but left in Russia, were then added to the overall database. Again, a photographic record of all fabric types for sample group 2 was compiled.

sectioning after a thorough analysis of the general database and an inspection of the photographic records of each sample group and fabric group. Sixty thin-sections in total were made, representing at least one sample of each previously distinguished fabric type from each site in the study. The sites included in the final thin-section study are Jeitun, Chopan, Togolok, Pessedjik, Chagylly, Chakmakli, Mondjukli, Djebel, and Tilkin. I initially selected, catalogued, and put on the database, samples from the site of Tilkin before it became apparent that Tilkin is not actually a Neolithic site. Tilkin is in fact an Anau 1A pre-Namazga Aeneolithic site. However, my communication with Kurbansakhatov during the reconnaissance survey of 1998 resulted in some confusion. Kurbansakhatov actually excavated the site himself, and unfortunately his findings have remained unpublished since his work there in the late 1970s. It seems clear now that, during our 1998 survey, Kurbansakhatov felt under pressure to show me as many archaeological sites as possible, despite the fact that some were not Neolithic in nature. Nevertheless, the ceramic and thin-section samples from Tilkin proved to be quite interesting and significantly more technologically advanced than the Jeitun Culture Neolithic assemblages, and thus provide a useful comparison between the late Neolithic and subsequent Anau 1A Aeneolithic period. An initial batch of 15 thin-sections was selected from sample group 1. These encompassed the fabric types represented in houses “A” and “B” at Jeitun. Once these samples were chosen, further preparation entailed obtaining a new profile slice from the original sherd (or if the sherd was too small for further sawing, the original profile slice was ground down to remove the lacquered surface). Then this surface was impregnated with epoxy resin to retain composition and hardness for the final grinding procedure. The remainder of the preparation, including the grinding of the sample to 30 microns thickness and the mounting of the sample on a glass slide, was carried out in the Earth Sciences Department, University of Oxford. The thin-sections were analysed using standard polarising microscopy with observations made in both plane polarised light (PPL) and with crossed polars (XPL) for mineralogical identification.

Sample group 3, collected at the Institute of Archaeology, Ashkhabad, and during the 1998 field reconnaissance survey was catalogued in similar fashion to sample group 2 with fabric types from each of the sites being selected and further analysed. The overall collection was catalogued in the field and later added to the database. Similarly, a photographic record of each fabric type in the sampled assemblage was compiled. Thin-Section Analysis

The thin-sections from sample groups 2 and 3 represent the geographical area of the central zone (western and central Kopet Dag), the eastern zone (east Kopet Dag/Meana-Chaacha district) and the site of Djebel in the Bolshoi Balkhan range (part of the western geological zone). The latter two batches of thin-sections, numbering 15 and 30 samples respectively, include samples from the sites of Chagylly, Chakmakli, Mondjukli, Pessedjik, Chopan, Togolok, Tilkin and Djebel. The same preparation and analytical procedure described above was followed.

Samples for thin-section analysis were chosen from each of the three sample groups. It should be noted that equal numbers of samples for thin-sectioning were not chosen from the three sample groups. This was due to the fact that sample group 1 is representative only of the type-site, Jeitun. Sample group 2 covers a range of sites, elsewhere duplicated in sample group 3. Therefore, thin-sections from sample groups 2 and 3 were selected on the basis of the range of fabrics occurring at each site. The number of thin-sections made reflects the range in fabrics, rather than the actual sample size of the site assemblage in the original database. Samples were selected for thin-

The results of these analyses were compiled in spreadsheet format using Microsoft Excel for ease in 98

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

tabulating and presenting results. The thin-sections were inspected using a polarising microscope prior to the creation of the spreadsheet parameters for the thin-section database. Again, this was undertaken in order to avoid “forcing” the data to fit a preconceived notion of the potential results of the petrographic study. Two spreadsheets were compiled (Figures 6.5, 6.6, 6.7, and 6.8). The first spreadsheet contains a comprehensive list of characteristics including: ware, existence and type of organic inclusions, existence of quartz, plagioclase feldspar, potassium feldspar, clay pellets, iron oxide staining or pellets, sandstone, limestone, muscovite mica, biotite mica, amphibole, epidote, chert, polycrystalline quartz, geothite/hematite, calcite, oolites, fossil fragments, and loess. The second spreadsheet contains potentially diagnostic observations pertinent to the provenancing study. These are ordered in a manner that allows for the most effective detection of the hypothesised regional differentiation we formulated in our initial separation of the study area into western, central and eastern geological zones. The fields included in the provenancing spreadsheet are restricted to ware, quartz, potassium feldspar, plagioclase feldspar, amphibole, muscovite mica, biotite mica, epidote and calcite.

tempered incised ceramics typical of the nomadic hunting, gathering, fishing and stockbreeding Keltiminar and Keltiminar-related Bolshoi Balkhan and Uzboi cultures of the northern super-zone. Let me stress that this cultural boundary is quite sharp, especially during the Neolithic, and is representative of a major cultural divide which incorporates not only southern Turkmenistan, but also Iran, Pakistani Baluchistan, Afghanistan and more broadly Mesopotamia, in the southern super-zone. Similarly, the Keltiminar and Keltiminar-related subsistence and cultural adaptations of the northern-super zone are not only pertinent to the Kara Kum and Kyzyl Kum deserts, but more broadly connect the desert area of northern Turkmenistan with the nomadic cultures of Uzbekistan, and the steppe cultures of Kazakhstan, western Siberia, and the Urals. The ceramics analysed in this study are predominantly of the chaff-tempered and painted “Jeitun Culture” southern super-zone variety. I found some evidence of cultural contact, for example, the greyware sand-tempered platter base found at Jeitun. Similarly, there is evidence of “Jeitun-type” pottery at the Bolshoi Balkhan cave site of Djebel. Both of these examples suggest cultural contact and exchange. During the Aeneolithic, it becomes apparent that pottery technology changes, and the sand-tempered ware begins to take precedence over the chaff-tempered ware even on the piedmont, as evidenced by the assemblages at Tilkin and Chakmakli. The thin-sections from Tilkin and the upper levels of Chakmakli (namely the greyware samples) are documented here, despite the chronological incongruity with the Jeitun Culture assemblages.

Results The results for the Kopet Dag case study are divided in three parts: fabric classification, production technology, and provenancing. In attempts to present the results of the study in as clear a manner as possible, textural fabric descriptions (Appendix 3) as well as the presentation of data in spreadsheet format (Figures 6.5, 6.6, 6.7, and 6.8) were used. The fabric classification is taken entirely from the analysis of the thin-sections, rather than from the general database (Appendix 5). I found that much of the differentiation I initially perceived through optical inspection was actually not supported by further thinsection analysis. I attribute these discrepancies to my initial inexperience with the discipline and my efforts at providing as comprehensive a thin-section sample as possible. Thus, it seems the thin-section analysis provides a more accurate view of the range in fabric types occurring at the Neolithic Jeitun Culture sites across the study area. The numbers used for the presentation of data are the laboratory numbers which were given to the thinsection samples (1-60). For clarity in the presentation of data, the fabric classification is described by site, and within that grouping, by fabric type. Hence, the laboratory numbers are non-consecutive.

The fabric classification is further divided by site and by ware for the assemblages at Jeitun, Chopan, Togolok, Pessedjik,. Chagylly, Chakmakli, Mondjukli, Djebel and Tilkin. The sites are grouped by region and time period. The textural descriptions in Appendix 3 are intended to provide a general perception of the fabric and its major characteristics, followed by a list of other inclusions which are inherent to the fabric but not central to its descriptive differentiation. These textural descriptions are intended to complement the spreadsheet data. I found it important and useful to analyse and present the data in both spreadsheet format, in terms of textural description, and through photography (photomicrographs at x40 magnification are presented in Figures 6.9a, 6.9b and 6.10) in order to convey an accurate sense of the range in fabric types. Too often, a spreadsheet alone does not demonstrate the intricacies of a particular dataset. Beyond the initial and fundamental division of the ceramics into chaff-tempered and sand-tempered fabrics, several other significant distinctions are immediately obvious. These differentiating factors are the existence or prevalence of loess or loess intraclasts, clay/sediment pellets, varying amounts of quartz sand (of varying grain size), sandstone, limestone, and calcite fragments. The array of other inclusions present are obvious at higher magnification, and are consequently listed separately after the initial fabric description. It should be noted that

Fabric Classification There are two basic types of ceramic in the Kopet Dag study area, broadly defined as chaff-tempered and sandtempered wares. The chaff-tempered painted ceramics typical of the agricultural sites along the Kopet Dag piedmont and the Meana-Chaacha district of the southern super-zone are profoundly different from the sand99

THE KOPET DAG CASE STUDY

after an initial inspection of all thin-sections, an overall trend towards mineralogical homogeneity was recognised. A simple spreadsheet does not allow for the accurate presentation of the quantity or abundance of the mineral constituents in each sample. Despite this, I decided not to employ point-counting to reach such determinations. Point-counting is an extremely arduous task which, in this particular case study and for the objectives of this thesis, seemed unjustified due to the discrepancy between the amount of supplementary data which would be obtained in comparison to the time such a task would require. Thus, as evidenced by the spreadsheet data, the mineral constituents noted at higher magnifications but not ranked in any hierarchical (or percentage-based) manner include mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, and chert.

of renewed homogeneity in ceramic ware and type during the Anau 1A period, there were at least the two abovementioned regional variants. Berdiev noted the influence of the Keltiminar-type ceramics on the assemblages of the far western site of Bami. It is with this preface that I present the classificatory data in appendices 3 and 5, Figures 6.5, 6.6, 6.7, and 6.8. Technology Study This section focuses on the pottery production technology used by the Jeitun potters of the Kopet Dag study area. I will initially detail the pottery production sequence and later provide a more detailed study of the chaff and sand tempers incorporated in the production of both the Jeitun and Keltiminar wares. The production sequence encompassed the following stages: gathering raw materials (clay, temper, and fuel), preparation of raw materials, preparation of the ceramic fabric, formation of the vessel, drying of the vessel, surface treatment, firing, and post-firing treatments. Archaeologically, it is often difficult or impossible to determine the details of task differentiation by age, sex, family unit member, or clan, until craft and economic specialisation develop and specialised potting quarters appear on archaeological sites. Therefore, archaeologists have traditionally used ethnographic or ethnoarchaeological studies of modern-day traditional potting communities to complement and supplement the missing details regarding the dynamics of pottery production. It appears that there was no full-time craft specialisation in this, as yet, unstratified society. Lollekova suggests that potting among the Jeitun people was a female activity, with help from males for the initial and extremely strenuous crushing and mixing of the clay (Lollekova 1988, 98).

Initially, we hypothesised a marked differentiation between the three geological zones, most strikingly between the mineralogy of the eastern zone and that of the central and western zones. The geological hypotheses are detailed above, under the section Geological Predictions. Subsequent to the creation of our initial hypotheses, I conducted a further geoarchaeological study with Marcolongo of the Institute for applied Geology, Padova, which utilised techniques of remote sensing for archaeological purposes. As reported already, this geoarchaeological study significantly altered the initial geological predictions Doherty and I postulated by illuminating the nature of the drift topography of the region in comparison to the bedrock solid geology. The differentiation in regional bedrock solid geology is not entirely manifested in the overlying drift topography, thus creating a situation of increased geological homogeneity across the study area.

Pottery production experiments were carried out by S. A Semenov, G. F. Korobkova, and O. Lollekova as part of the experimental expeditions of the Institute of Archaeology, Russian Academy of Sciences, Leningrad, between 1974-1978. The expeditions visited Lithuania, Moldavia, Crimea, and the Leningrad Oblast and carried out pottery production experiments intended to replicate the production of Jeitun-type ceramics (Lollekova 1988, 97). Lollekova details eight stages of experimental production, much the same as the above outline, including 1) crushing and mixing of the clay, 2) preparation of the paste, 3) vessel forming, 4) vessel smoothing, 5) drying, 6) polishing, 7) decoration, and 8) firing (Lollekova 1988, 98). I will endeavour to provide a basic overview of the production sequence and the technology likely to have been used by the Jeitun potters including insights from Lollekova’s description of the pottery production experiments.

Other major classificatory issues are temporal in nature. Whereas the Jeitun early and middle Neolithic central zone ceramics were generally chaff-tempered and painted, the middle and late Neolithic assemblages from the Meana-Chaacha eastern zone were initially chafftempered but unpainted. Eventually, these eastern zone ceramics became sand-tempered. Initially, during the late Neolithic, the sand-tempered wares of the east were unpainted. Gradually, painted decoration was reintroduced, but using predominantly geometric designs rather than various configurations of wavy and straight parallel horizontal and/or vertical lines. These geometric designs typify the pottery of the subsequent Aeneolithic and early Bronze Age periods. The late Neolithic and pre-Namazga Anau 1A Aeneolithic wares from the central zone also changed from being chaff-tempered to sand-tempered, but with this shift, they did not exemplify the unpainted stage found in the east. To summarise, by the Anau 1A period, all ceramics across the piedmont and Meana-Chaacha district were sand-tempered and decorated using geometric motifs. In reaching this stage

In assessing the homogeneity of the Neolithic economy and mode of production for the Jeitun Culture, Berdiev points out similarities in ceramic production and technology across the Kopet Dag piedmont (Berdiev 100

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

1969, 74). He suggests that the ceramics were produced on-site of locally-sourced raw materials, either in individual households or in a communal production area. Let me make clear here that during the Soviet period, archaeological research centred around the analysis of economy and productive capability. This occurrence was also noted by Kohl (1984, 246). Just as there is a specific formula for writing archaeological articles and texts, there are almost always references made, irrespective of author, to Marx and Engels prefacing conclusions about the origins of communal behaviour. Repetitive and often indirect or subtle, these were conclusions about the origins of communism. To a western scholar reading these articles and texts completely removed from the previous reality of Soviet Marxist propaganda, this constantly recurring theme and the interpretations that stem from it are quite stark.

men crushed enough clay for 20 vessels using a very large (1.5 metre) wooden grinding club in 1-1.5 hours (1988, 98). The temper also requires a certain amount of preparation. In the case of the majority of Jeitun Culture ceramics, chaff temper was used. This was presumably readily available on-site on a seasonal basis due to the agricultural activities of the village. Below, I will further discuss the implications of a non-specialised economy and the possibility that pottery production took place predominantly at certain times of the year. It is possible, but not probable, that the chaff temper used at Jeitun, which includes fine chaff, gramineae, and cereal grains, and a mixture thereof (Charles and Bending pers. comm. 2001), was collected at harvest time and stored for later use. It is my impression that any seeds found in the ceramics were not intended for use as temper, due to their infrequent and haphazard appearance in the fabric, but were accidental inclusions that were in fact man-added but unintentionally so. There was little or no preparation necessary for the chaff temper or gramineae, essentially a waste product of the processing of cereal crops and grass. The sand tempers added to the Keltiminar and Keltiminar-related Balkhan, Uzboi, and later piedmont Aeneolithic wares show considerable intra and inter-site variation, and were most likely collected in the environ of the production location. Occasionally, these would have been crushed, as is evidenced by samples of oolitetempered ceramic from Tilkin and Chakmakli. It is most probable that the potter generally selected naturallygraded sands suitable for their particular purposes, in order to avoid expending excess labour necessary in crushing non-plastic and non-organic tempering materials.

To return to Berdiev’s conclusions about pottery production, he suggests that potting was most likely a communal activity (Berdiev 1969, 74). It is unclear whether he suggests this only for the firing phase of the production sequence, or for the entire sequence. In terms of the ceramic fabrics used, Berdiev provides only the briefest mention, and concludes that the ceramic fabric with chaff temper was generally similar across the entire piedmont (1969, 74). I interpret this as the categorisation of a genre, meaning that all vessels were of a coarse hand-made coil-construction, produced using a ceramic paste with chaff temper. Collection of Raw Materials Rye (1981, 16) points out that the factors affecting the collection of clay, temper, and fuel are ease of access, depth of the clay deposit, and distance of travel. Unfortunately, it proved impossible during the 1998 reconnaissance survey to locate any clay deposits in the vicinity of the Jeitun Culture sites of the piedmont or the Meana-Chaacha district due to modern-day agricultural manipulation of the land. Irrigation canals, ditches, and agricultural fields now obscure the ancient landscape and waterways. The occupants of the Jeitun Culture sites probably collected their raw materials within close proximity to the sites themselves, due to the littledeveloped nature of pottery production during the Neolithic, and the importance of efficiency and ease of production in a non-specialised economy.

Preparation of the Ceramic Fabric Following the preparation of the raw materials, the potter prepares the ceramic fabric. In the simplest production sequence, this involves blending waterlogged clay, which has not been previously dried and crushed, with the possible addition of natural sand (Rye 1981, 19). However, it is more likely that the Jeitun potters employed the above-mentioned techniques of drying and crushing the clay to remove gross impurities and in order to have control over the re-hydration and subsequent workability of the clay. Lollekova reports that after the initial crushing of the clay, a bucket (of unknown volume) of water was added to the clay and this mixture was again pounded with the wooden grinding club by men working in shifts. Subsequently, women completed the kneading and addition of temper by hand in approximately 25-30 minutes (Lollekova 1988, 98). Kneading is typically done either by foot or by hand, depending on the amount of fabric to be blended. If initial blending is done by foot, usually further blending is also accomplished by hand (Rye 1981, 19). This kneading process insures the even distribution of moisture, the elimination of airpockets, and the thorough mixing of the clay fabric with the natural inclusions and tempering materials.

Preparation of Raw Materials After obtaining the raw materials, at least the most basic forms of preparation were necessitated. These would have included the initial removal of coarse matter such as roots or large rocks from the clay. This process usually requires the sun-drying and subsequent crushing of the clay into small lumps so that unwanted materials are easily extracted. Lollekova reports that this initial crushing stage was extremely labour-intensive, and required the assistance of men working in shifts. Two 101

THE KOPET DAG CASE STUDY

Vessel Forming

covered storage space available. Lollekova reports that during her experiments, the vessels were dried inside for 3 days (to avoid cracking caused by rapid drying in the sun) followed by 7 days in the sun (1988, 99). During the early stages of the Jeitun Culture, it is likely that pottery production was limited to small-scale household production. In this case, the output would have been small and would not have necessitated much storage space. During the middle and late-Jeitun phases, there is some evidence at Chopan for a central pottery production workshop which would have produced a larger output than the previous household production, and therefore necessitated space exclusive to the pottery production activity, at least periodically. The archaeological literature (Berdiev 1972a) does not offer an indication as to whether or not the pottery production area at Chopan was used for other purposes as well.

After the ceramic fabric or paste is prepared, the correct water content and plasticity/workability must be obtained (Rye 1981, 21). This is vital for vessel formation because problems in manipulation of the fabric will occur when the fabric is excessively moist, or in contrast, excessively dehydrated. Plasticity is founded on the basis of the clay/water system, thus when water is added to the clay it becomes more malleable, and when clay is dehydrated it shrinks and plasticity disappears (Rice 1987, 54). The Neolithic Jeitun Culture ceramics were made using the coil-construction technique. For the purposes of this thesis, “coil-construction” refers both to what is typically known as coil-construction and to what is known as ringbuilding. The former technique utilises coils joined together in a spiral, whereas the latter technique utilises coils the length of the vessel circumference which are not continuous from one coil layer to another. The forming of ceramic vessels usually takes place in two phases: primary and secondary forming techniques (Rye 1981, 67). Primary forming techniques include the rolling of coils (which are generally double the thickness of the eventual vessel wall and can be between 10 centimetres to 1 metre in length), the formation or pinching of the base (from a globular piece of clay), and the building of the vessel walls by laying successive layers of coils around the circumference of the base (Rye 1981, 67). In most cases, the base will be formed and allowed to partially dry so that it will more effectively support the vessel walls in their plastic state (Rye 1981, 21). Lollekova reports that during the experimental pottery production, the formation of vessels was accomplished on a square wooden board supported on the knees of the potter. Flat-bottomed vessels were made starting with coils from the base while pointed and rounded-bottom vessels were constructed starting with the rim. To widen the vessel profile, a coil was attached to the outside of the existing form, and to narrow the vessel profile, a coil was attached to the inside of the existing form (Lollekova 1988, 98). Lollekova notes that the Jeitun Culture pottery was flat-bottomed, but that variations in form representative of the research areas of other members of the experimental team were also attempted. Finally, the vessel walls were smoothed to effectively join the coil walls and produce an all-over smoothed appearance using a wooden knife and water. The entire production sequence to this stage was accomplished in 4-4.5 hours (Lollekova 1988, 98).

Surface Treatment Subsequent to the complete drying of the vessel but before firing, various forms of surface treatment may take place. For the Jeitun Culture ceramics, this entailed the application of a slip and painted decoration in many cases. If this slip or painted decoration is applied before the vessel is completely dry, the slip and pigments applied will smear. An example of this occurrence at Jeitun is sample J.93.123 (Appendix 5). Firing The purpose of firing is to subject pottery to sufficient heat for a sufficient time to insure sufficient fusion of the clay minerals to give the vessel requisite strength. The processes affecting the firing of a vessel are four-fold and include: rate of heating, maximum temperature attained, atmosphere of the firing environment, and length of firing. All of these factors are critical to the firing process. The variables in temperature include the type of fuel burned, the amount of heat produced by that particular type or combination of fuels, and the insulation of the firing pit or kiln (Rye 1981, 25). The effects of varying amounts of air on the firing atmosphere are as follows: insufficient air will create a reducing environment, a balanced ratio of air to fuel will create a neutral environment, and excessive air in the firing environment, as is the case for the Jeitun Culture pottery, will create an oxidation environment. In many cases, it is possible to have both reducing and oxidising stages in the firing process depending on the management of firing.

Drying of the Vessel Interestingly, Berdiev suggests that there were two methods of vessel decoration. The first was a process of double-firing in which vessels were fired, then painted, and subsequently fired again. The second process was painting before a single firing event (Berdiev 1969, 79). The paint was analysed and was determined to be a mineral paint made of red ochre lacking organic constituents, the final firing colour of which was determined by the temperature and duration of firing (Berdiev 1969, 75).

After forming is complete, the vessel is allowed to dry. Initially, the vessel reaches a leather-hard state at which point it will no longer succumb to deformation. The complete drying process may take between several days and several weeks depending on the temperature of the drying environment and the season. Drying may take place inside a structure or in the sun, depending on the warmth and dryness of the climate or the amount of 102

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Lollekova’s pottery production experiments involved firing the vessels in a firing-pit that measured 3 x 3 x 1 metre and paved with stone. The pit was pre-heated by lighting a fire. Once this pre-heating fire burned out, the vessels were stacked on a platform built on the base of the firing-pit over the smouldering ashes. A second fire was subsequently lit next to the vessels and charcoal and ash from the previous fire were piled around the vessels. The fire was continually stoked to provide a gradually increasing level of heat. Once the fire was extinguished, the vessels were left in place and the firing-pit retained heat for approximately 24 hours. Lollekova determined that 10-11 people-days were necessary to construct 20 vessels, or that each vessel required approximately 5 hours 30 minutes of man-labour from start to finish (1988, 100).

for cooking or for warming two rooms at the same time. Berdiev therefore suggested the household production of ceramics as a possible tertiary feature of these hearths (1969, 77). Sarianidi (Masson 1971, 97) also proposed that each household produced its own pottery in the household hearth. Ershov noted, in his explanation of the kiln structure at Chopan, that the existence of slag points to the use of kilns and purpose-built facilities for firing (Ershov 1956; Berdiev 1969, 78). In attempting to summarise the inconclusive evidence and the controversy over method of pottery firing, Berdiev concludes that the issue remains open for further research. It is likely that the firing method used by the Jeitun potters was open-firing, either in a hearth pit or an open fire. This is typical of early prehistoric cultures that utilised the least time-consuming or labour-intensive methods of pottery production. This open-firing may have been carried out in a location slightly removed from the main habitation area of the settlement due to the smoke and pollution created. The fact that it is not desirable to have open-firing occurring immediately adjacent to habitation areas perhaps provides some insight into the lack of archaeological evidence for pottery production found at the Jeitun Culture sites of the Kopet Dag study area. These open-firings may have been carried out very near the site, but outside the limits of the mound itself.

The firing temperature for the Jeitun Culture pottery was determined to be quite low, similar to most prehistoric pottery fired in open pits or proto-kilns. The Soviets noted the existence of unoxidised patches on the surface of vessels (Berdiev 1969, 75; Masson 1971, 38) and differences in oxidation between vessel core and margins. This led them to the assessment that the vessels were poorly-fired at relatively low temperatures. Masson states that, based on a thermodynamic test carried out by B. N. Piatkin, he determined the firing temperature of the Jeitun Culture pottery to be approximately 800° Celsius (Masson 1971, 97). Masson provides no reference for, or details of, Piatkin’s experiments or methodology. It is clear that dishes and vessels used for eating and storage were slightly better-fired than cooking vessels (Berdiev 1969, 75). There appears to be controversy among the Soviet investigators regarding the actual firing of ceramic vessels. Berdiev and Khlopin suggest open-firing in communal pits whereas Masson suggests both the use of the hearth within the house and the use of communal kiln facilities, and gives the example of “Platform A” at Jeitun (Figure 3.3) (Masson 1971, 35). Although Masson suggests the use of a primitive kiln on “Platform A” for pottery firing activities, he found no evidence of a kilnlike structure on this platform. Masson also suggests that “Platform A” was used for the communal baking of bread and cites the continuation of this practice among the modern-day Turkmen (1971, 97).

In open-firing, the type of fuel selected determines the rate of firing as well as the maximum temperature reached. A study of firing temperatures for the Jeitun Culture ceramic assemblages will not be addressed by this thesis, although such analyses are possible and were previously undertaken by B. N. Piatkin, but unfortunately not detailed beyond results in publication (Masson 1971). Potential sources of fuel for firing which would have been available to the Jeitun Culture potters include animal dung, grass, straw, twigs and wood. Worth noting are the varying burning rates of these materials. In general, animal dung is slow-burning, and is ideally combined with one or more of the other above-mentioned materials which are faster-burning. The use of a combination of slow and fast-burning fuels allows the fast-burning fuels to create a rapid increase in temperature which is then maintained by the slowerburning, longer lasting fuel. Usually, the slow-burning fuel is placed underneath the vessels being fired whereas the fast-burning fuel may be placed above and below (Rice 1987, 153). Sometimes, fuel is added as the firing process progresses. However, fuel added to the outside of a fire can decrease the temperature at the centre because the amount of air available at the centre will decrease in order to feed the newly-stoked outside areas. Rye reports (1981, 98) that for effective stoking, specialised methods of introducing the fuel to the central area of the fire must be employed. Finally, the entire pile of fuel and vessels may be covered by old vessel sherds, stones, wet grass, dung or mud (basically any noncombustible material) as a form of insulation for the

However, A. A. Marushchenko did unearth a kiln in the Anau 1A levels of Mondjukli in 1959, verifying the later use of more sophisticated pottery production techniques at this site. The kiln was of the two-chambered type also found at later sites in the Geoksyur oasis. Both Berdiev and Khlopin argue for the co-existence of open or pitfiring with the development of kiln-firing through the Namazga II period (Berdiev 1969, 76). After the discovery of the Mondjukli kiln, Masson (1971, 97) hypothesised that the hearths in the house structures at Jeitun, Chopan, and Chagylly were used for firing pottery as well as for cooking and as a source of heat. Berdiev noted their oval or rectangular proto-kiln shape, and the existence of one double-sided hearth at Jeitun reminiscent of the Mondjukli two-chambered kiln, but instead used 103

THE KOPET DAG CASE STUDY

retention of heat (Rye 1981, 25). It is unclear whether one type of fuel or a combination of types was used at Jeitun.

with the calcareous clays used to make the buffwares while the iron-rich red-firing clays used to make the redwares were tempered with the round variety of fine chaff. In a limited amount of cases a combination of the two types occurs, but this variation is represented primarily in the buff-firing wares.

Temper The temper used in the construction of Jeitun-type ceramics is primarily organic temper. However, Keltiminar ceramics as well as late Neolithic and Aeneolithic assemblages from the Kopet Dag piedmont (upper levels at Chakmakli and the general assemblage at Tilkin) exhibit the use of sand temper rather than organic temper. Berdiev makes reference to ceramic fabric temper in his discussion of the development of the Jeitun Culture in his concluding remarks (1969, 100). For the first and only time in any Soviet publication, Berdiev mentions the Jeitun Culture phase 3 transition from chafftempered to sand-tempered ceramics as evidenced at upper Chagylly and Mondjukli. He adds that similar sand-tempered wares also occurred in the uppermost levels at Chopan. Masson, however, refutes the existence of any sand tempers in the Jeitun Culture pottery, citing the long distance to the source area in the Kopet Dag mountains as the limiting factor for the collection of sand tempering material (1971, 39). Clearly, the current petrographic study has proven this conclusion incorrect. I will review both types of temper, their strengths and weaknesses, as well as the differentiation between the temper used in the Jeitun and Keltiminar pottery. I will also address the eventual shift from the use of organic to sand temper among the Kopet Dag peoples. The organic temper used in the construction of Jeitun Culture ceramics is of three main types: fine chaff, gramineae, cereal grains, and combinations thereof. The variation in the use of these three types of organic material is presented in Appendix 4, which was compiled by M. Charles and J. Bending of the Department of Prehistory, University of Sheffield. The differentiation in usage of the types of fine chaff appears to be deliberate. The fine chaff is essentially a finely chopped plant material, probably grass stems, and chaff from the grass ears (Charles and Bending pers. comm. 2001). The gramineae are seeds of grasses (not cereals), some of which still have chaff surrounding them. There were several types of grass seeds which can be divided into long and round. The round type could be similar to gramineae found in the charred assemblage analysed as part of the radiocarbon dating programme conducted at Jeitun. Research in this area is ongoing. This round type of gramineae, if it is the same as that found in the charred assemblage from Jeitun, flowers late in the year, over the summer rather than in the spring (Charles and Bending pers. comm. 2001). This information could be useful in determining the season of manufacture of the pottery. Nevertheless, gramineae and fine chaff could have been “saved” from one season to another, and once collected, used over an extended period of time (Charles and Bending pers. comm. 2001). It appears that certain combinations of round versus flat fine chaff were used for certain wares. The flat variety of fine chaff was mixed

The use of organic temper such as straw or stems of plant waste in ceramics “may be common when pottery making is a seasonal activity coinciding with the end of harvesting” (Rye 1981, 34). In general, the presence of organic temper improves workability during the forming process and reduces the shrinkage of the clay during the firing process. When fired, organic temper burns out creating voids which increase the insulating properties of the pot by reducing its ability to rapidly transmit heat by conduction. However, this makes the vessel more prone to thermal stress, a characteristic which is detrimental to the use of a particular vessel as a cooking vessel. The most common forms of sand temper used in the construction of ceramic vessels are calcium carbonate (CaCO3) and silica (SiO2). The sand temper used in the Keltiminar and late Neolithic/Aeneolithic piedmont pottery is of three main types: calcite, limestone, and quartz sand. The size of the temper varies depending on the size of the ceramic vessel and its wall thickness. There is usually also a correlation between the size of the sand temper grains or fragments and the hardness or porosity of the vessel. It appears that the Keltiminar peoples used sand temper in their pots due to availability and the relative lack of access to organic tempering materials that would have been prevalent on the piedmont. Traditionally, the existence of shell, calcite, and limestone temper has been interpreted as optimal for use in cooking pots because of the idea that calcium carbonate has thermal expansion properties similar to those of clay. However, there are many types of clay used in the construction of pottery, and these different types of clay have different thermal expansion properties. Traditionally, the use of various forms of calcium carbonate temper in a pottery fabric was interpreted as minimising the stress caused by differential thermal expansion (Rye 1981, 33). However, although many attempts have been made to categorise certain vessels as cooking pots based on an analysis of their properties, there are no hard and fast rules for what may be interpreted as a cooking pot and what may not. There are many examples of ceramic studies which break the traditionally accepted interpretations of the parameters of cooking wares. The use of quartz sand (silica), although common, was traditionally thought to be detrimental to the structure of low-fired cooking wares because of the differential thermal expansion created. Again, many studies have overridden these parameters, proving that the addition of quartz sand does not necessarily exclude the possibility of a vessel being used as a cooking pot. Quartz sand is extremely common in the Kopet Dag study area and readily available as a convenient tempering 104

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

material. It occurs widely in the Keltiminar and Keltiminar-related pottery and was incorporated in pots used for cooking. Occasionally, hydrated forms of silica such as chert or flint are also used as ceramic temper.

for the study area in question. A central production site would contribute to homogeneity in raw materials and production technology as well as standardisation of forms. If this were the case, there would be a distribution network in place covering the area of dispersal and ceramic ware similarity. The second possibility is that the Kopet Dag study area has less geological differentiation than initially hypothesised due to a discrepancy between the bedrock solid geology and the overlaying drift topography. Each of these possibilities will be addressed in turn.

For a potential variety of reasons, the agricultural piedmont dwelling potters of the Kopet Dag study area eventually curtailed their use of organic temper in favour of the adoption of sand temper during the late Neolithic and early Aeneolithic. One possibility is that as time went on and settlements grew, potentially necessitating more wide-scale pottery production, perhaps potting was no longer a strictly seasonal activity. The increase in potting activity would require a more readily available tempering source at any time of year. The adoption of sand temper could also have been due to cultural contact with either the northern Keltiminar peoples, or from contacts to the south in modern-day Iran and Afghanistan. Certainly, later Aeneolithic and Bronze Age wheel-turned wares necessitated a finer form of temper, perhaps more easily manipulated for the construction of finer wares. Whatever the reason for the change encompassing the continuance of handmade wares, it is clear that there was a marked shift from the use of organic to sand temper during the late Neolithic/early Aeneolithic.

In the case of a central production site for the Kopet Dag study area, certain characteristics would be expected. It would be important, but not absolutely critical for the substantiation of a central pottery production location, to find the production site itself. Unfortunately, no such site has been documented or excavated in the Kopet Dag study area. If the pottery were produced in a central workshop, a moderate level of craft specialisation and societal organisation is implied. The lack of general craft specialisation or a highly organised societal structure provides an initial stumbling block to a central production hypothesis, especially for the early and middle Neolithic, during which only the most basic mode of production, the initial advent of irrigation agriculture, and semi-sedentary settlement is evidenced. In order to substantiate or refute this initial hypothesis for fabric homogeneity, we will have to look more carefully at the Neolithic mode of production and its characteristics. To reach such a conclusion, it would be necessary to encounter less variability in the assemblages produced at a central pottery production site than those assemblages created through household production. This variability would pertain not only to production technology, but also to form, ware, decoration, and fabric. Fabrics were more homogeneous than expected. We need to look more carefully at the reasons for fabric homogeneity and at other characteristics which may substantiate one of the two main hypotheses suggested to explain fabric homogeneity. In looking broadly at the variability of the piedmont ceramic assemblages, there is simply too much variability present in the aspects of form, ware, decoration and fabric to substantiate such a possibility. This variability, the Neolithic mode of production, and the fact that no central pottery production location serving a range of sites is known for the piedmont, all point to the necessity to develop the second hypothesis for fabric homogeneity. This second hypothesis focuses on the differentiation between bedrock solid geology and the overlying drift topography.

Vessel Repair Masson reports evidence of the repair of broken vessels (Masson 1971). This was accomplished by drilling small holes on either side of the break and fastening the broken edges together with the use of plant fibre. Obviously, in this case, the vessel would no longer maintain its impermeable qualities, but could otherwise be used for the storage of dry goods. Provenancing The provenancing study is a comparison of the thinsection mineralogy with the solid geology and drift topography of the Kopet Dag study area. As I pointed out in the geological predictions, the mineralogy of the Kopet Dag study area exhibits more homogeneity than a study of the bedrock solid geology alone would suggest. This section will therefore rely on a comparison of the petrographic study with the bedrock solid geology and drift topography outlined above. The petrographic fabric descriptions outlined in the fabric classification are in accordance with what we would expect to find in pottery produced from locally-sourced raw materials. Let me preface this section by stating that in carrying out the petrographic and mineralogical study of the Kopet Dag piedmont Neolithic ceramic assemblages, it became apparent that there is more pottery fabric homogeneity than was initially expected based on the three-zoned hypothesis for geological differentiation across the study area. Initially, the two possibilities for this homogeneity will be addressed. Firstly, fabric homogeneity could be due to the existence of one central production workshop

Figures 6.5 and 6.6 represent the ware, mineralogical, and temper observations, while Figures 6.7 and 6.8 contain only those observations which are potentially diagnostic for the regional geological differentiation of the samples. The tables function best when reviewed in combination with the fabric textural descriptions in Appendix 3 and the photomicrographs in Figures 6.9a, 6.9b, and 6.10. 105

THE KOPET DAG CASE STUDY

Mineralogy Ware Organics Quartz Plagioclase K-spar Clay Pellets Iron Oxide Sandstone Limestone Muscovite Biotite Amphibole Epidote Chert Poly Quartz Goethite Calcite Oolites Fossil Frags Loess

1 R

2 B

3 R

4 B

5 B

6 R

7 W

8 B

9 R

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 Mineralogy R R B R B B W B G B G B B B B B B W R B W Ware Organics Quartz Plagioclase K-spar Clay Pellets Iron Oxide Sandstone Limestone Muscovite Biotite Amphibole Epidote Chert Poly Quartz Goethite Calcite Oolites Fossil Frags Loess

Figure 6.5 Overview of mineralogy Mineralogy Ware Organics Quartz Plagioclase K-spar Clay Pellets Iron Oxide Sandstone Limestone Muscovite Biotite Amphibole Epidote Chert Poly Quartz Goethite Calcite Oolites Fossil Frags Loess

31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Mineralogy O O O W O O W R R B BR R B G B R R O R R B B G B B G G B G BR Ware Organics Quartz Plagioclase K-spar Clay Pellets Iron Oxide Sandstone Limestone Muscovite Biotite Amphibole Epidote Chert Poly Quartz Goethite Calcite Oolites Fossil Frags Loess

Figure 6.6 Overview of mineralogy Mineralogy Ware Quartz K-spar Plagioclase Amphibole Muscovite Biotite Epidote Calcite

1 R

2 B

3 R

4 B

5 B

6 R

7 W

8 B

9 R

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 Mineralogy R R B R B B W B G B G B B B B B B W R B W Ware Quartz K-spar Plagioclase Amphibole Muscovite Biotite Epidote Calcite

Figure 6.7 Geologically diagnostic mineralogy

106

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Mineralogy Ware Quartz K-spar Plagioclase Amphibole Muscovite Biotite Epidote Calcite

31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50 51 52 53 54 55 56 57 58 59 60 Mineralogy O O O W O O W R R B BR R B G B R R O R R B B G B B G G B G BR Ware Quartz K-spar Plagioclase Amphibole Muscovite Biotite Epidote Calcite

Figure 6.8 Geologically diagnostic mineralogy

Jeitun PPL

Jeitun XPL

Jeitun PPL

Jeitun XPL

Chopan PPL

Chopan XPL

Togolok PPL

Togolok XPL Figure 6.9a Thin-section photomicrographs 107

THE KOPET DAG CASE STUDY

Togolok PPL

Togolok XPL

Pessedjik PPL

Pessedjik XPL

Pessedjik PPL

Pessedjik XPL

Chagylly PPL

Chagylly XPL

Chagylly PPL

Chagylly XPL Figure 6.9b Thin-section photomicrographs

108

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Chakmakali PPL

Chakmakali XPL

Chakmakali PPL

Chakmakali XPL

Mondjukli PPL

Mondjukli XPL

Djebel PPL

Djebel XPL Figure 6.10 Photomicrographs

109

THE KOPET DAG CASE STUDY

To summarise, the Jeitun Culture ceramic assemblages appear to be locally produced of locally-sourced raw materials. There were two basic types of Jeitun Culture ceramics, chaff-tempered wares and sand-tempered wares. Chaff-tempered wares were predominant during the early and middle Neolithic, whereas sand-tempered wares became more widely used (however, not exclusively) starting in the late Neolithic. The addition of sand temper was most likely linked to the advent of yearround workshop pottery production due to the convenient availability of raw materials. The Keltiminar Culture and Keltiminar-related groups utilised sand tempers exclusively throughout the Neolithic period. The geological predictions and findings of this case study are

consistent with what would be expected from locally produced wares. The variability of the assemblages, except at Chopan, evidence household production during the early and middle Neolithic. There was an eventual shift during the late Neolithic/early Aeneolithic (and during the middle Neolithic at Chopan) to a more centralised workshop mode of ceramic production. The Neolithic Jeitun Culture was a generally non-pottery exchanging culture, producing and utilising ceramics in a domestically oriented manner. This adaptation changed during the subsequent Aeneolithic and Bronze Age periods, where pottery was traded widely, evidencing cultural contact and communication networks with neighbouring groups to both the north and south.

110

CHAPTER 7 SYNTHESIS AND INTERPRETATION I will use this chapter to synthesise the results from Chapters 4 and 6 in order to provide an overview of the structure of the Neolithic in southern Turkmenistan. The fundamental research questions I will address are: 1. 2. 3. 4.

generalisations regarding vessel colour and decoration motifs for the three phases of the Jeitun Neolithic. Unfortunately, there is sometimes no correlation between vessel surface colour and ceramic fabric. Despite commenting on vessel colour, Soviet scholars made no attempt to categorise colour in comparison to form or function. Thus, it proves difficult to use what limited Soviet data exists to determine a correlation between form and fabric. Through my survey of 1319 vessel fragments, there appears to be little standardisation of the correlation between vessel form (or at least vessel function interpolated by sherd coarseness) and ceramic fabric type. The Jeitun Culture ceramic assemblages are generally categorised by a high level of variability in the characteristics of ware, form, decoration, and fabric. Thick-walled storage vessels and thin-walled domestic vessels are often made of a similar constituent clay matrix with similar mineral inclusions, but no hard and fast rules exist.

What can the ceramics tell us about the history of the Jeitun Culture (both in terms of early homogeneity and subsequent heterogeneity)? What can the ceramics tell us about modes of production and their changes through time? What can the ceramics tell us about distribution, exchange, and consumption? How can the above questions be synthesised into a cohesive picture of the Neolithic?

As a preface to this synthesis, it is important to mention several discrepancies in the ceramic data. First, and perhaps most striking, is the fact that many of the ceramic assemblages studied by the Soviet scholars cited and documented in Chapter 4 were destroyed due to lack of storage space. We are therefore left with the original, and sometimes controversial, interpretation of data (see below) rather than an objective presentation of results, and unfortunately no means for verification or falsification of the conclusions presented. Secondly, in relating the current petrographic study to the Soviet typologies for form and function, every effort was made to provide accurate comparisons based on general trends in form and fabric. However, I was prohibited from sampling material from complete vessels (of which only 12 survive). The thin-section samples in my petrographic study are from sherds often too small to discern the original vessel profile, but representative of the range in ceramic fabrics (for the surviving collections) of a specific site. The data presented in Chapter 6 is the best fabric classification possible, taking into account the restrictive conditions in availability of material for either analysis or verification.

The findings of the statistical analyses presented in Chapter 4 showed several major trends. Vessel function may be loosely interpreted by sherd consistency, i.e. coarse wares are representative of large storage vessels whereas fine wares are representative of domestic vessels. The intermediary category of vessels of medium consistency is more ambiguous, but most likely also applies to the category of domestic wares. Figure 4.21 is representative of the breakdown, by percentage, for vessel consistency and basic function group (either storage or domestic ware). Because there is considerably more ceramic data available from the typesite of Jeitun than for any of the other Kopet Dag piedmont Jeitun Culture sites, slightly more detailed analyses on the ceramic assemblage from the site were carried out. At Jeitun, small bowls make up 32%, storage vessels 30%, large bowls 26%, and plates or platters 12% of the overall assemblage (Figure 4.18). The distribution by ware for these four forms (Figure 4.19) breaks down as follows: 80% of small bowls are redware and only 20% buffware. This is broadly suggestive of a preference for finer redware, rather than buffware, domestic vessels. The percentages of the other three categories, large bowls, storage jars, and plates or platters, are more or less similar, representing an approximate 50% split, thus suggesting no clear preference for either redware or buffware in these vessel categories.

Statistical Analyses: Correlation of Form and Fabric It was possible to carry out a limited amount of statistical analyses based on the data from the ceramic assemblages in this study. These statistical analyses serve to synthesise the raw data compiled as part of the typological and petrographic case study. In so doing, the statistical analyses serve to provide some framework for the interpretation of the data and the structuring of answers to the above questions. The following is a summary of the comparisons compiled.

Finally, the trends for the distribution of consistency (roughly analogous to vessel function) by site assemblage through time (Figure 4.21) provide interesting insight into the changing and developing nature of the Neolithic.

Soviet data makes infrequent and inconsistent mention of the correlation between vessel form and fabric. As I will point out in the next section, the Soviets make 111

SYNTHESIS AND INTERPRETATION Figure 4.21 is presented with site assemblages listed in chronological order along the x-axis. There is a distinct increase in coarser wares through time from the early Neolithic (represented at Jeitun, Chopan and Togolok), through the middle Neolithic (represented at Chopan, Togolok, and Pessedjik), to the late Neolithic (represented at Chagylly, Chakmakli, and Mondjukli). This trend remains on a steady increase until the inhabitation of Chakmakli and Mondjukli, where there are significant increases in the amount of fineware evidenced. I interpret this as due to technological progression in tandem with outside influence and the increased dispersal of material cultural remains from neighbouring groups to the north and south.

to population pressure and limited water resources in the central zone. Berdiev provides comparative data regarding the percentage of decorated as opposed to undecorated ceramics for each regional assemblage for phases 2 and 3 as evidence for regional variation. Unfortunately, he gives no such data for phase 1. During Jeitun Culture phase 2 (middle Neolithic), 45% of ceramics from the western region, 25% of ceramics from the central region, and only 10% of ceramics from the eastern region were decorated (Berdiev 1969, 57). These figures show a more marked discrepancy during Jeitun Culture phase 3 (late Neolithic) with less than 2% of all ceramics in the eastern region and 45% of ceramics in the western region decorated. These statistics show that there was indeed considerable variation in ceramic typology during the late Neolithic.

In summary, both petrographic and typological variability show a broad trend toward decreasing variability consistent with the progression from household to workshop pottery production during the period between the early and the late Neolithic. Initially, in the early and middle Neolithic assemblages from Jeitun, Togolok, and Pessedjik, there is a high level of variability in the assemblages. Chopan (also middle Neolithic), however, represents a unique and relatively early shift to workshop production, most likely necessitated by the large size of the site and the demands of its population for domestic ceramic wares. Subsequently by late Neolithic times, a further shift occurred, this time caused by increasingly sophisticated technology, a preference for and the ability to produce finer wares. This shift to finer wares was most likely driven by outside influence, increased cultural contact, and proto-exchange networks that fostered an increasing dispersal of material goods. The abovementioned statistical analyses are complemented by the following broad-based conclusions derived from previous Soviet research carried out by Berdiev (1969) and Masson (1971).

Masson provides detailed tables which calculate the percentages of decorated and undecorated ceramics by region, but show no differentiation by time period or site (1971, 35). Presumably, the calculation of such statistical information was prescribed by the formulaic approach to archaeological interpretation and analysis. Masson’s statistics, however, do not prove particularly useful or accurate. Masson reports that for the western region, 116 vessel fragments or 10.7% of the overall excavated assemblage (by who, where, and in which years is unknown) were decorated while 962 vessel fragments or 89.3% of the assemblage were undecorated; for the central region, 61 vessel fragments or 8.8% of the assemblage were decorated while 634 vessel fragments or 91.2% of the assemblage were undecorated; for the eastern region, 117 vessel fragments or 19% of the assemblage were decorated while 472 vessel fragments or 81% of the assemblage were undecorated. Therefore, Masson gives the grand total for the Jeitun Culture as 294 vessel fragments or 12% of the entire excavated assemblage decorated and 2068 vessel fragments or 88% of the entire excavated assemblage undecorated. This table obviously does not agree with the findings reported above by Berdiev. It is impossible to determine where discrepancies lie because Masson chose to omit categorisation by time period and site. I am more inclined to accept Berdiev’s findings because of the detailed breakdown by time period and by site. I see Masson’s statistics as an overly broad attempt to answer an unframed question using potentially insufficient and misleading data.

The ceramic assemblages as evidence for Neolithic development Berdiev gives a regional and temporal summary for the development of the Jeitun Neolithic based on the typology of the ceramic assemblages and stone tool industries. Jeitun phase 1 (early Neolithic) existed solely in the central Kopet Dag piedmont. Jeitun phase 2 (middle Neolithic), however, covered the western, central and eastern piedmont watershed zones. Finally, during Jeitun Culture phase 3 (late Neolithic), settlement was limited to the western and eastern zones only. This overview points to the initial peopling of the central zone of the Kopet Dag piedmont. The initial population growth which occurred due the adoption of a more sedentary lifestyle combined with pressure on natural resources and the gradual aridification of southern Turkmenistan, caused the population to spread out across the entire piedmont during the middle Jeitun-phase. It is likely that more restricted inhabitation of the western and eastern zones, during the late Jeitun-phase, occurred due

To reiterate, Berdiev concludes that local variants were more marked during Jeitun Culture phase 3, as opposed to the relatively homogeneity of the assemblage across the entire Kopet Dag piedmont during Jeitun Culture phase 2 (1969, 58). He explains that as the Jeitun peoples were forced to migrate further afield from the original area of settlement, the central zone, their cultural adaptations developed increasingly independently. A

112

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

comparison of the ceramic assemblage from upper Bami (western archaeological zone) with those from upper Chagylly (eastern zone) and middle Mondjukli (eastern zone), during the late Neolithic, evidences regional typological variants (Berdiev 1969, 59). Berdiev attributes this differentiation to a higher level of aesthetic appreciation and cultural sophistication at Bami (1969, 58). This interpretation seems somewhat presumptuous in terms of the archaeological interpretation of aesthetic appreciation and cultural sophistication. However, if the people of Bami were actually more culturally sophisticated than the inhabitants of the Meana-Chaacha district, we should not simply be content to make such a statement, but we should rather ask ourselves why. Did the people of Bami maintain closer contacts with their Iranian neighbours or the Keltiminar-related groups to the south and west? Could the Meana-Chaacha Jeitun peoples have been similarly influenced toward a minimalist undecorated aesthetic by groups living upstream along the Tedjen river in northwestern Afghanistan? Noting the evidence of regional variation should not be a conclusion in itself, but rather a launching pad for future research to answer such questions. In its westernmost position on the Kopet Dag piedmont, Bami likely maintained contacts with and was influenced by its Jeitun Culture neighbours in the Gorgan valley of northeastern Iran and/or the Keltiminar-related groups of the Balkhan and Uzboi regions. In the same light, the Jeitun people of the Meana-Chaacha district likely maintained contacts with and were influenced by the nearby inhabitants of northeastern Iran and northwestern Afghanistan. There are two convenient corridors for the movement of people and the dispersal of knowledge which form a bracket-like shape on either end of the Kopet Dag range itself. Only future research will provide further evidence of the inner workings of these relationships and interconnections.

complexity, urbanism, and agricultural intensification. Rice suggests (1987, 182) that a “study of mode of pottery production is based on interrelated determinations of how the pottery is made, who makes it, and for whom it is made. Pertinent questions thus relate to manufacturing technology, the role and status of producers, the integration between tasks, the organization of producing units and their relation to the overall economic organization, and the relation between producers and consuming groups”. I addressed the issues of production technology for the Jeitun Culture, now I will focus on the changing nature of the Neolithic mode of production, the relationship of producers to consumers, and the local distribution or perhaps unintentional dispersal of ceramic wares across the Kopet Dag study area. During the early Neolithic, as evidenced at Jeitun, the high level of ceramic variability in ware, form, decoration and fabric points to the existence of household production. Rice (1987, 201) states that “production is a process of variety generation, whereas consumption is variety selection. Together these create cultural patterns distinctive of different sociocultural arrangements”. Sinopoli (1991, 126) believes that stylistic variability in ceramic decoration plays a role in signalling group identification and social boundaries. It seems that at Jeitun, the wide variety of forms, fabrics, and decoration reflects individual taste and household necessity. “Variability in locus, personnel, and resource is reduced or regularised in different productive modes” (Rice 1987, 201). Miller (1985, 1) theorises, in his study of ceramic variability, that “variability in artefacts, as objects created by people, embody the organisational principles of human categorisation processes”. His study of ceramic variation in a rural south Indian village “is intended to investigate the manner in which these organisational principles generate variability in material forms. The variability of objects is significant as a major source of evidence for the study of society, the artefactual environment being one of the main products of the social action”, he anticipates “therefore, that an understanding of the forces which create artefactual variability can also contribute towards an understanding of the social”.

Mode of Production The interpretation of the three petrographic datasets (classification, production technology and provenancing) serves to place the trends in regional and temporal ceramic variation of the Kopet Dag assemblages in terms of the developing Neolithic mode of production of the Jeitun Culture. Beyond these initial comparisons, I will look at the ceramics as evidence for cultural contact, distribution, and exchange. Prehistoric ceramics were often locally produced due to the difficulty in transporting unwieldy and fragile vessels. Moreover, during the early stages of sedentary or semi-sedentary subsistence, there was likely little or no economic specialisation due to the lack of societal organisation and stratification, and thus no means or necessity for the production or distribution of specialised goods. Craft specialisation in southern Central Asia developed gradually with the advent of larger settled agricultural communities during the Aeneolithic and Bronze Age. Production specialisation followed increased societal

Household production, as outlined by Peacock (1982), Van der Leeuw (1984) and Rice (1987), is the initial and most basic step in the production and economic sequences generally agreed upon by the contributing scholarly community. I will not reiterate the genesis of the entire economic sequence, because it seems that the literature is already rife with repetition, despite the authors’ acknowledgement of such and inherent promises to avoid further redundancy! I will, however, discuss the basic differentiation between the initial stage of household production and the later stage of workshop production evidenced in the Jeitun Culture. It appears that household production remained in place

113

SYNTHESIS AND INTERPRETATION during the early Neolithic through at least the middle Neolithic, represented at Jeitun, Togolok, Pessedjik and lower Chagylly. However, there was an eventual shift during the middle Neolithic at Chopan, and more generally during the late Neolithic toward workshop production. This increasingly complex late Neolithic mode of production eventually developed into the highly specialised Aeneolithic and Bronze Age modes of production. This is not to say that there was a complete absence of household pottery production during latter periods, but rather a predominance in workshop and possibly regionally-based pottery production. There is evidence that coarse handmade wares continued to be produced well into the Iron Age in southern Turkmenia (Hiebert pers. comm. 1997). The household production phase among the Jeitun Culture was marked by a high level of ceramic variability (ware, form, decoration, and fabric), most likely seasonal production, and little or no exchange, distribution or dispersal of pottery within or outside the community. During the early and middle Neolithic (with the exception of Chopan), the Jeitun peoples were creating pottery in their households for their own use. This was a distinctly non-exchanging and domestic mode of production.

Distribution and Exchange Before reviewing the nature of ceramic distribution and exchange for the Jeitun Culture, it will be helpful to lay out definitions and accepted theories about distribution and exchange in general. The mechanisms of distribution and exchange have many possible meanings in a variety of contexts. Rice (1987, 191-192, after Renfrew 1975) suggests that there are two major facets to distribution: reciprocity and redistribution. The notion of reciprocity involves the movement of goods based on symmetrical relationships between individuals. Reciprocity involves such arrangements as hospitality, socially-mandated gifts, or bride price. Redistribution, however, entails a pooling of goods and resources with a centralised control mechanism. Examples of redistribution include potlaching, share-out from a hunt, or the appropriation of goods and/or services by and for a separate entity. Exchange is defined (Rice 1987, 192) as “the process by which goods are moved and change hands in return for some other good, service or intangible”. Encompassed in this definition is the notion of a marketplace or market exchange. Market exchange is, however, exceedingly impersonal when compared to the abovementioned forms of reciprocity and redistribution. Rice (1987, 192) mentions that often, all three aspects of reciprocity, redistribution and exchange are broadly categorised as “trade”. For obvious reasons, exchange networks are important not only in supplying necessary goods, but also are crucial to a flow of information, technology and in the creation of societal or inter-group relationships.

The middle and subsequent late Neolithic shift to organised workshop production is evidenced at Chopan and documented by Berdiev. As the Neolithic Jeitun Culture progressed and gradually developed toward the Anau 1A pre-Namazga Aeneolithic and later urbanism of the Namazga civilisation, similar forms of workshop production would likely have been adopted elsewhere. Workshop production of pottery led to greater homogeneity in ware, form, decoration and fabric, preceding the introduction of the potter’s wheel during the Aeneolithic. This workshop phase in the ceramic mode of production was characterised by diminishing variability in ware, form, fabric and decoration in the assemblage of a particular site. It is also likely that with the advent of purpose-built pottery production areas, potting became a less seasonally dependent activity and was probably carried out on a year-round basis. The idea that potting shifted from being a seasonal activity to a year-round activity is supported by the shift in temper type from organic (fine chaff, gramineae, cereal grains) to sand. Charles and Bending (pers. comm. 2001) are currently carrying out further analyses on the gramineae to determine species. Provisionally, they interpret the gramineae as a grass species which flowers late in the harvest cycle, probably in summer (rather than during spring, the typical flowering time for grasses in the region). The shift to the use of sand temper instead of organic temper was documented by Soviet researchers, (Masson 1971, Berdiev, and Lollekova 1978b), but never researched. The shift may be attributed to technological development, but it is more likely that both the factors of seasonal access to agricultural waste materials and/or seasonally flowering grasses, and changing technology due to outside influence, were joint causes of this shift.

During the Neolithic period in southern Turkmenia, the Jeitun peoples maintained contacts not only across the Kopet Dag piedmont and with the neighbouring groups in northeastern Iran and northwestern Afghanistan, but also with the Keltiminar and Keltiminar-related groups of Kara Kum, Kyzyl Kum, Balkhan and Uzboi regions. This is evidenced by the (perhaps unintentional) dispersal and/or exchange of items of material culture such as ceramic vessels or stone tools. There is very limited evidence to suggest the dispersal or exchange of ceramics between the two groups. The limited nature of dispersal and/or exchange was probably in part due to the non-exchanging, domestic mode of production previously outlined for the Jeitun Culture. However, it is clear that the Jeitun peoples were aware of the Keltiminar and Keltiminar-related groups and vice versa, but for environmental and societal reasons did not overlap in their modes of subsistence and remained entirely separate as cultural groups. The ceramics of the two groups are distinct in form, fabric and decorative techniques. While Jeitun ceramics are chaff-tempered, flat-bottomed, and painted, Keltiminar ceramics are sand-tempered, round or pointed-bottomed, and have incised decoration. I discussed the implications for the use of organic versus sand temper and the eventual shift from organic to sand temper in the Kopet Dag piedmont in Chapter 6. The differences in vessel form, most 114

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

importantly bottom-shape, have wider implications for vessel usage and cultural tradition. The flat-bottomed vessels used by the Jeitun people are more stable when placed on a flat surface, however, they are less efficient for heating their contents if used over a fire because of the relatively low amount of surface area actually exposed to direct heat and the thermal differentials caused by the joint of the vessel wall at an angle to its base (Rice 1987, 241-242). But, as I noted in Chapter 4, Masson (1971) hypothesised that the Jeitun people did not cook directly over a flame, but used hot stones to heat the contents of the vessel, or otherwise cooked meat on a spit. Thus, the flat-bottomed form would have been useful in providing a stable platform for the vessel while its contents were being heated by hot stones. The round or pointed-bottomed vessels used by the Keltiminar, Bolshoi Balkhan and Uzboi Cultures require some sort of support, frame, or method of suspension over the fire. These vessels are efficient at heating their contents over an open fire because of the large surface area that is exposed to the flames while the vessel is propped or suspended over the fire. These systematic differences in the shape of the vessel base point to distinct cultural practices in terms of food preparation and consumption.

Creating a Cohesive Picture of the Jeitun Neolithic How can the above ceramic data for the Jeitun Culture specifically the history of regional and temporal variation, mode of production, distribution and exchange, be synthesised into a cohesive picture of the Jeitun Neolithic? During the early and middle phases, the Jeitun Neolithic represented a household mode of production in terms of ceramics, with an eventual shift to a more organised and standardised workshop mode of production. In light of distribution and exchange, the Jeitun Culture was characterised by a generally domestically-oriented non-pottery exchanging Neolithic adaptation. The shift to a standardised workshop mode of production and a complex set of exchange networks, which began in the late Neolithic and was later firmly established during the Aeneolithic period, is representative of the changing Neolithic economy and intensification of the subsistence patterns of the region. The Jeitun peoples of the Kopet Dag piedmont zone were the first agro-pastoralists of southwestern Central Asia, and as such, they determined the path of future economic intensification, population expansion, and the subsequent origins of urbanisation which occurred in the Bronze Age.

In terms of evidence for cultural contact, samples of limestone and calcite-tempered Keltiminar ceramics exist in the archived assemblage from the typesite of Jeitun, housed at the Hermitage Museum in St. Petersburg. Similarly, chaff-tempered Jeitun Culture ceramics were excavated by Soviets and surfacecollected by the British team at the Bolshoi Balkhan cave site of Djebel. Although these occurrences represent limited evidence, their existence across cultural and geographic boundaries proves some form of cultural contact and exchange. As mentioned above, such limited contact and exchange of ceramics points to the existence of a homogenous, domestic, generally non-pottery exchanging Neolithic adaptation among the Jeitun Culture agriculturalists. This does not exclude the possibility of the exchange of lithics or other items of material culture. Further research will be necessary in order to reach such conclusions.

In Chapter 8, I will look more closely at the nature of the relationships between the semi-sedentary Jeitun agriculturalists of the “southern super-zone” and the nomadic Keltiminar and Keltiminar-related hunters, fishers and stockbreeders of the “northern super-zone”, as well as the relations with northeastern Iran and northwestern Afghanistan. The focus of this comparison will be in terms of subsistence system, cultural tradition and the nature of contact, innovation and cultural exchange. Most importantly, I will address the issue of why such a stark cultural boundary existed, one which effectively separated the Keltiminar and steppe traditions to the north from the Jeitun, Iranian, Afghan, Baluchistani, and broader Mesopotamian traditions to the south. As an archaeological borderzone and the “northeastern frontier of the ancient Near East” (Tosi 1973-74), southern Turkmenia provides unprecedented opportunities for the comparative analysis of these disparate cultural traditions.

115

CHAPTER 8 NEOLITHIC CONTRASTS AND COMPARISONS Chapter 7 synthesised the results from the typological and petrological ceramic studies presented in Chapters 4 and 6. In interpreting this data, I outlined the structure of the Jeitun Neolithic in terms of mode of production and as a domestically oriented generally non-pottery exchanging culture. Using these conclusions as a baseline for the current chapter, I will explore the following main points in conclusion: 1) the idea of Turkmenia as a significant archaeological borderzone and a “hard frontier” 2) the Jeitun Culture as a potentially semi-sedentary rather than completely sedentary subsistence adaptation 3) the relationships between the Jeitun people of southern Turkmenia, northeastern Iran, and northwestern Afghanistan 4) the relationship between the Jeitun Culture and the Keltiminar Culture and Keltiminar-related groups 5) the shift from the Jeitun agricultural Neolithic to the pre-Namazga Aeneolithic Anau 1A period 6) topics for future research Southern Turkmenia as an Archaeological Border Zone I think it is ultimately crucial to lay a broader framework to convey the significance of Turkmenia as an archaeological borderzone during prehistory. Turkmenia lies at a major juncture, a significant cultural boundary where traditions to the north, beginning with the Keltiminar, stretched for thousands of miles northward, northwestward and northeastward. Similarly, cultural traditions contemporaneous with and analogous to the agro-pastoral adaptation of the Jeitun Culture stretched for thousands of miles southwestward through Iran, Mesopotamia, and the Near East; and southwastward through Afghanistan, Pakistani Baluchistan and the Indus valley. I am attempting to create an image of greater Eurasia painted with two broad swathes of colour, one representing the settled or semi-sedentary agriculturalists of the south and the other representing the nomadic stockbreeders of the north. Because so much of the archaeological research during the Soviet period was limited to the investigation of individual sites as insular entities, only more recently have Soviet scholars (Masson in particular) joined western scholars (Tosi, LambergKarlovsky, Dolukhanov and others), in their attempts to create broader generalisations covering large geographical areas during prehistory. Although I have discovered many smaller-scale comparisons between areas immediately adjacent to one another, I prefer to address this issue of cultural similarity and differentiation as broadly as possible. The geographical and

climatological hypotheses laid out by D. Christian (1998) serve to explain the subsistence systems and economies of inner Eurasia, the territory inhabited by the Keltiminar and Keltiminar-related groups. Although much study has focused on the Jeitun Culture and the prehistoric settled or semi-settled agricultural adaptation in southern Turkmenia as existing on the edge or frontier of the ancient Near East, I believe it is equally pertinent to view the Keltiminar and Keltiminar-related groups in terms of their position on the border of the greater steppeland nomadic stockbreeding adaptation of inner Eurasia. Previous Soviet work has never defined the Keltiminar or Keltiminar-related groups as similarly marginal to a broad-ranging cultural/geographical adaptation. This is in part due to the Soviet-style investigation of archaeological sites or cultures as isolated entities. It was due to Tosi’s analysis of the Jeitun Culture and the subsistence adaptation of the Kopet Dag piedmont, that southern Turkmenia was first termed “the northeast frontier of the ancient Near East”. This was not a Soviet analysis. In looking at the issues of differences in culture, economy, and exchange networks, I regard the analysis of both the Jeitun Culture and the Keltiminar Culture as marginal to the broader subsistence regimes of their respective regions is critical. The differentiation in subsistence systems between the Jeitun and Keltiminar Cultures was caused not only by cultural tradition, contact and dispersal, but also by environmental and geographical restrictions. This “hard frontier” characterised by dryness (Sherratt pers. comm. 2001), functioned as a barrier to cultural and economic synthesis until the Islamic period. The Keltiminar and related steppe, semi-steppe and desert adaptations were governed by environmental processes altogether different from those dictating the subsistence potential of the fertile oases of the Near East and Mesopotamia. “Inner Eurasia’s geographical coherence appears most clearly on a physical map of the world. Its dominant geographical feature is a vast plain, the largest unified area of flatlands in the world...Three main features define Inner Eurasia ecologically. These are: (1) interiority; (2) northerliness; and (3) continentality. First of all, Inner Eurasia is inner. It is remote from the sea to the west, south, and east, and its long northern coastline is frozen for most of the year. So interiority has meant aridity...too dry to support farming without irrigation...Second, Inner Eurasia is northerly...The southern parts of Inner Eurasia lie in the same latitudes as the northern Mediterranean and Central Europe, but, with less rainfall, and more extreme climates, they are largely regions of steppe and cold desert. Though these were the first parts of Inner Eurasia to be settled in prehistoric times, even these lands 117

NEOLITHIC CONTRASTS AND COMPARISONS lie north of latitudes in which the first civilizations emerged in Egypt, Mesopotamia, Northern India, and China...There is an important sense in which Inner Eurasia is colder than Outer Eurasia...Latitude also affects rainfall. Warm equatorial winds carry moisture upwards, which they shed as they travel north and south away from the equator. Eventually, having shed most of their rain, they descend at around the 30° parallel, creating most of the world’s desert lands...The flatness and size of Inner Eurasia explain a third main feature: the continentality of its climates...Aridity, northerly latitudes and continental climates combined to create a harsh environment for human settlement...the most accessible parts of Inner Eurasia are more impoverished ecologically than most parts of Outer Eurasia...The most obvious consequence of Inner Eurasia’s harsh ecology was low population density...Low population densities might simply have ensured that Inner Eurasia remained marginal to the dominant regions of world history...However, Inner Eurasia’s centrality in the Eurasian landmass ensured it a more prominent historical role. What happened in Inner Eurasia affected Outer Eurasia because most of the land routes connecting the various civilizations of Outer Eurasia [the Near East, Mesopotamia, the Indus valley, and China] passed through Inner Eurasia. As a result, the history of this region had a considerable impact on the rhythms of Outer Eurasian history” (Christian 1998, 4-9). In this light, it is crucial to look at the Jeitun Culture and the Keltiminar Culture as individual entities, and to understand them as such, but also to look at the dynamics at work between the two cultures, their respective geographical interaction spheres, their differing trajectories of economic and social development, and the borderzone between them as a locus for cultural exchange and the dispersal of traditions and knowledge. In order to create a useful baseline for the analysis of the Jeitun and Keltiminar Cultures, a clear understanding of the parameters of the Jeitun agricultural adaptation is crucial. Contrary to previous Soviet interpretation, Gosden proposed (pers. comm. 2001) that the Jeitun Culture may not have been as entirely sedentary as previous Soviet research and publication indicates. The Jeitun Culture as a semi-sedentary adaptation It seems entirely possible that the Jeitun Culture could have functioned (especially in the early and middleNeolithic) as a semi-sedentary agricultural adaptation, part of a seasonal-round subsistence system incorporating the exploitation of resources at different times of the year in the upland intermontane valleys, the piedmont, the floodplain, and the desert. Of course, such a hypothesis will remain unclear until further substantiating research can be conducted. Semi-sedentism is initially indicated by the multiple alternating floor layers and destruction layers found in the house structures at Jeitun and the other sites of the Kopet Dag piedmont. Soviet research documents multiple floor layers at various sites, but fails

to offer theories as to how and why there were so many successive layers or what length of time was represented by each one. A recent study by N. Boivin (2000) comparing ethnographic data from rural Rajasthan with archaeological data from the Neolithic Anatolian site of Çatal Höyük led to the following conclusions. “...the temporal rhythms of human life articulate with changes in the nature of village floors. Various temporal cycles, including the annual cycle, the lifecycle of the individual and the development cycle of the domestic group, find expression in this particular form of material culture. The varying experiences of the lifecycle by different groups within society are also, on some occasions, linked to variation in the changes undergone by house floors” (2000, 367). Boivin suggests that the anthropological study of floor layers and their association with differing time cycles can and should be applied to archaeological time in order to create a better understanding of the life rhythms and use versus disuse of particular households. Boivin states that “The particular aspect of my research that is of relevance concerns the variation through time that is commonly observed in floor sequences. My ethnoarchaeological research has shown that such variation derives primarily from two factors: 1) the laying down of floors of varying composition..., and 2) variation in the activities carried out on floors. Furthermore, and of pivotal importance to the discussion at hand, I have observed that the activities that generate this variation relate to the distinctive temporality of life, and in particular to ritual activities...” (2000, 369). For instance, in the Rajasthani ethnoarchaeological case study, there were special ritual floor layers laid down to mark births, marriages, deaths, widowhood, the harvest, and religious festivals. There was also a regular schedule of repair, maintenance and cleaning which produced further variation in the thickness and accumulation of the floor layers. The variation between floor layers laid down on a seasonal or yearly basis as opposed to those which were dictated by the lifecycle of the individual or the family group created considerable differentiation between the make-up and thickness of each layer, producing a complex deposit. Although I do not suggest the situation at Jeitun was as complicated as that recounted by Boivin for the rural Rajasthani village, her observations serve at least to partly explain the great variability in floor layers encountered at Jeitun. Clearly, site abandonment is not an issue in Boivin’s interpretation of floor layers in rural Rajasthan or at Çatal Höyük. I am attempting, in constructing this analogy, to partly explain the possible reasons for the excessive number of floor layers apparent at Jeitun and the other sites of the Kopet Dag piedmont in light of the relatively short occupation at those sites. There were two sets of cycles at work in the construction of the multiple floor layers and house destruction layers evidenced in the Jeitun Culture house structures. The first was a seasonal set of floor layers within each house while the second was the actual destruction and rebuilding of the houses themselves. It is difficult to interpret the destruction and wind-blown sand layers recounted by both Soviet researchers and the British team 118

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

at the piedmont sites, but it seems plausible that these indicate intermittent periods of disuse and abandonment. Otherwise, what would justify inhabiting a debris and rubble-scattered or sandy living surface at some times rather than others? One answer may be that during the periods of general disuse, there was “camping” taking place at the site. Recent radiocarbon dating carried out by the British team at Jeitun provided useful insight into the relatively short occupation of the site in comparison to the multiple cultural layers and construction/destruction layers. Much shorter than previously hypothesised by Soviet investigators, the dating carried out at the Oxford Radiocarbon Accelerator Unit points to an occupation spanning no more than 760 years, but most likely less due to the calibrations and low probability of the range presented for certain dates. “The deposits at Jeitun appear to have accumulated very rapidly and it is impossible to sustain a detailed phasing of the site on the basis of the radiocarbon chronology...At two standard deviations the dates span some 600 years, so we can say with 95% probability that all the dates for occupation at the site fall within this timespan” (Harris et al. 1996, 436, 440). Regarding the British analysis of the multiple floor level phenomenon, Harris et al. report the following. “A further conundrum arises when we consider the complexity of the occupational history of the two houses excavated in 1993 and 1994. Each house contained a large number of layers, composed either of living floors with gypsum surfaces or layers of sand or mudbrick destruction. Such a complicated history raises the possibility that each house was not occupied continuously (until natural decay took its toll and a successor was built), but that they were occupied discontinuously, i.e. that although the site was ‘permanent’ the people in it were not. Seasonal transhumance is often included in models of prehistoric patterns of mobility and , in the Jeitun case, the fact that the mountains of the Kopet Dag are within 50 km of the site might seem to increase the likelihood of such a pattern of life” (1996, 440). They also report the zooarchaeological analysis of sheep and goat slaughter patterns carried out by Legge in 1992, remarking that the small size of the sample analysed may have skewed the results of the study, but nevertheless, no seasonal kill patterns were evidenced (1996, 440). The existence of seasonal kill patterns would have supported the theory that the site was inhabited on a semi-sedentary basis. Harris et al. conclude that “the evidence (discussed above) for the use, intermittent abandonment, and reuse of at least parts of the site within a short period of (radiocarbon) time, suggests that all of it was not occupied continuously through many centuries” (1996, 440). Interestingly, Kohl and Heskel (1979) do suggest such a seasonal-round semi-sedentary subsistence system for the Darreh Gaz plain (figure 8.1), the area immediately to the south of the Kopet Dag piedmont. Although the majority of known Jeitun Culture sites

actually lie in the piedmont zone, floodplain, and desert, the site of Gademi is particularly significant in light of the question of semi-sedentism because of its position in the lower mountain terraces of the east Kopet Dag. Gademi is the highest altitude Jeitun Culture site known to date and evidences intermittent occupation (periodic destruction and building layers) with similar Jeitun-type mud-brick house structures. It is likely that other sites also exist in the low mountain areas, and that they could have provided convenient bases from which to exploit seasonal resources. Further research is needed both in Turkmenistan and northern Iran to clarify to what extent the Jeitun Culture practised a semi-sedentary lifestyle. The determination of semi-sedentism rather then total sedentism would provide valuable insight into the structure of the Neolithic in southern Turkmenia. An understanding of this structure would ultimately lead to more accurate comparison of the Neolithic adaptation of southwestern Central Asia with surrounding regions, and would serve to better place the Jeitun Culture and prehistoric Turkmenia in the broader context of the Near East, Mesopotamia, the Indus valley, and the steppelands of inner Eurasia. Relationships with Iran and Afghanistan Due to modern-day political boundaries, the contiguous study of the area inhabited by the Jeitun people has been significantly and unnaturally divided, an artificial border was effectively created with the demarcation of the Soviet Union. The sheer fact that we refer to these adjacent areas by different geographical names serves to separate them in a subconscious manner. These two archaeological zones, the mountains and mountain valleys (northern Iran) and the piedmont and desert floodplain (southern Turkmenistan), should instead be viewed as different subsistence adaptation areas of the same archaeological culture. Unfortunately, despite significant Soviet research in southern Turkmenistan, restrictions during the Soviet period led to the study of Turkmenian archaeology in complete isolation from that of northern Iran. It must also be said that archaeological research in northern Iran was much less widespread or intensive than the research undertaken within the confines of the former Soviet Union. Masson (1971), Masson and Sarianidi (1972), Tosi (1973-1974), and LambergKarlovsky (1973) addressed the broader issues of commonality in the background of the Jeitun peoples of southern Turkmenistan, northeastern Iran and northwestern Afghanistan. Masuda (1974, 1976), Deshayes (1967), Sarianidi (1970), and Kohl and Heskel (1979) reported on individual sites in relation to the overall picture of the Jeitun Culture as one unit crossing modern-day political boundaries. Tosi states that “It is indeed in the second half of the 6th millennium that southern Turkmenia and most of eastern Iran are definitively included in the Near Eastern zone of interchange” (1973-1974, 31). It is my opinion that cultural contact and interchange occurred significantly earlier than Tosi suggests. 119

NEOLITHIC CONTRASTS AND COMPARISONS

Figure 8.1 Settlement distribution on the Darreh Gaz plain (after Kohl & Heskell 1979) For the purposes of this study, we are concerned with the northern territory of Iran encompassing the Iranian plateau, the Gorgan valley, and the Darreh Gaz and Meshed-Misrian plains. Sarianidi (1970) provides an analogy for the geography of northern Iran and southern Turkmenistan. He suggests that the area inhabited by the Jeitun Culture people was shaped like a large flattened triangle. The sites on the Iranian side of the modern-day border included Sang-e Caxamaq, Tureng Tepe, and Yarim Tepe. The geographical area was created by the intrusion of the Turkmeno-Khorassan mountains into the

territory of southern Turkmenistan. Sarianidi (1970, 22) imagines the area as spanning from the Gorgan valley in the southwest (bottom left corner of the triangle), across the Kushan-Meshed valleys (the baseline of the triangle), to the Meana-Chaacha district in the southeast (bottom right corner of the triangle), and finally to Kyzyl Arvat in the north (the top of the triangle). Sarianidi hypothesises that the oldest aceramic Neolithic sites lie in the mountains. He does not mention Sang-e Caxamaq on the Iranian plateau, originally interpreted (Masuda 1974) as containing early aceramic levels, but suggests that the 120

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Jeitun people migrated initially from the Elburz mountains, through the Gorgan valley of northern Iran, moving northeast along the sub-tropical corridor along the southeast Caspian shore, through Kyzyl Arvat, and on to the Kopet Dag piedmont (Sarianidi 1970, 22-23). Sarianidi does not, however, exclude the possibility of a local origin and a separate late Neolithic cultural variant for the inhabitants of the Meana-Chaacha district sites (1970, 23). Although Sarianidi does not explicitly state exactly what he means by “local origin”, I assume he refers to the possibility that people settled the eastern Kopet Dag piedmont and the Meana-Chaacha district from Mundigak and Quetta in northern Afghanistan and Pakistani Baluchistan via the upper reaches of the Tedjen river above Serakhs. Although the main known sites in northern Iran are Sange Caxamaq, Tureng Tepe and Yarim Tepe, Kohl and Heskel (1979) also recorded a variety of smaller sites in the Darreh Gaz plain. Sarianidi hypothesised that the sites of northern Iran evidence the transition from the Caspian Mesolithic to the Jeitun Neolithic (1970, 21). As I mentioned in chapter 2, I find it unlikely that the Jeitun Neolithic developed directly from the Caspian Mesolithic due to the strong parallels between the Caspian Mesolithic and the Keltiminar and Keltiminar-related Neolithic, and the stark differentiation between the Caspian Mesolithic and the Jeitun Neolithic both in terms of economy and mode of production. It is possible that earlier interaction between the proto-Jeitun Culture and the proto-Keltiminar Culture was already in place as early as the Mesolithic/Neolithic transition. Further research will be necessary in order to draw any firm conclusions. Kohl and Heskel’s survey of the Darreh Gaz plain (Figure 8.1) north of Kuchan in November 1978, reports numerous sites from the Aeneolithic through the Islamic period in this intermontane valley of the Kopet Dag. The following is an outline of the pertinent prehistoric sites, of which there are four: DG2 (also called Yarim Tepe, but different from the Yarim Tepe excavated by D. Stronach near Gunbad-e Quabus), DG14 (Kara Khuyulu), DG19 (Chapishlu), and DG28 (Nowkhandan) date to the Aeneolithic and Bronze Age. “The purpose of these reconnaissance visits was to locate and provisionally date prominent archaeological sites in the plain and map in greater detail the major prehistoric settlement(s), the surface remains of which could be related to Aeneolithic and Bronze Age materials from Soviet Turkmenistan” (Kohl and Heskel 1979, 160). The fertile Darreh Gaz valley is separated from the Kopet Dag piedmont by a low mountain range called the Zarin Kuh, but sits north of the main ridges of the Kopet Dag. Kohl and Heskel state that “the Darreh Gaz plain simply represents an upper extension of the fertile foothills zone of the northern Kopet Dag” (1979, 161). Kohl and Heskel cite the archaeological research carried out by R. Invernizzi of Turin University as evidencing the ties between the prehistoric sites of the Iranian Khorassan and southern Turkmenistan. These investigations showed prehistoric

connections between the upper Atrek valley and Kopet Dag piedmont, the Murghab delta, and the Bactrian plain. In line with my previous argument that the Jeitun Culture was most likely semi-sedentary, Kohl and Heskel state that the “four [prehistoric Darreh Gaz] sites are distributed throughout the major [environmental] subregions of the [Darreh Gaz] plain, a fact preliminarily suggesting a dispersed settlement pattern throughout the Aeneolithic and Bronze Age periods” (1979, 163) and thus supporting the idea of a diversified subsistence round. The ceramics from the sites in Darreh Gaz are Jeitun-type assemblages also reminiscent of the ceramic finds at Yarim Tepe in the Gorgan valley. Kohl and Heskel suggest that occasional abandonments on the Kopet Dag piedmont proper were interspersed with occupations at sites in the Darreh Gaz plain (1979, 163). Although they remark on Aeneolithic and Bronze Age occurrences, it seems likely that a similar pattern of movement, use, disuse and reuse could have been in place during Neolithic times. Kohl and Heskel conclude with the following statement: “during prehistoric times settlements in northern Khorassan interacted most intensively with sites to the north in southern Turkmenistan...In other words, two major cultural areas stretching north to south tentatively can be drawn. The first links sites on the Gorgan plain with Hissar and related sites stretching, at least, as far west as the Teheran oasis. This area also includes ancient Daghistan, the Sumbar valley, and sites in the western atak [Kopet Dag piedmont] almost as far east as Ashkhabad. A second major zone, which in turn can be subdivided into western and eastern components, links the lowland fans of the Tedjen and Murghab rivers, the fertile piedmont belt of the northern Kopet Dagh mountains, the Darreh Gaz plain, and the intermontane Atrek and, presumably, Kashaf Rud valleys to form a highly integrated area of cultural interaction” (1979, 169). Kohl and Heskel favour the theory of constant cultural interaction between northern Iran and southern Turkmenistan during prehistory, and in particular the initial interdependence of upland, piedmont, and lowland components of the subsistence regime which later grew into a complex trade and exchange network between the increasingly settled inhabitants of these differing environmental zones. In accordance with LambergKarlovsky, Kohl and Heskel refute hypotheses suggesting cultural influence and dispersal of traditions as a onesided phenomenon from south to north (Kohl and Heskel 1979, 169). Lamberg-Karlovsky is a fierce advocate not only of the study of the interdependent nature of development in southern Turkmenia, northern Iran and Mesopotamia, but 121

NEOLITHIC CONTRASTS AND COMPARISONS also of active research comparison of the interaction between these areas, especially southern Turkmenistan and northern Iran. Lamberg-Karlovsky states that “Turkmenia is almost always depicted as a recipient of ‘civilization’ from Iran or Mesopotamia. It appears on the contrary, to me, that from the times of the Djeitun Culture to the end of the Iron Age, Turkmenia appears as an independent interaction sphere of primary and parallel development, which not only derived much from the west but also contributed to it. Only through an attempt to isolate the interdependence of Turkmenia, Iran, and Mesopotamia shall we begin to shed light on the processes which gave rise to that first great ‘internationalization’ which so characterizes Western, Central and Southern Asia” (1973, 45). In light of Aeneolithic and Bronze Age comparisons which may be used as a basis for previous similar interaction, Tosi states that during the Aeneolithic “initial population expansion thus leads to a widening of the territory which is exploited by the production system. This territory is later extended so as to cover the whole piedmont area and the delta of the Tedzen, a river whose upstream section above Serakhs links southern Turkmenia directly with the oasis of Herat in northwest Afghanistan” (Tosi 19731974, 33). Lamberg-Karlovsky points out that Bronze Age “Excavations at Shar-i-Sokhta [Iranian Seistan] have shown the strong influence of Namazga III correlations to Quetta and Mundigak [Afghanistan and Pakistani Baluchistan]. Tosi remarks that “It must be assumed that the south Turkmenian populations succeeded in producing a new expansionist phase in the period Namazga III at the end of the 4th millennium, i.e. when the Tedzen delta underwent a diminution and a further shift towards the west thereby forcing the populations settled there gradually to abandon a fertile territory. South Turkmenistan cultural elements appear simultaneously in the Hilmand and Quetta valleys, that is to say, 1,000 kilometres further south, with such a high frequency as to suggest that a true migratory flow was occurring” (Tosi 1973-1974, 35). Thus, recent work suggests that Turkmenistan was influencing the areas of northeastern Iran and northwestern Afghanistan and was not entirely the cultural backwater and continuous recipient of external influences that previous Soviet publication had argued. The strong correlations between Namazga III and Shar-i-Sokhta I, Shar-i-Sokhta II/III and Yahya IVB, and Yahya IVB and Early Dynastic Mesopotamia, provide not only a new chronological framework but also a new understanding of the complex cultural interaction between these distinctive interaction spheres” (Lamberg-Karlovsky 1973, 45). It seems much more plausible to accept such an interpretation of two-way cultural influence and interaction between southern Turkmenistan and northern Iran. LambergKarlovsky concludes that “...there can be little doubt that Turkmenia, rather than an ‘outpost’ of Mesopotamian civilization, is to be seen as an independent centre interacting with Mesopotamia and the Indo-Iranian borderlands form earliest Neolithic times” (1973, 46).

Jeitun Culture relations with the Keltiminar Now that I have created an image of Eurasia dominated by two separate trajectories of development and subsistence, I will address the interaction between the inhabitants of these two geographical zones. Above, I presented an image of the Jeitun Culture as a potentially semi-sedentary agro-pastoral regime on the edge of, but also exercising influence on, the adjacent cultures of northern Iran, and more broadly, Mesopotamia and the Near East. It is clear from the documentation of the Keltiminar Culture that the Keltiminar and Keltiminarrelated groups practised a nomadic seasonal-round subsistence system, with slight variation depending on physical geography and environmental constraints. Contrary to previous Soviet conclusions regarding southern Turkmenia as the categorical recipient of culture, innovation, and technology, I see the region instead as a crossroads of cultural contact and influence which flowed in both directions. Markov, often one to criticise the trend in standard Soviet-style formulaic publication and an advocate of generalist theories, hypothesised that the movement of people would have occurred from the north along the Uzboi corridor; from the south along the southeast coast of the Caspian Sea, the Aktam and Kelkyor rivers; and from the east along the Kopet Dag piedmont (Markov 1971, 62). It is likely that the main routes for the movement of people and the spread of ideas were river systems. Tosi suggests that exchange networks between the Jeitun Culture area and the Keltiminar Culture area existed from earliest Keltiminar times and traversed Turkmenia en route to the Zeravshan, north to the Aral Sea, and east to the Ferghana valley (1973-1974, 66). Until further research is carried out and secure radiocarbon dating chronologies are established, it is impossible to determine exactly to what extent each cultural area influenced the other. It is my feeling though, that due to the intermittent abandonment evidenced in the floor layers of the Jeitun Culture house structures, the Jeitun people were most likely semi-sedentary. It is therefore also likely that in such a borderzone, cultural interaction, flow of knowledge, and innovation were mixed to such an extent that an intermediate subsistence adaptation occurred. This intermediate subsistence adaptation would have included strong aspects of both sedentism and nomadism, a true mixture of south and north. In this way, Turkmenia was a focal point of interaction with influence coming from both directions. The idea of southern Turkmenia as a focal locus, rather than a cultural backwater, is to an extent a simple expansion on Lamberg-Karlovsky’s theory of interaction between southern Turkmenistan and northern Iran presented above. As I mentioned in chapter 2, Soviet research, until very recently, failed to create generalist hypotheses, and in most cases failed to present analogies or theories of interaction even for immediately adjacent archaeological cultures.

122

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Of course, a full program of radiocarbon dating must be carried out for the Keltiminar and Keltiminar-related sites before sound conclusions may be drawn, but it seems clear that interaction between the Jeitun Culture and the Keltiminar Culture existed. It is plausible that the flow of knowledge was not one-sided, and that some form of intermediate subsistence adaptation existed at this borderzone among the Jeitun Culture people. If it can be explicitly proven in the future that the Jeitun people led a semi-sedentary lifestyle, then their subsistence adaptation could be seen as marginal to and influenced by both the settled agricultural zone of the Near East and Mesopotamia and the nomadic stockbreeding desertsteppeland adaptation of inner Eurasia. Toward the Aeneolithic The shift to the Aeneolithic on the Kopet Dag piedmont represents a significant shift in technology, subsistence, and societal organisation. This period was initially characterised by the Anau 1A pre-Namazga Aeneolithic adaptations found at Anau, Gavoutch, Koushut, Tilkin, Chakmakli and Mondjukli. There are two main theoretical camps for the interpretation of the genesis of the Anau 1A period and more generally the Aeneolithic as a whole. The shift from the Jeitun Neolithic to the Anau 1A and subsequent Namazga Aeneolithic is striking because there is, as yet, no clear chronological progression between the two. Instead, the archaeological record points to a considerable technological leap forward in ceramic and lithic traditions, the introduction of metallurgy, increasing complexity in architecture, site layout and societal organisation. Masson and Sarianidi (1972, 47) report that the original stratigraphic sequence was based primarily on changes in ceramic typology. They also suggest that the Anau 1A adaptation found in the Meana-Chaacha district may represent the northeasternmost extension of the Hassuna-Sialk archaeological complex (Masson and Sarianidi 1972, 51). In postulating such a migration, Masson and Sarianidi state that “It is still hard to say when and why these migrations occurred. The very fact that in Mondjuklidepe the pre-Anau layers overlap the middle Djeitun layers, shows that this process [migration] took place as early as the Neolithic period. Moreover, it has been suggested that, for a certain time, the new-comers coexisted with the local tribes of the Djeitun culture, but further support is needed for this hypothesis” (1972, 51). Furthermore, they state that “One should not, however, oversimplify the process which established the early Chalcolithic communities in southern Turkmenia. This complicated process was, on the one hand, determined by the progressive economic development of the entire preceding Neolithic period, and, on the other, by the innovations which were brought by the new-comers themselves...in general the new-comers were soon absorbed into the local population and adopted many of its cultural traits” (Masson and Sarianidi 1972, 51-52). Tosi (1973-1974, 32) agrees with the above-outlined

theory of Masson and Sarianidi, and argues that the site of Mondjukli provides a perfect example of the “mingling of the traditional and innovatory elements”, thus representing the dispersal of knowledge from the Iranian plateau and Afghanistan. Tosi suggests that the Tedjen delta was first occupied during the Anau 1A period, due to increasing population pressure on the environment and population expansion into a widening territory. This ever-widening territory eventually encompassed “the whole piedmont area and the delta of the Tedzen, a river whose upstream section above Serakhs links southern Turkmenia directly with the oasis of Herat in northwestern Afghanistan” (Tosi 1973-1974, 33). Tosi’s comments support the growing evidence for the earlystage interconnection between Turkmenia, northeastern Iran and northwestern Afghanistan. Lamberg-Karlovsky vehemently disagrees with Masson and Sarianidi and their theory for the peopling of southern Turkmenia during the Aeneolithic and states that “We simply cannot agree with the authors [Masson and Sarianidi] that what appear to be changes in internal cultural processes were the result of ‘tribal migrations’, even less that such population movements provide explanation for the origins of the Geoksyur Culture and even later for the Quetta Culture. Migrations continue into the late Chalcolithic [late Aeneolithic] (Namazga III) when it would appear that tribes throughout the Near East were scrambling for entry into Turkmenia” (LambergKarlovsky 1973, 44). In terms of the change in ceramic assemblages during the Anau 1A period, there is marked differentiation in both production technology and decoration between the ceramic assemblages from Jeitun Culture phase 3 (late Neolithic) and Anau 1A. Whereas the Jeitun ceramics were chaff-tempered, flat-bottomed, and exhibited organic design motifs, the Anau 1A ceramics were sandtempered, concave-based, and exhibited geometric design motifs. The Anau 1A ceramic assemblages also represent a much wider variety of vessel forms than those found in the previous period. Kohl (1984, 65) states that “the relatively abundant metals [of the Anau 1A period], distinctively decorated sand-tempered ceramics, an architectural tradition clearly advanced from Djeitun times, new agricultural implements, and the appearance of non-indigenous semi-precious stones constitute a classic problem in archaeological interpretation: continuity or discontinuity with the previous period”. Hiebert (pers. comm. 1999) suggests that foreign populations from Iran may have settled the MeanaChaacha district and founded the Anau 1A sites there. It remains unclear whether the Neolithic/Aeneolithic transition represents the diffusion of foreign knowledge and technology through indigenous populations, or whether the Anau 1A sites were actually founded by settlers from Iran. Future research will be needed in order to refute or confirm these possibilities. As Masson and Sarianidi (1972, 52) aptly state, it is clear, however, 123

NEOLITHIC CONTRASTS AND COMPARISONS that “The accumulation of practical knowledge, improvement of agricultural techniques, the resulting higher crop yields and important developments in cattle breeding all tended to cause a sharp increase in the chalcolithic population” which in turn led to the subsequent urbanisation of the Namazga sequence commencing with the Anau 1B and Namazga I periods. Future Research Southwestern Central Asia played a vital role in both the origins of agriculture and urbanism in Eurasia. I have outlined the main features of culture, economy, and societal organisation for prehistoric Turkmenia. In writing this thesis, it became clear that what is most needed in terms of imminent research is the creation of a secure radiocarbon dating chronology for both the Jeitun and Keltiminar Cultures. As mentioned already, much of the Soviet literature has unfortunately focused on drawing conclusions from relative chronologies. What is needed is a baseline for future research, a sequence that will allow accurate comparison and that will silence the controversy surrounding relative chronologies. Imminent research should investigate the internal development of the Neolithic as a springboard for subsequent urbanisation during the Aeneolithic and Bronze Age. I envisage a long-term project with three main sets of objectives, the most immediate of which are: 1) to establish a secure absolute dating chronology for prehistoric Turkmenia through key-hole excavation of both Jeitun and Keltiminar sites 2) to understand the distribution of Neolithic sites through the analysis of satellite imagery, geomorphological analysis and surface survey 3) to understand the Neolithic adaptations, subsistence systems, and interaction between the Jeitun and Keltiminar Cultures, especially how the interaction and influence sphere created at this archaeological borderzone fits into the broader Eurasian archaeological context. Such a study would not only focus on southern Turkmenia, but also would include a concentrated campaign of excavation in northern Iran to ascertain the extent of Jeitun Culture sites immediately across the modern-day political border. A geoarchaeological project with a strong emphasis on landscape study using the analysis of satellite imagery and GIS will allow us to develop a predictive model of where sites may be visible in the present day landscape that can be tested through fieldwork. We will use high resolution satellite imagery (Soyuz KFA 1000, SovInforSputnik, SPOT, IKONOS, IRS, ERS, KVR and Landsat) to construct hypotheses for

verification including the possible location of unknown archaeological sites, the geomorphology of the Neolithic landscape, and the relationship of known and unknown sites to this Neolithic landscape. The methodology will include geoarchaeological and geomorphological analyses of remotely-sensed images including spectral bands in the visible-infrared-radar range of differing geometrical resolution and differing times of acquisition. These images will be elaborated through geometric and radiometric corrections, thematic classification and principal components analysis. The piedmont region of southern Turkmenistan and the desert oases are landscapes that have undergone extreme environmental change in the last 8,000 years because of climatic fluctuations and human influence. It is important to understand the impact these changes had on the environment. We will also attempt to verify to what extent deposits from the Bronze Age and later may have covered earlier sites. On the basis of the satellite imagery, we will gain both a general overview of the region, but also pick out smaller areas in which to carry out more detailed work in the future. Geomorphological ground-truth survey will be carried out in tandem with the archaeological fieldwork. It will consist of the surface mapping and sampling of deposits for OSL dating. The overall aim of the analysis of satellite imagery and ground-truth verification is to create the first reliable model of landscape change in Turkmenistan for the last 8,000 years and to understand the relationship of the Neolithic sites to their landscapes. The compilation and synthesis of satellite imagery and surface survey will entail the integration of data from previous research (both British and Soviet), remote sensing analysis, and ground survey in a GIS environment. The creation of a GIS will allow for the analysis of spatial and temporal aspects of the data produced, the reconstruction of diachronic settlement models (archaeological sites in their geomorphological setting), and archaeological interpretation regarding the production and trade of artefacts. In this way we will gain a clearer understanding of the extent of both the Jeitun and Keltiminar Neolithic adaptations. This landscape study will be supported by a secure absolute dating chronology, rather than the previous reliance on typological comparison. I see such a broad-based project as essential for laying a spatial and temporal framework that may be used as a stepping stone for future investigation of the Neolithic and subsequent periods in Turkmenia. Further study of prehistoric developments in southwestern Central Asia would then be set within the wider Near Eastern and Eurasian trends towards settled agriculture, nomadic pastoralism, and subsequent urbanism.

124

APPENDIX 1 GLOSSARY aeolian transport: the movement of sediment by wind allochems: organised aggregates of carbonate which form within the basin of deposition. They include ooids, bioclasts, peloids, and intraclasts. See oolite. alluvial fan: a fan or cone of material deposited by a stream where it debouches from a mountain front onto a plain, with the apex of the fan at the point of emergence from the mountains. amphibole: a group of rock-forming minerals that have silicon-oxygen tetrahedra linked to form a double chain. Amphibole is pleochroic in plane polarised light (PPL), exhibits medium/low relief, and contains two sets of cleavages. anticline: an arch-shaped fold into which rock strata have been compressed, the oldest rocks occurring in the core. barchan: a crescentic dune characterised by an oval shape upwind, with a slip face and two wings spreading out downwind. Barchans occur in desert with a uni-directional wind regime. bentonite: clay formed by the alteration and weathering of tuffs and volcanic ash. biotite mica: a mineral that has a layered structure in which cations are sandwiched between sheets of silicate tetrahedra and hydroxyl ions. Micas have a perfect basal cleavage which reflects their layered structure. Biotite is common in schists, gneisses, and granitic rocks. It is pleochroic in plane polarised light (PPL) and may be red, brown or black. birefringence: the property of anisotropic crystals to split a beam of light into two beams that pass through the crystal at different speeds, producing characteristic optical effects seen under crossed polars (XPL) in a polarising microscope; the difference between the largest and smallest refractive indices. boulder clay: the unstratified material deposited by glaciers and ice sheets. burnish: a method of producing a luster on an unfired clay surface by rubbing it while leather-hard with a hard, smooth object to compact and align the surface particles. calcite (calcium carbonate CaCO3): is usually colourless or white and cleaves into perfect rhombs. Its extreme birefringence is apparent in crystals of the transparent variety. Most limestones consist largely of calcite, which may be a primary precipitate or in the form of fossil shells. The main sources of calcite are limestone, shell, and limescale caused by post burial alteration of archaeological materials. Calcite also occurs in ore mineral veins. carnelian: is a variety of chalcedony which is red in colour. Chalcedony is found in hydrothermal veins,

amygdales and sediments. Other forms of chalcedony include agate, jasper, chert, and flint. chalcedony: a uniformly coloured cryptocrystalline silica mineral. chert: and flint are opaque, dull-coloured, or black varieties of chalcedony, the former being the massive or stratified form and the latter being found as nodules in chalk. cleavage: the tendency of a rock to break into closely spaced planar structures or fractures as a result of deformation or metamorphism. Cleavage in minerals refers to the tendency to split along planes of weakness in the molecular structure. coiling: the method of hand-building an object of clay by successive additions of ropes or coils of clay (variants include ring-building, spiral coiling, and segmental coiling). conglomerate: a coarse-grained sedimentary rock in which the constituent clasts or fragments are more or less well-rounded. crystal habit: the shape of a crystal arising from the shape and size of faces and the development of different forms. diabase: a microgabbro containing calcic plagioclase and augite. In addition, diabase may contain olivine, hypersthene, quartz, or feldspathoids. Diabase occurs mainly as dykes, plugs and sills. diagenesis: post-depositional compaction and chemical reaction processes by which loose accumulated sedimentary material becomes sedimentary rock. dolomite: is a colourless, white, or grey calciummagnesium carbonate mineral. It results from the substitution of calcium ions in calcite by divalent cations of magnesium and iron. drift: Glacial and fluvioglacial deposits. Great thicknesses of drift accumulated during the Pleistocene period, though much has subsequently been removed by erosion. drift topography: an area of glacial deposition. Solid geology, as opposed to drift topography, refers to the actual underlying rock. efflorescence: the process by which soluble salts in solution in the wall of a vessel are deposited on the surface during evaporation of the water. epidote: a green to yellow-green monoclinic mineral found in the medium-grade regionally metamorphosed rocks of the greenschist and amphibolite faces. Epidote is produced during retrograde metamorphism and may be found on joint surfaces and along fractures. Plagioclase feldspar, clinopyroxenes, and hornblendes can all be replaced by epidote minerals and basic igneous rocks, in particular, may suffer extensive epidotisation. 125

APPENDIX 1

fabric: the constituents of fired pottery, including matrix, inclusions (solids and voids), and pores, but excluding surface coatings; often used synonymously with body, paste or ware. fault: a fracture in the earth’s crust along the plane of which there has been displacement of rock on one side relative to the other, either in a horizontal, vertical or oblique sense. feldspar: the most abundant minerals in igneous rocks consisting of four main components: calcium, sodium, potassium, and barium. There are two feldspar series, plagioclase (Na-Ca) and potassium (K-Na) feldspar. Plagioclase feldspars are milky white or colourless, triclinic and exhibit multiple lamellar twinning. Plagioclase feldspar has a range of sodium and calcium variants. Potassium feldspar is a form of alkali feldspar that exhibits perthitic twinning. Alkali feldspars are colourless, white, pink or red. firing: heating pottery to a temperature high enough to cause permanent destruction of the clay minerals; rendering pottery hard and durable by heating; heating and cooling in a kiln. firing core: a zone of contrasting colour in the crosssection of a vessel wall, caused by and indicative of variations in atmosphere and rate of heating and cooling during firing or incomplete/short firing. flint: a nodular variety of chalcedony formed in chalk. glauconite: a bright green authigenic mica-like mineral which is found in marine sediments, particularly greensands. gneiss: a coarse-grained metamorphic rock in which quartz and feldspar predominate over micas. The schistosity is poorly defined and segregation banding is irregular and discontinuous. Course granular bands of quartz and feldspar alternate with thin often undulating bands in which micas and amphiboles are concentrated. grog (chamotte): crushed pottery used as an additive to clay. gypsum: a monoclinic evaporite mineral found in shales and limestones. haematite (goethite): a ferric iron oxide often occurring as a cement in sandstones producing a red coloration. Hercynian: a phase in the Variscan (mountain-building during the late Palaeozoic) orogeny affecting Europe and characterised by a northwest fold trend. The term is used without time significance and covers the Carboniferous and Permian. incised decoration: a common surface treatment used throughout prehistory and later. It involves dragging a sharp instrument through leather-hard clay. inclusion: any constituent of a fired pottery vessel or fragment that was not originally a clay mineral, such as voids, plant remains, rock fragments, temper, clay pellets, etc...

interference colours: the colours visible in minerals viewed through crossed polars in a polarising microscope when the mineral causes the weakening or destruction of certain wavelengths of a beam of light. intraclast: an eroded fragment of calcareous or noncalcareous material not derived from pre-existing limestones. Clastic sediments are sediments consisting of fragments of broken rock (clasts), eroded and transported generally to a different site to be deposited. jasper: a red opaque variety of chalcedonic quartz. kaolin: a clay mineral produced during the weathering and hydrothermal alteration of feldspars under acid conditions. The large-scale production of kaolinite by the alteration of granite gives rise to china clay deposits. leather hard: stage of drying when plasticity is minimal but water content is still sufficient to allow some kinds of modifications; the intermediate stage between plastic and bone dry. limestone: a rock formed from carbonate minerals, principally calcite but including others such as dolomite. Organic limestones are formed from the calcareous skeletons of living organisms. limonite: the yellow to brown amorphous and cryptocrystalline oxidation and hydration products of iron. The constituents include haematite, goethite, colloidal silica and clay minerals. marble: a rock composed largely of calcite or dolomite, produced by the regional or contact metamorphism of limestones. matrix: the part of a pottery fabric originally composed of clay minerals. micrite: chemically precipitated calcite with microscopic grains and present in some limestones. muscovite mica: a mineral that has a layered structure in which cations are sandwiched between sheets of silicate tetrahedra and hydroxyl ions. Micas have a perfect basal cleavage which reflects their layered structure. Muscovite mica is common in schists, gneisses, and granitic rocks. It is colourless in plane polarised light. monocline: a bending of rock strata produced in sedimentary sequences that have deformed under conditions favouring the development of a normal fault. Two areas of horizontally bedded sediments are left at different elevations but still connected by a steeply inclined series of the same beds. oolite: an allochemical limestone formed predominantly from ooliths. Ooliths are spherical bodies formed by the precipitation of carbonate in concentric layers around a nucleus. oxidation: the chemical weathering process involving the reaction between rocks and atmospheric oxygen often resulting in the production of iron oxide. oxidising: atmosphere during firing in which the amount of oxygen is more than required to combust the fuel.

126

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

paste: a clay or mixture of clay and added materials, often used synonymously with fabric, body, or ware. Technically paste differs from fabric because it does not include pores, and differs from ware because it excludes surface treatment. petrography: the study of mineralogical and textural relationships within rocks revealed by observation of thin-sections and hand specimens. petrology: the study of rocks involving mineralogy, chemistry, petrography, and petrogenesis. phenocryst: large generally euhedral crystals contained in many igneous rocks and set in a fine-grained matrix. piedmont: a gentle slope leading from the foot of a mountain range down to comparatively flat land. plagioclase: one of several soda-lime feldspar minerals. plasticity: the property of clay that allows change of shape when pressure is applied and retention of the new shape when the pressure is removed. pleochroism: the property of a material of differentially absorbing light that vibrates in different directions as it passes through a crystal; the change in colour of a mineral viewed under plane polarised light (PPL) in a polarising microscope. point count: a method of determining the quantity of grains or inclusions in a material by counting individual grains of different phases along a line (or within a transect) as the specimen is viewed in a microscope. polarising microscope (petrographic miscroscope): a microscope that incorporates filters to modify the vibration direction of light as it passes through a thinly-ground section of a mineral and permits characterisation of that mineral by its optical properties. It employs transmitted light, but its distinguishing features are a rotating stage and a pair of polarising filters. These filters limit the vibrations of light to one plane and are mounted in the microscope at right angles to each other. When both are in position (crossed polars or XPL), the field of view is black until a thin-section is placed on the microscope stage; many of the minerals present will have the ability to rotate the light so that it can pass through the top polariser (analyser) and so will appear coloured. When the stage is rotated, minerals or clay matrices that are anisotropic (optically active) will change colour, going black at intervals of 90º, while isotropic (optically inactive) minerals and clay matrices will always appear dark. Such properties (isotropism, anisotropism) may help in the identification of the minerals present. When the lower polariser only is in position (plane polarised light or PPL), the field of view is white and many of the minerals present will be transparent. Others, however, may show colour and will exhibit pleochroism, i.e. they will change colour when the stage is rotated. provenance: 1) refers to the actual site of origin of clay or inclusions; 2) an analysis carried out with the aim of discovering the geographical origin of an artefact,

usually by chemical and/or mineralogical characterisation of the composition of the artefacts in question and comparison with raw materials in a particular area of interest (e.g. pottery constructed from Nile clay with a provenance on the Red Sea coast with diagnostic inclusions with a provenance or rock outcrop in the Ethiopian and Sudanian highlands. In this example, the clay and inclusion provenances are not the same). pumice: highly vesicular, usually acid, volcanic rock. pyroxenes: a group of ferromagnesian rock-forming minerals. Sub-divisions include clinopyroxenes (monoclinic) and orthopyroxenes (orthorhombic). Pyroxenes have a continuous chain structure of silicate tetrahedra linked by sharing two of the four corners. quartz: is a mineral made entirely of silica dioxide (SiO2). This is trigonal without cleavage and may appear colourless (rock crystal), yellow (citrine), grey-brown to black (smoky quartz), pink (rose quartz), and violet (amethyst). Quartz is stable over a wide range of physical conditions and extremely common. It is an essential constituent of silica-rich igneous rocks such as granite. quartzite: a rock consisting almost entirely of quartz and having a pale colour. In its true sense, quartzite is the product of metamorphism of a pure sandstone during which process the quartz grains recrystallise and become interlocking. However, the term is used loosely, and may refer to other massive quartz such as vein quartz or silica-cemented sandstones. reducing: atmosphere during firing in which carbon monoxide is present because insufficient oxygen is available to combust fuel. sandstone (terrigenous or land-based deposit): a lithified arenaceous deposit of clastic sedimentary rock which may be bound by a secondary cement such as calcite or various iron minerals or welded together by pressure. scarp (escarpment): a steep cliff-like slope often of considerable size, which rises above the surrounding land surface. A fault scarp results from faulting. schist: a strongly foliated coarse-grained rock in which mica minerals are abundant and their sub-parallel orientation produces a marked schistosity. Schists are produced through regional metamorphism and are characterised by the segregation of minerals into thin layers alternately rich in micaceous minerals and quartz/feldspar. schistosity: a variety of slatey cleavage that is well developed in schists and gneisses. shale: a well-laminated argillaceous sedimentary rock that is fissile and splits easily along bedding planes. The fissility is related to the disposition of clay minerals within the rock. slag: glassy material formed in a kiln as a by-product of heating. slip: a fluid suspension of clay or clay-body in water; a non-vitreous coating applied to a pottery vessel. 127

APPENDIX 1

solid geology: the geology of rocks underlying glacial and other superficial deposits. sorting: the process by which materials forming a particular sediment are graded according to size by natural processes. syncline: a basin-shaped fold in which the beds dip towards each other. temper: inclusions (either organic, sand, chamotte, or unfired clay) which are added to the pottery fabric by the potter (and may be in addition to sand occurring naturally in the clay) rather than occurring naturally in the pottery clay. Tethys Sea: prior to the northward movement of Africa relative to Europe, a large sea existed, of which the Mediterranean sea is the surviving remnant. Within the Tethys sea, the sediments of the AlpineHimalayan orogenic belt were deposited. The Tethys sea closed as the result of sea-floor spreading and the wedge of sediments was compressed into the present-day mountain chains. thermal expansion: increase in the size (volume) of a ceramic material (e.g. a ceramic body) when heated. thermal shock resistance: the ability to withstand sudden changes in temperature or cycles of heating and cooling without damage, such as cracking.

thin-section: a thin slice of rock or ceramic cut and ground to a uniform standard thickness of 30 microns and mounted on a glass slide for petrographic study under a polarising microscope. tuff: consolidated ash, where ash is defined as tephra fragments less than 2 mm in diameter. Ash and tuff are pyroclastic rocks formed by the accumulation of fragmented materials thrown out by volcanic explosions. typology: a theoretically oriented classification directed toward solving a problem. ware: a ceramic material in the raw or fired state (e.g., greenware, earthenware, stoneware, etc.); a class of pottery whose members share similar technology, fabric, and surface treatment. waster: vessel damaged during the manufacturing process, particularly during firing; also, sherds from such vessels. workability: the suitability of a plastic clay body for forming pottery, as judged by its feel to the potter. Not synonymous with plasticity (a clay may be too plastic to be workable).

128

APPENDIX 2 CERAMIC TYPOLOGY

Figure A2.1 - Jeitun ceramic vessels

129

APPENDIX 2

Figure A2.2 – Jeitun ceramic vessels

130

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.3 Jeitun ceramic forms and decoration (after Masson 1971) Figure A2.5 Jeitun ceramic forms, decoration and artefactual assemblage (after Masson 1992)

Figure A2.4 Jeitun vessel forms (after Masson 1971)

Figure A2.6 Jeitun ceramic forms and decoration (after Masson 1971)

131

APPENDIX 2

Figure A2.9 Jeitun ceramic forms and decoration, grinding stones (after Masson 1971) Figure A2.7 Jeitun ceramic forms and decoration (after Masson 1971)

Figure A2.8 Jeitun ceramic forms and decoration (after Masson 1971)

Figure A2.10 Jeitun ceramic decoration motifs (after Masson) 132

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.11 Jeitun ceramic motifs (illustration James Reeve)

Figure A2.13 Jeitun ceramic motifs (illustration James Reeve)

Figure A2.12 Jeitun ceramic motifs (illustration James Reeve)

Figure A2.14 Jeitun ceramic motifs (illustration James Reeve)

133

APPENDIX 2

Figure A2.15 Jeitun ceramic motifs (illustration Jennifer Coolidge)

Figure A2.17 Chopan ceramic forms and decoration (after Berdiev 1972)

Figure A2.16 Chopan ceramic forms and decoration (after Berdiev 1972)

Figure A2.18 Chopan ceramic forms and decoration (after Berdiev 1972) 134

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.19 Chopan ceramic forms and decoration (after Berdiev 1972)

Figure A2.21 Togolok ceramic forms and decoration (after Berdiev 1964)

Figure A2.22 Togolok ceramic forms and decoration (after Berdiev 1964) Figure A2.20 Chopan ceramic motifs (after Berdiev 1972)

135

APPENDIX 2

Figure A2.23 Togolok ceramic motifs (illustration Kathy Allinson)

Figure A2.25 Togolok, Pessedjik, and Chakmakli ceramic motifs (illustration Kathy Allinson)

Figure A2.24 Togolok, Pessedjik, and Chakmakli ceramic motifs (illustration Kathy Allinson)

Figure A2.26 Pessedjik ceramic forms and decoration (after Berdiev 1970) 136

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.29 Pessedjik ceramic forms and decoration (after Berdiev 1968)

Figure A2.27 Pessedjik ceramic forms and decoration (after Berdiev 1970)

Figure A2.28 Pessedjik ceramic forms and decoration (after Korobkova)

Figure A2.30 Pessedjik ceramic forms and decoration (after Korobkova, Berdiev 1968)

137

APPENDIX 2

Figure A2.33 Pessedjik ceramic motifs (after Berdiev 1970)

Figure A2.31 Pessedjik ceramic forms and decoration (after Korobkova, Berdiev 1968)

Figure A2.34 Pessedjik ceramic motifs (after Lollekova 1978) Figure A2.32 Pessedjik ceramic motifs (after Berdiev 1970) 138

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.37 Pessedjik ceramic motifs (illustration Jennifer Coolidge)

Figure A2.35 Pessedjik ceramic motifs (after Lollekova 1978)

Figure A2.38 New Nisa ceramic forms and decoration (after Berdiev 1965)

Figure A2.36 Pessedjik ceramic motifs (illustration Kathy Allinson)

139

APPENDIX 2

Figure A2.41 Chagylly ceramic forms and decoration (after Berdiev 1966) Figure A2.39 Kelyata, Kantar, Kepele, Naiza ceramic forms and decoration (after Berdiev 1971)

Figure A2.42 Chakmakli ceramic forms and decoration (after Berdiev 1971)

Figure A2.40 Chagylly ceramic forms and decoration (after Berdiev 1966) 140

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.45 Chakmakli ceramic motifs (after Berdiev)

Figure A2.43 Chakmakli ceramic forms and decoration (after Berdiev 1971)

Figure A2.46 Chakmakli ceramic motifs (after Masson and Sarianidi 1972)

Figure A2.44 Chakmakli ceramic forms, decoration, and artefactual assemblage (after Berdiev 1968)

141

APPENDIX 2

Figure A2.47 Chakmakli ceramic motifs (after Berdiev 1971)

Figure A2.49 Chakmakli ceramic motifs (after Berdiev 1971)

Figure A2.48 Chakmakli ceramic motifs (after Berdiev 1968) Figure A2.50 Mondjukli ceramic forms and decoration (after Berdiev 1972) 142

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.53 Bami ceramic forms and decoration (Berdiev 1963)

Figure A2.51 Bami ceramic forms and decoration (after Berdiev 1963)

Figure A2.54 Djebel ceramic typology by occupational phase (Okladnikov 1956)

Figure A2.52 Bami ceramic forms and decoration (after Berdiev 1963)

143

APPENDIX 2

Figure A2.57 Djebel ceramic forms (after Okladnikov 1956) Figure A2.55 Djebel ceramic vessel with decoration (after Okladnikov 1956)

Figure A2.58 Djebel ceramic vessel with decoration (after Okladnikov 1956)

Figure A2.59 Djebel ceramic vessel with decoration (Okladnikov 1956)

Figure A2.56 Djebel ceramic forms and artefactual assemblage (after Okladnikov 1956) 144

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.60 Djebel ceramic forms and decoration (after Okladnikov 1956) Figure A2.62 Djebel ceramic forms (after Okladnikov 1956)

Figure A2.61 Djebel ceramic forms and decoration (Okladnikov 1956)

Figure A2.63 Djebel ceramic forms and decoration (after Okladnikov 1956)

145

APPENDIX 2

Figure A2.64 Djebel ceramic forms (after Okladnikov 1956) Figure A2.66 Djebel ceramic forms and decoration (after Okladnikov 1956)

Figure A2.65 Djebel ceramic forms (after Okladnikov 1956)

Figure A2.67 Djebel ceramic forms (after Okladnikov 1956) 146

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

Figure A2.68 Dam Dam Cheshme I ceramic profile and motifs (after Markov 1981)

Figure A2.70 Oyukli ceramic profiles and motifs (after Markov and Khamrakuliev 1980)

Figure A2.69 Oyukli ceramic profiles and motifs (after Markov and Khamrakuliev 1980)

Figure A2.71 Oyukli ceramic profiles and motifs (after Markov and Khamrakuliev 1980)

147

APPENDIX 3 FABRIC CLASSIFICATION TEXTURAL DESCRIPTIONS mixing. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote and chert. 8) Similar to samples 4 and 12. Chaff-tempered fabric with conspicuous large, angular, quartz sand. Other inclusions are quartz, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert and secondary carbonate. 12) Similar to samples 4 and 8. Chaff-tempered fabric with conspicuous large, angular quartz sand. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, sinoclast pyroxene, and secondary carbonate. 14) Chaff-tempered very fine-grained fabric with trace quartz sand in the clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, amphibole, and secondary carbonate. 15) Chaff-tempered coarse-grained fabric with abundant large, angular quartz sand in the clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole epidote, chert, limestone and secondary carbonate.

Neolithic central zone: Jeitun redware: samples 1, 3, 6, 9, 10, 11, 13 1) Chaff-tempered loess-rich fabric with the following inclusions: loess intraclasts, mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, chert, clay pellets and sandstone. Secondary carbonate is entirely absent. A high shrinkage clay as evidenced by high angle fractures to the outer surface of the vessel. 3) Chaff-tempered clay/sediment pellet-rich fabric with conspicuous quartz sand in the clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone and secondary carbonate. 6) Chaff and sand-tempered fabric with abundant, wellsorted, large, angular quartz sand in the clay matrix. Due to the well-sorted nature of the quartz sand, it is interpreted as man-added temper. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, occasional limestone grains and secondary carbonate. 9) Chaff-tempered loess-rich fabric with clay/sediment pellets and sandstone grains. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, biotite mica, amphibole and chert. 10) Chaff-tempered fine-grained clay matrix with occasional large, angular quartz sand. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone and secondary carbonate. 11) Chaff-tempered loess-rich fabric with clay/sediment pellets and a pronounced burnish. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, and chert. 13) Chaff-tempered fabric with occasional clay/sediment pellets and large, angular quartz sand in the clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert and secondary carbonate.

whiteware: sample 7 7) Chaff-tempered fine-grained calcareous fabric with the following inclusions: mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, amphibole, epidote, chert, goethite/haematite, and secondary carbonate. The fabric exhibits pronounced clay mixing. greyware: sample 44 44) Sand-tempered loess-rich fabric with loess intraclasts, large limestone grains (one with a calcite vein), very fine feldspathic sandstone or coarse siltstone grains, and sediment pellets in the clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, amphibole, chert, oolitic limestone, biomarle (fossil shell), and mudflakes. The clay exhibits very high shrinkage. Neolithic central zone: Chopan

buffware: samples 2, 4, 5, 8, 12, 14, 15 2) Chaff-tempered loess-rich fine-grained fabric with occasional large, angular quartz sand, occasional loess intraclasts, and occasional clay/sediment pellets. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote and chert. 4) Similar to samples 8 and 12. Chaff-tempered mediumgrained fabric with conspicuous large, angular quartz sand. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone and secondary carbonate. 5) Chaff-tempered calcareous fine-grained fabric with occasional large, angular quartz sand. Pronounced clay

redware: samples 47, 49, 50 47) Chaff-tempered loess-rich coarse-grained fabric with conspicuous large, angular quartz sand, clay/sediment pellets and sandstone (siltstone) inclusions in the clay matrix. The quartz sand contains the same mineral constituents as those observed in the sandstone (siltstone) inclusions. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, biotite mica, amphibole, chert, and secondary carbonate. 49) Chaff-tempered fine-grained fabric with large, angular quartz sand in the clay matrix. Other inclusions 149

APPENDIX 3 are mono and polycrystalline quartz, potassium feldspar, biotite mica, amphibole, epidote, and secondary carbonate. The fabric exhibits some clay mixing. The vessel is slipped on one side. 50) Chaff-tempered loess-rich fabric with loess intraclasts and clay/sediment pellets. Other inclusions are quartz, muscovite mica, biotite mica, amphibole, epidote, chert, and secondary carbonate. The vessel is slipped.

brownware: sample 41 41) Chaff-tempered fabric with clay/sediment pellets, loess intraclasts, and sandstone grains in a fine-grained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, amphibole, epidote, and chert. Neolithic central zone: Pessedjik

orangeware: sample 48 48) Chaff-tempered loess-rich fabric with loess intraclasts, occasional large, angular quartz sand, finegrained siltstone pellets, clay/sediment pellets, and sandstone inclusions. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, biotite mica, amphibole, and chert.

redware: samples 31, 33, 36 31) Chaff-tempered fabric with clay/sediment pellets and abundant large, angular quartz sand in the clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone grains, and secondary calcite. The fabric exhibits some clay mixing. The vessel has slip/self slip on one side. 33) Chaff-tempered fabric with conspicuous large quartz sand in a fine-grained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, and secondary carbonate. The fabric exhibits some clay mixing. The vessel is slipped. 36) Chaff-tempered fabric with occasional clay/sediment pellets and abundant large, angular quartz sand in a finegrained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, and limestone grains.

Neolithic central zone: Togolok redware: samples 38, 39, 42 38) Chaff-tempered loess-rich fabric with clay/sediment pellets, loess intraclasts, and occasional quartz sand in a fine-grained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone and secondary carbonate. 39) Chaff-tempered fine-grained matrix with conspicuous large, angular quartz sand and one very large clay/sediment pellet. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, and chert. 42) Chaff-tempered fabric with clay/sediment pellets, sandstone grains, and loess intraclasts in a fine-grained matrix. Other inclusions are mono and polycrystalline quartz, plagioclase, potassium feldspar, amphibole, epidote and chert.

orangeware: samples 32 and 35 32) Chaff-tempered fabric with abundant large, angular quartz sand in a fine-grained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, and chert. 35) Chaff-tempered fabric with abundant medium quartz sand in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, limestone grains, and secondary carbonate. The vessel is slipped on one side.

buffware: samples 40 and 43 40) Chaff-tempered fabric with occasional large, angular quartz sand in a very fine-grained clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, and calcite. The presence of calcite is notable. 43) Chaff-tempered fabric with conspicuous small, angular quartz sand and limestone inclusions. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, chart, and calcite. The presence of calcite is notable.

whiteware: sample 34 34) Chaff-tempered calcareous fabric with conspicuous small quartz sand in a fine-grained matrix. Other inclusions are quartz, muscovite mica, biotite mica, amphibole, epidote and chert. The fabric exhibits clay mixing. The vessel is burnished. Neolithic eastern zone: Chagylly

whiteware: sample 37 37) Chaff-tempered very fine-grained calcareous fabric with conspicuous small quartz sand in the clay matrix. Other inclusions are mono and polycrystalline quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, and chert. The fabric exhibits some clay mixing.

orangeware: samples 24 and 28 24) Chaff-tempered fabric with occasional quartz sand in a very fine-grained and pure clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, amphibole, and chert. The vessel is slipped. 28) Chaff-tempered fabric with conspicuous small quartz sand in a fine-grained clay matrix. Other inclusions are 150

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

quartz, muscovite mica, amphibole, chert, and secondary carbonate. The vessel is slipped.

plagioclase feldspar, potassium feldspar, muscovite mica, amphibole, and epidote. 20) Sand-tempered fabric with loess intraclasts, abundant large, angular quartz sand, conspicuous limestone fragments, and sandstone fragments in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, epidote, and chert.

buffware: samples 25, 26, 29 25) Chaff-tempered fabric with trace small, angular quartz sand in a very fine-grained clay matrix. Other inclusions are quartz, potassium feldspar, muscovite mica, amphibole, and chert. The vessel is slipped. 26) Chaff-tempered fabric with abundant large, angular quartz sand and occasional limestone grains in a finegrained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, fossil fragments, and secondary carbonate. The vessel is slipped. 29) Chaff-tempered fabric with conspicuous large quartz sand in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, and chert. The vessel is slipped.

Neolithic eastern zone: Mondjukli buffware: samples 21, 22, 23 21) Chaff-tempered fabric with clay/sediment pellets, abundant large, angular quartz sand, and occasional calcite and fossil fragments in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, and oolites. The fabric exhibits clay mixing. 22) Chaff-tempered fabric with conspicuous large, angular quartz sand and occasional calcite and fossil fragments in a fine-grained clay matrix. Other inclusions are quartz, potassium feldspar, muscovite mica, amphibole, and chert. The vessel is slipped on one side. 23) Chaff-tempered fabric with abundant large, angular quartz sand, occasional sandstone fragments, and conspicuous fossil fragments in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, epidote, and chert. The fabric exhibits some clay mixing.

whiteware: sample 27 27) Chaff-tempered loess-rich calcareous fabric with conspicuous large, angular quartz sand in a very finegrained, very well-sorted clay matrix. Other inclusions are quartz, muscovite mica, biotite mica, amphibole and chert. This whiteware is much more fine-grained and well-sorted than calcareous fabrics from other sites in the study area. The vessel is slipped on one side. Neolithic eastern zone: Chakmakli buffware: samples 17 and 19 17) Chaff-tempered fabric with conspicuous large, angular quartz sand and occasional limestone grains in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, epidote, chert, and secondary carbonate. The fabric exhibits some clay mixing. 19) Chaff-tempered calcareous fabric with abundant wellsorted medium quartz sand in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, and chert. The vessel has a fine red slip.

Neolithic western zone: Djebel redware: sample 46 46) Chaff-tempered fabric with clay/sediment pellets, loess intraclasts, limestone fragments, sandstone fragments, and conspicuous large, angular quartz sand in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, muscovite mica, biotite mica, amphibole, and chert. brownware: sample 60 60) Sand-tempered fabric with abundant large limestone fragments, abundant medium, angular quartz sand, abundant fossil fragments and oolites. Other inclusions are quartz, muscovite mica, and chert.

whiteware: sample 16 16) Chaff-tempered calcareous fabric with conspicuous poorly-sorted quartz sand and occasional limestone grains in a fine-grained clay matrix. Other inclusions are quartz, plagioclase feldspar, potassium feldspar, muscovite mica, biotite mica, amphibole, and chert. The fabric exhibits some clay mixing.

Aeneolithic Anau 1A central zone: Tilkin buffware: samples 51, 52, 54, 55, 58 51) Sand-tempered oolite and fossil-rich (bioosparite) fabric with occasional large, angular quartz sand. The oolitic limestone grew on siliclastic grains. Other inclusions are plagioclase feldspar, sandstone and limestone fragments, and calcite. 52) Similar to samples 54 and 58. Chaff-tempered fabric with conspicuous medium, angular quartz sand in a finegrained clay matrix. Other inclusions are plagioclase

greyware: samples 18 and 20 18) Sand-tempered broosparite-rich fabric with abundant oolite fragments (most likely man-added). Broosparite with oolitic limestone fragment nuclei or sand and plagioclase feldspar nuclei which suggests volcanic ash input. Immaline siliclasts in oolitic-forming environment. Quartz sand derives from arkose sediments. Other inclusions are monocrystalline quartz, 151

APPENDIX 3 Other inclusions are plagioclase feldspar, sandstone and limestone fragments, muscovite mica, epidote, and chert. The vessel has a self-slip. 56) Sand tempered calcite-rich fabric with poorly-sorted quartz sand inclusions. Other inclusions are sandstone fragments. The vessel has a fine-grained slip. 57) Sand-tempered loess-rich fabric with small quartz sand in a fine-grained clay matrix. Other inclusions are amphibole. 59) Sand-tempered fossil and oolite-rich fabric in a finegrained clay matrix. The oolitic limestone (which grew on siliclastic grains) is extremely weathered and rotten, suggesting a highly mature sediment. Other inclusions are potassium feldspar, sandstone fragments, chert, calcite, and fossil fragments. The vessel is slipped.

feldspar, muscovite mica, biotite mica, amphibole, epidote, and chert. 54) Similar to samples 52 and 58. Chaff-tempered fabric with conspicuous medium, angular quartz sand in a finegrained clay matrix. Other inclusions are plagioclase, sandstone fragments, amphibole, chert, and calcite. 55) Chaff-tempered fabric with large, angular quartz sand in a homogenous clay matrix. Other inclusions are plagioclase feldspar, muscovite mica, amphibole, epidote, and chert. 58) Similar to samples 52 and 54. Chaff-tempered fabric with conspicuous medium, angular quartz sand in a finegrained clay matrix. Other inclusions are muscovite mica, biotite mica, amphibole, and chert. greyware: samples 53, 56, 57, 59 53) Sand and calcite-tempered loess-rich fabric with abundant oolite and fossil fragments in the clay matrix.

152

APPENDIX 4 CHAFF OBSERVATIONS small grass seeds occurring commonly. The Jeitun pottery Type B is consistent in having impressions of small grass seeds. More work is required to establish whether the grass seeds are identifiable and subsequently of any use regarding seasonality of pottery production.

Jeitun period sites: impressions in pottery, Mike Charles & Jo Bending (June 2001) Overall there is a fair degree of consistency in the impressions. Fine chaff was present in most sherds with Key fine chaff –

fine chaff

very finely chopped up plant material, probably grass stems, and chaff from the grass ears. Its difficult to tell whether or not the grass is a cereal

gramineae

seeds of grasses (not cereals), some of which still have chaff surrounding them. There were several types of grass seeds which can be divided into long and round. The round type could be a type that we had in the charred assemblage and which flower late in the year, over the summer. This could be useful in determining season of manufacture, though of course the grain/chaff could be ‘saved’ from one season to another

grain

SITE

cereal grain

YEAR

CODE

IMPRESSIONS

CHAGYLLY

98

12

no plant impressions on surface - cross section - fine chaff

CHAGYLLY

98

15

fine chaff + small gramineae

18

fine chaff + gramineae of mixed sizes (may be idable)

CHAGYLLY CHAKMAKLI

98

4

fine chaff

CHAKMAKLI

98

6

fine chaff

CHAKMAKLI

98

7

fine chaff ? grain fragment

CHOPAN

98

SURFACE 1

fine chaff + gramineae

CHOPAN

98

SURFACE 7

Grampian (?x2) may be idable + small Gramineae

CHOPAN

98

JEITUN

93

ANOMALY J93.123

fine chaff + mixed Gramineae

JEITUN

93

TYPE A J93.17.6

shallow impressions of ?fine chaff +?idable Gramineae

JEITUN

94

TYPE A J94.127.2

fine chaff (below damaged surface)

JEITUN

94

TYPE A J94.155.4

small Gramineae (+ few larger Gramineae)

JEITUN

94

TYPE A J94.155.5

fine chaff (below damaged surface) ?Gramineae (shallow impressions)

JEITUN

98

TYPE A

fine chaff (below damaged surface)

fine chaff + Gramineae

JEITUN

TYPE A ????.2

fine chaff (below damaged surface)

JEITUN

TYPE A J9?.34.1

?fine chaff + shallow ?Gramineae impressions

JEITUN

93

TYPE B J93.04.3

small Gramineae + few larger Gramineae (may be idable) few shallow fine chaff impressions

JEITUN

93

TYPE B J93.4/14?

small Gramineae few shallow fine chaff impressions

153

APPENDIX 4 SITE

YEAR

CODE

IMPRESSIONS

JEITUN

93

TYPE B J93.97.1

small Gramineae and ?fine chaff - shallow impressions 3x Gramineae may be idable

JEITUN

94

TYPE B J94.137.2

small Gramineae + few larger Gramineae few shallow fine chaff impressions

JEITUN

94

TYPE B J94.175.2

small Gramineae + fine chaff

JEITUN

94

TYPE B J94.181.1

small Gramineae + one Gramineae few shallow fine chaff impressions

JEITUN

94

TYPE B J94.205.2

small Gramineae + ?one Gramineae/cereal few shallow fine chaff impressions

MONDJUKLI

98

12

fine chaff

MONDJUKLI

98

14

fine chaff + distorted Gramineae

PESSEDJIK

1967 25

?fine chaff

PESSEDJIK

1968 32

indet. impressions

PESSEDJIK

1968 37

?Gramineae-sized impression

TOGOLOK

1967 TYPE A 3.0.2.372

fine chaff

TOGOLOK

1967 TYPE A 20

fine chaff (below damaged surface)

TOGOLOK

98

TYPE A 10

fine chaff

TOGOLOK

1967 TYPE B KC 13

shallow impressions

TOGOLOK

1967 TYPE B KC 65

shallow impressions of ?fine chaff

TOGOLOK

1967 TYPE B KC 86

fine chaff

154

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY

APPENDIX 5 GENERAL DATABASE no. 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

site

field no J.93.6.1 J.93.1.1 J.93.4/11.1 J.93.17.1 J.93.4/11.4 J.93.1.2 J.93.6.6 J.93.4/11.5 J.93.17.3 J.93.18.1 J.93.17.7 J.93.4/11.8 J.93.4/11.9 J.93.6.2 J.93.17.6

consist. coarse coarse coarse coarse coarse coarse coarse medium medium medium medium medium medium medium medium

colour whiteware buffware buffware redware redware redware indet/unox buffware buffware buffware buffware buffware buffware buffware buffware

16 17 18 19 20

Jeitun Jeitun Jeitun Jeitun Jeitun

J.93.4/11.7 J.93.6.4 J.93.17.4 J.93.4/11.2 J.93.4/11.3

medium medium medium medium medium

redware redware redware redware redware

21 22 23 24 25 26 27 28

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

J.93.6.3 J.93.6.5 J.93.17.5 J.93.4/11.6 J.93.29.2 J.93.29.7 J.93.29.5 J.93.29.6

fine fine fine fine coarse coarse coarse medium

buffware redware buffware buffware buffware redware redware redware

29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

J.93.29.4 J.93.29.1 J.93.29.3 J.93.35.2 J.93.34.1 J.93.39.1 J.93.35.1 J.93.38.2 J.93.38.1 J.93.32.1 J.93.37.2 J.93.37.1 J.93.37.7 J.93.39.3 J.93.39.2 J.93.30.5 J.93.30.6 J.93.37.9

medium medium fine coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse

buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware redware redware 155

inclusions abundant organics abundant organics abundant organics common organics organics and clay pellets organics and fine sand organics and fine sand undifferentiated with organics undifferentiated with organics undifferentiated with organics undifferentiated with organics undifferentiated with organics undifferentiated with organics organics and grey sand organics, occasional grey sand and clay pellets abundant organics abundant organics abundant organics and grey sand abundant organics and grey sand organics and very abundant clay pellets organics organics organics very occasional and small organics common organics common organics common organics organics, occasional grey rock frags and sand common organics common organics common organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics and clay pellets moderate organics and clay pellets moderate organics moderate organics moderate organics moderate organics

status thin section binocular binocular thin section thin section binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular thin section binocular binocular binocular binocular binocular thin section binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular thin section thin section binocular binocular binocular binocular binocular

APPENDIX 5 no. 47 48 49 50 51 52 53

site

field no J.93.30.7 J.93.30.12 J.93.32.2 J.93.30.10 J.93.37.5 J.93.35.8 J.93.38.3

consist. medium medium medium medium medium medium medium

54 Jeitun

J.93.30.8

medium redware

55 Jeitun

J.93.37.10

medium redware

56 Jeitun

J.93.32.3

medium redware

57 Jeitun 58 Jeitun

J.93.38.7 J.93.30.18

medium redware medium redware

59 Jeitun

J.93.30.4

medium redware

60 61 62 63 64 65 66 67 68 69 70 71 72 73 74 75 76 77 78 79 80 81 82 83 84 85

J.93.30.21 J.93.30.14 J.93.38.5 J.93.35.5 J.93.37.6 J.93.37.4 J.93.30.20 J.93.30.19 J.93.35.3 J.93.35.6 J.93.30.15 J.93.37.3 J.93.30.11 J.93.30.2 J.93.32.4 J.93.30.1 J.93.30.3 J.93.30.13 J.93.30.16 J.93.30.17 J.93.38.4 J.93.30 J.93.35.7 J.93.38.6 J.93.35.4 J.93.37.8

medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium medium fine fine fine fine

86 Jeitun 87 Jeitun

J.93.40.1 J.93.40.2

medium buffware medium redware

88 Jeitun

J.93.40.3

medium redware

89 Jeitun 90 Jeitun 91 Jeitun

J.93.54.1 J.93.54.2 J.93.54.3

coarse coarse coarse

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

colour redware buffware buffware buffware buffware buffware redware

redware redware redware redware redware redware redware redware redware redware redware redware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware redware redware

buffware redware buffware

156

inclusions abundant organics abundant organics and clay pellets abundant organics and clay pellets abundant organics and clay pellets moderate organics moderate organics moderate organics and occasional clay pellets moderate organics and occasional clay pellets moderate organics and occasional clay pellets moderate organics and occasional clay pellets moderate organics moderate organics, grey sand and clay pellets moderate organics, grey sand and clay pellets moderate organics and grey sand moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics very occasional organics very occasional organics occasional organics occasional organics, grey sand and clay pellets moderate organics moderate organics and grey clay pellets moderate organics and grey clay pellets abundant organics and grey sand abundant organics and grey sand abundant organics and grey sand

status binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular thin section binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular thin section binocular binocular thin section binocular thin section binocular binocular binocular binocular binocular binocular

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY no. site 92 Jeitun 93 Jeitun

field no J.93.52 J.93.54.2

consist. colour coarse buffware medium redware

94 Jeitun

J.93.55.1

medium redware

95 Jeitun

J.93.51

medium buffware

96 Jeitun 97 Jeitun

J.93.97.3 J.93.97.1

medium redware medium redware

98 Jeitun 99 Jeitun

J.93.97.2 J.93.110

fine buffware medium redware

100 101 102 103 104

Jeitun Jeitun Jeitun Jeitun Jeitun

J.93.110 J.93.126.2 J.93.122.1 J.93.126.1 J.93.127

medium coarse coarse coarse coarse

105 Jeitun 106 Jeitun

J.93.124 J.93.123.2

coarse redware medium redware

107 Jeitun 108 Jeitun

J.93.123.4 J.93.123

medium buffware medium redware

109 110 111 112 113 114 115 116 117

J.93.126 J.93.124 J.93.122.2 J.93.124 J.93.123.3 J.93.123.1 J.93.126.3 J.93.130 gjv2&2

medium medium medium medium medium medium medium fine medium

118 Jeitun

gjv2&1

medium redware

119 120 121 122

Jeitun Jeitun Jeitun Jeitun

gjv2&3 gjv3&7 gjv3&8 gjv3&4

coarse coarse coarse medium

123 Jeitun

gjv3&5

medium redware

124 Jeitun 125 Jeitun 126 Jeitun

gjv3&11 gjv3&9 gjv4&12

fine redware medium redware coarse redware

127 Jeitun 128 Jeitun 129 Jeitun

gjv4&14 gjv4&20 ldjh gjv4&19

medium redware coarse buffware coarse redware

130 Jeitun 131 Jeitun

gjv4&15 gjv4&16

fine buffware medium redware

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

buffware buffware buffware buffware redware

buffware buffware redware buffware buffware buffware buffware buffware redware

redware buffware buffware redware

157

inclusions abundant organics and grey sand moderate organics and grey clay pellets moderate organics and grey clay pellets undifferentiated with moderate organics moderate organics moderate organics and grey clay pellets common organics moderate organics and grey clay pellets moderate organics and grey sand abundant organics abundant organics abundant organics moderate organics and very abundant clay pellets moderate organics occasional organics and grey clay pellets abundant organics and grey sand moderate organics, white clay pellets and grey sand moderate organics and grey sand occasional organics occasional organics and grey sand very occasional organics very occasional organics very occasional organics very occasional organics very occasional organics moderate organics and grey inclusions moderate organics and grey inclusions moderate organics abundant organics abundant organics moderate organics and occasional clay pellets

status thin section thin section

moderate organics and occasional clay pellets occasional organics moderate organics and grey sand abundant organics and very occasional clay pellets abundant organics abundant organics abundant organics and occasional clay pellets moderate organics moderate organics

binocular

binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular thin section binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular

binocular binocular binocular binocular binocular binocular binocular binocular

APPENDIX 5 no. 132 133 134 135 136 137 138 139 140 141 142 143 144 145 146 147 148 149 150 151 152 153 154 155 156 157 158 159 160 161

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

site

field no gjv4&11 gjv4&18 gjv5&20 gjv5&20 gjv6&25 gjv6&21 gjv6&27 gjv6&23 gjv6&26 ldjh gjv6&30 ldjh gjv6&29 ldjh gjv6&28 ldjh gjv6&31 gjv9&34 gjv9&33 gjv9&33 gjv10&35 gjv10&36 gjv10&35 gjv12&43 gjv12&39 gjv12&42 gjv12&41 gjv12&38 gjv12&46 gjv15&50 gjv15&50 gk,52 gjv18&58 gjv18&55

consist. coarse medium medium medium coarse coarse medium medium medium medium medium coarse medium coarse medium fine medium medium medium medium coarse medium fine medium medium medium medium medium coarse coarse

colour buffware redware buffware redware redware redware redware redware buffware buffware redware buffware redware redware buffware buffware redware redware buffware redware buffware buffware buffware buffware buffware redware buffware buffware buffware redware

162 163 164 165 166 167 168 169 170 171 172 173 174 175 176 177 178 179 180 181 182 183

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

gjv18&56 gjv18&56 gjv18&55 gjv18&51 gjv18&53 gjv18&59 gjv18&54 ldjh gjv15-18&60 ldjh gjv15-18&61 gjv20&63 ldjh gjv20&66 gjv23&66 gjv23&66 gjv23&65 gjv24&69 gjv24&68 gjv26&70 gjv28&71 gjv28&74 gjv28&72 gjv31&75 gjv31&75

medium medium coarse coarse medium coarse medium coarse medium coarse coarse medium medium medium medium medium medium medium medium coarse medium medium

redware redware redware buffware buffware redware buffware buffware redware buffware buffware buffware buffware redware redware buffware redware redware buffware redware buffware redware

158

inclusions abundant organics moderate organics moderate organics moderate organics abundant organics abundant organics abundant organics moderate organics abundant organics abundant organics moderate organics abundant organics moderate organics abundant organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics and grey sand abundant organics moderate organics occasional organics moderate organics moderate organics abundant organics moderate organics moderate organics abundant organics moderate organics and grey clay pellets moderate organics moderate organics moderate organics abundant organics moderate organics abundant organics moderate organics abundant organics moderate organics abundant organics abundant organics moderate organics abundant organics moderate organics moderate organics abundant organics moderate organics moderate organics moderate organics abundant organics abundant organics moderate organics

status binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular binocular optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY no.

site

field no

consist.

colour

184 Jeitun

ldjh gjv32&78

coarse

redware

185 Jeitun 186 Jeitun

ldjh gjv32&78 ldjh gjv32&80

coarse buffware medium redware

187 Jeitun 188 Jeitun 189 Jeitun

ldjh gjv32&80 ldjh gjv32-35&80 ldjh gjv32-35&80

medium redware medium redware medium redware

190 191 192 193

Jeitun Jeitun Jeitun Jeitun

vt; gjv32-66 gjv36&82 gjv41&86 gjv41&83

medium coarse coarse fine

redware buffware buffware buffware

194 195 196 197 198 199 200 201 202 203 204 205 206 207 208 209 210 211 212 213 214 215 216 217 218 219 220 221 222 223 224 225 226 227 228 229 230 231 232

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

gjv41&83 gjv41&88 gjv41&86 ldjh gjv41&92 ldjh gjv41&92 ldjh gjv41&87 ldjh gjv41&88 ldjh gjv41&87 ldjh gjv41&88 gjv50&94 gjv50&95 ldjh gjv50&98 ldjh gjv50&97 gjv56&99 gjv56&471 gjv56&100 gjv56&101 gjv56&103 gjv56&102 vt; gjv50-56&104 vt; gjv50-56&105 vt; gjv50-56&105 ldjh gjv61&106 ldjh gjv61&107 ldjh gjv61&107 gjv62&109 gjv62&108 gjv64&111 gjv64&110 gjv66&112 gjv66&115 gjv66&113 gjv69&121 gjv69&120 gjv69&118 gjv69&119 .&g&26&125 .&g&26&125 .&g&26&123

medium medium medium medium coarse medium fine medium medium coarse medium coarse medium coarse coarse medium medium medium medium coarse medium medium fine medium fine medium medium coarse medium coarse medium medium medium medium medium medium medium medium medium

buffware indet/unox indet/unox redware buffware buffware buffware buffware redware redware redware buffware buffware buffware redware buffware redware redware redware redware redware redware buffware redware redware buffware redware buffware redware buffware redware buffware redware redware redware buffware whiteware buffware redware 159

inclusions abundant organics and red and grey clay pellets abundant organics moderate organics and grey clay pellets abundant organics abundant organics moderate organics and common grey sand moderate organics abundant organics abundant organics very occasional organics and grey clay pellets moderate organics moderate organics moderate organics and grey sand abundant organics abundant organics moderate organics moderate organics moderate organics moderate organics abundant organics moderate organics moderate organics moderate organics abundant organics abundant organics abundant organics abundant organics moderate organics moderate organics and grey sand abundant organics abundant organics moderate organics moderate organics abundant organics occasional organics moderate organics moderate organics abundant organics moderate organics abundant organics moderate organics moderate organics moderate organics abundant organics moderate organics abundant organics moderate organics moderate organics moderate organics

status optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical

APPENDIX 5 no. site 233 Jeitun 234 Jeitun 235 Jeitun

field no .&g&26&124 .&g&26&125 .&g&26&125

consist. colour coarse redware medium redware medium redware

236 237 238 239 240 241 242 243 244 245 246

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

.&g&27&127 .&g&27&132 .&g&27&130 .&g&27&131 .&g&27&128 rd&,tb,7&136 rd&,tb,7&138 rd&,tb,7&136 rd&,tb,7&137 rd&,tb,7&139 u-7&3&145

coarse coarse medium medium medium coarse coarse medium medium medium fine

buffware redware redware buffware buffware redware redware redware buffware buffware redware

247 248 249 250 251 252 253 254 255 256 257 258 259 260 261 262 263 264 265 266 267 268 269 270 271 272 273 274 275 276 277 278 279 280 281 282 283 284

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

u-5&3&141 u-7&3&144 u-6&4&143 u-6&3&142 l-14&3&154 l-9&3&148 l-9&4&153 l-9&3 l-4&3 l-16&3 l-9&3&150 l-15&3&156 l-15&3&156 l-16&3&158 t-12&4&174 t-18&4&172 t-18&4 t-10&4&169 t-10&3&168 t-10&3&167 t-18&4&174 t-6&3&160 t-8&4&166 t-18&4&179 t-19&4&177 b-8&4 k-4&3&184 k-4&3&180 k-4&3&183 v-17&197 v-15&186 v-20&191 v-8&3&193 v-17&187 v-16&194 v-6&185 v-17&188 y-17&196

medium medium medium medium medium medium medium medium medium medium medium medium fine coarse coarse coarse medium medium medium coarse medium medium fine medium medium coarse coarse medium medium medium medium medium medium medium medium medium coarse coarse

redware buffware redware buffware buffware buffware redware buffware buffware buffware buffware redware buffware redware buffware redware redware redware buffware redware redware buffware redware orangeware buffware buffware whiteware redware buffware buffware buffware buffware buffware redware redware redware redware redware 160

inclusions abundant organics moderate organics moderate organics and grey clay pellets abundant organics abundant organics moderate organics moderate organics moderate organics abundant organics abundant organics moderate organics moderate organics moderate organics occasional organics, grey sand and clay pellets moderate organics and grey sand moderate organics moderate organics abundant organics abundant organics moderate organics and clay pellets moderate organics abundant organics abundant organics abundant organics abundant organics abundant organics occasional organics abundant organics abundant organics abundant organics moderate organics moderate organics moderate organics and clay pellets abundant organics moderate organics abundant organics moderate organics abundant organics moderate organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics abundant organics abundant organics

status optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY no. 285 286 287 288 289 290 291 292 293 294 295 296 297 298 299 300 301 302 303 304 305 306 307 308 309 310 311 312 313 314 315 316 317 318 319 320 321 322 323 324 325 326 327 328 329 330 331 332 333 334 335 336 337 338

site Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

field no y-16&195 v-17 v-18 v-6&4 v-19&190 k-4&3 k-4&3&181 v-11 g-15&207 g-6&4&210 g-9&215 g-8&212 g-9&213 g-15&208 l-5&4&198 g-6&4&210 g-8&112 g-10&216 g-8&4&206 g-9&5&214 l-17&204 g-10&216 g-8&4&206 g-8&4&206 l-6&4&200 l-16&204 g-9&214 g-16&209 l-18&205 l-16&203 g-10&217 g-10&216 l-16&202 l-6&4&200 l-6&4&199 g-7&211 n-9&266 c-10&220 t-8&3 b-8&4&228 c-11&222 c-11&222 c-11&221 t-8&218 o-21&229 n-9&227 n-6 n-9&225 n-8 c-10 i-1&4&231 i-1&4&230 i-1&4&235 i-1&4&231

consist. coarse coarse coarse medium medium coarse coarse medium medium medium coarse coarse fine medium medium medium medium medium medium medium medium medium coarse coarse medium coarse medium medium medium medium medium medium medium medium medium coarse medium medium coarse coarse medium coarse medium medium medium medium coarse coarse coarse coarse coarse coarse coarse coarse

colour redware redware redware redware redware redware redware redware redware redware redware redware buffware buffware whiteware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware redware buffware redware redware redware redware redware redware redware redware buffware buffware buffware buffware buffware buffware redware redware redware redware buffware buffware buffware buffware redware redware redware redware 161

inclusions abundant organics abundant organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics and grey sand moderate organics moderate organics abundant organics abundant organics occasional organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics abundant organics abundant organics abundant organics and blackening abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics moderate organics abundant organics abundant organics moderate organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics moderate organics abundant organics abundant organics abundant organics abundant organics

status optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical

APPENDIX 5 no. 339 340 341 342 343 344 345 346 347 348 349 350 351 352 353 354 355 356 357 358 359 360 361

site

field no i-1&4&231 i-1&6&239 i-1&6&238 i-1&6&239 i-1&7&245 i-1&6&239 i-1&4&231 i-1&8 i-1&4&232 i-1&8&292 i-1&5&239 i-1&6&238 i-1&4&235 i-2&4 i-1&7&244 i-1&5&234 i-1&8 i-1&8&289 i-1&7&242 i-1&8&282 i-1&4&233 i-1&7&246 i-1&6&240

consist. coarse medium medium medium medium medium medium medium medium medium coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse coarse medium fine

colour redware redware redware redware redware redware redware redware redware redware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware buffware

inclusions abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics occasional organics

status optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical

362 Jeitun

i-1&6&237

coarse

redware

optical

363 364 365 366 367 368 369 370 371 372 373 374 375 376 377 378 379 380 381 382 383 384 385 386 387 388 389 390

i-1&8&293 i-1&6&237 i-1&6&237 i-1&8&293 i-1&7&242 i-1&6&270 i-1&7&247 i-1&5&236 i-1&6&240 i-1&8&293 i-1&8&290 i-1&7&243 ;-12&250 b-15&254 [-9&256 u-13&250 [-14&258 p-16&212 i-20&264 [-20&258 t-18&250 x-20 [-12&257 i-16&262 n-23&255 n-23&255f n-23&255 n-23&255

coarse coarse coarse coarse coarse medium medium medium medium medium fine fine coarse medium medium medium medium medium medium medium fine coarse medium coarse coarse coarse medium medium

redware redware redware redware redware redware redware redware redware redware redware redware redware redware redware buffware buffware buffware buffware buffware buffware buffware whiteware whiteware redware redware redware redware

abundant organics and occasional grey clay pellets abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics and grey sand moderate organics and grey sand moderate organics and grey sand moderate organics and grey sand moderate organics and grey sand occasional organics occasional organics abundant organics abundant organics moderate organics moderate organics moderate organics moderate organics moderate organics abundant organics occasional organics abundant organics abundant organics abundant organics abundant organics abundant organics moderate organics moderate organics

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun

162

optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical optical thin section optical optical optical optical optical

SOUTHERN TURKMENISTAN IN THE NEOLITHIC: A PETROGRAPHIC CASE STUDY no. 391 392 393 394 395 396 397 398 399 400 401 402 403 404 405 406 407 408

site Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Jeitun Chopan

field no n-23&255f i-20&267 i-20&268 e-10&269 e-10&270 id&271 dj&273 dj&272 dj&273 gkf276 gkf274 gkf281 gkf277 gkf gkf gkf vt;61-32 xjgfy

consist. coarse coarse coarse coarse medium medium coarse medium medium medium medium medium medium medium coarse coarse medium medium

colour buffware buffware buffware redware redware orangeware redware redware buffware redware redware redware redware greyware buffware redware redware redware

409 410 411 412 413 414 415 416 417 418 419 420 421 422 423 424 425 426 427 428 429 430 431 432 433 434 435 436 437 438 439 440 441 442 443

Chopan Chopan Chopan Chopan Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik Pessedjik

xjgfy xjgfy xjgfy xjgfy ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh1 ldjhF ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 gjv4 ujh2 ldjh< ujh2 ldjh< ujh2 ldjh< ujh2 ldjh< ujh2 ldjh< ujh2 ldjh