Climate changes in the Holocene: impacts and human adaptation 9781351260244, 1351260243, 9780815365938

Climate change in the current geological era, from approximately 11,000 years ago to the present. Abstract: Climate chan

653 59 78MB

English Pages 406 [419] Year 2019

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Climate changes in the Holocene: impacts and human adaptation
 9781351260244, 1351260243, 9780815365938

Citation preview

Climate Changes in the Holocene Impacts and Human Adaptation

Climate Changes in the Holocene Impacts and Human Adaptation

Edited by

Eustathios Chiotis

CRC Press Taylor & Francis Group 6000 Broken Sound Parkway, NW, Suite 300 Boca Raton, FL 33487-2742 ©  2019 by Taylor & Francis Group, LLC CRC Press is an imprint of Taylor & Francis Group, an Informa business No claim to original U.S. Government works Printed on acid-free paper International Standard Book Number-13: 978-0-8153-6593-8 (Hardback) This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been made to publish reliable data and information, but the authors and publishers cannot assume responsibility for the validity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this form has not been obtained. If any copyright material has not been acknowledged, please write and let us know so we may rectify in any future reprint. Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or utilized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying, microfilming, and recording, or in any information storage or retrieval system, without written permission from the publishers. For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923, 978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged. Trademark Notice : Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe.

Library of Congress Cataloging-in-Publication Data Names: Chiotis, Eustathios.Title: Climate changes in the Holocene / [edited by] Eustathios Chiotis. Description: Boca Raton : CRC Press, Taylor & Francis Group, 2018. | Includes bibliographical references and index. Identifiers: LCCN 2018029042 | ISBN 9780815365938 (hardback : alk. paper) Subjects: LCSH: Paleoclimatology. | Climatic changes. Classification: LCC QC884.2.C5 C5745 2018 | DDC 551.609/01--dc23 LC record available at https://lccn.loc.gov/2018029042 Visit the Taylor & Francis Web site at  http://www.taylorandfrancis.com  and the CRC Press Web site at  http://www.crcpress.com 

Contents Preface..............................................................................................................................................vii Editor.................................................................................................................................................ix Contributors.......................................................................................................................................xi

Section I  Advances in Climate Reconstruction Chapter 1 Reconstructing the Environment as a Scenery of Human History and Civilization.........3 Eustathios Chiotis Chapter 2 Proxy Indicators of Climate in the Past...................................................................... 41 Marie-Michèle Ouellet-Bernier and Anne de Vernal Chapter 3 Pleistocene Glaciations................................................................................................ 77 Michel Crucifix Chapter 4 Solar Irradiance Variability and Earth’s Climate..................................................... 107 Natalie Krivova Chapter 5 High Resolution Climate Reconstruction of the Last 2,000 Years........................... 121 Sebastian Wagner and Eduardo Zorita

Section II  Tracing Major Human Migrations Chapter 6 Migration of Homo Sapiens Out of Africa............................................................... 143 P. Nick Kardulias Chapter 7 Ancient-DNA and Modern-DNA Genetics Can Reveal Past Population Movements................................................................................................................ 157 Konstantinos Voskarides

Section III  H  uman Responses to Climate throughout the Holocene Chapter 8 Climate Change, Mesoamerica, and the Classic Maya Collapse.............................. 165 Lisa J. Lucero and Jean T. Larmon v

vi

Contents

Chapter 9 From “Green” to “Brown”: The Archaeology of the Holocene Central Sahara....... 183 Savino di Lernia Chapter 10 Eastern Borders of the Sahara and the Relations with the Nile Valley and Beyond............................................................................................. 201 Barbara E. Barich Chapter 11 Human Adaptation in Arabia: The Role of Hydraulic Technologies........................ 221 Julien Charbonnier Chapter 12 Hydraulic Cultures and Hydrology under Climatic Change: North Arabian Mid-Holocene Pastoral and Proto-Oasis Land Use.................................................. 247 Hans Georg K. Gebel and Kai Wellbrock Chapter 13 Collapse of Bronze Age Civilizations....................................................................... 271 Guy D. Middleton Chapter 14 The Iranian Plateau and the Indus River Basin........................................................ 293 Cameron A. Petrie and Lloyd Weeks Chapter 15 Interaction of Climate, Environment and Humans in North and Central Asia during the Late Glacial and Holocene...................................................................... 327 Renato Sala

Section IV  Challenges Ahead Chapter 16 Perspectives of Climate Monitoring in the Satellite Era........................................... 363 Mika G. Tosca Chapter 17 Perspectives of Clean Energy and Carbon Dioxide Capture, Storage and Utilization.............................................................................................. 373 Nikolaos Koukouzas, Vasiliki Gemeni, and Nikolaos Tsoukalas Chapter 18 What Lies Ahead?: The Future of the Earth and Society as an Adaptive System.......... 387 Timothy Karpouzoglou and Feng Mao Chapter 19 Epimetron.................................................................................................................. 397 Michel Crucifix Index............................................................................................................................................... 401

Preface Although global warming is not overlooked as an anthropogenic climate change, this book is focused on natural climate changes and their environmental and societal implications in the last ten thousand years. The motivation for this collective work originated in the impressive scientific progress in climate sciences, in the evaluation techniques of archaeological material, and in genetics and modelling in the last decade. It is astonishing that detailed past environmental information is recorded in natural archives, which can now be retrieved and analyzed and this necessitates the closer synergy, the consilience, of natural sciences and humanities. The book is practically an application of the Unity of Knowledge, as highlighted by Wilson,* and the current transition of the focus of the natural sciences from the search for new fundamental laws toward new kinds of synthesis in order to understand complex systems based on coherent cause-and-effect explanations. The system of our interest is the climate, in the frame of the Earth system, over the short— but critical for our civilization— period of the current interglacial. The first section deals with the climate system which is fundamentally outlined in a broader than usually sense in the first introductory chapter, which is completed with a glossary of definitions. In particular, emphasis is given in the first chapter on the Cenozoic geological background and temperature trend, the global climate changes in the Holocene, the “global” monsoon and the carbon cycle. The second chapter refers to the most powerful tool of paleoclimatology, the proxy indicators and their quantitative application in environmental reconstruction, considering in tandem the necessary limitations. Chapter 3 on Pleistocene glaciations highlights the complexity of Earth system’s dynamics, the astronomical reasons underlying the alternation of glacial and interglacial conditions in the last three million years, the mechanisms of glaciations and deglaciations, and finally explains why the available data advance the hypothesis that we are now en route for an exceptionally long interglacial. Chapter 4 on solar irradiance variability investigates the potential mechanisms of solar influence on climate and the final chapter of the section on the climate system focuses on the factors of climate change over the last two millennia, the climate reconstruction methods, the main characteristics of the climate of the last 2,000 years and the implications on societal changes. This first section, and the whole book as well, should not be considered simply as a repository of functional knowledge; the scope extends further into the inter-relationships of diverse sources of evidence and the methods of their integration into a system approach, so that dogmatic application of knowledge is avoided, a case often encountered with abrupt climate events. The following second section includes a narrative of the dispersal of modern humans out of Africa and a case study for the application of DNA in tracing migrations in Eurasia in the last ten thousand years. It is demonstrated in this Section that migration was a mainstay of human adaptation from early in our evolutionary development and that our ancestors possessed an ability and willingness to venture far afield, very likely encouraged, if not forced, by altered environmental conditions. Astonishing conclusions can be drawn from DNA studies, as substantiated in Chapter 7 in a genetic study referring to Cyprus. The most extensive third section deals with the human responses to climate throughout the Holocene over a huge belt from Mesoamerica, Northern Africa, Arabia, Mediterranean, the Iranian Plateau, and South Asia, Central Asia and North Asia. A common denominator over this belt is the translocation of the Intertropical Convergence Zone (ITCZ) and the concomitant shift of monsoon rainfalls, ending up occasionally in desertification. The difficulties in the exploration of past human– climate interactions are typically highlighted in Chapter 14: “ Looking to the future of research, it appears that addressing these difficulties will *

Edward O. Wilson, 1998. Consilience: The Unity of Knowledge. New York: Vintage Books.

vii

viii

Preface

require not only an expansion of archaeological and paleoclimatic field and laboratory research, but also the development of new practical and theoretical approaches to the exploration of causality. Whereas previous research has focused on abrupt events and the identification of chronological correlations between instances of climate and cultural change, it is becoming increasingly important to make use of methods that allow modelling of the links between climate change, resource variation, and processes of human demographic and socio-economic change” . Finally, in the fourth section, the perspectives of technical and social adaptation are considered for the constraint of the anthropogenic environmental impact in the near future. The initial proposal for this book was submitted to CRC Press in January 2017 and the final one in April 2017, following a review stage by five experts, whose recommendations are thankfully acknowledged. The invited contributors were advised on the scope and the structure of the book and kindly and diligently responded in the difficult task of harmonizing their contributions in a co-ordinated reasoning. I gratefully acknowledge their contributions and the pleasure of collaboration with distinguished scholars and researchers. Dr. Marcia Glaze Wyatt’s support is thankfully acknowledged for her encouragement and recommendations. The top right picture on the cover, showing a hint of survival in the desertified savanna of the Tadrart Acacus Mountains, is thanks to the kind courtesy of Filippo Gallino, Savino di Lernia and the Archaeological Mission in the Sahara, Sapienza University of Rome. The fruitful cooperation and guidance of the CRC Press staff is also acknowledged, particularly of Joseph Clements, Joette Lynch and Lisa Wilford, as well as the creative contribution by Lara Silva McDonnell in copy editing. Joseph Clements in particular is credited with the overall coordination and the compilation of the cover. Eustathios D. Chiotis  Institute of Geology and Mineral Exploration of Greece

Editor Eustathios Chiotis  graduated from the National Technical University of Athens (NTUA) in Mining Engineering (1966), received a first MSc degree in Mineral Exploration at the Imperial College (IC), London, a second one in Petroleum Engineering also from IC, and his PhD from NTUA (1990). He served as Director of Geophysics at the Public Petroleum Corporation of Greece, Athens, and Director of Mineral Resources Evaluation at the Institute of Geology and Mineral Exploration of Greece, Athens. He has published papers on plate tectonics and the lithospheric structure in the Aegean Sea, oil exploration, archaeometry and ancient aqueducts in particular. He co-edited the Underground Aqueducts Handbook  (CRC Press, 2017) and investigated climate deterioration as a motive for the innovative exploitation of groundwater. His areas of interest include mineral exploration and mining, petroleum engineering and geophysics, geothermal energy and drilling, archaeometry and Ancient Greek technology, paleolithics and quaternary geology.

ix

Contributors Barbara E. Barich  ISMEO, International Association for Mediterranean and Oriental Studies Formerly Sapienza University of Rome Rome, Italy

P. Nick Kardulias  Program in Archaeology College of Wooster Wooster, Ohio

Julien Charbonnier  CEPAM, ‘ Cultures et Environnements : Pré histoire, Antiquité , Moyen Â ge’ University Cô te d’ Azur Provence-Alpes-Cô te d’ Azur, France

Timothy Karpouzoglou  Public Administration and Policy Group Wageningen University and Research Wageningen, The Netherlands

Michel Crucifix  Earth and Life Institute Université  catholique de Louvain Louvain-la-Neuve, Belgium

Nikolaos Koukouzas  Centre for Research and Technology Hellas Chemical Process and Energy Resources Institute Maroussi, Greece

Anne de Vernal  Department of Earth and Atmosphic Sciences Geotop research centre University of Québec in Montréal Montréal, Quebec, Canada

Natalie Krivova  Max Planck Institute for Solar System Research GÖttingen, Germany

Savino di Lernia  Dipartimento di Scienze dell’Antichità Sapienza University of Rome Rome, Italy

Jean T. Larmon  Department of Anthropology University of Illinois at Urbana– Champaign Urbana–Champaign, Illinois

Hans Georg K. Gebel  Institute of Near Eastern Archaeology and ex oriente Berlin Free University Berlin, Germany

Lisa J. Lucero  Department of Anthropology University of Illinois at Urbana– Champaign Urbana–Champaign, Illinois

Vasiliki Gemeni  Centre for Research and Technology Hellas Chemical Process and Energy Resources Institute Thessaloníki, Greece

Feng Mao  School of Geography, Earth and Environmental Sciences University of Birmingham Birmingham, United Kingdom

xi

xii

Guy D. Middleton  Czech Institute of Egyptology Charles University Prague, Czech Republic and School of History, Classics and Archaeology Newcastle University Newcastle upon Tyne, United Kingdom Marie-Michè le Ouellet-Bernier  Institute of Environmental Sciences University of Québec in Montréal Université  du Qué bec à  Montré al Ouellet Montreal, Quebec, Canada Cameron A. Petrie  Department of Archaeology University of Cambridge Cambridge, United Kingdom Renato Sala  Laboratory of Geoarchaeology Faculty of History, Archeology and Ethnology Al-Farabi Kazakh National University Almaty, Kazakhstan Mika G. Tosca  School of the Art Institute of Chicago Chicago, Illinois

Contributors

Nikolaos Tsoukalas  Centre for Research and Technology Hellas Chemical Process and Energy Resources Institute Maroussi, Greece Konstantinos Voskarides  Medical School University of Cyprus Nicosia, Cyprus Sebastian Wagner  Institute for Coastal Research Helmholtz-Zentrum Geesthacht Geesthacht, Germany Lloyd Weeks  Archaeology, School of Humanities, Arts and Social Sciences University of New England Armidale, Australia Kai Wellbrock  Department of Architecture, Civil Engineering and Urban Design, Laboratory for Urban Water Management University of Applied Sciences Lü beck, Germany Eduardo Zorita  Institute for Coastal Research Helmholtz-Zentrum Geesthacht Geesthacht, Germany

Section I Advances in Climate Reconstruction

1

Reconstructing the Environment as a Scenery of Human History and Civilization Eustathios Chiotis

CONTENTS 1.1 Introduction................................................................................................................................3 1.2 Earth’s Climate System..............................................................................................................5 1.3 Geological Background of Climate Fluctuations.......................................................................6 1.3.1 Cenozoic Temperature Trend.........................................................................................7 1.3.2 Global Climate Changes in the Holocene......................................................................9 1.3.3 Holocene Subdivisions................................................................................................ 10 1.4 Recent Developments in Environmental Studies..................................................................... 11 1.4.1 Environmental Reconstructions and Modelling.......................................................... 11 1.4.2 Paleoclimatic Studies of Monsoon Rainfall and ITCZ Translocation in the Holocene�����������������������������������������������������������������������������������������������������������12 1.4.2.1 The Asian Summer Monsoon (ASM)........................................................... 12 1.4.2.2 The East Asian Summer Monsoon (EASM)................................................. 14 1.4.2.3 The Interchange of Monsoon and Westerlies in Arid Central Asia.............. 14 1.4.2.4 The African Monsoon................................................................................... 16 1.4.2.5 ITCZ and Monsoon Translocation in the Peninsula of Arabia..................... 19 1.4.2.6 Concluding Remarks on the Global Monsoon Belt in the Holocene............ 19 1.4.3 Progress in Understanding the Terrestrial Carbon Cycle.............................................20 1.4.4 Interpretations of the Recent Global Warming Hiatus................................................. 22 1.5 Applications of Ancient DNA Studies in Humanities.............................................................. 23 1.6 Applications of Archaeobotany in Paleoclimatology...............................................................24 1.7 Potential of Human Adaptation to Environmental Changes in the Holocene..........................25 1.8 Concluding Remarks................................................................................................................26 Acknowledgements........................................................................................................................... 27 Glossary............................................................................................................................................ 27 References......................................................................................................................................... 35 The central idea of the consilience world view is that all tangible phenomena, from the birth of stars to the workings of social institutions, are based on material processes that are ultimately reducible, however long and tortuous the sequences, to the laws of physics. Edward O. Wilson (Consilience: The Unity of Knowledge, 1998)

1.1 INTRODUCTION Aristotle’s Meteorologica, written around 340 BC, is the oldest treatise dedicated to meteorology. A good number of weather forecasts in this work are derived from Egyptian works, and many other aspects are of Babylonian origin, mainly those on the nomenclature and classification of winds. 3

4

Climate Changes in the Holocene

In Babylon, cuneiform records on clay plaques indicate a very diverse and sophisticated intellectual society, with meteorological records being associated with astronomical events, which founded astrometeorology, a widespread practice in Europe. The system established by Aristotle remained for two millennia as the standard of scientific texts (Neves et al. 2017). Routine observations of surface air temperature, precipitation and surface pressure began in Western Europe during the late seventeenth and early eighteenth centuries, and gradually spread to most of the rest of the world by the 20th century. Many countries set up meteorological agencies after the Vienna Meteorological Congress of 1873. Even now, coverage is sparse in both polar regions, and recording did not start in parts of northern Canada and the Antarctic until the 1940s and late 1950s respectively (Jones and Bradley 1992). Fortunately, past climate assessments are possible thanks to natural features of the earth that preserve clues about past climate and environmental change, and can be dated. Common climate archives are tree-rings, marine or lacustrine sediments, fossils, geological organic matter, stalactites, and polar ice cores. Geochemical, physical, geological, and biological indicators measured in these natural features provide quantitative estimates about atmospheric and oceanic temperature and circulation, sea levels, ice-sheet and sea-ice distribution, precipitation, atmospheric chemistry, biological productivity, and other climate parameters over all timescales. For this reason, these indicators are called “proxies” for climate variables. The proxies, dating techniques and the quantitative estimates of climate parameters based on transfer functions or analogous techniques are treated in depth by Marie-Michèle OuelletBernier and Anne de Vernal in Chapter 2 of this book. The benefits of consilience of Natural Sciences and Humanities are best exemplified in the calibration of radiocarbon dating of the mid-second millennium BCE eruption of Thera (Santorini, Aegean Sea), as pointed out in the second chapter (2.2.4). The findings shed new light on the long-running debate focused on a discrepancy between radiocarbon (late 17th–early 16th century BCE) and archaeological (mid16th–early 15th century BCE) dating evidence for Thera. Results indicate a shift in the calibrated age range for Thera toward the archaeological evidence. Furthermore, the volcanic eruption offers a critically important marker horizon to synchronize archaeological chronologies of the Aegean, Egypt, and the Near East and to anchor paleoenvironmental records from ice cores, speleothems, and lake sediments (Pearson et al. 2018). Fortunately, the variations in the principal greenhouse gases in the past (CO2, CH4 and N2O) can be derived directly from high-resolution ice cores in Antarctica (Schmidt et al. 2011). As emphasized by Middleton in the introduction to Chapter 13 in this book, Archaeologists and historians, amongst others, have long been interested in climate and its potential effects on human societies. …An early example of dendroclimatology and its application to a historical problem was the decline of Rome. …Climate change was linked with agricultural productivity, and with the economic, political and biological spheres. …Climate would also affect the

prevalence of diseases, such as malaria. As eloquently posited by Wyatt (2014), weather is a cooling process. The Sun heats the Earth’s surface; weather removes the surplus heat. Weather, averaged over time, is climate. Climate is what happens between the delivery of energy and the exit of its excess, the net result of which modulates the Earth’s average surface temperature. In geoscience research, the Earth system comprises the non-living and living parts of the Earth, especially through interactions of the lithosphere, biosphere, and atmosphere, as well as the other parts of the system, such as the asthenosphere, the Earth’s core, and extraterrestrial influences like solar variations and meteorite impacts. Earth system research considers a system that spans scales from microscopic (micrometer scale) to megascopic (many thousands of km scale), and from milliseconds to millions of years. The evolution of the lithospheric part of the Earth system has been affected by changes in life on Earth, which have had major effects on sedimentary environments, and the atmosphere/hydrosphere. For example, formation of carbonate rocks, carbon dioxide and oxygen production, terrestrial vegetation development, and organic material production are all

Environment, Human History and Civilization

5

biological processes that have radically changed parts and processes in the lithosphere. Through subduction, these biological changes have affected the asthenosphere as well. Plate tectonics is the overall global process that has shaped our planet, including its climate and biosphere. All surface phenomena are ultimately based on plate tectonics and the interaction between parts of the Earth system. Without plate tectonics, there would be no recycling of water into the mantle, or expelled into the atmosphere. CO2 would not be recycled-released through volcanoes, possibly leading to global cooling and the development of a snowball Earth. Van Wyk de Vries et al. 2017

The importance of orogenesis and paleogeography in glaciations in the long term is analyzed by Michel Crucifix (Section 3.3). In climate science, the climate system is defined in a narrower sense as required in climate understanding, modelling and centennial prediction and consists of five major, interacting components: the atmosphere, the hydrosphere, the cryosphere, the lithosphere and the biosphere, considering the pedosphere as an essential part of the biosphere. All components or subsystems of the climate system are intimately linked or coupled with all other components. Volcanic, anthropogenic and extraterrestrial influences are considered as external forcings on climate, whereas slow geological processes like subduction, ocean spreading, weathering, erosion, sedimentation and orogenesis are not taken into account in climate predictions. Similarly, the earth’s heat flow is neglected, even along the mid-oceanic ridges. The carbon cycle in geological processes and the importance of weathering in the carbon cycle and consequently in climate are outlined in this book by Michel Crucifix in Section 3.1. In extreme cases of rapid uplift, however, tectonic movements can divert the riverbed, with severe implications on ancient cities and civilizations (Berger 2006). In particular, it has been postulated that the Indus Valley Tradition declined in part because of a change in climate and as a result of neotectonic movements that changed the course of rivers away from the towns and cities (Belcher and Belcher 2000); this view is not yet broadly accepted.

1.2 EARTH’S CLIMATE SYSTEM The Sun is the dominant energy source to the Earth, and thus changes in its radiative output are expected to affect Earth’s climate system, as documented by Natalie Krivova in Chapter 4. The incoming energy from the Sun fuels the dynamical, chemical, and biological processes of the Earth system. The overarching external climate driver throughout the Holocene is the changing geometry of Earth’s orbit around the Sun, along with changes in the direction of the polar axis, which together, over the last 10 ka, have caused decreasing summer insolation in the Northern Hemisphere. This orbital cycle affects the climate on millennial, and longer, time scales by, for example, driving latitudinal translocations of the polar front and the Intertropical Convergence Zone (ITCZ). The fraction of the incoming solar energy scattered by Earth back to space, about 29%, is referred to as the planetary albedo. Radiation from a planet’s atmosphere warms the planet’s surface to a temperature above what it would be without its atmosphere. Greenhouse gases in the atmosphere radiate energy in all directions and part of this radiation is directed towards the planet’s surface, warming it. Computer modelling shows that without carbon dioxide, the terrestrial greenhouse would collapse and within 50 years the average Earth temperature would drop to −21°C from 15°C today. Changes in one climate component may involve compensatory changes throughout the entire climate system. These changes may amplify the initial disturbance, or dampen it. Interactions that tend to amplify the disturbance are termed positive feedback mechanisms or processes; they operate in such a way that the system is increasingly destabilized. Interactions that tend to dampen the initial disturbance are termed negative feedback mechanisms or processes; they provide a stabilizing influence on the system, tending to preserve the dynamic equilibrium.

6

Climate Changes in the Holocene The concepts behind feedbacks as applied to climate change are derived from concepts in control theory that were first developed for electronics. ….In this sense, the relationship between the magnitude of the climate forcing and the magnitude of the climate change response defines the climate sensitivity. Furthermore, a process that changes the sensitivity of the climate response is called a feedback mechanism. A feedback is positive if the process increases the magnitude of the response, and negative if the feedback reduces the magnitude of the response. Curry and Webster 1999, 351

Climate variations happen on all time scales, as well as on all spatial scales, from the regional and continental to the global. As an example, the oceanic El Niño phenomenon is most pronounced in the tropical Pacific off the coast of Peru, but the associated Southern Oscillation in the atmosphere has far-reaching, nearly global implications. Natural climate variability includes, in general, variations that are either directly driven by a purely periodic external force, like the seasonal cycle of insolation, and/or variations due to the non-linear interplay of feedbacks within the climate system and/or variations associated with random fluctuations in physical or chemical factors. Natural climate variability refers to natural fluctuations in climate, in averages and extremes about the mean on all temporal and spatial scales. On the basis of instrumental records of land surface air temperature, sea surface temperature and sea level pressure, decadal to multidecadal natural variability can be established from quasi-global or quasi-interhemispheric manifestation. Large-scale oceanic–atmospheric oscillations affecting regional weather are typically described by indices of differences of a certain atmospheric or oceanic parameter, such as sea level pressure or sea surface temperature, in a region. These temporally correlated occurrences of meteorological parameters in far-removed regions are called teleconnections (Lehr et al. 2012). The most prominent of these is the El Niño–Southern Oscillation (ENSO), which affects climate worldwide. El Niño is a large-scale oceanic warming event in the tropical Pacific Ocean that occurs every few years. The Southern Oscillation is characterized by an interannual oscillation in tropical sea level pressure between the western and eastern Pacific, consisting of a weakening and strengthening of the easterly trade winds over the tropical Pacific. It is recognized that there is a close connection between El Niño and the Southern Oscillation and they are two different aspects of the same phenomenon (Wang et al. 2016). In our current understanding the climate system is a most delicate fabric of interwoven planetary components, such as the atmosphere, the oceans, the cryosphere, the soils, and the ecosystems, that interact through intricate physical, chemical, geological and biological processes, such as advection, upwelling, sedimentation, oxidization, photosynthesis, and evapotranspiration. Schellnhuber and Martin 2014

As underlined by Mika Tosca in Chapter 16, climate-monitoring satellites have substantially improved our understanding of Earth’s changing climate. Satellites have allowed for a deeper understanding of the dynamics of decreasing Arctic sea ice, changing patterns of landscape fire, interannual variability of the atmospheric carbon dioxide growth rate, the vertical structure of clouds, global temperature warming, and much more. The once-emerging field of climate science has been transformed over the last several decades by the wealth of data derived from the satellites and instruments that have orbited—and continue to orbit—the planet.

1.3 GEOLOGICAL BACKGROUND OF CLIMATE FLUCTUATIONS Earth is a live planet with a 4.5 billion years long past and an unpredictably long future, and the Holocene is just a moment in Earth’s history. To simplify our understanding of Earth’s age and importance in all aspects of the human past, it is noted that if we simulate the Earth’s age to 24 hours, then the Holocene would reduce to about 0.2 seconds. Taking into account Earth’s climate evolution in the geological past

Environment, Human History and Civilization

7

FIGURE 1.1  After Hansen et al. (2013): (a) Global deep ocean δ18O from Zachos et al. (2008) and (b) estimated deep ocean temperature by Hansen et al. (2013).

is, therefore, indispensable in understanding climate changes in the Holocene, as briefly indicated here and described by Michel Crucifix in Chapter 3 on the Pleistocene glaciations (cf. Section 3.1).

1.3.1 Cenozoic Temperature Trend The long-term Cenozoic temperature trends show warming up to about 50 Ma (millions of years before present) and subsequent long-term cooling, extending into the Holocene (Figure 1.1). Superimposed on the long-term trends are occasional global warming spikes, “hyperthermals”, most prominently the Palaeocene–Eocene Thermal Maximum at approximately 56 Ma and the Mid-Eocene Climatic Optimum at approximately 42 Ma, coincident with large temporary increases of atmospheric CO2. In addition to occasional hyperthermals, superimposed on the long-term trends are continual high-frequency temperature oscillations, which are apparent in Figure 1.1 after 34 Ma, when the Earth became cold enough for a large ice sheet to form on Antarctica, and are still more prominent during ice sheet growth in the Northern Hemisphere (Hansen et al. 2013). Deep ocean temperature for the Pliocene and Pleistocene is shown in Figure 1.2. Across the Pliocene–Pleistocene transition, dramatic glacial advance occurred in the Pliocene Arctic, with the intensification of Northern Hemisphere glaciation, after which northern landmasses were permanently altered by the growth of large ice sheets (Keisling et al. 2016). During the past 800,000 years for which precise data are available from ice cores, atmospheric CO2, CH4 and N2O have varied almost synchronously with global temperature. It is generally admitted that the greenhouse gases provide an amplifying feedback that magnifies the climate change instigated by orbit perturbations (Hansen et al. 2013). This view extends the theory formulated by Milankovitch, which links changes in summer insolation at the high latitudes of the Northern Hemisphere to the glacial–interglacial oscillations (cf. Crucifix, Section 3.2). The slowly varying parameters of the Earth’s orbit modulate the solar radiation received at the top of the atmosphere and its distribution over latitudes. The periodicity of stable isotopes in

8

Climate Changes in the Holocene

FIGURE 1.2  After Hansen et al. (2013): Deep ocean temperature in (a) the Pliocene and Pleistocene and (b) the last 800,000 years.

sedimentary sequences from deep-sea-sediment cores have been successfully correlated with these astronomically forced cycles. A slightly different model, also compatible with these correlations, is that ice age cycles emerge from a self-maintained process, controlled by the orbital forcing. Continuing research is thus focused on the interconnections between insolation, ice sheets, greenhouse gas forcing, and climate, particularly for the last 800,000 years characterized by 100-kyr glacial cycles (Figure 1.2). As substantiated by Tzedakis et al. (2017), over the past one million years, fewer of the insolation peaks resulted in deglaciation—that is, more insolation peaks were “skipped”, implying that an energy threshold is required for the onset of deglaciation (cf. Crucifix, Section 3.6.2). Considering now a wider time frame, Vahlenkamp et al. (2018) analyzed stable oxygen and carbon isotopes from contourite sediments in the western North Atlantic. The Middle Eocene sediments from the Newfoundland Ridge reveal a unique archive of paleoceanographic change from the progressively cooling climate of the middle Eocene. The results reflect variations of bottom water temperature and deep current velocity on orbital timescales, dominated by obliquity. Combined with global circulation, model experiments show that enhanced overturning is associated with a strong cooling of surface waters in the Greenland–Norwegian Sea during obliquity minima, while the current formation is weaker with relatively warm temperatures in the Greenland–Norwegian Sea during obliquity maxima. The importance of the oceanic circulation in buffering Earth’s climate is presently demonstrated by the Atlantic meridional overturning circulation. This system of currents is an important component of the Earth’s climate system, regulating the distribution of heat as well as the cycling of carbon and other nutrients. It is characterized by the northward-flowing Gulf Stream of warm, salty water masses in the upper layers of the Atlantic Ocean, deep water formation in the North Atlantic and southward-flowing colder water masses in the intermediate to deep Atlantic Ocean. Owing to the large heat capacity of water, the northward-flowing near-surface waters contribute effectively to the heat transport from the Tropics to the mid- and high latitudes, and therefore affect the regional climate around the North Atlantic. Throughout the Pleistocene, circulation and convection in the North Atlantic have oscillated between stronger and weaker regimes, corresponding to generally warm and cold conditions. These changes occurred at the scale of glacial–interglacial cycles, but also within the glacial phases of these cycles.

Environment, Human History and Civilization

9

1.3.2 Global Climate Changes in the Holocene Although the Holocene climate is considered to be relatively stable, significant climate changes have been recorded in the natural archives, disproportional in frequency and magnitude to the short duration of this epoch. Among these changes, the Holocene Thermal Maximum and rapid deglaciation, decreasing insolation since five millennia ago and, of course, the monsoon variability associated with aridification in Africa and Asia, the Medieval Climate Anomaly and the Little Ice Age (cf. Chapter 5), are most noteworthy. Climate models’ ability to reproduce the observed paleoclimatic records can help understand the mechanisms behind the climate variability in the Holocene and provide context for current and future climate change. Alternatively, discrepancies between models and observations can help identify gaps in our understanding, or possible inaccuracies in the observational record, as demonstrated by S. Wagner and E. Zorita in Chapter 5, focusing on the Common Era. Marcott et al. (2013) reconstructed regional and global temperature anomalies for the past 11,300 years from 73 globally distributed records. They stacked the proxy records over the Extratropical Northern Hemisphere (30°N–90°N) and the low latitudes band (30°S–30°N), to investigate the general patterns of climate evolution and found that the high-latitude cooling trend is opposite to a warming trend in low latitudes during the last 11 ka. They conclude that Early Holocene warmth (ca. 10 to 5 ka ago) is followed by ca. 0.7°C cooling through the Middle to late Holocene (90%). For about a thousand years, food security is guaranteed by this animal and by the collection of tubers and other plants. Though pollen analysis has shown that wild cereals were present in the area (Mercuri 2008), the very limited presence of grinding equipment suggests that they were infrequently consumed or rarely processed. The territorial organization is based on residential camps, larger and used for longer periods, and smaller, highly specialized sites for specific purposes such as hunting, sourcing raw materials and stone processing. We have no dating older than 11.2 ka anywhere in the central Sahara: the entire population of this immense region thus coincides with the northwards ascent of the summer monsoon, the penetration of the Mediterranean winter rains and, simultaneously, the expansion of the human groups that had survived the last phases of the Pleistocene in refuges such as the Nile Valley and the highlands near the Mediterranean coast. Due to the scarcity and, in some cases, the total absence of human skeletal remains, we cannot evaluate their physical anthropological features. Based on the genetics of modern populations, the beginning of the Holocene appears to be also characterized by population drifts of European origin (Pereira et al. 2010). Regrettably, the persisting scarcity of intensive regional projects (partially mitigated only in recent times) makes it difficult to construct a chronological and cultural grid connecting different areas of the Sahara, which certainly saw the circulation of cultural features (as indicated by the strong similarities in the lithic toolkit). At present, alongside the central Sahara we have a reasonable knowledge of the oases west of the Nile and the north-western Maghreb: these are thousands of

From “Green” to “Brown”

189

kilometres away, with specific features and characteristics of their own (e.g., Linstädter et al. 2012; Barich et al. 2014; Lubell 2014; Mulazzani et al. 2016). Even less well known, and crucial to fully understanding the overall population dynamics, are the sites of the Southern Sahara (Breunig 2013; Haaland and Haaland 2013).

9.4.2 Diversification and Increased Sedentism: Late Acacus Foragers (10.2–8.0 ka) In the Acacus and the surrounding region, we record important environmental changes around 10.3 ka: in particular, a marked spread of grasslands, rich in wild cereals. This expansion is due to the strengthening and definitive stabilization of the monsoon with its summer rains. Plants such as sorghum and other cereals are systematically selected, stored and cultivated, as recorded at Takarkori (Mercuri et al. 2018). Similar behaviours have also been suggested at Nabta Playa in the Western Desert in Egypt (Wasylikowa 2001). The emphasis on a broader, delayed-return system of resource exploitation is also observed for animal species: no longer the selective hunting of Barbary sheep, as in the previous Early Acacus phase, but a broader range of prey, including gazelles, antelopes and even fish. Specifically, specimens of Ammotragus lervia are corralled and fed with fodder purposely collected by foragers at several sites, such as Uan Afuda and Takarkori (di Lernia 2001). All these changes reflect opportunistic choices, in turn mirroring a new way of using the landscape, the environment and resources. For this reason, these “complex” foragers are seen as a distinct cultural entity, locally termed Late Acacus (Figure 9.4). These groups are more sedentary, as directly and indirectly indicated by the size and organization of residential sites, such as Ti-n-Torha and

FIGURE 9.4  Late Acacus features. (a) alcareous tufa at Ti-n-Lalan, Tadrart Acacus; (b) Round Heads paintings from Sughd, Tadrart Acacus; (c) Dotted Wavy Lines and other decorated potsherds, Site A1150, Edeyen of Murzuq; (d) view of the Uan Afuda Cave, Tadrart Acacus, where the Late Acacus culture was first identified; (e) 3D rendering of a stone fence from Takarkori, southern Tadrart Acacus.

190

Climate Changes in the Holocene

Takarkori (Barich 1987; Biagetti and di Lernia 2013), the significant number of large grinding stones and the introduction of pottery (Barich 1987; di Lernia 1996; Garcea 2004). The presence of artefacts with traces of colour (grinding stones, bones, lithics, wooden spatulas, etc.) suggests a possible link with the rock art of the Round Heads style (di Lernia et al. 2016). As concerns pottery, based on the global radiometric information now available, it is evident that African ceramic production is original and independent from the other nuclear area – that is, East Asia (Jordan et al. 2016). The oldest pottery is produced at the Pleistocene-Holocene transition (if  not earlier) by the hunter-gatherers of present-day Mali. The excavations at Ounjougou have yielded potsherds dated between 11.1–10.7 ka, including some certainly older than 11.3 ka (Huysecom et al. 2009). Although geographically isolated, the Ounjougou material reveals important innovations, not just the potsherds but also the presence of small bifacial tools, material evidence of a new subsistence strategy aimed more at the exploitation of a warm and humid environment characterized by a great abundance of wild cereals (for this reason Huysecom now labels this culture as Early Neolithic: 2018). Apart from Ounjougou, the dating of the earliest pottery in the Sahara indicates a south-north chronological cline: the oldest sites date to 11.3–10.5 ka and are located in the southern parts of the central Sahara (Tagalal, Adrar Bous 10, Temet, etc.: e.g., Jesse 2010 for bibliographical references), whilst findings beyond latitude 23°N date to around 10.1–9.8 ka. It is very tempting to associate this chronological difference with the ascent of the monsoon and the consequent expansion of grasslands (di Lernia 1996). The Eastern Sahara yields similar dates, raising the question (currently still unresolved) of the relations between the two regions of the Sahara: is one pottery tradition more ancient than the other, conditioning its development? Or, alternatively, are both linked to older phenomena of which Ounjougou in Mali is just one expression? In other words, this is a fairly complicated picture: however, taking into consideration the large geographical area under examination, it is worth making some specific points. First, most stratigraphic contexts with early pottery are problematic, because they are surface contexts or were excavated decades ago, or both (particularly Temet and Tamaya Mellet). Studies of raw materials procurement and manufacturing techniques generally show that these were local productions, with a geographical range rarely exceeding 70 km (Echallier and Roset 1986; Livingstone Smith 2001; Eramo et al. 2014). Formal aspects are poorly defined: the state of preservation does not always allow for the reconstruction of the shape and, as far as we can see, globular/spherical forms prevail. Functional studies are also rare. The state of preservation has conditioned the study of decorations, both technical and syntactical (e.g., Jesse 2010). As concerns technical features, two major distinct traditions are commonly identified. Along the Nile Valley, and more generally in the Eastern Sahara, the predominant technique is incision, while the impressed technique is present in the Central Sahara where incised decorations are practically absent. From a syntactical point of view, the decoration is normally present on large areas of the pot and is made using simple tools such as combs, sticks and other tools. The “wavy lines” are narrow and impressed in the Central Sahara but much broader and incised in the Eastern Sahara and Nile Valley. There are significant differences in quantity, ranging from one or two fragments (in Egypt or Mali) to many thousands of pieces in the sites of the central Sahara and Sudan. The mobility of Early Holocene foragers and their exogamic social structure certainly favoured the circulation and exchange of knowledge and traditions linked to pottery production.

9.5 CLIMATE VARIATIONS AND CULTURAL TRAJECTORIES OF SAHARAN HERDERS 9.5.1 Few and Poorly Preserved: The First Herders of the Early Pastoral (8.3–7.2 ka) The African Humid Period is interrupted by a short and acute arid spell dated to 8.2 ka. This is a “central” date whose lower and upper limits vary locally, as do the effects on landscape and

From “Green” to “Brown”

191

environmental resources (Cremaschi et al. 2014). In the mountainous areas of the Central Sahara, this short dry interval is less intense than that recorded elsewhere in North Africa and in the periMediterranean basin (Kobashi et al. 2007; Berger and Guilaine 2009), but it is nevertheless of particular environmental significance. From this period onwards, the monsoon rains no longer seem able to saturate the rock strata, and the formation of tufaceous crusts in the Acacus region is interrupted. In the stratigraphic sequences of rock shelters, especially in the Tassili and Acacus, erosive phenomena are recorded that, though their extent is not fully defined, shape the landscape of these regions (Cremaschi and Zerboni 2011). In this environmental framework, the introduction of domestic livestock is a crucial turning point. Unfortunately, the archaeological documentation of North African sites with remains of domestic fauna is still very incomplete (di Lernia 2013). However, the distribution of radiocarbon datings of sites with remains of domestic fauna in North Africa shows that the north-eastern quadrant has the oldest dates (Figure 9.5), thus confirming a potential introduction from south-western Asia. In fact, the debate over a possible African

FIGURE 9.5  Map of principal North African sites with earliest presence of cattle (dot) and sheep/goat (triangle) (after di Lernia 2013, modified). Radiocarbon dates have been calibrated (OxCal 4.2) and age of each site is approximated in kilo annum (ka) and grouped in millennia. Sites are numbered according to country and by latitude. Key: 1, Merimde Beni Salama (7.2–6.7); 2, Fayyum (6.9–6.4); 3, El Awag (8.2–7.1); 4, Abu Gedar 2 (7.5–6.8); 5, Djara 90/1-Cl. 7 (7.8–7,8); 6, Farafra (8,1–7,9); 7, Tree Shelter (7,7–7.5); 8, Sodmein Cave (8,1–7,9); 9, Dakhla (7,9–7.7); 10, Kharga/E-76–7 (8.9–8,5); 11, Glass Area 81/61 (7.8–6.8); 12, Eastpans 95/2–1 (6.8–6.8); 13, Wadi Bakht (7.9–7.6); 14, Bir Kiseiba E-79–8 (? 10.4–9.7); 15, Nabta Playa E75–3 (? 10.1–9,6); 16, Nabta Playa E 75–8 (8.1–7.9); 17, Burg et Tuyur 85/73–2 (6.9–6,9); 18, Wadi Howar (6.2–5.7); 19, El Kadada (6.7–6.2); 20, Shaqadud (8.4–8.1); 21, Shaheinab (7.5–7.2); 22, Kadero (6.5–6.2); 23, Um Direiwa (7–6.6); 24, Kashm el Girba (5.7–5.6); 25, Haua Fteah (8–7.6); 26, Abu Tamsa (8.1–8); 28, Ti-n-Torha (8–7.7); 27, Murzuq 2 (6.1–5.9); 29, Uan Muhuggiag (8.7–8); 30, Takarkori (8.4–8.1); 31, Enneri Bardagué (8.6–7.9); 32, Gabrong (7.2–6.7); 33, Délébo (8.6–7.4); 34, Gueldaman-Akbou (5.6–4.8); 35, Grotte Capeletti (7.8–6.8); 36, Meniet (6.5–5.8); 37, Relidijem (6.4–5.9); 38, Ti-n-Hanakaten (8.3–7.7); 39, Adrar Tiouyine (6.3–5.7); 40, Adrar Bous (7.7–6.5); 41, Arlit (6.4–5.9); 42, Tamaya Mellet (6.1–5.9); 43, Kaf Taht el Ghar (7.4–7.1); 44, Ifri Oudadane (7.4–7.1); 45, Kehf-el-Baroud (6.5–4.8); 46, Karkarichinkat south (4.8–4); 47, Kobadi (3.8–3.3); 48, Windé Koroji W I (4.2–3.7); 49, Chami (4.5–4.4); 50, Khatt Lemaiteg (4–3); 51, Dhar Tichitt (4.8–3.5).

192

Climate Changes in the Holocene

domestication of cattle has now been settled thanks to the review of the archaeological, archaeological and genetic data (e.g., Brass 2017). Different datasets paint a homogeneous picture, with the first herders with Bos taurus arriving in NE Africa around 8.3 ka. These first small groups of herders enter regions densely inhabited by bands of hunter-gatherers, with a highly complex organization involving the sophisticated exploitation of environmental resources. It is therefore unsurprising that this penetration is intermittent: we should imagine encounters and negotiations between profoundly different human groups with radically different characteristics, especially in ideological terms (e.g., Smith 2005). As a result, the sites of early herders are patchily distributed, sometimes at great distances from each other. Hunter-gatherer clans sometimes resist for a considerable time, probably also leading to encapsulation processes, but maintaining their habits and organization, coexisting with herders in many regions of North Africa at least up to 6 ka (Barham and Mitchell 2008). As suggested years ago by Fekri Hassan (2002), the first herders probably moved quickly (“leapfrogging”), a particularly effective strategy for expanding into unknown places. Their rapidity of expansion through much of Northeast Africa may also be linked to the climatic and environmental characteristics mentioned above: the 8.2 ka “event” likely degraded environmental conditions in SW Asia and NE Africa and may have accelerated these dispersals. Interestingly, the pace of expansion of pastoral groups differs on a regional basis: it is faster in the north-eastern quadrant but slows abruptly west of the central Saharan massifs, true environmental refuges during the Holocene (Figure 9.5). To explain why the expansion of these pastoralists slows from east to west, we must consider elements other than “mere” ecological aspects. These include cultural factors, such as the dynamics of interaction between foragers and herders. It is no coincidence, therefore, that food production is “quickly” adopted where planned and delayed exploitation strategies and the sophisticated management of animal and plant resources already exist (di Lernia 2017c). In our study area, too, the earliest pastoral sites are poorly preserved. Only a handful of contexts provide some information, together with around forty radiocarbon dates that bracket the Early Pastoral between 8.3 and 7.2 ka. In the Acacus Mts. these include Takarkori, Uan Muhuggiag, Uan Tabu, Ti-n-Torha North, Imenennaden and Uan Telokat, alongside Mathendush cave in the Messak (Biagetti and di Lernia 2013). Only a few contexts have been identified in the dune fields of Murzuq and Uan Kasa, again due to the aforementioned problem of preservation. Early Pastoral groups culturally managed the marked alternation of wet and dry seasons through a peculiar form of settlement organization, with scattered settlements in the lowlands and longerlasting mountain camps (di Lernia 2002). Their belonging to a specific “place” is apparent from the funerary practices of these groups. At some sites (such as Imenennaden and Takarkori), burials of women, juveniles and children are kept inside, “under the floor”. Isotopic data show that these tombs represent the family burials of people who were born and lived locally (di Lernia and Tafuri 2013). Among the most radical changes in the material culture, it is worth noting the introduction of foliated pieces (especially arrowheads) and a transformation of ceramic production, with an increase of simple impressed decorations (chevrons, etc.).

9.5.2 Milking the Green Sahara: Cattle Herders of the Middle Pastoral (7.1–5.6 ka) The climate and environmental optimum falls in the middle of the Early Pastoral but is abruptly interrupted by a new short arid phase dated about 7.4–7.1 ka. This is signalled by geomorphological data (collapse of rock shelter vaults; eroded levels in stratigraphic sequences), archaeobotanical information (growth of xerophilous plants) and archaeological findings, in particular the analysis of the radiometric datings that suggest a significant reduction in human occupation (di Lernia 2002). Starting from 7.1 ka, we see a new increase in human occupation, corresponding to the archaeological phase that we term the Middle Pastoral, characterized by important innovations. Once again, palynological studies (Mercuri 2008) and indications of lake fluctuations (Cremaschi and Zerboni 2011) allow us to reconstruct the environmental dynamics of a landscape that is now highly variable, oscillating from grasslands with modest tree cover and small lakes during the wet

From “Green” to “Brown”

193

and rainy season (our summer) to a more steppe-like vegetation during the dry season (our winter). During the Middle Pastoral phase, herding reaches full maturity, exploiting the secondary products of cattle. Milking scenes are depicted in Saharan rock art, the outstanding evidence for this society (Figure  9.6). Cow milk residues have been identified in potsherds from Takarkori (Dunne et al. 2012): dated at least from 7.1 ka, the oldest evidence for milk processing anywhere in the African continent. The settlement organization exploits available resources on a seasonal basis: herders take their flocks to different pastures, sometimes distant or at a different altitude (vertical transhumance: di Lernia 2002). Fodder, analysed by studying the C and N isotopic signatures of animal bones, consists of C3 (dry environment) and C4 plants (indicators of wetter environments). These results, combined with the strontium signatures of the same animals, indicate the seasonal transhumance of herds from different places (di Lernia et al. 2013). Funerary practices illustrate the complex composition of these Neolithic groups. Over several decades of research, our excavations have brought to light a multi-faceted world (di Lernia and Manzi 1998; di Lernia and Tafuri 2013). Burials are found both in open-air contexts and immediately outside villages, as well as in rock shelters. In the latter case, the practice of burying only women and children in shelters – a specifically Pastoral tradition – survives, persisting until the end of the Middle Pastoral (ca. 5.6 ka). Funerary practices and palaeoanthropological data, although

FIGURE 9.6  Middle Pastoral features. (a) Excavation of a pit with ceramic fragments, restored; (b) from Site A492, Edeyen of Murzuq; (c) engraved wall from Wadi Tiksatine, Messak, and detail of milking scene (d); (e) oblique view of excavated corbeilles from Wadi Bedis, Messak; (f) a ~7000-year-old burial of an adult female from Takarkori, southern Acacus.

194

Climate Changes in the Holocene

scanty given the often poor preservation of human remains, reveal a significant mingling of human groups, a potential indicator of the integration of groups from different regions (Di Vincenzo et al. 2015). This incredible cultural complexity is clearly visible in the ceramic repertoire, more innovative and diversified than the previous production. The Middle Pastoral environment evolved over the centuries, marked by a progressive deterioration of the climate. Indications of greater environmental instability include the presence of stratigraphic contexts characterized by layers of dung. Around 6.4 ka and until the end of the Middle Pastoral (about 5.6 ka), many shelters in the Acacus Mts. present these stratifications. Dung is therefore a palaeoclimatic indicator of progressively more arid conditions, as are variations in the depth levels of lake basins. These are characterized by a further increase in seasonal fluctuations, with a complete, albeit temporary, desiccation during the dry season.

9.6 THE REVENGE OF THE DESERT: OASES, NOMADISM AND SOCIAL STRATIFICATION 9.6.1 The Onset of the Desert and Ethnic Fragmentation: Late Pastoral Herders (5.9–3.4 ka) Around 5.9 ka, the desert definitively reoccupies its lost spaces. The southwards migration of the Intertropical Convergence Zone (ITCZ) leads to the retreat of the African monsoon, and the Sahara basically takes on the forms we know today. The establishment of fully desert conditions obviously varies in its timing and features from region to region (Brooks et al. 2005; Kuper and Kröpelin 2006; Cremaschi and Zerboni 2011; Clarke et al. 2016). The disappearance of the African monsoon primarily affects the eastern Sahara. Water resources are drastically reduced in the oases of Farafra, Dakhla, Kharga – to name but a few – (Bubenzer and Riemer 2007; Barich et al. 2014; McDonald 2016). The mobility strategies of human groups change radically in favour of a nomadic frequentation with small settlements and “light” features. At Farafra, for instance, the presence of Neolithic herders is limited to the edges of the seasonal lakes, whereas in previous phases the same area benefitted from a bimodal rainfall regime that allowed for more stable forms of settlement (Barich et  al. 2014). The geographical, geomorphological and physiographical features of specific regions also affect how and when the African Humid Period ends, either abruptly or more gradually. Humans doubtless play a major role in shaping their own fate, accelerating the process of desertification especially by overexploiting resources and overgrazing, as suggested by recent research (Wright 2017). In our study area, though mitigated by the presence of the mountains, the establishment of a dry climate and the construction of what is basically a desert landscape, with scarce plant cover limited to a few tamarisks and rare acacia trees, also has enormous consequences for pastoral communities. The Late Pastoral (and to an even greater extent the Final Pastoral) phase differs significantly from that which preceded it. Cattle herding (the fulcrum of Middle Pastoral society) disappears almost entirely, unable to resist in an arid landscape with scarce permanent water resources. An ethnic fragmentation process that continues in the following centuries and that comes to characterize the human geography of the Final Pastoral (di Lernia and Merighi 2006) is now set in motion. People adopt diverse social and cultural strategies for coping with the increasing aridity. We witness the development of specialized nomadic pastoralism based on year-round mobility and an almost exclusive exploitation of small livestock such as sheep/goats, more resistant to the changed environmental conditions. In this way, ever-moving herders in search of pastures and water have less impact on resources, progressively creating a cultural landscape based on regional contacts and with a social exogamic structure (Tafuri et al. 2006). In the mountains, rock shelters are now used as genuine stables, as indicated by the hardened dung layers decimetres thick (Figure 9.7). On the other hand, human communities concentrate around the most abundant and reliable water resources: it is the

From “Green” to “Brown”

195

FIGURE 9.7  Late Pastoral features. (a) Site TH125, Acacus, a small cave used as stable; (b) a ~5000 yearold deposition of an adult male from In Aghelachem, Wadi Tanezzuft; (c) a typical Late Pastoral container, In Habeter III, Messak; (d) Painting of herders with cows from Ti-n-Lalan, Acacus.

formation of the desert that leads to the birth of the oases. In particular, the Wadi Tanezzuft valley, west of the Acacus escarpment and bordering the Algerian Tassili, fed by rains falling on the mountains to its south, guarantees a water supply and fertile soils throughout the Late and Final Pastoral (Cremaschi and di Lernia 2001). In this region, pastoral groups begin a process of greater sedentarization, in part supported by farming practices, with a high settlement density. The availability of resources is physically limited – the river valley rarely exceeds 10 km in width – yet the density of villages seems to increase steadily, resulting in a serious imbalance in the overall system. Access to pasture, water and other resources becomes an element of conflict: this imbalance is likely caused by an increasing stratification that produces the first forms of social hierarchy, tangibly expressed in funerary practices (di Lernia et al. 2002).

9.6.2 Funerary Practices and Social Stratification: From Late to Final Pastoral (3.7–2.7 ka) From the beginning of the Late Pastoral, the oases of the Tanezzuft Valley, the combined product of environmental changes and human actions, host the first great funerary monuments – mostly conical mounds, but also platforms, crescent-like and key-hole monuments and so forth – a formidable archive of the social and cultural dynamics of the last pastoral groups. The territorial study

196

Climate Changes in the Holocene

and stratigraphic excavation of these monuments has made it possible to reconstruct the biological and cultural history of the last Neolithic herders (di Lernia and Manzi 2002; di Lernia and Tafuri 2013; Mori et al. 2013). For the first time in the Central Sahara, we see a clear separation between places for the living (settlements) and places for the dead (necropolises). In the early stages of the Late Pastoral, the conical tumuli are large (some over 15 metres in diameter), in an isolated position dominating large portions of the landscape: inside, often protected by a lithic cist, are the individual burials of adult males, in a contracted position and lying on their side. Progressively, and especially from the transition to the Final Pastoral onwards, these same structures become the “family tombs” of clans, which sometimes reuse the same stone mound for centuries. The burials are not only men (as during the Late Pastoral), but also women, children and adolescents: their shared feature is the (social) potential to access this privilege. A demographic study of the Wadi Tanezzuft region shows that only part of Final Pastoral society is buried in these stone monuments: the others, those of lesser rank and power, are buried elsewhere (di Lernia et al. 2002).

FIGURE 9.8  Final Pastoral features. (a) The conical mounds (T1–T2) from Site 96/129 mark the transition from the concluding phases of the Late Pastoral to the mature development of the Final Pastoral, Wadi Tanezzuft; (b) a megalithic stone monument from Wadi Tanezzuft; (c) painting of different styles from Ti-n-Anneouin, Acacus; (d) a pot after restoration from a tomb, Site 00/195, dated ~3.7 ka, Wadi Tanezzuft.

From “Green” to “Brown”

197

The isotopic study of dental enamel and post-cranial skeletons (particularly the strontium ratio) suggests that these groups were increasingly mobile, with birthplace and burial place no longer coinciding. Women, especially in the Final Pastoral, seem to come from other regions outside the Wadi Tanezzuft, confirming the exogamic structure of these families, often formed thanks to the interregional contacts of the nomadic elites and warriors that characterize the last stages of Saharan prehistory (Tafuri et al. 2006). The material culture mirrors this process, with different pottery traditions and the presence of exotic materials (Figure 9.8). The architectural complexity of the Sahara’s stone monuments has always aroused the interest of scholars: the presence of these “megalithic” structures is a true marker of the desert landscape and its history. The construction of monuments for the burial of family members, alongside large structures with a symbolic and ritual significance probably belonging to ceremonial complexes, is part of that “three-dimensional” landscape described by Ingold (2000) and that specifically characterizes nomadic herders and especially those of desert environments. Ever since, the Sahara Desert has been continuously dotted with these stone structures: parts of a pastoral landscape, constantly revitalized, mnemonically constructed and socially transmitted to subsequent generations.

REFERENCES Aumassip, G., and G. Delibrias. 1982–1983. Ages des depots neolithiques du gisement de Ti-n-Hanakaten (Tassili-n-Ajjer. AlgeÌrie). Libyca XXX–XXXI: 207–211. Barham, L., and P. Mitchell. 2008. The First Africans. Cambridge: Cambridge University Press. Barich, B.E., ed. 1987. Archaeology and Environment in the Libyan Sahara. The Excavations in the Tadrart Acacus, 1978–1983. Oxford: BAR. Barich, B., G. Lucarini, M.A. Hamdan, and F. Hassan, eds. 2014. From Lake to Sand. The Archaeology of Farafra Oasis Western Desert, Egypt. Firenze: All’Insegna del Giglio. Barré, H. 2003. The Sahara of cultures and people. Towards a strategy for the sustainable development of tourism in the Sahara, in the context of combating poverty. Paris: UNESCO, http://unesdoc.unesco.org/ images/0013/001330/133046e.pdf. Berger, J.F., and J. Guilaine. 2009. The 8200 cal BP abrupt environmental change and the Neolithic transition: A Mediterranean perspective. Quaternary International 200(1–2): 31–49. Biagetti, S., and S. di Lernia. 2013. Holocene deposits of Saharan rock shelters: the case of Takarkori and other sites from the Tadrart Acacus Mts. (southwest Libya). African Archaeological Review 30(3): 305–328. Brass, M. 2017. Early North African cattle domestication and its ecological setting: A reassessment. Journal of World Prehistory 31: 81–115. Breunig, P. 2013. Pathways to food production in the Sahel. In The Oxford Handbook of African Archaeology, edited by P. Lane and P. Mitchell, 555–570. Oxford: Oxford University Press. Brooks, N., I. Chiapello, S. di Lernia, N. Drake, M. Legrand, C. Moulin, and J. Prospero. 2005. The climateenvironment-society nexus in the Sahara from prehistoric times to the present day. The Journal of North African Studies 10(3–4): 253–292. Bubenzer, O., and H. Riemer. 2007. Holocene climatic change and human settlement between the central Sahara and the Nile Valley: Archaeological and geomorphological results. Geoarchaeology 22(6): 607–620. Cancellieri, E., M. Cremaschi, A. Zerboni, and S. di Lernia. 2016. Climate, environment, and population dynamics in Pleistocene Sahara. In Africa from MIS 6–2: Population Dynamics and Palaeoenvironments, edited by S. Johnes and B. Steward, 123–145. Dordrecht: Springer Science. Cancellieri, E., and S. di Lernia. 2013. Middle Stone Age human occupation and dispersals in the Messak plateau (SW Libya, central Sahara). Quaternary International 300: 142–152. Cancellieri, E., and S. di Lernia. 2014. Re-entering the central Sahara at the onset of the Holocene: A territorial approach to Early Acacus hunter–gatherers (SW Libya). Quaternary International 320: 43–62. Clarke, J., N. Brooks, E.B. Banning, M. Bar-Matthews, S. Campbell, L. Clare, M. Cremaschi, et al. 2016. Climatic changes and social transformations in the Near East and North Africa during the ‘long’ 4th millennium BC: A comparative study of environmental and archaeological evidence. Quaternary Science Reviews 136: 96–121. Claussen, M., V. Brovkin, A. Ganopolski, C. Kubatzki, and V. Petoukhov. 2003. Climate change in northern Africa: The past is not the future. Climatic Change 57(1–2): 99–118.

198

Climate Changes in the Holocene

Cremaschi, M. 2002. Late Pleistocene and Holocene climatic changes in the central Sahara: the case study of the southwestern Fezzan Libya. In Droughts. Food and Culture, edited by F.A. Hassan, 65–82. New York: Kluwer Academic. Cremaschi, M., and S. di Lernia. 1999. Holocene climatic changes and cultural dynamics in the Libyan Sahara. African Archaeological Review 16(4): 211–238. Cremaschi, M., and S. di Lernia. 2001. Environment and settlements in the Mid-Holocene palaeo-oasis of Wadi Tanezzuft (Libyan Sahara). Antiquity 75: 815–825. Cremaschi, M., and A. Zerboni. 2011. Human communities in a drying landscape. Holocene climate change and cultural response in the central Sahara. In Landscape and Societies, edited by I.P. Martini and W. Chesworth, 67–89. Dordrecht/Heidelberg/London/New York: Springer. Cremaschi, M., A. Zerboni, A.M. Mercuri, L. Olmi, S. Biagetti, and S. di Lernia. 2014. Takarkori rock shelter (SW Libya): an archive of Holocene climate and environmental changes in the central Sahara. Quaternary Science Reviews 101: 36–60. Cremaschi, M., A. Zerboni, C. Spötl, and F. Felletti. 2010. The calcareous tufa in the Tadrart Acacus Mt. (SW Fezzan, Libya). An early Holocene palaeoclimate archive in the central Sahara. Palaeogeography, Palaeoclimatology, Palaeoecology 287: 81–94. deMenocal, P.B. 2011. Climate and human evolution. Science 331: 540–452. DeMenocal, P., J. Ortiz, T. Guilderson, J. Adkins, M. Sarnthein, L. Baker, and M. Yarusinsky. 2000. Abrupt onset and termination of the African Humid Period: Rapid climate responses to gradual insolation forcing. Quaternary Science Reviews 19(1–5): 347–361. de Menocal, P.B., and J.E. Tierney. 2012. Green Sahara: African humid periods paced by earth’s orbital changes. Nature Education Knowledge 3(10):12. di Lernia, S. 1996. Changing adaptive strategies: A long-term process in the Central Saharan Massifs from Late Pleistocene to Early Holocene. The Tadrart Acacus perspective (Libyan Sahara). In Colloquia of the XIII International Congress of Prehistoric and Protohistoric Sciences. Vol. 15: The Prehistory of Africa, edited by G. Aumassip, J.D. Clark, and F. Mori, 195–208. Forlì:A.B.A.C.O. di Lernia, S. 2001. Dismantling dung: Delayed use of food resources among early Holocene foragers of the Libyan Sahara. Journal of Anthropological Archaeology 20: 408–441. di Lernia, S. 2002. Dry climatic events and cultural trajectories: adjusting Middle Holocene Pastoral economy of the Libyan Sahara. Droughts, Food and Culture, edited by F. Hassan, 225–250. New York: Kluwer Academic/Plenum Publisher. di Lernia, S. 2013. The emergence and spread of herding in Northern Africa: A critical reappraisal. The Oxford Handbook of African Archaeology, edited by P.J. Mitchell and P.J. Lane, 527–540. Oxford: Oxford University Press. di Lernia, S. 2017a. Archeologia Africana. Preistoria, Storia Antica e Arte Rupestre. Roma: Carocci Editore. di Lernia, S. 2017b. Desert environments. The SAS Encyclopedia of Archaeological Sciences, Wiley. di Lernia, S. 2017c. Genesis and development of pastoralism in Sahara and North Africa. General History of Africa, edited by A. Holl, IX. Paris: UNESCO. di Lernia, S., S. Bruni, I. Cislaghi, M. Cremaschi, M. Gallinaro, V. Gugliemi, A.M. Mercuri, et al. 2016. Colour in context. Pigments and other coloured residues from the Early-Middle Holocene site of Takarkori (SW Libya). Archaeological and Anthropological Sciences 8(2): 381–402. di Lernia, S., and G. Manzi. 1998. Funerary practices and anthropological features at 8000–5000 BP. Some evidence from central-southern Acacus (Libyan Sahara). In Wadi Teshuinat: Palaeoenvironment and Prehistory in South-western Fezzan (Libyan Sahara), edited by M. Cremaschi and S. di Lernia, 217– 242. Milano: C.N.R. di Lernia, S., and G. Manzi, eds. 2002. Sand, Stones, and Bones: The Archaeology of Death in the Wadi Tanezzouft Valley (5000–2000 BP). AZA Monographs. Firenze: All’Insegna del Giglio. di Lernia, S., G. Manzi, and F. Merighi. 2002. Cultural variability and human trajectories in later prehistory of the Wadi Tanezzuft. In Sands, Stones and Bones. The Archaeology of Death in the Wadi Tanezzuft Valley (5000–2000 BP), edited by S. di Lernia and G. Manzi, 281–302. Firenze: All’Insegna del Giglio. di Lernia, S., and F. Merighi. 2006. Transitions in the later prehistory of the Libyan Sahara, seen from the Acacus Mountains. In The Libyan Desert: Natural Resources and Cultural Heritage, edited by D. Mattingly, S. McLaren, E. Savage, Y. al-Fasatwi, and K. Gadgood, 111–122. London: The Society for Libyan Studies. di Lernia, S., and A. Palombini. 2002. Desertification, sustainability, and archaeology: Indications from the past for an African future. Origini 24: 303–334.

From “Green” to “Brown”

199

di Lernia, S., and M.A. Tafuri. 2013. Persistent deathplaces and mobile landmarks. The Holocene mortuary and isotopic record from Wadi Takarkori (SW Libya). Journal of Anthropological Archaeology 32(1): 1–15. di Lernia, S., M.A. Tafuri, M. Gallinaro, F. Alhaique, M. Balasse, L. Cavorsi, P.D. Fullagar, et al. 2013. Inside the “African Cattle Complex”: Animal Burials in the Holocene Central Sahara. PLoS ONE 8(2): e56879. Di Vincenzo, F., I. Carbone, L. Ottini, A. Profico, R. Ricci, M.A. Tafuri, G. Fornaciari, et al. 2015. Modern beams for ancient mummies computerized tomography of the Holocene mummified remains from Wadi Takarkori (Acacus, South-Western Libya; Middle Pastoral). Medicina nei Secoli 27(2): 557–588. Drake, N.A., R.M. Blench, S.J. Armitage, C.S. Bristow, K.H. White, and O. Bar-Yosef. 2011. Ancient watercourses and biogeography of the Sahara explain the peopling of the desert. Proceedings of the National Academy of Sciences of the United States of America (2): 458. Dunne, J., R. Evershed, M. Salque, L. Cramp, S. Bruni, K. Ryan, S. Biagetti, et al. 2012. First dairying in ‘green’ Saharan Africa in the 5th millennium BC. Nature 486: 390–394. Echallier, J.C., and J.P. Roset. 1986. La Céramique des gisements de Tagalagal et de l’Adrar Bous 10 (Aïr, Republique du Niger). Cahier Sciences Humaines 22(2): 151–158. Eramo, G., A. Aprile, I. Muntoni and A. Zerboni. 2014. Textural and morphometric analysis applied to Holocene pottery from Takarkori rock shelter (SW Libya, Central Sahara): A quantitative sedimentological approach. Archaeometry 56: 36–57. Garcea, E.A.A., ed. 2001. Uan Tabu in the Settlement History of Libyan Sahara. Firenze: All’Insegna del Giglio. Garcea, E.A.A. 2004. An alternative way towards food production: the perspective from the Libyan Sahara. Journal of World Prehistory 18(2): 107–154. Haaland, R., and G. Haaland. 2013. Early farming societies along the Nile. In The Oxford Handbook of African Archaeology, edited by P. Mitchell and P. Lane, 541–553. Oxford: Oxford University Press. Hassan, F.A. 2002. Ecological changes and food security in the later prehistory of North Africa: looking forward. In Droughts, Food and Culture, edited by F. Hassan, 321–334. New York: Kluwer/Plenum. Hassan, F.A. 2002. Palaeoclimate, food and culture change in Africa: An overview. In Droughts, Food and Culture: Ecological Change and Food Security in Africa’s Later Prehistory, edited by F.A. Hassan, 11–26. New York: Kluwer Academic/Plenum Publishers. Huysecom, E. 2018. Arguments for an Early Neolithic in Sub-Saharan Africa. Retrieved January 9, 2018, from http://www.ounjougou.org/en/projects/mali/archaeology/arguments-for-an-early-neolithic-in-subsaharan-africa. Huysecom, E., M. Rasse, L. Lespez, K. Neumann, A. Fahmy, A. Ballouche, S. Ozainne, et al. 2009. The emergence of pottery in Africa during the tenth millennium cal BC: New evidence from Ounjougou (Mali). Antiquity 83(322): 905–917. IFAD. 2010. Desertification. Rome: International Fund for Agricultural Development. Ingold, T. 2000. The Perception of the Environment. Essays in Livelihood, Dwelling and Skill. London/New York: Routledge. Jesse, F. 2010. Early pottery in Northern Africa—An overview. Journal of African Archaeology 8(2): 219–238. Jordan, P., K. Gibbs, P. Hommel, H. Piezonka, F. Silva, and J. Steele. 2016. Modelling the diffusion of pottery technologies across Afro-Eurasia: Emerging insights and future research. Antiquity 90(351): 590–603. Kleindienst, M.R., H.P. Schwarcz, K. Nicoll, C.S. Churcher, J. Frizano, R.W. Giegangack, and M.F. Wiseman. 1996. Pleistocene geochronology and paleoclimates at Dakleh Oasis and Kharga Oasis, Western desert, Egypt, based upon uranium-thorium determinations from water-laid tufas. Nyame Akuma 46: 96. Kobashi, T., J.P. Severinghaus, E.J. Brook, J.M. Barnola, and A.M. Grachev. 2007. Precise timing and characterization of abrupt climate change 8200 years ago from air trapped in polar ice. Quaternary Science Reviews 26(9–10): 1212–1222. Kröpelin, S. 2002. Damage to natural and cultural heritage by petroleum exploration and desert tourism in the Messak Settafet (Central Sahara, Southwest Libya). In Tides of the Desert. Contributions to the Archaeology and Environmental History of Africa in Honour of Rudolph Kuper, edited by T. LenssenErz, U. Tegtmeier, and S. Kröpelin, 14: 405–423. Köln: Heinrich Barth Institute. Kröpelin, S., D. Verschuren, A.M. Lézine, H. Eggermont, C. Cocquyt, P. Francus, J.P. Cazet, et al. 2008. Climate-driven ecosystem succession in the Sahara: The past 6000 years. Science 320(5877): 765–768. Kuper, R., and S. Kröpelin. 2006. Climate-controlled Holocene occupation in the Sahara: Motor of Africa’s evolution. Science 313(5788): 803–807. Laureano, P. 2001. Atlante d’acqua: conoscenze tradizionali per la lotta alla desertificazione. Torino: Bollati Boringhieri.

200

Climate Changes in the Holocene

Linstädter, J., I. Medved, M. Solich, and G.-C. Weniger. 2012. Neolithisation process within the Alboran territory: Models and possible African impact. Quaternary International 274: 219–232. Livingstone Smith, A. 2001. Pottery manufacturing processes: Reconstruction and interpretation. In Uan Tabu in the Settlement History of the Libyan Sahara, edited by E. Garcea, 113–152. Firenze: All’Insegna del Giglio. Lubell, D. 2014. Northwest African prehistory: Recent work, new results and interpretations. Quaternary International 320: 1–2. Maley, J. 2010. Climate and Palaeoenvironment evolution in north tropical Africa from the end of the Tertiary to the Upper Quaternary. Palaeoecology of Africa 30: 227–278. McDonald, M.M.A. 2016. The pattern of Neolithization in Dakhleh Oasis in the Eastern Sahara. Quaternary International 410, Part A: 181–197. Mercuri, A.M. 2008. Human influence, plant landscape evolution and climate inferences from the archaeobotanical records of the Wadi Teshuinat area (Libyan Sahara). Journal of Arid Environments 72(10): 1950–1967. Mercuri, A.M., R. Fornaciari, M. Gallinaro, S. Vanin, and S. di Lernia. 2018. Plant behaviour from human imprints and the cultivation of wild cereals in Holocene Sahara.Nature Plants 4: 71–81. Mori, F. 1965. Tadrart Acacus. Arte Rupestre e Culture del Sahara Preistorico. Torino:Einaudi. Mori, L., F. Ricci, M.C. Gatto, E. Cancellieri, and C. Lemorini. 2013. The excavation of the Fewet Necropolis. In Life and Death of a Rural Village in Garamantian Times. Archaeological Investigations in the Oasis of Fewet (Libyan Sahara), edited by L. Mori, 253–318. All’Insegna del Giglio, Florence, Italy. Mulazzani, S., L. Belhouchet, L. Salanova, N. Aouadi, Y. Dridi, W. Eddargach, J. Morales, et al. 2016. The emergence of the Neolithic in North Africa: A new model for the Eastern Maghreb. Quaternary International 410, Part A: 123–143. Nicoll, K. 2001. Radiocarbon chronologies for prehistoric human occupation and hydroclimatic change in Egypt and Northern Sudan. Geoarchaeology 16(1): 47–64. Pereira, L., N. Silva, R. Franco-Duarte, V. Fernandes, J. Pereira, M. Costa, H. Martins, et al. 2010. Population expansion in the North African Late Pleistocene signalled by mitochondrial DNA haplogroup U6. BMC Evolutionary Biology 10(1): 390. Smith, A.B. 2005. African Herders: Emergence of Pastoral Traditions. Walnut Creek: AltaMira Press. Tafuri, M.A., R.A. Bentley, G. Manzi, and S. di Lernia. 2006. Mobility and kinship in the prehistoric Sahara: Strontium isotope analysis of Holocene human skeletons from the Acacus Mts. (southwestern Libya). Journal of Anthropological Archaeology 25(3): 390–402. UNEP. 2008. Africa: Atlas of Our Changing Environment. Nairobi: United Nations Environment Programme (UNEP). Van Peer, P. 2001. Observations on the Palaeolithic of the south-western Fezzan and thoughts on the origin of the Aterian. In Uan Tabu in the settlement history of Libyan Sahara, edited by E.A.A. Garcea, 51–62. Firenze: All’Insegna del Giglio. Wasylikowa, K. 2001. Site E-75–6: Vegetation and subsistence of the Early Neolithic at Nabta Playa, Egypt, reconstructed from charred plant remains. In Holocene Settlement of the Egyptian Sahara, edited by F. Wendorf, R. Schild and Associates 544–591. Kluwer Academic Publishers/Plenum Publishers, New York. Wright, D.K. 2017. Humans as agents in the termination of the African humid period. Frontiers in Earth Science 5(4).

10

Eastern Borders of the Sahara and the Relations with the Nile Valley and Beyond Barbara E. Barich

CONTENTS 10.1 Introduction: Climate of the Eastern Sahara in the Late Pleistocene.................................... 201 10.2 The Reoccupation of the Eastern Sahara: Climate Framework and the Early Settlements......... 203 10.2.1 The Climate............................................................................................................... 203 10.2.2 The Flimsy Structures............................................................................................... 205 10.3 The Problem of Cattle in the Nubian Desert.........................................................................207 10.4 The Bimodal Phase of Rainfall—Settlements of the Middle-Holocene at Farafra and Dakhla............................................................................................................209 10.4.1 Farafra........................................................................................................................209 10.4.2 Dakhla....................................................................................................................... 210 10.5 Mobility and Exchanges in the Late Holocene: From Desert to Nile................................... 211 10.6 The Near East Influences and the Spread of the Levantine “Neolithic Package”................. 213 10.7 Conclusion............................................................................................................................. 214 Acknowledgments........................................................................................................................... 214 References....................................................................................................................................... 215

10.1 INTRODUCTION: CLIMATE OF THE EASTERN SAHARA IN THE LATE PLEISTOCENE The importance of the environment as a main driving force of social transformations has always been acknowledged in the studies of the Sahara and North Africa in general. The environment played an important role and indeed, through adaptation processes, seemed to manage change. This model, easily imputable of environmental determinism, has often been opposed by the cognitive trends in archaeology aiming at diminishing the role of the environmental component tout court, in comparison to the way in which it is perceived by the society (Zubrow 1975). The alternation between cycles of wet phases, more favorable to human occupation, and cycles of arid phases that occurred repeatedly in the Sahara during the Late Pleistocene were governed by astronomical phenomena (primarily the precession of the equinoxes) which determine the summer radiation over the Earth’s surface and the relative advance and retreat of the glaciers in the boreal hemisphere (de Menocal and Tierney 2012). The Sahara was particularly arid during the Last Glacial Maximum (LGM), when ice in the Northern Hemisphere blocked the seasonal shift to the North of the Inter Tropical Convergence Zone (ITCZ), on which depends the onset, duration and termination of the monsoon-rainy season in the tropics and subtropics. Northern cold flows to the very southern latitudes went to reinforce the anticyclones of the subtropical belt that block monsoonal rainfalls: because of the cold weather, the monsoon could not go back North every summer. Furthermore,

201

202

Climate Changes in the Holocene

the polar mobile anticyclones blocked the cyclonic rainfalls coming from the Atlantic Ocean on the western face of North Africa (Leroux 1991, 1993). Snowfields were present on the northern slopes of the Ahaggar’s peaks and, according to the indications given by pollens, the climate in Algeria, northern Mali, Niger, and southern Libya experienced a drop up to 6°C (Vernet 1995). After the last interglacial (MIS5, between 125 ka BP and 80 ka BP),* during which Saharan territories benefited from a favorable climate with abundant rains also leading to the formation of lakes (Gasse 2000), the climate rapidly worsened. The crisis was particularly severe in the continental areas, while both the Nile Valley and the North African coast did not suffer equally drastic consequences. The Nile continued to flow during the Glacial Maximum, despite a water decrease in the region of the springs. Because of the reduced activity of the White Nile and the Blue Nile, the Nile river’s course was very slow since it was weighed down by the sediment load. Dunes from the Western Desert invaded the Nile Valley at several places, and the lakes newly formed behind such dams offered ideal conditions for human subsistence. In Sudan, the Jebel Marra lake dropped about twenty meters between 22 and 20/16 ka cal BP, with a minimum around 17 ka years ago (Williams et al. 1980; Nicoll 2004). MIS 4 (between ca 80 ka BP and 60 ka BP) was a particularly severe phase throughout the Sahara, and in part also the MIS 3 (between 60 ka BP and 30 ka BP). It is thought that the particularly arid conditions of the tropical areas in the phases that followed MIS 5 could have caused migrations to the North and abandonment of the territories (Cohen et al. 2007; Schulz et al. 2007; Carto et al. 2009). However, it is difficult to reconstruct a uniform palaeoclimatic framework in such a large geographical area, lacking data archives to be used comparatively. The absence of synchroneity in the environmental changes indicates rather the existence of regional situations that may have created favorable local conditions contradicting the general trend (Nicoll 2004). For the Eastern Sahara – the subject of this chapter – a fundamental archive of information is offered by the paleo-lakes region of Bir Tarfawi and Bir Sahara, in the southernmost part of the Egyptian Western Desert (Wendorf et al. 1993). An improved geochronology of the Middle Stone Age lakes based on nine SAR-OSL dates (Single Aliquot Regeneration Optically Stimulated Luminescence) made it possible to specify the expansion phases of the groundwater-supported Pleistocene paleo-lakes present at Bir Tarfawi and in the neighboring area of Bir Sahara East and to establish the respective inter-basinal correlations. These new results allow the highlighting of peaks of greater availability of water at 115 (MIS5d-e), 102 (MIS5c-d), 80 (MIS5a) and 41 (MIS3) ka BP (Nicoll 2018: Table 3). The presence of periods in which one of the currently most arid areas of the globe was favorable for human occupation, seems also important for reconstructing the “Out of Africa” paths of anatomically modern man (Nicoll 2018 cit.). The date of the last lake phase so far known (41 ka BP), falling in the MIS3, can offer a chronological reference to the Saharan Aterian collections found in the area. The fact that in other Saharan regions this same date is associated with hyperarid conditions (Cancellieri et al. 2016) clearly confirms what has been said above: the existence of diversified Saharan landscape where regional factors and hydroclimate variations may have influenced water availability. The presence of wet conditions in the Western Desert around 40 ka BP is also confirmed by geological records in the Wadi el Obeiyid Cave 1 (Farafra Cave) (Hamdan et al. 2014; Barich 2014d). Two U/th datings on speleothems from the Farafra Cave place flowstone and dripstone formations at 287 ± 67 ka (Early Stone Age, MIS8-7) and 45 ± 2 ka BP (MIS3). The latter dating, which may correspond to the last deposition of carbonates in the cave, could also provide the geochronological reference for the formation of the playa identified at Sheikh el Obeiyid, on one of the erosion surfaces of the Northern Plateau (Barich et al. 2012) and also for the MSA stone artifacts collected inside the playa sediments and in the surrounding area (Van Peer 2014). *

Radiocarbon dating presented in this work are uniformly indicated as calibrated dates by the present (cal BP), in which BP corresponds to the conventional 1950 value. They are expressed in ka: unit of measure corresponding to 1000 years. When useful, to the cal BP dating even the one calculated from the current era (BC) has been added. For those dates not available in a calibrated form, the Oxford University calibration system has been used by the author (Reimer et al. 2013).

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

203

Extensive lacustrine sediments in the form of marls (authigenic calcite silts) deposited in basins of perennial waters are also described from Dakhla Oasis (Smith 2010, 10). These paleo-lakes were active mainly during MIS5, and this is also indicated by the MSA artifacts stratigraphically associated with the lake sediments. To the same MIS5 – undoubtedly the wettest and most watered phase in the Egyptian Western Desert – can also be referred the paleo-lakes discovered south of Dakhla in the Oasis of Kharga, whose major expansion phases have been determined by Uranium-series datings on the tufas coming from three different regions of the oasis (Smith et al. 2007). These new dates place the lake phases between late MIS6 and MIS5 (between 127.9 ± 1.3 ka and 104 ± 14 ka), and also in this case are a reference for the associated MSA industries: Upper Levalloisian at Mata’na SiteG and Bulaq Wadi. The latter complex appears to be slightly more recent than 124 ka, while the tufas of Wadi Midauwara, not associated with industries, have yielded an intermediate date of about 133 ka (Smith et al. 2007). No presence of lakes is attested in the Eastern Sahara after MIS5 and before MIS1. The very low monsoonal indexes recorded between MIS4 and MIS2 suggest that any sporadic precipitations may have had only a short and low intensity character. Some evidence of such sporadic rains, recorded by Smith et al. (2004) in the territory of Kharga, are dated to about 50 ka, while two other small lake deposits in the Dakhla region have been dated respectively 62 ka (Brookes 1993) and 40 ± 10 ka (Churcher et al. 1999). The latter therefore appear to have been the most recent wet episodes of a certain importance in the Western Desert. During MIS2 (25–15 ka BP), the availability of water and the related peopling was further reduced throughout the Sahara, especially in parallel with the LGM and Younger Dryas (YD) harsher phases. The aeolian MIS2 deposits in north-western Libya (Jebel Gharbi region) are a clear indication of the expansion to the North and the Mediterranean of the northern boundary of the Sahara (Giraudi 2005). Record of human settlement dated to this period comes only from the Mediterranean coast and from the Nile Valley (Wendorf et al 1993). The reestablishment of the monsoon regime with the shift towards North of the ITCZ, started at 15 ka BP and arrested in relation to the YD, was then resumed and decisively established by 12–10,000 BP.

10.2 THE REOCCUPATION OF THE EASTERN SAHARA: CLIMATE FRAMEWORK AND THE EARLY SETTLEMENTS 10.2.1 The Climate Evident transformations in the social aggregations took place at the Pleistocene/Holocene transition. With the resuming of the monsoon regime, even the Saharan lands, so long abandoned, were progressively reoccupied by groups that tended to expand their resource catchment. The Eastern Sahara and the Nile Valley represented, then, a generalized ecosystem of hunting and hunting-fishing-gathering activities aimed at a broad exploitation of resources. Groups acquired a deep knowledge of their habitats that they exploited intensively, diversifying to the maximum the exploitation base. Since people used immediately available resources, or those at a short distance from the home base, they were not forced to long transfers. The reduced nomadism ended up with the establishment of residential or semi-residential attitudes, during the periods of greater environmental productivity. Based on these assumptions, it was suggested that these groups could represent a context pre-adapted to fully domestication activities (Clark 1976). During the Holocene, wetter conditions prevailed in most of Egypt’s and northern Sudan’s territories (Haynes 2001). Various temporary lakes (playas) filled the basins and deposited the lacustrine sediments that today yield copious remains – tools, hearths, fauna assemblages – of the old encampments, formerly settled on the beaches. On the other hand, dune formations and deflation activity, mainly recorded at 9.7, 9.2 and 8.2 ka cal BP, indicate strong climatic events that determined the abandonment of the inhabited sites (Wendorf and Schild 2001). The Egyptian Western Desert (Figure 10.1), a buffer zone between the Nile Valley and the Central Sahara, was inhabited since the Early Stone Age, but the most important settlement phase started in the

204

Climate Changes in the Holocene

FIGURE 10.1  Map of the Western Desert with location of sites cited in the text. (Image: B.E. Barich.)

Early Holocene (>10.0–8.0 ka cal BP i.e. 8000–6000 cal BC) and was characterized by wet and dry climatic fluctuations. Fieldwork carried out during the second half of the twentieth century highlighted with precision the palaeoclimate sequence of this large territory. As a whole, the extensive playa deposits recorded in the Farafra and Dakhla Oases (Brookes 1989; Hassan et al. 2014); the Gilf El Kebir (Kuper and Kroepelin 2006, 803), and the Kiseiba-El Nabta region (Wendorf and Schild 1980; Close 1987, 1995; Haynes 2001) were the basis for the reconstruction of the palaeoclimate and human occupation of the Egyptian Western Desert. In the southern region of Kiseiba-El Nabta, research in the 1990s (1990–1999) has, compared to previous investigations, allowed the drawing of a new picture (Wendorf et al. 2001). The Early Holocene comprises three main wet phases: El Adam, between 10.8 and 10.0–9.9 ka cal BP; El Ghorab 9.5–9.2 ka cal BP; El Nabta/Al Jerar 8.9–8.1 ka cal BP. The three phases are separated by as many dry intervals which lasted only a century or so and, therefore, were comprised in a time space highly evaluable on a human scale, such as to produce evident reorganization in the groups. Further North, inside what is today one of the most arid territories on earth, the five oases – Siwa, Bahariya, Farafra, Dakhla and Kharga – are at the center of limestone depressions. It seems that the deepest limestone basins, true reservoirs of water, were excavated through a process of karstic activity. All these areas have come to the fore for the outstanding evidence of the collection and early management of local wild plants and of their association to herding, thus dispelling the doubts of the

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

205

anthropologists about the way in which these two activities could coexist among primitive societies. In particular, such early ‘proto-cultivation’ experiences highlighted in these territories have allowed a new interpretation of the agricultural contexts of the Nile Valley, a few of which can be compared to the Neolithic development in the Near East. A four-step sequence has been proposed for the entire region, using the 14C dates known from various territories of the Eastern Sahara, mostly Libya, Chad and Egypt (Kuper and Kroepelin 2006, 803). A phase of reoccupation of the desert in the Early Holocene (8500 to 7000 BCE)* would be followed by a stable phase (Formation phase: 7000–5300 BCE) which ended abruptly. It would then be followed by a new wet phase with a shift in occupation towards the plateaus to exploit the water basins present there (Regionalization phase: 5300–3500 BCE). Finally, during the fourth phase (Marginalization phase: 3500–1500 BCE) the territory was increasingly abandoned until it became only a crossing area. Greater precision in comparison to this general outline have been offered by the long geoarchaeological investigation in the Farafra Oasis and by the results obtained thanks to the multidisciplinary project integrating paleoenvironmental, stratigraphic and archaeological data (Barich et al. 2014). At Farafra, the Early Holocene (>10.0 ka cal BP) experienced a wet climate suggesting wadi activity and the formation of ephemeral lakes inside the limestone basins (Hamdan 2014a). However, around c.8.5 ka cal BP the situation began to change with the establishment of a dry trend that in general caused the shrinking of the wetlands. Later, between c. 8.1 and 7.4 ka cal BP, the territory up to the latitude of Dakhla benefited from a favorable climate with a double rainfall regime: besides the tropical monsoon, a descent of moist air from the Mediterranean would have also brought winter rains there (Arz et al. 2003; Barich and Lucarini 2014). Despite this, the sediments of the Hidden Valley basin at Farafra show that the lake expanded and shrank repeatedly in relation to short climatic fluctuations. By c.7.0 ka cal BP, a brief period of arid crisis led to the desiccation of the lake (Hamdan cit.), which nevertheless experienced a new favorable phase around 6.2 ka cal BP. But starting from 5.5 ka cal BP, the site was completely abandoned. From 9.4 to 7.8 ka cal BP, even southward of the Nabta territory in northern Sudan, lacustrine conditions existed as evidenced by the remains of aquatic fauna (Crocodylus niloticus, Hippopotamus amphibius, Testudo sulcata) from various places (Nicoll 2004; Hoelzmann et al. 2001). Pollen records from Selima Oasis and Oyo depression show a steep vegetation gradient within the eastern Sahara during the Early to Middle Holocene (Haynes et al. 1989). In this same period, the Oyo depression was filled by a deep lake with stagnant deep water and laminated carbonate-rich muds containing fossil diatoms, gastropods and pollen (Haynes et al. cit.).

10.2.2 The Flimsy Structures E-79–8, one of the Bir Kiseiba’s oldest sites (SMU-858: 9820 ± 380/11,834–10,700 cal BP), has yielded a lithic assemblage with a strong microlithic component of the El Adam type, with backed bladelets, endscrapers (often on re-used Levallois flakes), notches and microburins. The El Adam communities already used ceramics decorated in the style of the Saharo-Sudanese Neolithic (Figure 10.2). The almost exclusive presence of tamarisk trees in the vegetation has led to the supposition that during the El Adam phase (dated between 10.8 and 10.0/8.9 ka cal BP) (Wendorf et al. 2001), the El Nabta basin had experienced a desert-type regime, with very scarce vegetation. The monsoon rains had to allow the presence of at least a sparse vegetation on the plateaus, beyond the limit of the depression (Wendorf et al. cit., 651). Compared to the ephemeral character of the El Adam sites, those of the El Ghorab horizon, between 9.5 and 9.2 ka cal BP, exhibited a more complex organization. Settlements related to this phase (such as site E-72-5 in the Dyke area or E-77-6 in the Nabta area) had more stable characteristics. In fact, the same houses, or at least the same areas, were often reoccupied by groups as soon as the playas were dry. *

The dates cited here are reported as they appear in the original publication (Kuper and Kroepelin 2006).

206

Climate Changes in the Holocene

FIGURE 10.2  Pottery sherds from Bir Kiseiba, Site E-79–8. (From Barich 2014a, Figure 3.2.)

The situation had to change radically a few hundred years later, with the onset of the new Nabta/Al Jerar wet phase (8.9–8.1 ka cal BP). Over 20,000 remains of plants were collected from sites E-76-6 and E-91-1, of the type that prefers the edges of swamps in dry savannah environments. It would therefore seem that a certain amount of water could survive in the form of pools in the deepest areas of the depressions. A greater abundance of plants was noticed at Site E-75–6 compared to E-91–1, belonging to the younger Al Jerar horizon. Moreover, while copious taxa of Gramineae grasses were recorded inside the former site, among which the Sorghum as well as tubers (Cyperus) and remains of trees prevail, in the other, only five taxa of grasses with almost irrelevant presence of Sorghum were recognized. In both sites the only tree species were Acacia (ehrenbergiana, nilotica, raddiana) and Tamarisk. Wendorf et al. (2001, 650) assume a considerable supply of water in the playa of Nabta during the rainy season, such as to feed a dense vegetation. The deposition of clay sediments mixed with minerals was minimal during the Nabta/Al Jerar phase, although the lake, at its maximum, had to reach the beaches where the settlements were located. This could indicate that the shores of the lake were covered by dense vegetation that strengthened the ground and prevented any slipping. Nabta E-75–6 presented alignments of structures – hut foundations, hearths, storage wells. Numerous comparisons have been made between the lithic industry found on this site with those recovered from other areas in the Western Desert, such as Wadi Bakht in the Gilf Kebir, Abu Ballas and Lobo in front of the Great Sand Sea, the Masara B phase at Dakhla and the Early Holocene horizon at Farafra. Some cattle remains at an incipient domestication stage were collected from E-75–6 as well as from other sites of the Kiseiba-Nabta region (Gautier 2001), while caprines appeared very late in the Nabta sequence. In the Farafra Oasis, the oldest Early Holocene occupation is represented by Ain e-Raml and other small complexes close to center of the Oasis, Qasr Farafra. These are dated between 9260 ± 110 (Hel-4130: 10,562–10,290 cal BP) and 9650 ± 190 (R-1983: 11,230–10,723 cal BP) (Barich and Hassan 1984–87; Barich 2014a). Other sites, related to both the El Adam and El Nabta phases, were recorded more to the North along the Wadi El Obeiyid: Bir El Obeiyid Playa (Gd-30181: 9420 ± 230/10,890–10,404 cal BP); Sheikh El Obeiyid (Gd-11648:7925 ± 60/8786–8638 cal BP); Northern Plateau (GdA-2260: 8580 ± 25/9546/9534 cal BP; GdA-2258: 8010 ± 30/8832–8735

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

207

FIGURE 10.3  Epipalaeolithic industry from Ain e-Raml (Farafra Oasis). (Image: B.E. Barich.)

cal BP) (Hamdan and Lucarini 2013; Barich and Lucarini 2014). From all these sites, representing short-term encampments on the shores of seasonal lakes or pools, rich collections of artifacts with a pronounced laminar character, such as bladelets from single platform cores, burins, backed bladelets and microliths, were recovered. At Ain e-Raml, the archaeological level is only a few centimeters thick, a sign of short-term frequentation, like for a seasonal stay. The archaeological complex is made up of a microlithic laminar industry with backed elements, microbladelets and burins with a clear Epipalaeolithic characterization (Figure 10.3). Pottery has also been collected, whose association with the Epipalaeolithic occupation is, however, doubtful, because it lacks diagnostic features. For their ephemeral character, these Early Holocene sites should be linked to hunter-forager entities still very mobile on the territory, as it is mirrored in their wide distribution and the shallow depth of the deposits. As a matter of fact, similar cultural features are found scattered over wide spaces, reaching even the inner Sahara. Influences from the El Ghorab (and also El Adam) lithic traditions have been detected in the Ti-n-Torha assemblage in the Libyan Tadrart Akakus (Close 1987) and also in the intermediate Libyan-Egyptian desert, such as Lobo in the Sand Sea (Klees 1989, Figure 5) and Wadi El Akhdar in the Gilf El Kebir region where El Ghorab sites can be found (Site 80/7–4, Kuper 1981, Figure 16). The Great Sand Sea is another region where the presence of interdune corridors, which became a point of water collection and thus the basis for the development of plants for grazing, encouraged hunter-gatherer occupation (Riemer 2006; Gehlen et al. 2002). Even in the other oases of the Western Desert, the Early Holocene saw the flowering of various industries with an accentuated microlithic base. As examples, we can cite the Siwan in the Siwa Oasis (Hassan and Gross 1987), the Masara A horizon at Dakhla (McDonald 1991, 93; 1999), and finally, the Bedouin microlithic industry already described by Caton-Thompson in the Kharga Oasis (Wendorf and Schild 1980, 189).

10.3 THE PROBLEM OF CATTLE IN THE NUBIAN DESERT The Holocene archaeology of Nubia, in the southern Western Desert, is linked to the recognition of an original management center for the cattle, which for a long time represented a very controversial

208

Climate Changes in the Holocene

subject in the studies (Gautier 1984; Wendorf and Schild 1994, 2001; Decker et al. 2014; Marshall and Weissbrod 2011). The Bir Kiseiba area yielded more than 10,000 faunal remains in which arid climate animals (hares and gazelle) dominate. Among these, in a dubious form, some remains of ox at an ‘incipient domestication’ stage (Gautier 1984) were identified. Gautier observed that bovids – ox and buffalo – could not survive in the wild with such a low rate of humidity. Wendorf and Schild (1980, 1994; Wendorf et al. 1984) amplified Gautier’s recognition by stating that the reoccupation of the Nubian Desert, a territory where the only meat resources were given by hunting hares and gazelle, would have suggested to the groups coming from the Nile Valley to travel with some cattle specimens which could guarantee survival in a desert environment by supplying milk and blood. More recently, Stock and Gifford Gonzales (2013, 56) have recognized the validity of the taxonomic arguments according to which Gautier differentiated the remains of Kiseiba/Nabta from the African Buffalo (Syncerus caffer). The question, however, remains that of the status of those fragmentary remains. Are they a few remains of a domestic species, then cattle, or are these specimens simply wild bovids? On this dilemma (which is the expression of a binary paradigm that definitely separates hunting from pastoralism and/or agriculture), two opposing ‘factions’ have continued to collide, each bringing arguments in favor of its position. With regard to the ‘ecological’ topic mentioned above, the discoverers insisted on the arid conditions of the area which at that time would not have allowed the presence of herbivores requiring at least 25 L/day of water (Gautier 1984; Wendorf and Schild 1994; Wendorf et al. 2001). But to this argument it was objected that with an estimated rate of 200 mm/year, it would have been possible for wild cattle to be present on their own (Smith 2005, 88). Even this observation, however, is attackable because if the cattle had been present in the natural state and, therefore, the object of hunting like gazelles and others, it would be strange that their presence was so limited (Stock and Gifford Gonzales 2013, 59). This rarity would be best matched with the same exceptionality of these animals and their preciousness that would have induced those early shepherds to use milk and blood and to sacrifice them only on rare occasions. From then on, in short, the ritual and cultural conception linked to the ox, to cattle burials and to ceremonial centers of the type discovered at Nabta (Wendorf et al. 1997) would have been established. Since the 1990s, even geneticists have entered the discussion. At first highlighting the mtDNA taurine haplogroup, an independent domestication of the North African auroch was postulated (Bradley et al. 1996). Then, in subsequent years, the theme was further deepened, leading to recognize at least five sub-sets with different geographical distribution within the T macro-haplogroup of Bos taurus. Among these, in Africa the T1 haplogroup is present without any relation to the mtDNA sequence of Bos indicus, which is completely divergent from that of Bos taurus, indicating a separate domestication of the respective prototypes (Magee et al. 2007). Furthermore, the haplogroups distributed in Africa and Eurasia are all variations of haplogroups (T1, T2, T3) documented only in the Near East. This, therefore, allowed the Near East to be looked at as the only primary center of domestication. In essence, the T1 mtDNA African haplotype would reflect an ancient dispersion in Africa of a Southwest Asian domestic population (Stock and Gifford Gonzales 2013, 65). However, on this point we are far from having reached a conclusion. The thesis that, in the process aimed at cattle domestication, Africa may have been an independent center from those of Southwest Asia and the Indus Valley continues to be discussed (Decker et al. 2014; Marshall and Weissbrod 2011). A further contribution to the debate, that may offer the solution to the puzzle, has come from a re-orientation of the analysis. Next to the mitochondrial DNA, prevalent in the early investigations, the Y-chromosomal contributions were studied. The latter highlight the presence of male-mediated introgressions of Bos indicus Y-chromosome lineage within the taurine populations exhibiting taurine-specific mtDNA haplotypes (Stock and Gifford Gonzales 2013, 66; Hanotte et al. 2002). This may indicate the presence of crossbreed domesticated taurine livestock with wild indicine cattle, possibly entered through the horn of Africa by repeated separate introductions. In conclusion, as suggested by Gifford-Gonzales (2013), while neither the mtDNA nor the autosomal diversity support a hypothesis of repeated domestication events in the Bos taurus (on the

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

209

contrary, the star-like shape of the network supports a scenario of a single primary domestication center and subsequent diffusion), Y-chromosome-related data may support a scenario of secondary introductions of local wild aurochsen into the domestic gene pool population.

10.4 THE BIMODAL PHASE OF RAINFALL—SETTLEMENTS OF THE MIDDLE-HOLOCENE AT FARAFRA AND DAKHLA The latest research in the Western Desert has shown that starting from about 8.1 ka cal BP, a descent of humid air would have brought winter rains at least until the latitude of Dakhla. Human groups were therefore allowed to remain in the territory throughout the year, with a much more stable model than previously known. In the southernmost region of Nabta, the same result was obtained by opening some wells to store the water for the driest months (Wendorf and Schild 2002). Thus, the groups could remain at the home bases, located in the deepest parts of the basins, from the late autumn up to the return of the summer rains. From then on, they could resume hunting activities and, above all, graze the first flocks available, venturing further away from their seasonal bases. Edible plants, reaching maturity in the first part of the winter, were then stored in silos to be used during the most difficult periods (Wasylikowa 2001). The increasing importance of vegetable resources in the diet ended up imposing a change in the annual cycle. It was, in fact, necessary to prevent the flocks from staying close to the plants too much, to avoid their over-exploitation. This forced communities to split into two segments, one of which, involved in the care and collection of plants, would have remained at the residential base for the entire period, between autumn and winter; the other, however, formed by those who had the task of grazing the flocks, would have walked not far away but mainly on the surrounding plateaus (Wendorf and Schild 2002).

10.4.1 Farafra An important occupation cycle in the Mid-Holocene, representing an original example of integration between herding and horticulture, has come to light in the northern part of the Farafra Depression. The Hidden Valley site, on the shores of a Holocene lake fed by two main wadis coming from the Northern Plateau (Hamdan 2014a) (Figure 10.4), experienced various phases of occupation alternating over a long phase. The fifty dates known from this settlement cover the period from

FIGURE 10.4  Playa formations in the Hidden Valley basin, Farafra Oasis. (Photo: B.E. Barich.)

210

Climate Changes in the Holocene

8.7 to 7.0 ka cal BP (c. 6800–5200 cal BC) and, together with the size of the inhabited area, the presence of stone structures, the thickness of the stratigraphic section, imply a repeated occupation and the beginning of a stable settlement. The beginning of a sedentarization process has been proposed for this archaeological context, linked to the systematic exploitation of the abundant Gramineae plants that grew around the lake (Barich et al. 2014). Layer III of the section, made up of sand, charcoals and ash deposited inside the numerous hearth stones, is the one that showed the greatest human intervention (Barich 2014b). Of great importance is the presence of grains of Sorghum and other spontaneous grasses evidently collected by the inhabitants of the village. They were used within the settlement and within a composite economic pattern that included both domestic and wild animals. A key element to understand exchanges and relations with external territories is the presence of goat/sheep since c. 8.2 ka cal BP (c. 6200 cal BC), one of the oldest presences in all of Africa. The Hidden Valley was made attractive for prehistoric shepherd-horticulturalists by the exceptional presence of spontaneous grasses. Besides the Sorghum, many other wild plants are known, such as Brachiaria, Digitaria, Echinocloa, Panicum, Setaria and Urochloa. Sorghum remains (grains, ears and fragments of spikes) are present in almost all of the soil samples taken locally (Fahmy 2014). The composition of the floristic sample confirms the presence of abundant water, as evidenced by the remains of aquatic plants such as Boerhavia, Juncus and Scirpus. According to Fahmy (2014 cit.), plants were certainly introduced into the settlement to be consumed as food. This testifies to a deliberate choice in the collection practiced by the prehistoric community, in analogy with what is still used by many modern tribes of Sahelian Africa. In particular, the Zaghawa and the Tuaregh, pastoralists whose economy includes an important component of harvesting and cultivation, offer an adequate example of the survival of the typical costumes of dry areas (Tubiana and Tubiana 1977; Lucarini 2014a). In general, the village of Hidden Valley can be compared to other contemporaneous large settlements in various locations of the Western Desert. One of the most ancient belongs to the Masara C cultural units at Dakhla and Midauwara at Kharga, followed by – from the ninth-eighth millennium cal BP – the slab structures at Farafra, the Dakhla Bashendi Units, the Kharga Early Baris unit, in addition to those of Abu Ballas, the Great Sand Sea, the Gilf Kebir and further south of Karkur Talh and Jebel Uweinat (McDonald 2002, 2006, 2009, 2016; Hendrickx et al. 2013; Zboray 2013).

10.4.2 Dakhla In the oasis of Dakhla, a few hundred years elapse between the Epipalaeolithic occupation (called Masara: McDonald 2002) and the beginning of the Bashendi unit, whose development occupies the entire Mid-Late Holocene. Within this chronological period, McDonald (2009, 2016) has distinguished two cultural units: Bashendi A, between 8.4 and 7.7 ka cal BP, and Bashendi B, from 7.4 to 5.9–6.0 ka cal BP. However, this is more than the attribution to two distinct cultural phases, since it represents the reconstruction of a process through which in the Dakhla Oasis (as at Farafra) we observe the transition from early experiences of intensive exploitation of plants, as well as semipermanent occupation (Bashendi A), to the beginning of a pastoral organization (Bashendi B). The early Bashendi A phase has appeared in the southeastern basin (Stake Hollow). Sites were distributed over a wide area and showed signs of the exploitation of plants, as evidenced by the abundant grindstones. The fauna consisted only of wild species; caprines and ceramics were absent. The late Bashendi A phase (8.1–7.7 ka cal BP) saw, instead, the formation of very large sites. Among these, site 270, located at the end of the southeastern basin, testified to a noticeable increase in human presence, similarly to what has been described for the Hidden Valley village. It was located near an outcrop of sandstone and within an area of 300 by 150 meters that included about 200 huts. These were circular or rectangular hut floors with a clear fence wall that looked northward. The lithic industry included arrowheads, denticulates, notches typical of Bashendi A, while other types – bifacial foliates, tranchets, side-blow flakes – instead recalled Bashendi B.

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

211

Numerous other sites can be compared to Site 270 and show similar characteristics, although being much less rich. At the beginning of the 8th millennium cal BP, they attest to an intensification of human presence and the formation of large assemblages in the southeastern basin. At that time even some ceramic sherds decorated in the Saharan style were found, to which also remains of goat/sheep and perhaps also cattle can be added. In particular, around 7.8 ka cal BP (c. 6000 cal BC), a settlement pattern that does not show a precise distribution of sites appears in the oasis. For example, some were out of the playa extension, while others were within these; some were typical dry season sites, while others were of wet season, and so on (McDonald 2016). This heterogeneous model may indicate a decline in the seasonal occupation and, on the other hand, the emergence of a more stable settlement. Also in this case the change in the settlement distribution could be related to the phase in which rainfalls were evenly distributed throughout the year, thus allowing settlers to live for a long time in the same territory. Several hundred years later, from 7.4 ka cal BP, Bashendi B witnesses an increase in the pastoral activities, although the millstones and the numerous sickle-blades prove that the intensive collection of plants was still practiced. The following period, called the Sheikh Muftah phase, represents the beginning of a new episode that will last until the Egyptian Old Kingdom and during which the exchanges with the Nile Valley will be consolidated. Relations with the Nile Valley can be detected in the ceramic production, in the way of working the chert, in the maceheads and finally in the first appearance of copper. The settlement model was then decidedly oriented within the oasis but, in parallel, it became much more evanescent than it was in the previous phase, especially if compared to the late Bashendi A unit.

10.5 MOBILITY AND EXCHANGES IN THE LATE HOLOCENE: FROM DESERT TO NILE After 7.4 ka cal BP, an arid trend is established in the Western Desert (Hamdan and Lucarini 2013, 165), which leads to an abrupt change in the settlement. Around 7.0 ka cal BP, the sediments at El Bahr, characterized by white sub-angular chalk rubble grading vertically to the underlying bedrock, indicate severe mechanical weathering under sub-aerial conditions corresponding to Unit IV in the Hamdan’s palaeoclimatic sequence (Hamdan 2014b; Barich 2014c). Sites afterwards became even more scattered and, although stone slab structures continued to be present, next to them another type of stone structure – the steinplätze – much lighter and linked to short presences of groups in the area were widespread. This change in the settlement type and distribution became particularly relevant after 7.4 ka cal BP. These evidences testify for a logistic model divided between larger locations still containing slab structures and occasional encampments, used only for a few nights, scattered along the most common and frequented tracks (Figure 10.5). At the same time, the needs for water supply and grazing drove the groups to move to greater distances and develop a more organized form of herding. Middle Nile on the one side and the Northern region with the depressions of Siwa, Qattara and Fayum became the main destinations during late Holocene. After the Al Jerar cycle and the sensible deterioration of the environment, the Nabta Playa region recorded the presence of mobile pastoral groups such as the El Baqar (7.5–6.6 ka cal BP), followed by El Ansam (6.5–5.2 ka cal BP) units. To these groups we can attribute the establishment of a wide network of exchanges extended from the central Egypt to the Khartoum region (Kobusiewicz and Kabaciński 2010). The diffusion and circulation of the most popular artifacts are obviously indicative for the reconstruction of these movements. Particularly useful to this end are the bifacial items (knives, gouges, scrapers, drills, arrowheads) that exhibit a very high standard of workmanship. Such materials have appeared in stratigraphic settings at Farafra/Hidden Valley since at least 7.8 ka cal BP (Lucarini 2014b) (Figure 10.6), but their use certainly increased in the last part of Phase A and during Phase B of the oasis sequence. The exchange of high-standard manufactured tools may have been used as a form of intergroup mediation and conciliation in times of stress due to environmental difficulties (Barich and Lucarini 2014).

212

Climate Changes in the Holocene

FIGURE 10.5  Hearth places at Sheikh el Obeiyid, Northern Farafra Oasis. (Photo: B.E. Barich.)

FIGURE 10.6  Late Neolithic bifacial artifacts from the Hidden Valley Area, Farafra Oasis. (Image: B.E. Barich.)

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

213

Unlike the desert, the Nile Valley continued to enjoy wet weather even in the period following 6.3 ka cal BP. Unfortunately, the deposition of the sediments of the Nile, particularly strong in the phase between 6.8 and 6.1 ka cal BP (4900–4400/4100 cal BC), ended up obliterating and destroying most of the archaeological remains. A fortunate exception to this are the sites between Qena and Armant in Upper Egypt. The oldest complex, the Tarifian, is located on the left side of the Nile near Qurna, with dates of c. 6310 ± 80 (Gd-1756: 5525–5000 cal BC [Ginter and Kozlowski 1994]). In this regard, it is important to emphasize the correlation with the dry episode at the end of Phase B of the Farafra Sequence (Wadi el Obeiyid B: Barich and Lucarini 2014, 477) that would have increased movements from the desert territories towards the Upper Nile. According to Ginter and Kozlowski (1994), between 4400 and 4000 cal BC the whole region of the left bank of the Nile in the area of Qurna was still characterized by a very wooded and uninhabited savannah environment, almost a blank ready to be settled. The Tarifian is described as a pre-agricultural complex based on a broad spectrum organization and a settlement pattern comparable to that of Farafra. It is also important to remember that both the Tarifian, but particularly the Badarian which will flourish in the same region, present a rich bifacial stone production in which the prototypes already known in the Western Desert are clearly recognizable. Other movements certainly were directed towards the North, where the Fayum’s and Delta’s communities were flowering. Here, however, the contributions from the desert peoples had to intermingle with those coming from the southwest Levant. To shed light on the multiplicity of these exchanges, a strong consideration of the recently reassessed Fayum’s chronology is needed (Holdaway et al. 2017). The latest research in that geographical context has shown that the first irruptions in the Fayum indigenous sphere, at the end of the Qarunian (around 8.9 ka cal BP), seemingly came from the desert. However, over the following millennia the cultural materials in the most famous Fayumian sites (such as Kom K and Kom W) attest new arrivals, this time from the southwest Levant, that introduced a true Neolithic economy.

10.6 THE NEAR EAST INFLUENCES AND THE SPREAD OF THE LEVANTINE “NEOLITHIC PACKAGE” The general instability related to the 8.2 ka cal BP climatic crisis is responsible for a first break-in of elements of Near East origin in the autochthonous North African framework. In this case, the oldest element are the caprines (goat/sheep), the first among the ‘Neolithic’ resources of undisputed Near Eastern origin to enter the North African context. I can make here reference to the structures identified in the plain of El Qaa in the southwestern Sinai, facing onto the Red Sea (Close 2002, 462–466). Sites of the Jebel Qabiliat have circular structures, built using stone slabs placed vertically in the ground, which can be compared to those of the Sheikh el Obeiyid at Farafra. Remains of caprines collected inside the structures clearly indicate that they were left by herders. Even the dates of these structures – 7410 ± 75 (AA33717: 8333–8175 cal BP) and 6255 ± 60 (AA33715: 7264–7156 cal BP) – compared to those of similar findings in the Western Desert. On a chronological basis it is possible to reconstruct a route that descends from the Sinai along the coast of the Red Sea and then turns towards the western territories. Dating from Sodmein on the Red Sea, Farafra, Abu Tamsa and Haua Fteah, the latter on the Cyrenaica coast, can support this path that from the latest centuries of the 9th millennium cal BP brings the goat/sheep into the African scenario (Barich 2014e, 2016). New arrivals will follow over the later centuries and, as regards Lower Egypt in particular, will lead to the establishment of a Neolithic culture both in the Fayum and in the Delta. In the Near East, the first agricultural organization appears in the framework of the PPNA (10,200–8800 cal BC), whose origin can be placed in southeastern Anatolia or in southern Levant (Bar-Yosef and Meadow 1995). Animal domestication appears later, during the PPNB (8800–7000 cal BC) and at the same time in different regions, from the southern Levant to the Zagros. Goat is firstly domesticated around 8500 cal BC and is followed by sheep and later by cattle and pig (Zeder

214

Climate Changes in the Holocene

2011, 222; Linseele et al. 2014; Marshall et al. 2014). From their original areas, the first domesticates had a rapid diffusion towards the neighboring territories of the Fertile Crescent (Syria, Jordan, Lebanon, Palestine). These first experiments were consolidated into a true Neolithic culture over the course of at least two millennia. It is believed that the 8.2 ka cal BP arid event marks the disintegration and collapse of the PPNB horizon in the whole Fertile Crescent (Shirai 2006; Berger and Guilaine 2009; Fuller et al. 2012). The Near East underwent then successive instabilities that broke the balance formerly achieved. This phenomenon had repercussions throughout the whole Mediterranean (Berger and Guilaine 2009) but particularly in Egypt. The Nile River seems to have acted as a crossroads on which movements and exchanges of peoples from the desert, to Southwest, and the Levant to the East, converged. In an attempt to identify the contexts of origin, great weight is attributed to the comparisons that can be established within the Neolithic and Predynastic contexts of the Nile Valley. These comparisons can be mainly addressed toward the Pottery Neolithic contexts flourishing between 7.5 and 6.9 ka cal BP mainly in Israel and Negev countries. This, therefore, tells us that the relations between northern Africa and the territories of the Near East had a long tradition and that the routes that opened in the Mid-Holocene continued to be traveled many times over the following millennia.

10.7 CONCLUSION After the climate re-establishment that followed the arid Pleistocene phase, the eastern edges of the Sahara with the Egyptian Western Desert oases showed a progressive cultural development. In the Western Desert, the Early Holocene saw the flowering of various industries with a strong microlithic base. The Masara A horizon at Dakhla (McDonald 1991, 93); the Ain e-Raml at Farafra (Barich 2014a); the Siwan in the Siwa oasis (Hassan and Gross 1987); the Bedouin microlithic industry already described by Caton-Thompson in the Kharga oasis (Wendorf and Schild 1980, 189): they are all contexts that we can relate to small groups of hunter-gatherers with a mobile settlement model. At the same time, the history of human occupation in the Nubian area of the Egyptian Western Desert is linked to the recognition of an original management center for the local wild ox. In the Mid-Holocene, the Egyptian desert became the center of an important phenomenon of occupation, linked to the stability and availability of water. Original was the economic arrangement that prevailed then: a “diffuse” economic model that, alongside the traditional foraging activities, saw the beginnings of herding practice, mainly directed to goats. However, the climate instability caused frequent reorganization in the settlement model, with movements towards new areas. In the Late Holocene, these movements coincided with contemporary, similar phenomena in the Near East, producing in the most favorable areas of Lower Egypt a mixture of influences coming from various directions. However, in no way should these arrivals be represented as invasions or replacement of people but, more likely, as influences that were integrated and re-elaborated within the local well-established context.

ACKNOWLEDGMENTS Data relating to the Farafra Oasis were collected as part of the Italian Archaeological Project in the Farafra Oasis, formerly carried out in the context of the Rome Sapienza University and now of the ISMEO, the International Association for Mediterranean and Oriental Studies. The project is codirected by Barbara E. Barich and Giulio Lucarini and benefits from grants from the Italian Ministry of Foreign Affairs and ISMEO. The Author intends to thank the Editor of the present Handbook, Eustathios Chiotis, for the kind invitation to participate in this undertaking.

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

215

REFERENCES Arz, H.W., F. Lamy, J. Pätzold, P.J. Müller, and M. Prins. 2003. Mediterranean moisture source for an EarlyHolocene humid period in the Northern Red Sea. Science 300: 118–121. Bar-Yosef, O., and R.H. Meadow.1995. The origins of agriculture in the Near East. In Last Hunters, First Farmers: New Perspectives on the Prehistoric Transition to Agriculture, edited by D.T. Price and A.B. Gebauer, 39–94. Santa Fe: School of American Research Press. Barich, B.E. 2014a. Early to mid Holocene archaeology of the Egyptian Western Desert. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 39–53. Firenze: All’Insegna del Giglio. Barich, B.E. 2014b. Hidden Valley: a 7000 year-old village in Wadi el Obeiyid. In From Lake to Sand— The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 167–207. Firenze: All’Insegna del Giglio. Barich, B.E. 2014c. The El-Bahr/Wadi el Obeiyid playa: its general setting and the main archaeological features. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 101–128. Firenze: All’Insegna del Giglio. Barich, B.E. 2014d. The Wadi el Obeiyd Cave 1: The Rock Art archive. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 385–405. Firenze: All’Insegna del Giglio. Barich, B.E. 2014e. Northwest Libya from the early to late Holocene: New data on environment and subsistence from the Jebel Gharbi. Quaternary International 320(1–2): 15–27. doi:10.1016/j.quaint.2015.11.138. Barich, B.E., and F.A. Hassan. 1984–87. The Farafra Oasis Archaeological Project (Western Desert, Egypt) 1987 Field Campaign. Origini XIII: 117–191. Barich, B.E., and G. Lucarini. 2014. Social dynamics in northern Farafra from the middle to late Holocene: changing life under uncertainty. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 467–484. Firenze: All’Insegna del Giglio. Barich, B.E., G. Lucarini, M. Gallinaro, and M. Hamdan. 2012. Sheikh/Bir El Obeiyid: Evidence of Sedentism in the Northern Farafra Depression (Western Desert, Egypt). In Prehistory of Northeastern Africa. New Ideas and Discoveries, edited by J. Kabaciński, M. Chlodnicki, and M. Kobusiewicz, 255–278. Poznań: Poznań Archaeological Museum. Barich, B.E., G. Lucarini, M.A. Hamdan, and F.A. Hassan, eds. 2014. From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt. Firenze: All’Insegna del Giglio. Berger, J.-F., and J. Guilaine. 2009. The 8200 cal. BP abrupt environmental change and the Neolithic transition: A Mediterranean perspective. Quaternary International 200: 31–49. doi:10.1016/j.quaint.2008.05.013. Bradley, D.G., D.E. MacHugh, P. Cunningham, and R.T. Loftus. 1996. Mitochondrial diversity and the origins of African and European Cattle. Proceedings of the National Academy of Sciences, USA 93(10): 5131–5135. Brookes, I.A. 1989. Early Holocene basinal sediments of the Dakhleh Oasis region, south central Egypt. Quaternary Research 32: 139–152. Brookes, I.A. 1993. Late Pleistocene basinal sediments, Dakhla Oasis region, Egypt: A non interglacial pluvial. In Geoscientific Research in Northeast Africa: Proceedings of the International Conference, edited by U. Thorweihe and H. Schandelmeier, 627–633. Rotterdam: Balkema. Cancellieri, E., M. Cremaschi, A. Zerboni, and S. di Lernia. 2016. Climate, environment, and population dynamics in Pleistocene Sahara. In Africa from MIS 6–2: Population Dynamics and Paleoenvironments, edited by S.C. Jones and B.A. Stewart, 123–145. Vertebrate Paleobiology and Paleoanthropology series. Dordrecht: Springer. Carto, S.L., A.J. Weaver, R. Hetherington, Y. Lam, and E.C. Wiebe. 2009. Out of Africa and into an ice age: on the role of global climate change in the late Pleistocene migration of early modern humans out of Africa. Journal of Human Evolution 56: 139–151. Churcher C.S., M.R. Kleindienst, and H.P. Schwarcz. 1999. Faunal remains from a middle Pleistocene lacustrine marl in Dakhleh Oasis, Egypt: palaeoenvironmental reconstructions. Palaeogeography Palaeoclimatology Palaeoecology 154: 301–312. Clark, J.D. 1976. The domestication process in sub-Saharan Africa with special reference to Ethiopia. In Origine de l’ élevage et de la domestication, edited by E. Higgs, 56–115. Nice: Union Internationale Sciences Préhistoriques et Protohistoriques, Colloque XX – IXème Congrès Nice 1976. Close, A.E. ed. 1987. Prehistory of Arid North Africa. Essays in Honour of Fred Wendorf. Dallas: Southern Methodist University Press.

216

Climate Changes in the Holocene

Close, A. E. 1995. Few and far between: Early ceramics in North Africa. In The Emergence of Pottery: Technology and Innovation in Ancient Societies, edited by W.K. Barnett, 23–37. Washington, DC: Smithsonian Institution Press. Close, A.E. 2002. Sinai, Sahara, Sahel: The Introduction of Domestic Caprines to Africa. In Tides of the Desert – Gezeiten der Wüste, Contributions to the Archaeology and Environmental History of Africa in Honour of Rudolph Kuper, edited by Jennerstrasse 8, 459–469. Köln: Heinrich Barth Institut. Cohen, A.S., J.R. Stone, K.R.M. Beuning, L.E. Park, P.N. Reinthal, D. Dettman, C.A. Scholz, et al. 2007. Ecological consequences of early Late Pleistocene megadroughts in tropical Africa. Proceedings of the National Academy of Sciences 104: 16422–16427. Decker, J.E., S.D. McKay, M.M. Rolf, J. Kim, A. Molina Alcalá, T.S. Sonstegard, O. Hanotte, et al. 2014. Worldwide patterns of ancestry, divergence, and admixture in domesticated cattle. PLoS Genet 10(3): e1004254. doi:10.1371/journal.pgen.1004254. De Menocal, P.B., and J.E. Tierney. 2012. Green Sahara: African humid periods paced by earth’s orbital changes. Nature Education Knowledge 3(10): 12–18. Fahmy, A.G. 2014. Plant food resources at Hidden Valley, Farafra Oasis. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 333–344. Firenze: All’Insegna del Giglio. Fuller, D.Q., G. Willcox, and R.G. Allaby. 2012. Early agricultural pathways: Moving outside the ‘core area’ hypothesis in Southwest Asia. Journal of Experimental Botany 63(2): 617–633. Gasse, F. 2000. Hydrological changes in the African tropics since the Last Glacial Maximum. Quaternary Science Reviews 19: 189–211. Gautier, A. 1984. Archaeozoology of the Bir Kiseiba region, Eastern Sahara. In Cattle-Keepers of the Eastern Sahara, assembled by F. Wendorf and R. Schild; edited by A.E. Close, 49–72. Dallas: Department of Anthropology, Institute for the Study of Earth and Man, Southern Methodist University. Gautier, A. 2001. The early to late Neolithic archeofaunas from Nabta and Bir Kiseiba. In Holocene Settlement of the Egyptian Sahara. Volume 1: The Archaeology of Nabta Playa, edited by Wendorf, F., R. Schild, and Associates, 609–635, New York: Kluwer Academic/Plenum Publishers. Gehlen, B., K. Kindermann, J. Linstädter, and H. Riemer.2002. The Holocene occupation of the Eastern Sahara. In Tides of the Desert—Gezeiten der Wüste, Contributions to the Archaeology and Environmental History of Africa in Honour of Rudolph Kuper, edited by Jennerstrasse 8, 85–116. Köln: Heinrich Barth Institut. Ginter, B., and J.K. Kozlowski.1994. The Tarifian and the origin of the Naqadian. In Origin and Early Development of Food-Producing Cultures in North-Eastern Africa, edited by L. Krzyńaniak, and M. Kobusiewicz, 247–260. Poznań: Poznań Archaeological Museum. Giraudi, C. 2005. Eolian sand in the peridesert northwestern Libya and implications for Late Pleistocene and Holocene Sahara expansion. Palaeogeography Palaeoclimatology Palaeoecology 218 (1–2): 161–173. Hamdan, M.A. 2014a. Geology of the Holocene playa sediments of Hidden Valley, Wadi el Obeiyid, Farafra. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 151–165. Firenze: All’Insegna del Giglio. Hamdan, M.A. 2014b. Sedimentological characteristics and geomorphic evolution of the Holocene playa of Wadi el Obeiyid. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 151–165. Firenze: All’Insegna del Giglio. Hamdan, M.A., F.A. Hassan, and A.M. Mahmoud, 2014. The Wadi el Obeiyid Cave 1 – Geological features. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 377–384. Firenze: All’Insegna del Giglio. Hamdan, M.A., and G. Lucarini. 2013. Holocene paleoenvironmental, paleoclimatic and geoarchaeological significance of the Sheikh El-Obeiyid area (Farafra Oasis, Egypt). Quaternary International 302: 154–168. Hanotte, O., D.G. Bradley, J.W. Ochieng, Y. Verjee, E.W. Hill, and J.E.O. Rege.2002. African pastoralism: Genetic imprints of origins and migrations. Science 296: 336–339. Hassan, F.A., and G.T. Gross. 1987. Resources and subsistence during the Early Holocene at Siwa Oasis, Northern Egypt. In Prehistory of Arid North Africa. Essays in Honour of Fred Wendorf, edited by A.E. Close, 85–103, Dallas: Southern Methodist University Press. Hassan, F.A., M.A. Hamdan, and A.A. Mahmoud. 2014. Desert and oasis: geomorphology and geomorphic evolution. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 63–78. Firenze: All’Insegna del Giglio.

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

217

Haynes, C.V. Jr. 2001. Geochronology and climate change of the Pleistocene transition in the Darb el Arba’in desert, Eastern Sahara. Geoarchaeology: An International Journal 16(1): 119–141. Haynes, C.V. Jr., C.H. Eyles, L. Pavlish, J.C. Ritchie, and M. Rybak.1989. Holocene palaeoecology of the Eastern Sahara: Selima Oasis. Quaternary Science Reviews 8: 109–136. Hendrickx, S., F. Förster, and M. Eyckerman. 2013. The Pharaonic pottery of the Abu Ballas Trail: ‘Filling stations’ along a desert highway. In Desert Road Archaeology, edited by F. Förster and H. Riemer, 339–379. Koln: Africa Praehistorica 27- Heinrich Barth Institut. Hoelzmann, P., B. Keding, H. Berke, S. Kroepelin, and H.J. Kruse. 2001. Environmental change and Archaeology: Lake evolution and human occupation in the Eastern Sahara during the Holocene. Palaeogeography Paleaeoclimatology Palaeoecology 169:193–217. Holdaway, S.J., and W. Wendrich eds.2017. The Desert Fayum Reinvestigated—The Early to Mid-Holocene Landscape Archaeology of the Fayum North Shore, Egypt. Monumenta Archaeologica 39. Los Angeles: UCLA Cotsen Institute of Archaeology Press. Klees, F. 1989. Lobo: a contribution to the prehistory of the eastern Sand Sea and the Egyptian oases. In Late Prehistory of the Nile Basin and the Sahara, edited by L. Krzyzaniak and M. Kobusiewicz, 223–231. Poznan: Poznan Archaeological Museum. Kobusiewicz, M., and J. Kabaciński. 2010. Conclusions. In Gebel Ramlah: Final Neolithic Cemeteries from the Western Desert of Egypt, edited by M. Kobusiewicz, J. Kabacinski, R. Schild, J. D. Irish, M. C. Gatto, and F. Wendorf, 249–258. Poznań: Institute of Archaeology and Ethnology, Polish Academy of Sciences. Kuper, R. 1981, Untersuchungen zur Besiedlungsgeschichte der östlichen Sahara. Beitrage zur Allgemeinen und Vergleichenden Archäologie 3: 215–275. Kuper, R., and S. Kropelin. 2006. Climate-controlled Holocene occupation in the Sahara: Motor of Africa’s evolution. Science 313: 803–807. Leroux, M. 1991. Paléométérologie de la region de Taoudenni. In Paléoenvironnements du Sahara. Lacs holocènes à Taoudenni (Mali), edited by N. Petit-Maire, 197–203. Paris: CNRS. Leroux, M. 1993. The mobile polar high: A new concept explaining present mechanisms of meridional airmass and energy exchanges and global propagation of palaeoclimatic changes. Global and Planetary Change 7(1–3): 69–93. doi:10.1016/0921-8181(93)90041-L. Linseele, V., W. Van Neer, S. Thys, R. Phillipps, R. Cappers, W. Wendrich, and S. Holdaway. 2014. New archaeozoological data from the Fayum neolithic with a critical assessment of the evidence for early stock keeping in Egypt. Plos One 9(10): e108517. doi:10.1371/journal.pone.0108517. Lucarini, G. 2014a. Exploitation and management of wild grasses at Hidden Valley, Farafra Oasis. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 345–367. Firenze: All’Insegna del Giglio. Lucarini, G. 2014b. The bifacial products from Hidden Valley and neighbouring areas in Wadi el Obeiyid. In From Lake to Sand—The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 265–283. Firenze: All’Insegna del Giglio. Magee, D., H. Mannen, and D.G. Bradley. 2007. Duality in Bos indicus mtDNA diversity: Support for geographical complexity in zebu domestication. In Interdisciplinary Studies in Archaeology, Biological Anthropology, Linguistics and Genetics, edited by M.D. Petraglia and B.A. Allchin, 385–391. New York: Springer. Marshall, F.B., K. Dobney, T. Denham, and J.M. Capriles. 2014. Evaluating the roles of directed breeding and gene flow in animal domestication. PNAS Early Edition: 1–6. doi:10.1073/pnas.1312984110. Marshall, F., and L. Weissbrod. 2011. Domestication processes and morphological change. Through the lens of the donkey and African pastoralism. Current Anthropology 52, Supplement 4: S397–S413. doi:10.1086/658389. McDonald, M.M.A. 1991. Origins of the Neolithic in the Nile Valley as seen from Dakhleh Oasis in the Egyptian Western Desert. Sahara 4: 41–52. McDonald, M.M.A. 1999. Neolithic cultural units and adaptations in the Dakhleh Oasis. In Reports from the Survey of the Dakhleh Oasis Western Desert of Egypt 1977–1987, edited by C.S. Churcher and A.J. Mills, 117–132. Oxford: Oxbow Monograph 99. McDonald, M.M.A. 2002. Dakhleh oasis in predynastic and early dynastic times: Bashendi B and the Sheikh Muftah cultural units. In Actualité de la Recherche Prédynastique: les Terrains de fouilles, edited by S. Hendrick and B. Midant-Reynes, Archéo-Nil 12: 109–119. McDonald, M.M.A. 2006. Holocene prehistory of the Wadi el Midauwara above Kharga Oasis, Egypt. In Archaeology of Early Northeastern Africa. In Memory of Lech Krzyźaniak, edited by K. Kroeper, M. Chłodnicki, and M. Kobusiewicz, 479–492. Poznan: Poznan Archaeological Museum.

218

Climate Changes in the Holocene

McDonald, M.M.A. 2009. Increased sedentism in the central oases of the Egyptian Western Desert in the early to mid-Holocene: Evidence from the peripheries. African Archaeological Review 26: 3–43. McDonald, M.M.A. 2016. The pattern of Neolithization in Dakhleh Oasis in the Eastern Sahara.Quaternary International 410 (Part A): 181–197. Nicoll, K. 2004. Recent environmental change and prehistoric human activity in Egypt and Northern Sudan. Quaternary Science Reviews 23: 561–580. Nicoll, K. 2018. A revised chronology for Pleistocene paleolakes and middle stone age—Middle Paleolithic cultural activity at Bîr Tirfawi - Bîr Sahara in the Egyptian Sahara. Quaternary International 463: 18–28. Reimer, P.J., E. Bard, A. Bayliss, J.W. Beck, P.G. Blackwell, C.B. Ramsey, C.E. Buck, et al. 2013. Intcal 13 and Marine13 radiocarbon age calibration curves 0-50,000 Years cal BP. Radiocarbon 55 (4): 1869–1887. Riemer, H. 2006. Out of Dakhla: Cultural diversity and mobility between the Egyptian Oases and the Great Sand Sea during the Holocene humid phase. In Archaeology of Early Northeastern Africa. In Memory of Lech Krzyźaniak, edited by K. Kroeper, M. Chłodnicki, and M. Kobusiewicz, 493–526. Poznan: Poznan Archaeological Museum. Scholz, C.A., T.C. Johnson, A.S. Cohen, J.W. King, J.A. Peck, J.T. Overpeck, M.R. Talbot, et al. 2007. East African megadroughts between 135 and 75 thousand years ago and bearing on early-modern human origins. Proceedings of the National Academy of Sciences 104: 16416–16421. Smith, A.B. 2005. African Herders: Emergence of Pastoral Tradition. Walnut Creek: AltaMira Press. Shirai, N. 2006. Origins and development of bifacial stone tools and their implications for the beginning of animal herding in the Egyptian Western Desert. In Archaeology of Early Northeastern Africa. In Memory of Lech Krzyźaniak, edited by K. Kroeper, M. Chłodnicki, and M. Kobusiewicz, 355–374. Poznan: Poznan Archaeological Museum. Smith, J.R. 2010. Palaeoenvironments of eastern North Africa and the Levant in the late Pleistocene. In SouthEastern Mediterranean Peoples Between 130,000 and 10,000 Years Ago, edited by E.A.A. Garcea, 6–17. Oxford and Oakville: Oxbow Books. Smith, J.R., R. Giegengack, H.P. Schwarcz, M.M.A. McDonald, M.R. Kleindienst, A.L. Hawkins, and C.S. Churcher. 2004. A reconstruction of Quaternary pluvial environments and human occupations using stratigraphy and geochronology of fossil-spring tufas, Kharga Oasis. Egypt. Geoarchaeology 19: 407–439. Smith, J.R., A.L. Hawkins, Y. Asmerom, Y. Polyak, and R. Giegengack. 2007. New age constraints on the Middle Stone Age occupations of Kharga Oasis, Western Desert, Egypt. Journal of Human Evolution 52: 690–701. Stock, F., and D. Gifford-Gonzales. 2013. Genetics and African Cattle Domestication. African Archaeological Review 30: 51–72. Tubiana, M.J., and J. Tubiana. 1977. The Zaghawa from an Ecological Perspective. Rotterdam: A.A. Balkema. Van Peer, Ph. 2014. Hints at Middle Stone Age occupation in the Farafra Oasis. In From Lake to Sand— The Archaeology of Farafra Oasis, Western Desert, Egypt, edited by B.E. Barich, G. Lucarini, M.A. Hamdan, and F.A. Hassan, 25–38. Firenze: All’Insegna del Giglio. Vernet, R. 1995. Les paleoenvironnements du Nord de l’Afrique depuis 600.000 ans. Dossier et Recherches sur l’Afrique 3:149–271. Meudon: Centre National de la Recherche Scientifique. Wasylikowa, K. 2001. Site E-75–6: Vegetation and subsistence of the Early Neolithic at Nabta Playa, Egypt, reconstructed from charred plant remains. In Holocene Settlement of the Egyptian Sahara, Volume 1, The Archaeology of Nabta Playa, edited by F. Wendorf, R. Schild, and Associates, 544–591. New York: Kluwer Academic/Plenum Publishers. Wendorf, F., and R. Schild.1980. Prehistory of the Eastern Sahara. New York: Academic Press. Wendorf, F., and R. Schild. 1994. Are the Early Holocene cattle in the Eastern Sahara domestic or wild?. Evolutionary Anthropology 1994: 118–128. Wendorf, F., and R. Schild. 2001. Conclusions. In Holocene Settlement of the Egyptian Sahara, Volume 1, The Archaeology of Nabta Playa, edited by F. Wendorf, R. Schild, and Associates, 648–675. New York: Kluwer Academic/Plenum Publishers. Wendorf, F., and R. Schild. 2002. The Role of Storage in the Neolithic of Egyptian Sahara. In Tides of the Desert—Gezeiten der Wüste, Contributions to the Archaeology and Environmental History of Africa in Honour of Rudolph Kuper, edited by Jennerstrasse 8, 41–49. Köln: Heinrich Barth Institut. Wendorf, F., R. Schild, and Associates. 2001. Holocene Settlement of the Egyptian Sahara, vol. I: The archaeology of Nabta Playa. New York: Kluwer Academic/Plenum Publishers. Wendorf, F., R. Schild, A. Applegate, and A. Gautier. 1997. Tumuli, cattle burials and society in the Eastern Sahara. In Dynamics of Populations, Movements and Responses to Climatic Change in Africa, edited by B.E. Barich and M.C. Gatto, 90–104. Roma: Forum for African Archaeology and Cultural Heritage.

Eastern Borders of Sahara and Relations with Nile Valley and Beyond

219

Wendorf, F., R. Schild, and A.E. Close, eds. 1993. Egypt During the Last Interglacial: The Middle Paleolithic of Bir Tarfawi and Bir Sahara East. New York: Plenum Press. Wendorf, F., and R. Schild, assemblers, and A.E. Close, ed.1984. Cattle-Keepers of the Eastern Sahara, Dallas: Department of Anthropology, Institute for the Study of Earth and Man, Southern Methodist University. Williams, M.A.J., and D.A. Adamson. 1980. Late Quaternary depositional history of the Blue and White Nile rivers in Central Sudan. In The Sahara and the Nile, edited by M.A.J. Williams and H. Faure, 281–304. Rotterdam: A.A. Balkema. Zboray, A. 2013. Prehistoric trails in the environs of Karkur Talh, Jebel Uweinat. In Desert Road Archaeology, edited by F. Förster and H. Riemer, 339–379. Koln: Africa Praehistorica 27- Heinrich Barth Institut. Zeder, M. A. 2011. The origins of agriculture in the Near East. Current Anthropology 52, Supplement 4: S221–S235. Zubrow, E.B.W. 1975. Prehistoric Carrying Capacity: A Model. Menlo Park: Cunnings.

11 The Role of Hydraulic Technologies Human Adaptation in Arabia Julien Charbonnier CONTENTS 11.1 Introduction........................................................................................................................... 221 11.2 A Land of Starvation?............................................................................................................ 222 11.3 From Green Arabia to the Desert: Environmental Change in the Holocene........................ 223 11.4 The Prehistory of Water Management at the Northern End of Arabia................................. 223 11.5 From Pastoralism to Agriculture: Northwest Arabia During the 6th–5th Millennium B.C. ....... 226 11.6 Early Farming Communities: The Bronze Age..................................................................... 227 11.6.1 The Genesis of Floodwater Harvesting in Southwest Arabia.................................... 227 11.6.2 Early Water Management in Southeast Arabia......................................................... 229 11.7 Blooming of Hydraulic Techniques in Arabia: The Iron Age............................................... 231 11.7.1 The Oases of Northern Arabia During the Iron Age................................................. 232 11.7.2 Southeast Arabia During the Iron Age: Groundwater Exploitation Through Wells and Aflāj.......................................................................................................... 232 11.7.3 Southeast Arabia During the Iron Age: Runoff Harvesting in the Oasis of Masāfī........ 236 11.7.4 The Harnessing of Floodwaters in the South Arabian Kingdoms............................ 237 11.8 Discussion: Phasing the Development of Water Management in Arabia.............................. 239 11.9 Conclusion.............................................................................................................................240 References....................................................................................................................................... 241

11.1 INTRODUCTION During the Holocene, human populations inhabiting Arabia have followed a singular path, which differs in many ways from the one pursued by the neighboring regions of the Near East and Mesopotamia. While during the first half of Holocene farming communities developed in the Fertile Crescent, which extends from the Levant to the Taurus and Zagros mountains, Arabian populations adopted agriculture much later, probably during the Late Chalcolithic (4th millennium B.C.) in Northwest Arabia and at the beginning of the Bronze Age in Southwest Arabia (end of the 4th/beginning of the 3rd millennium B.C.) and Southeast Arabia (3rd millennium B.C.) (Bouchaud, Dabrowski, and Tengberg 2016; Tengberg 2003; Ekstrom and Edens 2003; Charbonnier 2008). Before that date, between the 6th and the 4th millennium B.C., the Arabian Peninsula was occupied by mobile hunter-pastoralists. These groups were hunting and herding sheep, goat and cattle in a savanna-like environment, which existed thanks to the Holocene moist phase (see below). Along the eastern coasts, groups specialized in fishing, gathering shells and exploiting mangroves (Biagi and Nisbet 2006; Cleuziou 2005: 134). As far as we know, these populations never developed any hydraulic technologies, apart from the Northwest of Arabia during the 5th millennium B.C., as we shall see. In biology, the term ‘adaptation’ refers to the dynamic evolutionary process that fits organisms to their environment as well as the state reached by the population during that process. Can we apply similar notions to the human societies? As we will see, the environment of Arabia has changed 221

222

Climate Changes in the Holocene

dramatically during the Holocene, with the incremental development of arid conditions to which populations have had to adapt. They did, in many different ways. It is beyond the scope of this chapter to present all the aspects of this process. Instead, I will try to demonstrate that Arabian populations were able to adapt over time to their changing environment by developing hydraulic technologies that allowed them to take the best advantage possible of available water resources. Regarding the second half of the Holocene, I will also argue that population adapted their environment in order to make it more favorable and suitable to their needs. In some ways, it is a sort of Human Niche Construction process (Kendal, Tehrani, and Odling-Smee 2011). In order to achieve my objectives, I will review early archaeological evidence of water management systems in Arabia, from the Neolithic to the end of the Iron Age (9th millennium B.C. to the end of the 1st millennium B.C.). I will focus on three regions for which we have the relevant data: Northwest Arabia (Northern Saudi Arabia and Southern Jordan), Southwest Arabia (Yemen and Southwestern Oman) and Southeast Arabia (Northern Oman and the United Arab Emirates). The hydraulic systems will be presented chronologically by distinguishing the Pre-Pottery Neolithic (PPN), the Chalcolithic period, the Bronze Age and the Iron Age. First of all, we shall expose the geography of Arabia and the state of our knowledge on climate change during the Holocene.

11.2 A LAND OF STARVATION? Arabia appears today as a dry place. In the Western collective imagination, it is a place of sandy desert punctuated by oases (water and date palms) and hostile rocky mountains. It is actually a more diverse and richer environment. The Arabian Peninsula, which is part of the Arabian-Nubian Shield, extends about 2000 km north-south, and up to about the same distance east-west, and it covers an area of 3.5 million sq. km (Figure 11.1) (Sanlaville 2000: 117). The peninsula is crossed by two ergs (i.e. sandy deserts), the Nafūd to the north (c. 100,000 sq. km) and the Rub‘al-Khālī to the south (c. 600,000 sq. km), separated by a zone of cuestas (i.e. low ridge with a steep slope and a gentle back slope formed by the differential erosion of strata of different hardness). While

FIGURE 11.1  Physical map of the Arabian Peninsula ((c) Charbonnier 2018).

Human Adaptation in Arabia

223

these regions are arid to hyper-arid, wetter areas exist. To the west, mountain ranges located along the Red Sea, from the Gulf of Aqaba to the Gulf of Aden, benefit from higher precipitations. In Southwest Arabia, these mountain chains are extended by the plateaus of H ̣ad ̣ramawt and Dhofār. In Southeast Arabia, the mountains of al-H ̣ajar, which follow the Gulf of Hormuz, also benefit from higher rainfalls. In the north and the southeast of the peninsula, winter rainfalls are related to the Mediterranean climatic system (Sanlaville 2000: 117; Fleitmann et al. 2007: 173). The later can also bring moderate rains in winter in Southwest Arabia, but in this region, the precipitations are mainly related to the monsoon and occur in summer (July–August). Additionally, the Red Sea convergence zone contributes to a shorter rainy season in spring (April–May). However, despite the occurrence of rainfall events, the Arabian Peninsula is characterized by the absence of permanent water streams. This is the major difference with the Near East.

11.3 FROM GREEN ARABIA TO THE DESERT: ENVIRONMENTAL CHANGE IN THE HOLOCENE During the Holocene, Arabia has experienced major climatic changes. We have several palaeoenvironmental records at our disposal for the early-mid Holocene period (until the 5th millennium B.C.), which include speleothems and lacustrine archives. They indicate that this period corresponded to a moist phase, the Early Holocene climatic optimum, related to monsoonal incursions (Fleitmann, Burns, Neff, et al. 2003; Fleitmann et al. 2007; Preston et al. 2015). Wetter climatic conditions mean higher rainfall, more vegetation and probably more permanent watercourses than today (Fleitmann and Matter 2009: 640), resulting in the development of lakes and a savannah-like landscape in some areas, such as Northwest Arabia (Parker et al. 2006; Parker et al. 2004; Engel et al. 2012; Engel et al. 2017). Starting from the end of the 5th/beginning of the 4th millennium B.C., palaeoenvironmental records indicate an increase in aridity across much of Arabia. It seems to be the consequence of a decline in summer rainfall following the southward retreat of the Intertropical Convergence Zone (ITCZ). With the retreat of the monsoon towards the south, summer rains ceased to reach North- and Southeast Arabia from the 5th millennium B.C. (Fleitmann and Matter 2009: 640); peaks of aridity then occurred at the beginning and end of the following millennium. The 4th millennium B.C. corresponded to a slightly moister period that was interrupted by an arid event around 3200–3000 B.C. The climate became slightly wetter again but remained arid during the 2nd millennium B.C. (Parker et al. 2006: 472–473). Aridity increased around 1000 B.C.: during the Iron Age, the climate was as arid as it is today (Fleitmann et al. 2007: 180). Regarding Southern Arabia, which still benefited from the monsoon, speleothems from Qunf Cave (Dhofār, Southwestern Oman) indicate the last climatic optimum lasted until around the 6th century B.C. and was followed by hyper-arid conditions between 600 B.C. and 1550 A.D. (Fleitmann, Burns, Mudelsee, et al. 2003).

11.4 THE PREHISTORY OF WATER MANAGEMENT AT THE NORTHERN END OF ARABIA Water is the most precious resource to humankind, both for consumption and for agriculture in arid regions. The Neolithic is a turning point in the history of water management. More than ever before, people had to rely on water. This was due to the increased sedentariness of populations, demographic growth and the development of farming. As soon as population settled, in the Neolithic, it was necessary to both get and store water. Potteries could be used to carry and store water in the villages near springs, rivers and lakes. When surface water resources were only temporarily available, however, populations had to developed hydraulic techniques to harvest them or to exploit groundwater. The oldest wells were found in

224

Climate Changes in the Holocene

Pre-Pottery Neolithic villages in Cyprus and the Levant and date back from the 9th–8th millennium B.C. (Peltenburg et al. 2000: 846–848; Galili and Nir 1993). These structures were probably used for human and animal consumption, as they were found inside villages. Regarding Arabia, the oldest evidence of hydraulic activity was found in the Jordan desert, at the northern edge of the peninsula (Figure 11.2). In the Jafr basin, a series of so-called ‘barrages’ were investigated by a Japanese project. Although this is still debated, these structures seem to be dated from the Pre-Pottery Neolithic B (PPNB, c. 9000–7000 B.C.). This region is now hyper-arid. In the Early Holocene, however, as rainfalls were more abundant, the Jafr basin may have been wetter. However, it seems the rainfalls were not strong and/or reliable enough to allow dry-farming, at least not every year. In Wādī Abu Tulayha, three successive water-management walls (named ‘barrages’ by Fujii) were found along a small wādī. Wādī Abu Tulayha is a small settlement, about 1.5 hectares, made of semi-subterranean houses. The economy of the site was probably oriented toward more agriculture than animal breeding, as no evidence of animal pens were found, while evidence of agricultural tools (grinding stones, flint sickles) and cultivated crops (cereal and pulse seeds) were collected in

FIGURE 11.2  Map of Northeast Arabia with hydraulic structures mentioned in the text. ((c) Charbonnier 2018).

Human Adaptation in Arabia

225

the nearby PPNB settlement. According to Fujii, these PPNB settlements were seasonally occupied by transhumant populations. The largest and upstream-most one, Barrage 1, is located close to a PPNB settlement. It is a V-shaped masonry structure, with a length of 120 m (Fujii 2011: 15–17). The proposed PPNB dating for this structure is based on stratigraphic correlation as well as architectural similarities with the PPNB site, which is, in addition, the only settlement in the vicinity (Fujii 2011: 17). Fujii interprets Barrage 1 as a ‘water catchment facility’ and hypothesizes that it was used for basin-irrigation; that is, it would have been built to retain runoff water in order to facilitate its infiltration and then the growing of crops directly in the reservoir. The flatness of the location and the permeability of the ground would be incompatible with water storage over a long period of time. The area that would have been flooded by this retention is estimated to be a few hectares. Additionally, it is suggested that the reservoir could have been used as a pasture land during dry years. Two other walls (Barrage 2 and 3) built across the same subsidiary wādī were identified about 200 and 250 m downstream of Barrage 1. They have been interpreted as retention dams dedicated to the storage of drinking water for humans and animals due to the fact that they were built on a rocky impermeable ground as well as their small scale compared to Barrage 1. A PPNB date is proposed on the basis of the fact the PPNB site is the only settlement in the vicinity and on the presence of a semi-circular reinforcement wall in the center of the wall, which recalls the one visible on Barrage 1 (Fujii 2011: 17). Two more walls similar to Barrage 1 of Wādī Abu Tulayha were investigated at Wādī Ruweishid al-Sharqi. These long masonry walls seem to have been used to support field systems, as they were located in a flat permeable zone. A PPNB dating is proposed based on architectural similarities with Wādī Abu Tulayha dams and settlements. Two similar masonry walls were identified in Wādī Ghuwayr, located about 70 km east of Wādī Abu Tulayha. Both of them are incurved and constructed at the lowest end of a semi-open playa that they closed off. The uppermost structure, ‘Barrage 1’ was 72 m long and ‘Barrage 2’, located 130 m downstream, was 74 m long. Their plan and architecture seem to suggest that they were used for basin-irrigation, like Barrage 1 of Wādī Abu Tulayba. No settlement was found in proximity to Wādī Ghuwayr’s structures, and the PPNB dating proposed by Fujii is based on the presence of a protruded reinforcement in the central part of Barrage 1, similar to the one in Wādī Abu Tulayba Barrages 1 and 3, as well as, again, architectural similarities (Fujii 2011: 28–29). In addition to the dam systems, a cistern was found near the settlement of Wādī Abu Tulayha (Structure M). This cistern was 18 m long, made of several rooms and was semi-subterranean, partly cut in the bedrock and partly stone-built. Radiocarbon analyses have allowed dating it from the Middle/ Late-PPNB (Fujii 2011: 26). It was probably used to store runoff waters. As already noted by Finlayson et al., a strong argument for dams of Wādī Abu Tulayha is the fact that findings from the site indicate an agriculture-oriented economy and not pastoralism. Indeed, given the fact that rainfalls were not important enough to sustain dry-farming, at least on a regular basis, the implication is that the local population had to manage water (Finlayson et al. 2011: 201– 203). Absolute dating would, however, be needed to safely date these dams from the Pre-Pottery Neolithic. According to Fujii, the use of dams was related to the fact that rainfall was too scant to allow dry-farming in the area. One could also argue, in absence of detailed enough data for the area in the Neolithic, that this structure was used for enhancing the productivity of fields or to balance the unreliability of rainfall. Furthermore, similar terrace walls built in wādī beds were used in Southern Arabia where they are named h. arra (Wilkinson refers to them as cross-valley walls, Wilkinson 2003: 192–193). Interestingly, these walls are used to retain not only moisture but also sediments. Indeed, soils are as rare as water in an arid environment, so h. arra-s are therefore used to create soil for cultivation and also to prevent erosion (Wilkinson 2003: 193). In Yemen, the oldest structures of this kind are dated from the Himyarite period (1st century B.C.–6th century A.D.; Wilkinson and Edens 1999: 10–11).

226

Climate Changes in the Holocene

Such a technology is quite simple in its principles, and I assume that there is no need to search for a direct relationship between the structures used in the mountains of Yemen and the walls of Wādī Abu Tulayha. They reflect more probably comparable responses to similar environmental constraints. If the datings are confirmed, the dams and cistern found in the Jafr basin would show that human populations were able to manipulate surface water in order to settle in arid environments at an early stage, since the first half of the Holocene (Finlayson et al. 2011: 206).

11.5 FROM PASTORALISM TO AGRICULTURE: NORTHWEST ARABIA DURING THE 6TH–5TH MILLENNIUM B.C. After the Neolithic, the earliest evidence of water management in Arabia comes from the Northern part of the peninsula, at the edge of the Nafūd. During the 6th millennium B.C., the area was roamed by mobile pastoralists. This ‘colonization’ represents the second large-scale occupation of Arabia, after the establishment of the Middle/Late PPNB-Late Neolithic groups discussed above (Gebel 2017: 20). It was made possible by the mid-Holocene climate optimum, as extensive steppe grasslands covered the area at that period. At that time, rainfall must have been more reliable, and the denser vegetation promoted reduced or decelerated surface runoff and higher groundwater recharge. Numerous watering places (lakes and aquifers with high water tables) therefore allowed the crossing of dry lands. During the 5th millennium B.C. (the period corresponding to the Chalcolithic in the Levant), groups of mobile pastoralists developed a distinctive domestic, ritual and funerary architectural tradition characterized by standing stones. The best-known sites from this culture are Rajājil, Rasif and Qulbān Banī Murra; nowadays the first two are located in the north of the Kingdom of Saudi Arabia, and the third is located in southern Jordan. These three seem to have been water-favored spots where nomadic groups regularly gathered. These pastoralists groups had hydraulic knowledge, as they built wells near lake shores and in wādī beds, sometimes associated with troughs in order to water their herd, and they harnessed runoff water thanks to small dams. A good example of these technologies is the ‘well/watering complex’ D15 from Qulbān Banī Murra, which has been dated from the second half of the 5th millennium B.C. by radiocarbon (sample obtained from the coating of the troughs) (Gebel 2013: 113). Several structures of this type have been recorded on the site, but D15 is the only one excavated. This large structure is composed of a well, located approximately in the center of the complex, surrounded by compartmented and paved troughs of different shapes, whose walls are made of vertical slabs and single-course walls. The well, which was dug through wādī gravels, measures about 1.2 m in diameter. Its depth is unknown, as excavation had to stop about 4 m below the ground for safety reasons. The coping of the well, which could be accessed by a small staircase, was made of horizontal ashlars (Gebel 2013: 116–117; Gebel and Mahasneh 2013: 137). Fifth-millennium-B.C. populations apparently had an excellent knowledge of groundwater, as the complex D15 was located in an area where the main aquifer received additional influx from two local aquifers and was pressed up thanks to a sub-surface narrowing of the bedrock (Gebel 2013: 116; Gebel and Mahasneh 2013: 136). Similar wells associated with troughs were discovered in Rasif, a site located in al-Jawf province of Saudi Arabia, occupied from the Late Neolithic to the Early Bronze Age. Dwellings and cairn graves are attested on the site, which is characterized by a succession of endorheic depressions trapping sediments and water. Excavation 3 and 4, both revealed a well associated to a rectangular trough delineated by a row of standing stone slabs and paved. Although badly preserved, the well of Excavation 4 clearly recalls, by its architecture and size (c. 1 by 1 m for the well-coping), the one of complex D15 in Qulbān Banī Murra and its upper part’s corbelling masonry. Based on this evidence, a 5th millennium B.C. dating is suggested by the excavators. The well was recently partially excavated, up to 5 m below the ground (Zielhofer et al. 2018: 131–132). The coping of the well, from

Human Adaptation in Arabia

227

Excavation 3, c. 0.8 by 0.5 m, is surrounded by a sub-circular row of standing stone slabs, and the associated trough measures c. 2 by 0.6 m. A 5th or 4th millennium B.C. site is proposed for this structure (Gebel 2016: 94–97). As the investigation of the site was just started few years ago, further data and better dating should be available in the coming years. Recent surveys and archaeohydrological studies have demonstrated that most Rasif wells were located in the endorheic depressions and were likely tapping a shallow aquifer, recharged by flooding, located in their sediments filling (Zielhofer et al. 2018: 131–132). After the Chalcolithic, some of these shallow wells were probably abandoned due to the depletion of the shallow aquifer related to the decline of precipitation and aridification; others were deepened in order to reach the underlying fossil aquifer (Zielhofer et al. 2018: 137). After the 5th millennium B.C., the exploitation of groundwater probably became more difficult because of the aridification, and the team working at Rasif has demonstrated that the endorheic depressions were improved by building dams that retained the runoffs in large pools. This runoff water harvesting system seems to have been modified and used until recent times (Zielhofer et al. 2018: 137). It could have been used by Bedouins to store water and perhaps also to create ephemeral cultivation areas after rainfall events. The work in Rasif illustrates the adaptation of human populations to the aridification of the climate and the increasing water shortage in the mid-Holocene (Zielhofer et al. 2018: 138). H.-G. Gebel hypothesizes that the decrease of water resources (lowering of water table, disappearance of lakes) during the 4th millennium B.C. led the population to settle in more favorable spots in terms of water resources and to establish sedentary oases (“Oasization Model”), while some groups remained mobile pastoralists. Gebel assumed that local populations used their experience of water manipulation to develop these oases (Gebel 2013: 123–124; Gebel 2016: 108).

11.6 EARLY FARMING COMMUNITIES: THE BRONZE AGE While the Chalcolithic and Bronze Age correspond to the rapid development of hydraulic systems in the Near East and Mesopotamia, in parallel to the development of city-states and urban centers (Wilkinson 2003: 72–74), this period is far less known in Arabia. The Bronze Age of Arabia ranges from the end of the 4th/beginning of the 3rd to the end of the 2nd millennium B.C., depending on the places. This period saw the ‘neolithization’ of Arabia, in the sense that permanent villages become the norm, and agriculture is adopted by populations (Charbonnier 2008). Local pottery productions also appear during the period in Southern Arabia. It has been debated whether these evolutions were caused by internal or external factors; that is, did local populations adopt agriculture to meet new needs and the evolution of the social structure, or were they influenced by direct contacts with agricultural societies from the north? The question remains open, but we can guess that it was a little bit of both. Very few water systems are attested for the Bronze Age in Arabia and many of the hydraulic structures attributed to this period are badly dated. Indeed, most of them were found during surveys and are not in a stratified context. Furthermore, absolute dating is not always available. Some data come from the end of 4th/3rd millennium B.C. (Early Bronze Age), while information is scanty for the 2nd millennium B.C. (Middle/Late Bronze Age). This is at least partly due to the fact that the 2nd millennium has been under-investigated compared to previous and later phases.

11.6.1 The Genesis of Floodwater Harvesting in Southwest Arabia Although our data is incomplete, it is becoming clear that Early Bronze Age populations took advantage of surface water. In the western mountains of Yemen, it has been suggested that early farming communities living in Khawlān region, in between the wet plateaus and the arid desert fringe, were growing crops in humid areas near wādī beds, after the floods (Figure 11.3a) (De Maigret 1990): 29–30).

228

Climate Changes in the Holocene

FIGURE 11.3  Map of Southwest (a) and Southeast Arabia (b) with hydraulic structures mentioned in the text ((c) Charbonnier 2018).

An alleged dam, retaining sediments and water, was located near a Bronze settlement, but its dating remains hypothetical (De Maigret 1990): 25–26). In the neighboring plateaus, which were wetter and probably supported rainfed agriculture, hillsides and wādī bottoms were terraced since the 4th millennium and during the 3rd millennium B.C. (region of Dhamār, Wilkinson and Edens 1999; region of Khawlān: Ghaleb 1990: 134, 139). According to Wilkinson, these terraces were mainly used to prevent erosion of soils, in a context of population growth and increased human impact on the landscape (Wilkinson 1999: 183–184). It is likely that some of these terraces were also used to catch and control runoffs. In the interior regions of Southwest Arabia, unlike the western mountains, rainfalls were already probably too low and irrigation was necessary. During the Holocene and until today, wādīs born in the mountains concentrate in this area and dry up into the erg of Ramlat as-Sab‘atayn, an endorheic basin extending the Rub‘al-Khālī. During the Pleistocene, however, as the precipitation was higher, these wādīs joined and crossed the basin and then followed the H ̣ad ̣ramawt/Masīla valley up to the sea (Cleuziou, Inizan, and Marcolongo 1992: 8). The main issue for understanding pre-1stmillennium-B.C. settlements and irrigation systems in this area is the sedimentation of the valley floor. Indeed, the flood irrigation entailed the regular accumulation of silty sediments, sometimes several meters in height, has covered older structures. If we take into account the great mastery of floodwater during the South Arabian period (see below), it is clear that their manipulation started

Human Adaptation in Arabia

229

much earlier, at least during the Bronze Age. Evidence of the early domestication of floods comes from Wādī Ṣanā, a tributary of Wādī Masīla in H ̣ad ̣ramawt. Ancient hydraulic structures were identified in the valley floors of Wādī Ṣanā and its tributaries. These structures, embedded in sediments, were interpreted as ‘check dams’, used to slow down floodwater in order to foster the deposition of sediments and the infiltration of water, and ‘diversion channels’, that diverted runoff water on the neighboring valley slopes. They were dated from around the 3rd millennium B.C. (Harrower 2008b; Harrower 2008a). However, no geoarchaeological investigations were carried out in the sediment fill and in the surroundings of the structures to confirm their nature. In the H ̣ad ̣ramawt region, however, a more recent geoarchaeological and paleoenvironmental study suggests that the environment was wet enough between the 3rd millennium B.C. and the beginning of the 1st millennium B.C. to allow dry-farming in valley bottoms thanks to a high groundwater table and permanent surface flows (Berger et al. 2012: 158–159). In the other valleys leading to the Ramlat as-Sab‘atayn, no Bronze Age hydraulic structure has been discovered yet. Some absolute dating, conducted on the irrigation sediments in the old field systems, suggests that the earliest systems were developed during the 3rd millennium B.C. in Wādī Markha (radiocarbon, Brunner 1997: 196) and the beginning of the 2nd millennium B.C. in Wādī Bayh ̣ān (Balescu et al. 1998). In the coastal plains of Yemen, the first evidence of wādī irrigation dates back to the end of the 3rd/beginning of the 2nd millennium B.C. (Brunner 2013). This period saw the development of Ṣabir culture, from the eponymous site, located near Lah ̣j, along Wādī Tuban. During that period, the Ṣabir culture saw the development of social stratification, tasks specialization and a process of urbanization, evidenced in particular by the appearance of monumental architecture, besides dwellings, and the differentiation of quarters depending on activities. The neighboring archaeological site of Ma‘layba belongs to the same culture as Ṣabir and was occupied between the end of the 3rd millennium and the 13th century B.C. (Buffa 2007). During the excavation of the settlement, canals dated between 1900 and 1400 B.C were discovered at different levels running between the dwellings. They were following the same orientation, some seeming to correspond to primary canals (1.5 to 2 m wide), while others were secondary canals (less than 0.4 m wide) (Buffa 2002: 25–30). The successive canals were interspaced with layers of sand and gravel witnessing regular destructive flooding episodes. The canals were excavated only on a small surface, so the plan of the canal network is unknown. It is, however, likely that they were part of an irrigation system, most probably fed by the neighboring Wādī Tuban. The regime of Wādī Tuban is still debated for the 2nd millennium B.C. V. Buffa has argued that Ma‘layba’s canals were initially fed by a perennial water stream which then became intermittent due to the environmental degradation (Buffa 2002: 171). U. Brunner, on the other side, thinks that from the beginning the irrigation system conveyed both floodwaters, during the monsoon season, and permanent water streams during the rest of the year (Brunner 2013: 64).

11.6.2 Early Water Management in Southeast Arabia In the second half of the Holocene, Southwest Arabia was a favored spot in term of precipitation thanks to the summer rains brought by the monsoon. Conversely, with the retreat of the monsoon towards the south, summer rains ceased to reach Southeast Arabia from the 5th millennium B.C. (Fleitmann and Matter 2009: 640). Interestingly, however, recent research led in the area of Bāt (Oman) – an important Early Bronze Age settlement – suggests that floodwaters were diverted to irrigate the crops from the second half of the 4th millennium B.C. in the area (Figure 11.3b). This information was obtained through the micromorphological analysis of a naturally exposed crosscut in Wa d̄ i ̄ al-Sharsa ,̄ in which Bāt is located. According to this study, the wādī was an active floodplain before the 4th millennium B.C. and then became less active. A subsequent gradual silting of the plain, interspaced with episodes of rapid flooding, testifies to anthropic flood diversion, that is, fields were irrigated with wādī flows. Irrigation then became more intensive, with

230

Climate Changes in the Holocene

evidence of tillage in the 3rd millennium B.C. (Desruelles et al. 2016). We must remain cautious with these results, as they were not confirmed by the study of other crosscuts or test-pits and as no related hydraulic structure was found. Furthermore, what crops could have been grown at such an early date remains a mystery, as domestic cereals and pulses are attested only from the 3rd millennium B.C. In the 1980s, a series of walls located near the archaeological sites, at the bottom of a slope and perpendicular to a wādī bed, were interpreted as diversion dams of Bronze Age date (Frifelt 1985: 99). This is, however, questionable since no absolute dating was carried out and since pottery sherds recovered during the excavation of one of the walls only provided a terminus post quem (Brunswig 1989: 22–25). The same applies to their function, as they are not slanted compared to the wa d̄ i ̄ bed and they are too close to each other. It was assumed that floodwaters were exploited at the same period in Wādī Samad (Oman), where several protohistoric settlements are attested. Two levees of stone blocks, interpreted as diversion dams, were identified next to a wādī channel. It has been suggested that they were used to trap water and sediments (Hastings, Humphries, and Meadow 1975: 11), although no detailed geoarchaeological and/or geomorphological study was conducted to confirm this assertion. The Bronze Age date being based only on their proximity to an Umm an-Nar period (2700–2000 B.C.) settlement, these structures are therefore poorly dated. In the area of Bahla ’̄ (Oman), another structure that could be related to surface water management was found near the village of al-Āqir. This 300-m-long low wall is blocking a small valley can be quite securely dated from the Bronze Age since several plano-convex bronze ingots, generally dated to between the 3rd and 2nd millennia B.C., have been recovered from its masonry (Weisgerber and Yule 2003: 48–51). Its function, however, was not clearly identified, since no geoarchaeological study was conducted. We can hypothesize that it was used to retain water and sediments for agriculture or grazing. Unfortunately, no data regarding surface water exploitation is available for the 2nd millennium B.C. in Southeast Arabia. The evidence for groundwater exploitation is sparse for this period. In the North of the peninsula, wells were dug since at least the 5th millennium B.C., and one may imagine that they were still used during the Bronze Age. The use of wells has not been evidenced yet in Southwest Arabia during the Bronze Age. In Southeast Arabia, wells are attested since the Umm an-Nar period, but only in a settlement context. To be more specific, the known wells were located in the middle of massive structures built with mud-bricks or stone blocks, mostly referred as ‘towers’ in the literature even though their nature is still debated. They are notably attested on the sites of Bāt, Maysār (Wādī Samad), Ṣa lūt and Hīlī 8 (oasis of al-‘Ayn) (Frifelt 2002; Weisgerber 1981; Degli Esposti 2011; Cleuziou 1989). A well was also excavated in the Early Bronze Age settlement Maysār 1. It is therefore incontestable that these structures are not related to irrigation. They may have been used for domestic and perhaps religious purposes. For obvious security reasons, most of these wells were not completely excavated. In Hīlī 8, both attested wells were associated to a ‘tower-like’ building (Figure 11.4). The oldest was used during the first half of the 3rd millennium B.C., the latest in the second half of the same millennium until c. 1800 B.C. Their upper part was made of stone blocks and roughly circular in plan, with a diameter of c. 0.5 m. While the oldest well was about 4 m deep and had a rectangular drainage chamber, larger than the access shaft, the latest was 8.5 m deep and was cylindrical (Cleuziou 1989: 67–68). Given the facts that the climate was already arid during that period, and that this technology was used for human consumption, we can assume that wells were also used for irrigation during the Bronze Age in Southeast Arabia. They would have supplied annual crops throughout the year. I have also suggested that, given the fact that the environment was slightly more favorable during the 3rd millennium B.C., some plants, such as date palms, might have been cultivated without irrigation, their roots taking advantage of water tables close to the surface, as it is sometimes the case in Northern Africa (Charbonnier 2017: 65–66).

Human Adaptation in Arabia

231

FIGURE 11.4  Sections of the Early Bronze Age wells excavated at Hīlī 8 (from Cleuziou 2009: Figure 4).

11.7 BLOOMING OF HYDRAULIC TECHNIQUES IN ARABIA: THE IRON AGE Between the end of the 2nd millennium and the beginning of the 1st millennium B.C., Arabian societies experienced considerable changes. This was the time of the ‘Iron Age’ cultures, although this term, imported from European archaeology, does not fit with the Arabian cultures. In Southeast Arabia, the Iron Age II period (1100–600 B.C.) seems characterized by a rapid demographic growth, witnessed by the increase of the sites and their widespread geographic distribution (Magee 2014: 215–220). In Southwest Arabia, the South Arabian culture flourished around 1000 B.C. with, notably, the development of several kingdoms, important urban centers and the introduction of writing. The domestication of the camel around 1000 B.C. (Magee 2015) could have played a role in this evolution, as it allowed for improved land transportation, especially accross the desert. It clearly supported the development of the incense road between Southwest Arabia and the Near East. In Southeast Arabia, it could have improved connections between sites and the movement of goods among more specialized sites (Magee 2004). In Northeast Arabia, the 1st millennium B.C. saw the blooming of important oases such as Taymā’ and al-‘Ulā (capital of the Lih ỵ ān kingdom) that were involved in caravan trade. Regarding Southeast Arabia, it was long stated that the demographic and economic growth of the Iron Age II was related to the adoption or invention of underground draining galleries (qanāt). More recent research suggests, however, that strategies of water management were diverse during that period. Groundwater was accessed by both wells and qanāt-like structures. The harvesting of runoffs is also clearly evidenced for this period. Very few irrigation systems are known extensively, but we have some evidence that they remain small scale. In Southwest Arabia, on the other hand, large-scale irrigation systems, mainly based on the diversion of floodwaters, developed in the valley surrounding the Ramlat as-Sab‘atayn and in

232

Climate Changes in the Holocene

H ̣ad ̣ramawt, in parallel with the flourishing of South Arabian kingdoms. The knowledge for managing surface flows developed since the beginning of the Bronze Age had given birth to complex and elaborate systems.

11.7.1 The Oases of Northern Arabia During the Iron Age Data is unfortunately limited on water management during the Iron Age to the North of the Arabian Peninsula. However, we have some evidence of well-fed irrigation for this period. An irrigation system dated from this period was recently investigated in Taymā’, an important oasis that was occupied by a sedentary population since the 3rd millennium B.C. A Sabkha, a depression with a high salinity, in which surface water accumulates after rainy episodes, is located north of the oasis. The geological setting of the area in which the site was established accounts for the presence of artesian groundwater that could be easily be exploited thanks to wells (Wellbrock, Grottker, and Gebel 2017: 36). South of the oasis, in the Compound A (a large area delineated by a section of the antique city wall), an 8 ha irrigation system was recently identified (thanks to a geomagnetic prospection) and investigated by a German mission. A series of canals of different sizes was partially excavated by means of test-pits. Some correspond to simple ditches cut in the ground while others are stone-built structures, opened or covered stones slabs. Beside the canals, distributors, basins or plot borders were also discovered (Wellbrock et al. in press). Although this was not proven by archaeology, it seems that this irrigation system was fed by one of the wells that was located to the south of Compound A. In Area H, it was demonstrated that an Iron Age building was constructed on top of one of the canals. Radiocarbon dating obtained in the building suggest that the irrigation system was in operation before the 8th–6th century B.C. (Hausleiter 2016: 149). It therefore can be dated from at least the Early Iron Age (Wellbrock et al. in press). An Iron Age irrigation system taking advantage of runoffs and floodwater is also mentioned in the oasis of Qurayyah in Saudi Arabia (Wellbrock, Voss, and Grottker 2012: 42; Luciani 2016: 33–35; Parr, Harding, and Dayton 1970). I should mention here, as a parallel, the settlement of Jawa, located in the basalt desert of northeastern Jordan, where floodwaters and runoffs were exploited since the Early Bronze Age (3500–3000 B.C.) (Meister et al. 2016). This evidence proves that surface management was developed at a large scale since the Bronze Age.

11.7.2 Southeast Arabia During the Iron Age: Groundwater Exploitation Through Wells and Aflāj The Iron Age seems to correspond to a peak of hydraulic activity in Southeast Arabia, as many more hydraulic structures dating from that period are known. Although we do not have any detailed information on the evolution of the environment during this period, it is clear that the environment was already arid at that time. Populations had to rely on water-rich areas or hydraulic technologies to irrigate their fields. The exploitation of groundwater was a key for the development of Iron Age settlements, both for demographic and craft reasons. For the first time in the region, we have good evidence of wells used for irrigation and craft activities. A well was located inside the Iron Age village of al-Thuqaiba, composed of mud-brick houses, located in al-Mādam (Emirate of Sharjah, U.A.E.). It is a 7-m-deep structure, rectangular with rounded corners in plan, and its bottom is dug into the bedrock. A small basin located next to the edge of the well was probably intended to provide water to the livestock (Córdoba and Del Cerro 2005: 518–519). More interesting, two other wells were identified in another area, located southwest of the settlement, that were clearly used to feed a wide area devoted to the manufacturing

Human Adaptation in Arabia

233

of mudbricks. This workshop, dated stratigraphically from the Iron Age, was made of several small basins, fed by the well through a network of canals (Córdoba and Del Cerro 2005: 520–521; Córdoba 2013: 144–146). This discovery highlights the fact, often forgotten by archaeologists working in Arabia, that water was not only useful for human consumption or agriculture but also for manufacturing many items, such as mud-bricks or potteries, and for processing raw material, such as animal and plant fibers or copper ores. Further evidence of water management in a crafting context might come from Bayt Bin ‘Ātī al-Darmakī’s excavation, located in Qaṭtạ̄ rah oasis (al-‘Ayn, U.A.E.). A possible workshop area was found at the bottom of the ‘Workshop trench’, north of the excavation. Although it can be stratigraphically dated from the beginning of the Iron Age, its function remains unknown. This installation was made of square tanks, c. 0.45 on one side and c. 0.5 m deep, connected by small canals, c. 0.20 m wide and c. 0.03 m deep, dug into the bedrock. Thanks to the sloping surface of the bedrock, the water was brought by gravity from a 3-m-deep well located north of the tanks (Power and Sheehan 2011: 271). Several hypotheses have been made regarding the use of this installation: dying, tanning or copper processing, but for now we have no evidence to support or reject these hypotheses. After the abandonment of this workshop, the area was devoted to agriculture during the Iron Age II. Evidence of gardens was found, associated with wells. To date, this is the oldest clear evidence of well-irrigation in Southeast Arabia. In the ‘Energy center trench’ (southern part of the excavation), a deep canal with a ‘V’-shaped section, oriented northwest-southeast, was surrounded by four subcircular basins (c. 2 m in diameter and 1 m in depth). Close to some of them were discovered small curvilinear channels. According to Power and Sheehan, the ditch was providing water to the basins from which water was brought to the crops by means of channels. They were contemporary with a series of small irrigation canals, earthen-made, found in the Workshop trench (Power and Sheehan 2011: 272). We have evidence that the area remained devoted to agriculture for a relatively long period of time, as tree pits were found in the overlying stratum in the Workshop trench. They were surrounded by circular earthen-made watering basins, c. 3 m wide and 0.12 m deep, arranged in rows. The tree pits themselves were about 0.5 m in diameter and 0.35 m in depth and were filled with organic/ashy material. They seem to have been irrigated by wells, since a 3 m in diameter and 4.5-m-deep well was found in the Energy center trench (Power and Sheehan 2011: 272–273). Analyses are ongoing to identify what tree species was grown during the Iron Age II in Qat ̣t ̣ā rah. More recently, Iron Age II wells associated to ancient gardens were found in the oasis of Masāfī, located north of the H ̣ajar mountains (Emirate of Fujairah). These structures were found in a stratified context, within test-pits. Two circular wells were identified in the section of test-pit 9, dug in the center of the Masāfī palm grove (Figure 11.5). They were dug on top of a fine dark brown silt, rich in charcoal, probably corresponding to a past agricultural soil dated to 795–540 cal BC by radiocarbon. The wells were filled in burnt gravels and charcoals. One of them was partly excavated and proved to be 1.2 m in diameter. The wells were probably exploiting an ancient permanent water table, indicated by a layer of gravels and blocks of gabbro, cemented in a carbonated green matrix, visible at the bottom of the test-pit. Pottery sherd collected in the strata covering or associated with the wells and in their filling also point to an Iron Age II dating (Charbonnier et al. 2017: 59). In our present state of knowledge, the qanāt technology was developed or introduced during the Iron Age II in Southeast Arabia. Qanāt are subterranean galleries intended to drain groundwater resources in their upstream section and channel the water to the surface by gravity, the gradient of the underground gallery being lower than that of the natural terrain (Figure 11.6). They are generally ventilated by shaft holes that are also used to remove the spoil when digging underground. Depending on the context, different groundwater sources can be exploited, such as watercourse underflows, perched water tables or deep aquifers. In Southeast Arabia, the local term for these structures, regardless of the type of groundwater drained, is falaj (pl. aflāj) dā‘ūdi. For the sake of clarity, I will only use the term falaj to designate these structures.

234

Climate Changes in the Holocene

FIGURE 11.5  Test-pit 9 from Masāfī with Iron Age wells ((c) Purdue, Charbonnier 2018).

FIGURE 11.6  Schematic section of a qanāt ((c) Charbonnier 2018).

A few years ago, an Early Bronze Age date was suggested for the earliest falaj in Southeast Arabia (Jocelyn Orchard and Orchard 2007) but the evidence put forward has been criticized (Charbonnier 2015a: 47). By contrast, many aflāj have been identified in or next to Iron Age settlement sites in northern U.A.E. and Oman, of which only three have been partly excavated, all in the U.A.E.: al-Mādam 2, Bida Bint‘Sa‘ūd and Hīlī 15. Although there is no doubt about the fact that they were really structures designed to drain groundwater, the dating of most of them remains problematic, as absolute dating remain rare (Charbonnier 2015a: 66). The most spectacular discovery was made in al-Mādam. In this area, AM-2 falaj is visible from the ground over a distance of about 1 km. It has been partly excavated by a Spanish team, revealing an underground gallery followed by an open canal that fed an extensive irrigated area located c. 500 m south-east of the Iron Age settlement of al-Thuqaiba (Figure 11.7). The underground gallery was c. 4 m deep from its flat bottom to its barrel vault and c. 0.5 m wide. It also adopted a meandering course. Five sub-circular access shafts were excavated. They were strengthened with mud and mud-bricks in their upper part (Córdoba and Del Cerro 2005: 522–523; del Cerro and Córdoba 2018: 90–91). When approaching the surface, the gallery transformed into an open canal simply dug into the ground. This primary canal was not completely excavated, but it seems to have followed its straight course up to the downstream end of the irrigation system. On either side of this primary canal, secondary canals branched off at right angles, at variable intervals. They were therefore roughly parallel. Small circular to sub-square basins and larger rectangular and elongated basins

Human Adaptation in Arabia

235

FIGURE 11.7  Aerial view of the Iron Age irrigated area associated with falaj al-Mādam 2 (from Del Cerro and Córdoba 2018: fig. 10).

were distributed along the secondary canals. The entire irrigation system was cut into the bedrock and filled with sand, on which the crops were, in all likelihood, growing. Two date stones were found during the excavation (Córdoba 2013: 147–149; del Cerro and Córdoba 2018: 96). It is likely that this system was devoted to some form of horticulture, as the small basins are probably tree pits. Date palm could have dominated, as it grows naturally on sandy soils. The elongated basins could have been water tanks, but alternatively we can hypothesize that they were devoted to the cultivation of annual crops, such as vegetables, pulses or cereals. This irrigation systems differs greatly in terms of spatial organization from the one we know today in the region. The latter tend to form arborescence with a hierarchy of subcanals (see, for example, Ādam oasis, Charbonnier 2014: Figure 2). An Iron Age date is likely for this system: it is supported by the Iron Age II sherds recovered as well as by a radiocarbon date obtained on one shell of a gasteropod from the Thiaridae family, which live in humid condition. It was dated between 1160 and 808 B.C. with 95% probability (Córdoba 2013: 147–148). Unfortunately, the stratigraphy of the canals’ filling and the exact strata in which the dated shell was collected was never published. More interestingly, AM-2 shafts exhibit small steps, allowing access to the underground gallery, similar to those discovered in the well located in the Iron Age village of al-Thuqaiba (Córdoba and Del Cerro 2005: 523). The depth of the underground gallery and evidence of deepening of the canals and basins, uncompleted, suggest that the water table dropped, leading eventually to the abandonment of the falaj and of the irrigation system (Córdoba and Del Cerro 2005: 525; Córdoba 2013: 148). Located in the area of al-‘Ayn, the downstream end of falaj Hīlī 15 was identified over a distance of 500 m and partly excavated. It is situated at close proximity to several Iron Age settlements. The investigated section of the falaj corresponds to an open-air canal, about 0.5 m wide, faced with stone blocks. It is dug below the ground surface and is about 0.4 m deep at the level of Area B. Upstream, the canal is covered with stone slabs (from Area E), and its bottom is located up to 2 m below the ground. In this area, a 0.5 m in diameter access shaft, made of mudbrick, allowed access to the slab-covered gallery (al-Tikriti 2002: 112). Due to the urbanization of al-‘Ayn, the upstream end of the falaj and its origin could not be recognized. W. al-Tikriti mentions an underground gallery, discovered accidentally 1.5 km northeast of the portion excavated, that can perhaps be associated with Hīlī 15 (al-Tikriti 2010: 228). A water distributor (named ‘sharī‘a’ by W. al-Tikriti), of a simple type, has been recovered along the main canal. At that point, two channels split up at right angles from the primary canal. Sluice gates could have been set against vertical supports located on either side of each canal, allowing

236

Climate Changes in the Holocene

for the flow of water to be diverted into the secondary canals or maintained in the primary canal. Downstream, subsidiary earth canals were recovered that seem to have fed some cultivated plots (al-Tikriti 2002:120). The Iron Age date for this structure has been proposed on the fact that Iron Age sherds, some in situ, were recovered during the excavation and due to its position in the middle of the Iron Age settlement. This dating could have been strengthened by establishing a more direct spatial relationship between the Iron Age sites and the falaj, through excavation. Absolute dating would also be needed. It was suggested that Iron Age aflāj drained only shallow water tables or wādī underflows (Boucharlat 2003), although most mother-wells were not recovered and no geomorphological/ hydrological studies have been conducted to confirm it. Evidence from al-Mādam-2 falaj shows that aquifers were exploited during the Iron Age as the underground gallery was cut into the bedrock. Iron Age wells excavated proved to be quite shallow, the deepest being 7 m, suggesting that groundwater, including that contained in aquifers, was quite close to the surface during the 1st millennium B.C., at least in some areas. Falaj irrigation could have been important during the Iron Age, but the use of wells for irrigation should not be underestimated. Until recently, many oases of Southeast Arabia were fed exclusively by wells (Costa and Wilkinson 1987: 43–53).

11.7.3 Southeast Arabia During the Iron Age: Runoff Harvesting in the Oasis of Masa- fiThe oasis of Masāfī has been populated since at least the 2nd millennium B.C. (Benoist 2013; Degli Esposti and Benoist 2015). During the Iron Age, a fortified settlement (Masāfī-2), a large building with a columned-hall (Masāfī-2) and a sanctuary (Masāfī-3) were distributed around a garden area that was partly irrigated with the help of wells (see above). At the eastern edge of this area, an irrigation system was found north of Masāfī-1 (Figure 11.8). The primary canal was about 0.15 m

FIGURE 11.8  Aerial view and plan of Masāfī with the Iron Age runoff harvesting system on top (from Charbonnier, Purdue, and Benoist 2017: Figure 2).

Human Adaptation in Arabia

237

wide and was delimited by single-faced stone walls. It fed subsidiary earthen canals, which were about 0.10 to 0.2 m wide. The latter were directly connected, and partly covered, by a silty stratum, light brown and rich in charcoal, interpreted as an ancient cultivated layer. Microcharcoal sampled in this layer provided a date of 897–801 BC, dating this system to the Iron Age II. The cultivated layer and the subsidiary canals were located north of a low wall made of vertical stone slabs that seems to have separated the gardens from the Iron Age building. The gardens are unfortunately not preserved, as the area has been bulldozed in the last decade. The upstream end of the primary canal could not be found, as the area was already urbanized, but the micromorphological study of its filling (0.15 m of coarse sediments) suggests that it used to channel runoff water that was harvested in a broad valley to the east of the oasis (Charbonnier, Purdue, and Benoist 2017). This is the oldest runoff harvesting system attested in Southeast Arabia, and it shows that surface water was exploited in this region during the Iron Age in addition to groundwater.

11.7.4 The Harnessing of Floodwaters in the South Arabian Kingdoms As we have seen, the manipulation of surface water started during at least the Early Bronze Age in Southwest Arabia. With the development of the South Arabian kingdoms at the margin of the desert, huge irrigation systems were developed around the main urban settlements. The size of these systems that were, at least partly, controlled by the rulers, is generally between a hundred and several thousand hectares. They were mainly fed by wādī floods, caused by the monsoon, that occurred in spring and summer. Runoff seems to have been also exploited, and wells are attested in many flood-fed irrigation systems where they allowed year-round cultivation (Bowen 1958: 58; Maraqten 2017: 125). The great similarity of hydraulic practices from the western edge of the Ramlat as-Sab‘atayn to the H ̣ad ̣ramawt is striking and reflects both comparable environmental constraints and the belonging of all these groups to the same cultural complex. The floods were diverted toward primary canals and gardens with the help of dams, whose role was not to store but to tame and slow down the water flow. The purpose of these structures was, in all likelihood, to divert only a portion of the floods and to evacuate the surplus of water downstream. Two main reasons for this: avoiding excess water to damage the irrigation systems and allowing part of the flood to reach other irrigation systems located downstream. Indeed, at the scale of a wādī, some sort of coordination was necessary to share water among different communities (Mouton 2009). Apart with few exceptions, the floods were never retained nor stored. Water was immediately distributed on the largest possible area of fields (Gentelle 1991: 51). At that point, it was partly retained by low earth levees in order to seep into the ground (Brunner 2004: 135). Overflows evacuated the excess water toward the downstream fields, allowing the movement of water from plot to plot. This type of irrigation was intended to allow the silty sediments brought by the floods to settle in the fields. Indeed, soils are rare in arid environments and rapidly depleted (Brunner and Haefner 1986: 82). Furthermore, if the sediments deposited in the canals, the systems would be rapidly clogged in. This system also prevented soils salinization (Brunner 2004: 135). A negative feedback of this process was the slow but consistent elevation of field systems, which, after some time, would prevent the water flowing by gravity. At that point, water intake had to be moved upstream in the wādī or irrigation systems relocated downstream. Diversion dams are rarely preserved, as they have been exposed to the floods for a very long time. Some were stone-built, while in many cases, they corresponded to earth levees. Diversion dams were generally slanted across wādī beds and sometimes situated in a bend. Masonry dams are, for example, mentioned in Shabwa or Wādī d ̣urā’ (Gentelle 1991: 22–24; Breton and Roux 2002). One exception to this model is the so-called Great Dam of Ma’rib, capital of the Sabaean kingdom, located at the western edge of the desert. The function of this structure, as well as its dating, have been much debated. The Great Dam was associated to a very large oasis, about 10,000 hectares (Brunner and Haefner 1986: 81), surrounding the different components constituting the antique

238

Climate Changes in the Holocene

city, on both side of Wādī Dhana. This important intermittent water course, born in the western mountains of Yemen, has a 10,000 sq. km watershed and an average discharge of one million cubic meters of water (Brunner 2000: 171). The Great Dam, positioned at the level of two rocky hills, was made of three main components. The dam itself corresponded to a 620-m-long earth levee, about 16 m high and triangular in section, covered by stone blocks (Brunner 2004: 136). It is nowadays preserved over a length of 250 m long on the northern bank of the wādī. At each end of the levee, a massive water intake, built with dressed stone blocks, provided water to a large primary canal. The northern water intake (or ‘sluice’ in the literature) was providing water to the northern part of the oasis through two outlets (Figure 11.9). An overflow allowed excess water to be discharged into the neighbouring Wādī al-Jufayna. The northern water intake roughly measured 150 m long, up to 50 m wide and up to 7 m high (Vogt 2004: 378). The Southern water intake (or ‘sluice’) was about 60 m long and was also equipped with a water intake and an overflow. Several building phases have been evidenced on the dam and the water intakes. The visible remains of the Great Dam can be dated from the mid-1st millennium A.D. (Darles et al. 2013: 13; Vogt 2004: 385). However, it is very likely that it had replaced older dams. Although it seems similar to a storage dam, the Great Dam was acting more as a threshold, that is, it retained water in order to allow part of the flow to reach the level of the irrigation and field systems, which had been slowly raised due to the continuous accumulation of silts over the centuries (Brunner and Haefner 1986: 81). The South Arabian irrigation systems are most of the time well visible on the aerial and satellite imagery. They can be mapped, and their spatial organization can be studied. Two paradigmatic organizations can be noticed. Some irrigation systems, such as the one in Wādī Bayh ̣ān or around Shabwa, consist in a central or lateral primary canal extending along the field system and from which secondary canals gradually branch off (Gentelle 1991; Bowen 1958). Others, such as Ma’rib, are organized like an arborescence, with a hierarchy of secondary and tertiary canals (Hehmeyer 1989). The canals were delineated by earth levees, while other hydraulic structures (water distributor, sluices, overflows, etc.) were generally masonry built. It is obvious that major irrigation and hydraulic structures were built by South Arabian kings (see Darles et al. 2013 about the Great Dam or Charbonnier 2015b about Wādī Bayh ̣ān). However, the corpus of inscription also indicates that individual farmers or tribal groups owned land and were taking care of some hydraulic structures. Thus, I have suggested that local and central management coexisted at the scale of each irrigation system. The kings probably did not make all the decisions concerning irrigation, but inscriptions show they were in charge of building and maintaining major structures, such as dams and primary canals, benefiting a broad community (Charbonnier 2015b: 483). It is still debated whether building hydraulic structures and legislating water brought prestige

FIGURE 11.9  View of the Northern water intake of the Great Dam of Ma’rib ((c) Charbonnier 2008).

Human Adaptation in Arabia

239

and power to the South Arabian kings (Harrower 2009: 65) or whether the latter did not initially derive power from the control of water but became involved in water management subsequently, in order to increase their legitimacy within and their influence on civil society (Mouton 2009: 90; Charbonnier 2015b: 485). The South Arabian irrigation systems were abandoned between the end of the 1st millennium B.C. and the first half of the 1st millennium A.D. This phenomenon seems to be linked to the progressive disintegration of South Arabian kingdoms and the shift of part of the population in the western mountains where the heart of the h ̣imyarite kingdom was located. The accumulation of silts and the raising of fields might also have caused the abandonment of some systems in a time of social restructuring. One must not forget that many other types of irrigation techniques were used in Southwest Arabia during the 1st millennium B.C. In the western mountains, terrace fields, dry-fed or irrigated, are attested (Wilkinson 1999: 187). We also have epigraphic evidence of the exploitation of the few permanent stream flows (ghayl) in the mountains (Maraqten 2017: 125–126).

11.8 DISCUSSION: PHASING THE DEVELOPMENT OF WATER MANAGEMENT IN ARABIA Although, over the Holocene, the Arabian environment became drier and drier and water resources more difficult to exploit, the region has been continuously occupied. Hydraulic strategies changed over time, and new techniques were developed in order to take the best advantage possible of hydric resources. One should, however, be cautious: linking direct relationship between climate change and evolution of societies and water technologies should be avoided. Water mastership went through different stage in Arabia. The management of surface water seems to appear at an early date, since the PPNB in Southern Jordan, with the harnessing of runoffs with walls for agriculture/herding and their storage into cistern for drinking. These practices are also attested during the Chalcolithic in Northwestern Arabia, in a context of a drying environment. Similar techniques were perhaps used in Southeast Arabia since the 4th–3rd millennium B.C., although the few structures that were found are difficult to interpret. At about the same period, the first agricultural communities of Southwest Arabia also started to manipulate surface flows. In this area, however, populations probably benefited from spring and summer rainfalls, while in Southeastern and Northwestern Arabia, surface water was then probably limited to winter. The use of wells is attested at a slightly later date in Arabia, although this technology was known in the Levant and Cyprus since the PPNB. Shallow wells were dug during the 5th millennium B.C. in Northwest Arabia, in order to provide drinking water to humans and livestock. Well technology is known since at least the Early Bronze Age in Southeast Arabia in a domestic context. It is still unclear if wells were used for irrigation also, but it is likely given the environmental context (Charbonnier 2015a: 66). These examples illustrate the capacity of prehistoric Arabian population to developed technical solutions to compensate for the lack of surface water (rains limited to winter) after the end of the climatic optimum. The slow aridification of Arabia might have triggered the adoption of hydraulic techniques, but other socio-economic factors must have played as well. In Southeast Arabia, for example, the adoption of agriculture seems to be related to cultural changes associated with external stimuli, that is, integration of the region into an extra-regional exchange network with the export of raw materials – mainly copper – to Mesopotamia and the Indus valley (Cleuziou 1999: 99). The fact that there is no direct link between climate change and water technologies is better illustrated by the Iron Age societies. At the end of the 2nd millennium – beginning of the 1st millennium B.C., while the climate was becoming even drier than before, Northwest, Southeast, and Southwest Arabia were characterized by increased social complexity. These evolutions were accompanied by the development of new hydraulic technologies and the greater sophistication of

240

Climate Changes in the Holocene

irrigation systems. Qanāt allowed tapping deeper groundwater in the Oman peninsula, and floodwater harvesting reached a very important scale with systems of thousands of hectares in Southwest Arabia. This proves a great resilience of Arabian societies to environmental change. It seems that hydraulic techniques were also more diverse during this period. The use of wells and runoff harvesting is clearly attested in Southeast and Southwest Arabia, in addition to the diversion of perennial and intermittent watercourses in the latter region. At this point, it is important to stress the fact that the different hydraulic strategies used in the second half of the Holocene in Northwest and Southeast Arabia, on one hand, and Southwest Arabia, on the other hand, is related to the temporal distribution of precipitation. The important role of flood and runoff harvesting in Southwest Arabia is related to the fact that, due mainly to the monsoon, surface water is available in spring and summer and allows several harvests a year. As flood episodes were short in time, the main problem the local population had to solve was to allocate water as soon as possible. Managing water spatially and solving the issues between communities along each water course was the key (Mouton 2009). In Northwest Arabia and in the Oman peninsula, however, rainfall was available only in winter after the 5th millennium B.C. The exploitation of groundwater allowed securing access to water during the other seasons and adapting to the limitations of the environment, hence the important role of cisterns, wells, and qanāt.

11.9 CONCLUSION In terms of water management, the Arabian Peninsula proved to be an innovative place. The climate change over the Holocene stimulated the development of innovative hydraulic technologies such as runoff and flood harvesting systems or qanāt. The cradle of qanāt technology is not known, and this technique could actually have several cradles (Boucharlat 2016; Charbonnier and Hopper 2018), but Arabia seem to have played an important role in its genesis. These hydraulic technologies would have been useless without the deep knowledge of their environment Arabian populations had developed over the millennia, as finding water in arid and semi-arid areas is a difficult task. This knowledge allowed them to identify available water resources and to implement appropriate techniques. The synthesis of water techniques proposed in this chapter is preliminary. As can be noticed on a map, the hydraulic systems discussed here are all located in the Northern part of Saudi Arabia, in Yemen or in the Oman peninsula. Our knowledge of other regions is unfortunately limited, and we must hope that further research will shed some light on them. The same remark applies to Bronze Age hydraulic systems, which remain badly understood compared to later periods. To move forward and to broaden our knowledge of ancient water systems in Arabia, it is necessary to develop new methodological approaches. Detailed stratigraphical analyses remain rare; the fillings of hydraulic structures were not systematically studied. However, a geoarchaeological and micromorphological analysis would allow addressing the functioning of the systems and linking them to both environmental and social processes (see for example Charbonnier, Purdue, and Benoist 2017). Absolute dating is also needed to understand the history of hydraulic techniques. Optically stimulated luminescence (OSL), for instance, could be really helpful (Bailiff et al. 2018). Issues remain regarding hydraulic structures that have been dated only because of pottery sherds or spatial association with an archaeological site. While many irrigation systems were discovered because of remote sensing or field surveys, recent research in Southeast Arabia highlights the necessity to carry on systematic test-pits to detect ancient hydraulic structures or field systems (Charbonnier et al. 2017; Power and Sheehan 2011). Modelling hydric resources and water systems is another promising approach that has been recently applied in the north of the Arabian Peninsula (Wellbrock, Voss, and Grottker 2012). Such an approach will allow understanding the functioning of these systems in a very detailed way. Furthermore, these approaches should be combined within the framework of a new discipline: archaeohydrology (Wellbrock, Grottker, and Gebel 2017; Gebel 2017). The transdisciplinary study of water systems in the

Human Adaptation in Arabia

241

Arabian Peninsula will be fundamental to understanding the adaptation of human populations to increasing aridity, and this will be a meaningful contribution in the current debate about climatic and environmental change.

REFERENCES al-Tikriti, W.Y. 2002. The South-East Arabian origin of the Falaj system. Proceedings of the Seminar for Arabian Studies 32: 117–138. al-Tikriti, W.Y. 2010. Heading north: An ancient caravan route and the impact of the Falaj system on the iron age culture. In Eastern Arabia in the First Millennium B.C., Arabia Antica 6, edited by A. Avanzini, 227–247. Rome: “L’Erma” di Bretschneider. Bailiff, I. K., N. Jankowski, L. M. Snape, C. M. Gerrard, A. Gutiérrez, K. N. Wilkinson. 2018. Luminescence dating of Qanat Technology: Prospects for further development. Water History 10(1): 73–84. doi:10.1007/ s12685-018-0213-x. Balescu, S, J.-.F Breton, B. Coque-Delhuille, and M. Lamothe. 1998. La Datation Par Luminescence Des Limons De Crue: Une Nouvelle Approche De L‘Étude Chronologique Des Périmètres D’Irrigation Antiques Du Sud-Yémen. Comptes Rendus De lAcadémie Des Sciences - Earth and Planetary Science 327(1): 31–37. Benoist, A. 2013. A green paradise. economic strategies, collective practices, and local ancestors of the iron age community of Masa fi ̄ ̄ (Emirate of Fujairah, UAE). Proceedings of the Seminar for Arabian Studies 43: 47–61. Berger, J.-F., J.-P. Bravard, L. Purdue, A. Benoist, M. Mouton, and Frank Braemer. 2012. Rivers of the Hadramawt watershed (Yemen) during the Holocene: clues of late functioning. Quaternary International 266: 142–161. doi:10.1016/j.quaint.2011.10.037. Biagi, P., and R. Nisbet. 2006. The prehistoric fisher-gatherers of the western coast of the arabian sea: a case of seasonal sedentarization? World Archaeology 38(2): 220–238. doi:10.1080/00438240600708188. Boucharlat, R. 2003. Iron age water-draining galleries and the Iranian ‘Qanāt’. In Archaeology of the United Arab Emirates, edited by D. Potts, H. Naboodah, and P. Hellyer, 161–172. London: Trident Press. Boucharlat, R. 2016. Chapter 17 Qanāt and Falaj: Polycentric and multi-period innovations. Iran and the United Arab Emirates as Case Studies. In Underground Aqueducts Handbook, edited by A. N. Angelakis, E. D. Chiotis, S. E., and H. Weingartner 279–301. Boca Raton: CRC Press. Bouchaud, C., V. Dabrowski, and M. Tengberg. 2016. Etats Des Lieux De La Recherche Archéobotanique en Péninsule Arabique. Routes De l’Orient Hors série n°2: 21–37. Bowen, R.L. 1958. Irrigation in Ancient Qatabân (Beiḥân). In Archaeological Discoveries in South Arabia, edited by R. L. Bowen and F. P. Albright, 43–131. Baltimore: The Johns Hopkins Press. Breton, J.-F., and J.-C. Roux. 2002. Le Wâdî Durâ’: Un Modèle D’irrigation Antique? Chroniques Yéménites 10: 1–11. Brunner, U. 1997. Geography and human settlement in ancient southern Arabia. Arabian Archaeology and Epigraphy 8(2): 190–202. Brunner, U. 2000. The great dam and the Sabean oasis of Ma’rib. Irrigation and Drainage Systems 14(3): 167–182. Brunner, U. 2004. The great dam of Ma’rib as a part of the hydraulic culture of southern Arabia. In Men of Dikes and Canals. The Archaeology of Water in the Middle East, edited by H.D. Bienert and J. Häser, 135–143. Rahden: Verlag Marie Leidorf. Brunner, U. 2013. The history of irrigation in Lahj oasis. Raydān 8: 51–66. Brunner, U., and H. Haefner. 1986. The successful floodwater farming system of the sabeans, Yemen Arab republic. Applied Geography 6(1): 77–86. Brunswig, R. H. 1989. Cultural history, environment, and economy as seen from an Umm an-Nar settlement: Evidence from test excavations at Bāt, Oman, 1977/78. The Journal of Oman Studies 10: 9–50. Buffa, V. 2002. The stratigraphic sounding at Ma’layba, Lahj Province, Republic of Yemen. Archäologische Berichte aus dem Yemen. Vol. IX: 1–7. Mainz: Philipp Von Zabern. Buffa, V. 2007. Ma’layba Et L’Âge Du Bronze Du Yémen. Archäologische Berichte aus dem Yemen. Vol. XII. Wiesbaden: Ludwig Reichert Verlag. Charbonnier, J. 2008. L’agriculture en Arabie du Sud avant l’Islam: une reconstitution des Paysages et des systèmes de culture antiques. Chroniques Yéménites 15: 3–30. Charbonnier, J. 2014. In the shadow of the palm trees: Time management and water allocation in the oasis of Ādam (Sultanate of Oman). Proceedings of the Seminar for Arabian Studies 44: 83–98.

242

Climate Changes in the Holocene

Charbonnier, J. 2015a. Groundwater management in southeast Arabia from the bronze age to the iron age: A critical reassessment. Water History 7(1): 39–71. doi:10.1007/s12685-014-0110-x. Charbonnier, J. 2015b. The control of water in the Kingdom of Qataba n̄ (Yemen): From local to central management of the irrigation systems during antiquity. Water History 7(4): 473–487. doi:10.1007/ s12685-015-0143-9. Charbonnier, J. 2017. The genesis of oases in Southeast Arabia: Rethinking current theories and models. In Oases and Globalization. Ruptures and Continuities, edited by É. Lavie and A. Marshall, 32: 53–72. Springer Geography. Cham: Springer International Publishing. Charbonnier, J., and K. Hopper. 2018. The Qanāt: A multidisciplinary and diachronic approach to the study of groundwater catchment systems in archaeology. Water History 10(1): 3–11. doi:10.1007/ s12685-018-0214-9. Charbonnier, J., L. Purdue, and A. Benoist. 2017. Taming surface water in pre-Islamic southeast Arabia: Archaeological, geoarchaeological, and chronological evidence of runoff water channeling in Masa fi ̄ ̄ (UAE). Journal of Field Archaeology 42(1): 13–28. doi:10.1080/00934690.2016.1256172. Charbonnier, J., L. Purdue, C. Calastrenc, E. Régagnon, T. Sagory, and A. Benoist. 2017. Ancient agricultural landscapes in Southeast Arabia: Approach and first results of an archaeological, geo-archaeological, and spatial study of the Masa fi ̄ ̄ Palm Grove, Emirate of Fujairah. In Proceedings of Water & Life in Arabia Conference 14th–16th December, 2014, edited by W. Y. al-Tikriti and P. A. Yule, 45–65. Abu Dhabi: Abu Dhabi Tourism and Culture Authority. Cleuziou, S. 1989. Excavation at Hili 8: A preliminary report on the 4th to 7th campaigns. Archaeology in the United Arab Emirates 5: 61–87. Cleuziou, S. 1999. Espace Habité, Espace Utilisé, Espace Vécu Dans Les Communautés d’Arabie Orientale À l’Âge Du Bronze. In Habitat Et Société. XIXe Rencontres Internationales d’Archéologie Et D‘Histoire d’Antibes, edited by F. Braemer, S. Cleuziou, and A. Coudart; 83–108. Antibes: Éditions APDCA. Cleuziou, S. 2005. Pourquoi si Tard? Nous avons pris un autre chemin. L’Arabie des chasseurs-cueilleurs de lHolocène à lAge du Bronze. In Aux Marges Des Grands Foyers Du Néolithique, edited by J. Guilaine, 123–148. Paris: Errance. Cleuziou, S. 2009. Extracting wealth from a land of starvation by creating social complexity: A dialogue between archaeology and climate? Comptes Rendus Geoscience 341: 726–738. Cleuziou, S., M.-L. Inizan, and B. Marcolongo. 1992. Le Peuplement Pré- Et Protohistorique Du Système Fluviatile Fossile Du Jawf-Hadramawt Au Yémen. Paléorient 18(2): 5–29. Córdoba, J. M. 2013. New perspectives about iron age and the oasis culture in the Oman Peninsula. Two conclusive seasons at Al Madam (Sharjah, UAE). Isimu 16: 139–151. Córdoba, J. M., and M. del Carmen Del Cerro. 2005. Archéologie De L‘Eau Dans Al Madam (Sharjah, Emirats Arabes Unis). Puits, Aflaj et sécheresse pendant l’Âge du Fer. Iranica Antiqua 40: 515–532. Costa, P., and T. J. Wilkinson. 1987. The Hinterland of Sohar: Archaeological Surveys and Excavations within the Region of an Omani Seafaring City, 9th ed. Muscat: Ministry of Heritage and Culture. Darles, C., C. J. Robin, J. Schiettecatte, and G. El Masri. 2013. Contribution À Une Meilleure Compréhension De L‘Histoire De La Digue De Ma’rib Au Yémen. In Regards Croisés d’Orient Et D‘Occident. Les Barrages Dans l’Antiquité Tardive, edited by F. Baratte, J. C. Robin, and E. Rocca, 9–70. Paris: De Boccard. Degli Esposti, M. 2011. The excavation of an early bronze age tower near Salut (Bisyah, Sultanate of Oman): The iron age levels. Egitto E Vicino Oriente 34, 198–224. Degli Esposti, M., and A. Benoist. 2015. More on Masafi ancestors: The late bronze age site of Masafi-5. Proceedings of the Seminar for Arabian Studies 45: 57–74. del Cerro, C., and J.M. Córdoba. 2018. Archaeology of a Falaj in Al Madam Plain (Sharjah, UAE); a study from the site. Water History 10 (1): 85–98. doi:10.1007/s12685-018-0210-0. de Maigret, A. 1990. The Bronze Age Culture of Hawlān at-Tiyāl and Al-Hadā (Republic of Yemen): a First General Report. Rome: IsMEO. Desruelles, S., E. Fouache, W. Eddargach, C. Cammas, J. Wattez, T. Beuzen-Waller, C. Martin, et al. 2016. Evidence for early irrigation at bat (Wadi Sharsah, Northwestern Oman) before the advent of farming villages. Quaternary Science Reviews 150(C): 42–54. doi:10.1016/j.quascirev.2016.08.007. Ekstrom, H., and C. M Edens. 2003. Prehistoric agriculture in highland Yemen: New results from Dhamar. Bulletin of the American Institute for Yemeni Studies 45: 23–35. Engel, M., H. Brückner, A. Pint, K. Wellbrock, A. Ginau, P. Voss, M. Grottker, et al. 2012. The early Holocene humid period in NW Saudi Arabia - sediments, microfossils and palaeo-hydrological modelling. Quaternary International 266(C): 131–41. doi:10.1016/j.quaint.2011.04.028.

Human Adaptation in Arabia

243

Engel, M., A. Matter, A. G. Parker, A. Parton, M. D. Petraglia, G. W Preston, and F. Preusser. 2017. Lakes or wetlands? A comment on ‘the middle Holocene climatic records from Arabia: Reassessing lacustrine environments, shift of ITCZ in Arabian sea, and impacts of the southwest Indian and African monsoons’ by Enzel Et Al. Global and Planetary Change 148, 258–267. doi:10.1016/j.gloplacha.2016.11.001. Finlayson, B., J. Lovell, S. Smith, and S. Mithen. 2011. The archaeology of water management in the Jordan valley from the epipalaeolithic to the nabataean, 21,000 BP (19,000 BC) to AD 106. In Water, Life and Civilisation, edited by S. Mithen and E. Black, 191–217. Cambridge: Cambridge University Press. Fleitmann, D., S. J. Burns, A. Mangini, M. Mudelsee, J. Kramers, I. Villa, U. Neff, et al. 2007. Holocene ITCZ and Indian monsoon dynamics recorded in stalagmites from Oman and Yemen (Socotra). Quaternary Science Reviews 26(1–2): 170–188. doi:10.1016/j.quascirev.2006.04.012. Fleitmann, D., S. J. Burns, M. Mudelsee, U. Neff, J. Kramers, A. Mangini, and A. Matter. 2003. Holocene forcing of the Indian monsoon recorded in a stalagmite from southern Oman. Science 300(June): 1–10. Fleitmann, D., S. J. Burns, U. Neff, A. Mangini, and A. Matter. 2003. Changing moisture sources over the last 330,000 years in Northern Oman from fluid-inclusion evidence in speleothems. Quaternary Research 60(2): 223–232. doi:10.1016/S0033-5894(03)00086-3. Fleitmann, D., and A. Matter. 2009. The speleothem record of climate variability in Southern Arabia. Comptes Rendus Geoscience 341(8–9): 633–642. doi:10.1016/j.crte.2009.01.006. Frifelt, K. 1985. Further evidence of the third millennium BC town at Bāt in Oman. The Journal of Oman Studies 7: 89–104. Frifelt, K. 2002. Bat, a Centre in third millennium Oman. In Essays on the Late Prehistory of the Arabian Peninsula, edited by S. Cleuziou, M. Tosi, and J. Zarins, 101–110. Rome: Instituto Italiano per l’Africa et l’Oriente. Fujii, S. 2011. Domestication of runoff surface water: Current evidence and new perspectives from the Jafr pastoral Neolithic. Neo-Lithics 2(10): 14–32. Galili, E., and Y. Nir. 1993. The submerged pre-pottery Neolithic water well of Atlit-Yam, Northern Israel, and its palaeoenvironmental implications. The Holocene 3(3): 265–270. doi:10.1177/095968369300300309. Gebel, H.G.K. 2013. Arabia’s fifth-millennium BCE pastoral well cultures: Hypotheses on the origins of oasis life. Proceedings of the Seminar for Arabian Studies 43 (August): 111–126. Gebel, H.G.K. 2016. The socio-hydraulic foundations of oasis life in NW Arabia. the 5th millennium BCE shepherd environs of Rajajil. In The Archaeology of North Arabia Oases and Landscapes, edited by M. Luciani, 79–113. Vienna: Austrian Academy of Sciences Press. Gebel, H.G.K. 2017. The origins of oasis life in NW Arabia a model based on the Qulban Beni Murra and Rajajil case study regions, and the need of archaeohydrology as a discipline for studying Arabia’s past. In Proceedings of Water & Life in Arabia Conference 14th–16th December, 2014, edited by W. YA alTikriti and P. A. Yule, 1–26. Abu Dhabi: Abu Dhabi Tourism and Culture Authority. Gebel, H.G.K., and H.M. Mahasneh. 2013. Disappeared by climate change the Shepherd cultures of Qulban Beni Murra (2nd half of the 5th millennium Bc) and their aftermath. Syria 90 (October): 127–158. Gentelle, P. 1991. Les Irrigations Antiques À Shabwa. Syria 68(1): 5–52. doi:10.3406/syria.1991.7261. Ghaleb, A.O. 1990. Agricultural Practices in ancient Radman and Wadi al-Jubah (Yemen). Unpublished PhD thesis. Philadelphia: University of Pennsylvania. Harrower, M.J. 2008a. Hydrology, ideology, and the origins of irrigation in ancient southwest Arabia. Current Anthropology 49(3): 497–510. Harrower, M.J. 2008b. Mapping and dating incipient irrigation in Wadi Sana, Hadramawt (Yemen). Proceedings of the Seminar for Arabian Studies 38: 187–202. Harrower, M.J. 2009. Is the hydraulic hypothesis dead yet? Irrigation and social change in ancient Yemen. World Archaeology 41(1): 58–72. doi:10.1080/00438240802668354. Hastings, A., J.H. Humphries, and R.H. Meadow. 1975. Oman in the third millennium BCE. The Journal of Oman Studies 1: 1–55. Hausleiter, A. 2016. Tayma, Saudi-Arabien. Multidisziplinäre Forschungen in Der Oase. German Archaeological Institute, Research Reports. Hehmeyer, I. 1989. Irrigation farming in the ancient oasis of Mārib. Proceedings of the Seminar for Arabian Studies 19: 33–44. Kendal, J., J.J. Tehrani, and J. Odling-Smee. 2011. Human niche construction in interdisciplinary focus. Philosophical Transactions of the Royal Society B: Biological Sciences 366(1566): 785–792. doi:10.1098/rstb.2010.0306. Luciani, M. 2016. Mobility, contacts and the definition of culture(S) in new archaeological research in Northwest Arabia. In The Archaeology of North Arabia Oases and Landscapes, edited by M. Luciani, 21–56. Vienna: Austrian Academy of Sciences Press.

244

Climate Changes in the Holocene

Magee, P. 2004. The impact of southeast Arabian intra-regional trade on settlement location and organization during the iron age II period. Arabian Archaeology and Epigraphy 15: 24–42. Magee, P. 2014. The Archaeology of Prehistoric Arabia. Cambridge: Cambridge University Press. Magee, P. 2015. When was the dromedary domesticated in the ancient near east? Zeitschrift Fur OrientArchaologie 8: 252–277. Maraqten, M. 2017. Typology of irrigation systems in ancient Yemen in the light of the epigraphic evidence. In Proceedings of Water & Life in Arabia Conference 14th–16th December, 2014, edited by W. Yasin al-Tikriti and P.A. Yule, 115–136. Abu Dhabi: Abu Dhabi Tourism and Culture Authority. Meister, J., Jan Krause, B. Müller-Neuhof, M. Portillo, T. Reimann, and B. Schütt. 2016. Desert agricultural systems at EBA Jawa (Jordan): Integrating archaeological and paleoenvironmental records. Quaternary International 434(B): 33–50. doi:10.1016/j.quaint.2015.12.086. Mouton, M. 2009. L’eau en partage: territorialité, réseaux d’irrigation et formation des sociétés antiques dans les Basses-Terres du Yémen. In Stratégies d’acquisition de l‘eau et sociétés au Moyen-Orient depuis l’Antiquité: Études de cas, edited by M. Al-Dbiyat and M. Mouton, BAH 156: 79–95. Beyrouth: Presses de l’IFPO. Orchard, J., and J. Orchard. 2007. The third millennium BC oasis settlement and the first evidence of their irrigation by Aflaj From Bahla, In Archaeology of the Arabian Peninsula Through the Ages. Proceedings of the International Symposium (7th–9th May 2006), 143–173. Muscat: Ministry of Heritage and Culture. Parker, A.G., L. Eckersley, M.M. Smith, A.S. Goudie, S. Stokes, S. Ward, K. White, et al. 2004. Holocene vegetation dynamics in the northeastern Rub’ Al-Khali Desert, Arabian Peninsula: a phytolith, pollen and carbon isotope study. Journal of Quaternary Science 19(7): 665–676. doi:10.1002/jqs.880. Parker, A.G., A.S. Goudie, S. Stokes, K. White, M.J. Hodson, M. Manning, and D. Kennet. 2006. A record of holocene climate change from lake geochemical analyses in southeastern Arabia. Quaternary Research 66(3): 465–476. doi:10.1016/j.yqres.2006.07.001. Parr, P., G.L. Harding, and J.E. Dayton. 1970. Preliminary Survey in N.W. Arabia, 1968. London: Institute of Archaeology. Peltenburg, E., S. Colledge, P. Croft, A. Jackson, C. McCartney, and M.A. Murray. 2000. Agro-pastoralist colonization of Cyprus in the 10th millennium BP: Initial assessments. Antiquity 74(286): 844–853. doi:10.1017/S0003598X0006049X. Power, T., and P. Sheehan. 2011. Bayt Bin ‘Ātī in the Qattārah Oasis: A prehistoric industrial site and the formation of the oasis landscape of Al-’Ain, UAE. Proceedings of the Seminar for Arabian Studies 41: 267–282. Preston, G.W., D.S.G. Thomas, A.S. Goudie, O.A.C. Atkinson, M.J. Leng, M.J. Hodson, H. Walkington, et al. 2015. A multi-proxy analysis of the Holocene humid phase from the United Arab Emirates and its implications for southeast Arabia’s Neolithic populations. Quaternary International 382: 277–292. doi:10.1016/j.quaint.2015.01.054. Sanlaville, P. 2000. Le Moyen-Orient Arabe. Le Milieu Et L’homme. U. Géographie. Paris, France: Armand Colin. Tengberg, M. 2003. Archaeobotany in the Oman peninsula and the role of eastern Arabia in the spread of African crops. In Food, Fuel and Fields. Progress in African Archaeobotany, edited by K. Neumann, A. Butler, and S. Kahlheber, 229–237. Cologne: Heinrich Barth Institut. Vogt, B. 2004. Toward a new dating of the great dam of Mārib. Preliminary results of the 2002 fieldwork of the German Institute of Archaeology. Proceedings of the Seminar for Arabian Studies 34: 377–388. Weisgerber, G. 1981. Mehr Als Kupfer in Oman. Ergebnisse Der Expedition 1981. Der Anschnitt 33: 174–263. Weisgerber, G, and P. A. Yule. 2003. Al-Aqir Near Bahlā’—An early bronze age dam site with planoconvex ‘copper’ ingots. Arabian Archaeology and Epigraphy 14: 24–53. Wellbrock, K., M. Grottker, and H.G.K. Gebel. 2017. Archaeohydrological investigation in NW arabia potentials, problems, needs and goals. In Proceedings of Water & Life in Arabia Conference 14th–16th December, 2014, edited by W.Y. al-Tikriti and P.A. Yule, 27–44. Abu Dhabi: Abu Dhabi Tourism and Culture Authority. Wellbrock, K., P. Voss, and M. Grottker. 2012. The evolution of water management methods in north-western Arabia and the south- Ern Levant from the Neolithic age through antiquity. Schriften Der Deutschen Wasserhistorischen Gesellschaft 20(1): 29–46. Wellbrock, K., F. Weigel M. Strauss, A. al-Souliman, C. Külls, and M. Grottker, in press. Archaeohydrological investigations in the oasis of Tayma, Saudi-Arabia, during seasons 2013 to 2015. Atlal.

Human Adaptation in Arabia

245

Wilkinson, T.J. 1999. Settlement, soil erosion and terraced agriculture in highland Yemen: A preliminary statement. Proceedings of the Seminar for Arabian Studies 29: 183–191. Wilkinson, T.J. 2003. Archaeological Landscapes of the Near East. Tucson: The University of Arizona Press. Wilkinson, T.J., and C.M. Edens. 1999. Survey and excavation in the central highlands of Yemen: Results of the Dhamār survey project, 1996 and 1998. Arabian Archaeology and Epigraphy 10: 1–33. Zielhofer, C., K. Wellbrock, A.S. al-Souliman, M. von Grafenstein, B. Schneider, K. Fitzsimmons, A. Stele, et al. 2018. Climate forcing and shifts in water management on the northwest Arabian Peninsula (Mid-Holocene Rasif Wetlands, Saudi Arabia). Quaternary International 473(Part A): 120–140. doi:10.1016/j.quaint.2018.03.001.

12

Hydraulic Cultures and Hydrology under Climatic Change North Arabian Mid-Holocene Pastoral and Proto-Oasis Land Use Hans Georg K. Gebel and Kai Wellbrock

CONTENTS 12.1 Introduction........................................................................................................................... 247 12.2 Vulnerability: Interacting Climate, Hydrology, Land Use and Culture............................................................................................................ 249 12.3 Degradation: Mid-Holocene Climate in Northern Arabia.................................................... 253 12.4 Adaptation: From Mobile Pastoralism to Sedentary Gardening........................................... 255 12.4.1 The 5th Millennium BCE Sepulchral and Hydraulic Habitation Landscapes.......... 257 12.4.2 The 4th Millennium BCE Hydraulic and Sepulchral Oases Sites............................. 259 12.5 Engineering: Water Management and Hydraulic Features.................................................... 259 12.5.1 (Pre-) Historic Arid Land Water Management.......................................................... 259 12.5.1.1 Preparing Catchments/Modifying Landscape/Changing Watersheds.......260 12.5.1.2 Qas, Khabrats, Sabkhas, etc. as a Parts of Artificial Landscape Alteration���������������������������������������������������������������������������������������������������260 12.5.1.3 Groundwater Management..........................................................................260 12.5.2 (Pre-) Historic Arid Land Hydraulic Structures........................................................260 12.5.2.1 Wells and Related Techniques....................................................................260 12.5.2.2 Dams, Diversion Canals and Similar.......................................................... 261 12.5.2.3 Troughs....................................................................................................... 261 12.5.3 Rasif Case Study........................................................................................................ 261 12.6 Sustainability: Lessons from the Past.................................................................................... 262 12.6.1 Mugur Systems: e.g. Bweitat 1a................................................................................. 263 12.6.2 Khabra Systems: Open Ponds for Retaining Surface Runoff....................................264 12.6.3 Mshash or Thamil’: Temporary Tapping of Shallow Groundwater...........................264 12.6.4 General Statement on Traditional Systems................................................................264 12.7 Archaeohydrology: Integrated and Holistic Approaches to Holocene Adaptations.............. 265 12.8 Summary...............................................................................................................................266 References.......................................................................................................................................266

12.1 INTRODUCTION What people of discrete generations might experience as a severe ecological disaster affecting or terminating their lifeways and (re-) production may be understood by following generations as an advantageous ecological adaptation that created more sustainable lifeways and safer (re-) production. 247

248

Climate Changes in the Holocene

This chapter describes such a trajectory in Mid-Holocene human history, finally establishing the sustainable sedentary oases lifeways in the Arabian lands: Resulting in wide-spread degradation and vulnerability of human, animal and plant systems, the major climatic shift towards drier conditions from 4200 BCE* caused human adaptation into oasis life and related hydraulic engineering. This adaptation made the arid lands for the first time and for good suitable for a sustainable sedentary life. While the socio-economic and economic frameworks of the oasis paradigm was new, and later became subject of migration to other arid regions of the Old World, it may not have altered the deeply relational dispositions of semi-arid and arid land societies: The extensive but partial transformation of the 5th millennium’s BCE mobile shepherd societies into sedentary horticultural societies from the late 5th millennium appears to have preserved the mental traits of early tribalism, as expressed by their sepulchral behaviour and traditions. The sociohydraulic trajectory from mobile pastoralism to oasis horticulturalism is the subject of two transdisciplinary archaeohydrological projects† which follow the question how social and economic life and related water management and land use adapted to vanishing steppes, lakes and lowering of water tables after 4200/4000 BCE. Since this new research field still is grounded on yet meagre evidence and few absolute dates, research is following a system of testable theses, steadily improved by new results from field and lab work. The projects’ negotiations of the transdisciplinary approaches exposed the need and demand to establish Archaeohydrology as a formal discipline without which arid Arabia’s archaeology would not develop properly. In addition, field work exposed the imperative (Gebel n.d. 1) to document the sustainable traditional Bedouin water management as a mean of future sustainable water management for Arabia’s deserts; this would supplement Archaeohydrology in terms of an Applied Archaeohydrology (Figure 12.1). The major research objectives/questions of the projects are, aside from documenting the material cultures, 1) to evaluate if and how the pastoral progenitor cultures and their hydrological and hydraulic competence have become the socioeconomic substratum of oasis life; 2) to reconstruct how the supra-regional climate shift from around 4200/4000 towards drier affected the pastoral steppes and their palaeohydrology, and how this process could have encouraged a sedentarization process based on horticultural oases socioeconomies; 3) to identify proto-oases potentially developing at water-favoured localities during the 5th millennium BCE; 4) to describe the character of social and cogntive alteration taking place by this sedentarization; and 5) to understand to what extent mobile pastoral life co-existed and interacted with early oasis life during the 4th millennium. By the 2015 season at Rajajil, solid stratigraphical evidence proved the hitherto-claimed thesis that mobile shepherd cultures were followed by permanent oasis settlements, in the shape of ashlar-line graves getting rebuilt into domestic structures, most likely in a proto-oasis. However, the aforementioned trajectory is based on a model using the previously mentioned system of testable theses and meta-theoretical approaches connecting research from several disciplines (for the development of the updated theses, cf. Gebel 2013, 2016, 2017a). We are far from the needed comparative high-resolution environmental and societal reconstructions for the regions and greater regions under study to confirm this trajectory. As a matter of fact, the deflated arid lands hardly offer preserved sedimentary environments and organic matter to easily do such reconstructions; here, we may have to invest more and more time and persistence than in the more favoured regions. All dates used in this contribution refer to calBC, respectively calBCE. This contribution contains summarised information for the purposes of a handbook; detailed information on the topics and findings discussed here is published in: Gebel 2013, 2016, 2017a; Gebel and Mahasneh 2012, 2013; Wellbrock, Voß, and Grottker 2012; Wellbrock, Gottker, and Gebel 2017; Zielhofer et al. 2018. The goals of our research are also guided by other research results, for example, the archaeohydrological research at Tayma conducted by M. Grottker and K. Wellbrock (Wellbrock, Voß, and Grottker 2011; Wellbrock et al. 2017), or the palaeoenvironmental research carried out by Dinies and others (Dinies, Neef, and Kürschner 2011; Dinies et al. 2015, 2016) and Engel et al. (2012). References to these publications are only made in this contribution’s text when specific information is referred to. † The two related projects in Northern Arabia are the Saudi-German Rajajil/Standing Stones Joint Archaeological Project/ Rajajil (RJJ 2012–2017) near Sakakah in the Kingdom of Saudi Arabia, and the Eastern Jafr Archaeological Project/ Qulban Beni Murra (EJP 2001 ff.) in the southeastern Kingdom of Jordan. *

Hydraulic Cultures and Hydrology under Climatic Change

249

FIGURE 12.1  Towards an integrated and holistic research in water history: The Archaeohydrology Aggregate. (Graph: H.G.K. Gebel.)

More and more the modelling of climate/environment – society/culture co-evolution – is subject of disciplinary cooperation (cf. also Chiotis, this volume, Chapter 1). Indeed, this is a most-needed research attitude if mankind should be given the chance to receive and use lessons from the past. Researchers of this “new era in the progress of earth sciences and humanities” and their “strong message of holistic approach in research” (wording by E. Chiotis, in letter) became the “translators” of the past’s messages, aiming to serve mankind’s future.

12.2 VULNERABILITY: INTERACTING CLIMATE, HYDROLOGY, LAND USE AND CULTURE Contrary to the stratified hydraulic cultures in historic Mesopotamia and Egypt, the Mid-Holocene development of human occupations, water use and interaction with nature in Arabia’s (hyper-) arid water-deficit regions followed other conditions and frameworks. Vulnerability patterns of the latter were more a matter of formal structures applying complex adaptive ideocratic systems to support social and economic balance by, for example, using innovation on all levels, large-scale reorganisation of land use, organised territorial conflict/warfare, including forced migrations/deportations, and so on. On the contrary, arid Arabia’s hydraulic societies* always seem to have operated on flat hierarchies applying traditional/conservative water management systems and conservative technologies, using the social frameworks of habitus apparatuses (cf. Gebel 2017b) for managing and balancing their basically relational social systems; at least, we so far have no evidence for other *

The term “hydraulic societies” is to be understood sensu Gebel 2016 and 2017a (arid Arabia’s hydraulic societies, and not sensu Wittfogel’s hydraulic societies (Wittfogel 1957). Hydraulic societies (or hydraulic empires, hydraulic despotism) sensu Wittfogel mean political and social structures using administrative power to distribute irrigation water, as is characteristic for the hydraulic (proto-) historic Nile and Mesopotamian civilisations.

250

Climate Changes in the Holocene

forms of social organisation than chiefdoms in the Mid-Holocene. Such cognitive and “frugal” socioeconomic frameworks, of course, are the direct result of, and reason for, the different adaptive strategies for coping with the different vulnerability patterns in arid lands. A recent work illustrates, by the example of the 8.2 ka calBP Event in the Western Fertile Crescent, that “even periods with recurrent short-lived hazards can influence human behaviour and cultural development” in moderate regions (Clare 2016: 212). While this statement is a most important reminder that climate deterministic approaches should not be neglected on account of the “cultural deterministic” approaches characteristic for humanities, one should not underestimate the potentials of socioeconomic adaptations in complex productive systems in moderate zones. This is especially true for the past hydraulic civilisations developing along the world’s great rivers. However, in regions easily vulnerable to desiccation, climate determinism has to dominate in the research. Different types of vulnerability permanently interact during periods of stability and instability. Basically, environmental, social and ideological vulnerabilities form a vulnerability system. When water resources are substantial for all development, we may speak of hydraulic vulnerability systems. Figure 12.2 shows the major interacting fields and agents of vulnerability in Arabia’s MidHolocene societies. In general, one may conclude from evidence that mobile pastoral societies would have continued their conservative lifeways, water management, cognitive systems and belief systems if not dramatically forced to adapt into sedentary socioeconomies by the onset of drier conditions from 4200/4000 BCE onwards. As stated before, the crucial cogwheel for change in these societies was climate forcing Gebel 2016, 2017a; Zielhofer et al. 2018); once it started to rattle or move, the other fields of vulnerability were affected (Figure 12.2). Climate forcing was the direct key agent of change in arid Arabia’s past: The purely nature-related steppe pastoralists of Arabia’s MidHolocene could react only within the limits of their environmental frameworks, as is true for the following oasis economies which show a similar conservative character and even similar traits in cultural and socioeconomic expression. Once other risk-buffering options were available, like the socioeconomic and innovation strategies, including high-level aggression of the stratified complex societies in the contemporary Fertile Crescent and Nile Valley (flourishing also under different climatic regimes), climate change had a less direct impact.

FIGURE 12.2  Water vulnerability research aggregate: Driving forces and agents of vulnerability, and related archaeohydrological research fields of the arid land hydraulic/(archaeo-) hydrological topics. (Graph: Gebel.)

Hydraulic Cultures and Hydrology under Climatic Change

251

Land use and culture in Arabia’s Mid-Holocene 5th millennium BCE sepulchral and hydraulic steppe landscapes were based on extensive long-distance networks with regional hubs in the shape of (often megalithic) burial and well fields with ritual and domestic structures (Figures 12.3 through 12.5); smaller places with sepulchral and watering functions as well as niche agriculture* were part of the networks. Here, wells also tapped intermediate and deep groundwater. Regional hubs like Qulban Beni Murra, Rajajil and Rasif (Figure 12.6) appear to have played the major role in social transaction and tribal identity, assisted by ancestor commemoration. Increasing aridification after 4200/4000 BCE caused hydrologically suitable locations to host more (sensu “in situ

FIGURE 12.3  Sepulchral and watering landscapes: Panoramic view of the megalithic pastoral centre of Qulban Beni Murra (2nd half of the 5th millennium BCE) with its well-trough structures at the wadi bottom (Area D) and its burial fields (Area E) and ritual structures (Area A) on the slopes. (EJP Project; photo: H.G.K. Gebel.)

FIGURE 12.4  Wells and troughs: Example of watering Place D15 of Qulban Beni Murra (2nd half of the 5th millennium BCE), consisting of well mouth in well room with staircase, two canal-type trough lines ending in roundish basins, spaces of unknown functions. (EJP Project; graph: C. Purschwitz.) *

So far, we could not identify in our research areas sites with large fields irrigated by rainwater harvesting systems, as reported by Müller-Neuhof (2014) from Northern Jordan. We cannot exclude the existence of such land use and economies for the 6th–4th millennia BCE while we met this kind of agri-/horticulture for (sub-)recent times. In our survey areas, we so far identified only smaller spots with terraced soil-catching drainages which we address as niche agriculture. Several of them are in the neighbourhood of 5th–4th millennium pastoral camp sites and show an ill-preserved system of succeeding terrace walls in drainages leading off from a potential or suitable water harvesting catchment (flat and large surfaces above drainages).

252

Climate Changes in the Holocene

FIGURE 12.5  Gradient troughs: Watering Place D15 of Qulban Beni Murra (2nd half of the 5th millennium BCE), (top) canal-type trough line with descending compartments, (bottom) interior of a trough’s/compartment’s pavement delineated by ashlars. (EJP Project; photos: H.G.K. Gebel.)

FIGURE 12.6  Key locations: Sepulchral and watering centres of the 5th–early 4th millennium BCE in the Greater Wadi Sirhan Region, Northern Arabia. (EJP/RJJ Projects; graph/cartography: P. Voss.)

adaptation”, cf. Chiotis, this volume, Chapter 1). This “oasisation” of the former pastoral hydraulic and sepulchral landscapes might have started, as mentioned before, already in the 5th millennium BCE by the establishment of proto-oases (e.g. Rajajil) at hydrologically favourite spots. However, this trajectory does not mean that pastoral steppe use and the associated hunting of migrating ungulates and niche agriculture ever disappeared during oasis times: All evidence from the 4th millennium BCE onwards indicates shifting reliances between both economies/lifeways, controlled by climate oscillations and potential human (mis-) management of water and land while oasis lifeways were progressing. In all these shifts, migration to other locations most likely was the last option for the late pastoralists/early oasis horticulturalists. Rather, they were better able to

Hydraulic Cultures and Hydrology under Climatic Change

253

adapt to water deficiency than to social turbulence in the settlements. Mobile pastoralism, on the other hand, offered a broader range of adaptation and niche economy under climate change (cf. also Pei et al. 2017 with his example from medieval China). Another type of vulnerability is represented by the state of preservation of the open land occupations. Dry-land archaeology/archaeohydrology and land-use research in Arabia’s deflated environments not only requires one to identify hardly preserved pockets of sedimentary environments and installations, but also the geographical extensions of the (pre-) historic landscape cultures make it difficult to trace related network elements and catchments.

12.3 DEGRADATION: MID-HOLOCENE CLIMATE IN NORTHERN ARABIA Arid Arabia’s environments always reacted most sensitively to even minor climate oscillations, which had major impacts on population dynamics, land use, water management and thus cultural developments. Until recent times, degradation through aridification did not cause major unsustainable developments, since the limited ecosystems did not allow unsustainable adaptations or innovations. Rather, the severe environmental conditions forced sustainable adaptation. In modern times, however, brought-in technologies allow overcoming these limits (water pumps, deep wells, exploitation of fossil waters, the various irrigation systems, fertiliser, etc.) and establishing agri- and horticultures on industrial scales (cf. Section 12.6). Extensive olive yards exist in sand dunes (Figure 12.7), and gladiola grow in greenhouses: The natural consequences of climate and hydrological conditions are suspended. For the Mid-Holocene, here formally defined on supra-regional levels by following Walker et al. (2012) as the period between the 8.2 ka BP Event (cooling episodes) and the 4.2 ka BP Event (aridification episodes), we hardly have relevant and non-contradicting data on climate developments for Northern Arabia. Thus, we need to rely on supra-regional information for our research areas (following Zielhofer et al. 2018), including the comparison with Early Holocene conditions: The major changes in Holocene climate and hydrology are driven by alternating influences of monsoonal air masses due to insulation forces. For the Early Holocene, palaeoclimatic and palaeoenvironmental studies from the Arabian Peninsula (Arz et al. 2003; Fleitmann et al. 2003; Radies et al. 2005; Fuchs and Buerkert 2008) as well as from the Western (Tierney, Pausata, and deMenocal 2017;

FIGURE 12.7  A common picture in Northern Arabia: Bulldozed land surfaces preparing new fields, and an olive yard on the dunes sustained by drippler irrigation with water from deep wells; near Rajajil 2010. (Photo: H.G.K. Gebel.)

254

Climate Changes in the Holocene

Zielhofer et al. 2017a–b) and Eastern Saharan region (Kröpelin et al. 2008) reveal more humid conditions during the Early Holocene because of the northward shift of monsoonal air masses at that time. Today’s hyperarid regions were characterised by steppe environments in the Early Holocene, although climatic deterioration with increasing aridity towards the Mid-Holocene is in evidence (e.g. Migowski et al. 2006; Weninger et al. 2006; Zielhofer et al. 2017 a-b); these deteriorations correspond with global Rapid Climate Changes (RCCs) (e.g. Mayewski et al. 2004; Wanner et al. 2015). However, the 6th and earlier parts of the 5th millennium of the Arabian Peninsula still were rather humid; lacustrine and steppe environments are well attested (e.g. Masry 1974; McClure 1976; Whitney, Faulkender, and Rubin 1983; Schultz and Whitney 1986; Gebel et al. 1989; Arz et al. 2003; Fleitmann et al. 2003, 2011; Lézine et al. 2007; Dinies, Neef, and Kürschner 2011; Dinies et al. 2015; Engel et al. 2012) and were the base of the aforementioned mobile pastoralism with regional hubs and extensive networks. In hydrological and vegetational terms, the 6th to early 5th millennium’s moister climate promoted a denser vegetation (steppe grasslands) sustained by positively interrelated factors: reduced or decelerated surface runoff volumes, reduced sediment transport, enhanced groundwater recharge triggering higher aquifers and soil formation promoted by reduced sediment transport. Rainfall reliability must have been higher, making water sources more permanent and periodical, less episodical. Reliable data on Holocene annual precipitation from Northern Arabia are very limited (Enzel et al. 2015, 2017; Engel et al. 2017). Palaeo-rainfall during the Early Holocene (8000–6500 BCE) has been estimated by Wellbrock, Voß and Grottker (2011) and Engel et al. (2012) for the oasis of Tayma (c. 350 km Southwest of Rasif/Rajajil) to have reached 150 ± 25 mm per year and thus 2–3 times higher than today. The RCC interval, or climatic deterioration with a significant decrease of moisture, relevant for this presentation, started in the 2nd half of the 5th millennium BCE in Mediterranean North Africa (e.g. Ibouhouten et al. 2010), the Eastern Mediterranean and Middle East (Bar-Matthews and Ayalon 2011; Benito et al. 2015; Preston et al. 2015; Dinies et al. 2016); for the neighbourhood of one of our immediate research areas, a significant shift towards more arid conditions from around 4300 BCE is attested (Dinies et al. 2016). Among other consequences, populations started to decrease in the Southern Levant (Weninger et al. 2009) and the Arabian Peninsula (Staubwasser and Weiss 2006; Gebel 2010a, 2013, 2016; Gebel and Mahasneh 2012, 2013). Our recent sedimentological and geoarchaeological findings (Zielhofer et al. 2018) from the Rasif deposits (Figures 12.8 and 12.9) confirm this progressive aridification. Well and trough building appears to be the reaction in a location that formerly was an area of natural and artificial ponds. The initial results of our archaeohydrological study at Rasif as well as the site’s huge potential to understand land use and water management in Mid to Late Holocene Northern Arabia should not “entrap” us to ignore a basic objection to be made for this kind of research in arid environments: results from even long-term palaeoenvironmental and archaeohydrological archives like Rasif (6th millennium BCE up to recent times) may not be representative for the region, period and beyond; they may rather reflect the high levels of variability we have to expect for the spectrum of adaptive behaviour in Arabia’s “frugal” hydraulic cultures. The environmental knowledge of these palaeoBedouins, and the related adaptive hydrological competencies, might have created specific and timeless local adaptations characteristic for the conditions at such spots. Rasif, for example, remained a watering place throughout the millennia, just hosting occasionally other occupational evidence throughout time (Gebel 2016; Zielhofer et al. 2018; Figures 12.8 and 12.9). It also should be mentioned that it possibly is too simplistic to expect a one-way development towards oases economies on the Arabian Peninsula, starting from the 4200/4000 BCE degradation of steppes. Instead, oscillations may rather have caused a more complex and polycentric formation of Arabia’s oases in the Early Bronze Age – respectively after 4000 BCE – (Dinies et al. 2016; Tengberg 2012; Beech and Shepherd 2001), after a larger part of the pastoral population had to abandon mobile life modes and continued to use their well and canal technologies as permanent oases settlers in remaining areas with higher water tables.

Hydraulic Cultures and Hydrology under Climatic Change

255

FIGURE 12.8  Rasif Case Study: Major occupation areas in the site’s central part and its hydrological system. (RJJ Project; graph: K. Wellbrock.)

12.4 ADAPTATION: FROM MOBILE PASTORALISM TO SEDENTARY GARDENING The key thesis* of our research is that mobile 5th millennium pastoral hydraulic communities experienced in steppe water management (especially well and canal-type trough building) were establishing the first sedentary proto-oases and oases as a result of climate forcing/environmental degradation†. From 4200/4000 BCE and during the 4th millennium, the pastoral people of today’s Arabian deserts increasingly concentrated at water-favoured locations which partly were occupied by wild date palm stands. Due to their ecological makeup, not all regions may have successfully participated in this transition, and parts of the Arabian Peninsula may also have continued to It is out of this contribution’s focus to present our Oasisation Model which is published in its latest update in Gebel (2017a), Frame 1. This model represents an ever-updated set of testable theses on the hypothetical socioeconomic and land-use trajectory from well-based mobile pastoralism to sedentary oasis life in Arabia (5000–3500/3000 BCE) constantly steered by fresh field data and accordingly re-worked meta-theoretical considerations. † A similar trajectory is described by di Lernia, this volume (Chapter 9), for the Saharan lands as the “onset of the desert and ethnic fragmentation: Late Pastoral herders (5.9–3.4 ka)”: this study, and the other quoted there, illustrates how advanced the Saharan Mid-Holocene research already developed if compared with that on North Arabia. While we should not expect that the regional diversity of the Saharan lands, of the Sinai/an-Naqab and of the Arabian lands allow direct comparisons of regional climate-culture co-evolutions – as also pointed out by di Lernia – we possibly have to expect overall similar adaptive strategies and developments in this cross-continental Mid-Holocene web of polycentric reaction and interaction during these “green to brown desert” processes (wording by di Lernia). *

256

Climate Changes in the Holocene

FIGURE 12.9  Rasif Case Study: Various phases of water management throughout Rasif’s occupation over millennia. (RJJ Project; graph: K. Wellbrock.)

sustain on much less mobile pastoralism, witnessing a regression of their population (Gebel 2010a, 2013, 2016; Gebel and Mahasneh 2012, 2013). This trajectory appears to have happened throughout the 5th millennium’s “Mid-Holocene Green Saharo-Arabian Pastoral Belt” (Gebel and Mahasneh 2012, 2013; Harrower 2008). The pastoral traditions of steppe hydraulic engineering most likely date back to the beginnings of pastoralism, that is, the 8th millennium BCE, when parts of the Late Pre-Pottery Neolithic B populations in the Southern Levant became mobile as a consequence of social pressure in their villages (Gebel 2014). Well building is known from sedentary villages since the middle of the 9th millennium BCE (Peltenburg 2001). However, to our knowledge wells have not yet been identified, excavated and dated for North Arabia’s 7th and 6th millennium BCE, since little archaeological experience is around to identify such wells in landscapes. The latter is one of the reasons why archaeology should not enter hydraulic landscapes without archaeohydrological expertise. Little hydraulic innovation investment has to be expected for the adaptation into oasis hydraulics when shifting from pastoral well and trough building to well and canal building. Actually, the demanded hydrological and hydraulic competence is the same, and this might be the reason why – in these terms – the transition to oasis appears smooth. The chains of troughs connected by gradient canal-type units (e.g. Figure 12.5) used by the mobile pastoralists are similar to the canals and basins expected to have been used in the early oases from the beginning. Little is known from the supposed proto-oases and early oases; all hints that this was not a linear trajectory and that Arabia’s regions established oases at different times by different ecological

Hydraulic Cultures and Hydrology under Climatic Change

257

adaptations with different operational modes for gardens and fields; the classical date palm oasis must not have been the common type. This overall trajectory of adaptation is described in more detail in the following two subsections, supported by Table 12.1, which summarises available basic information as well as evidence to be expected for the developments from the Late Neolithic to the oasis times (6th–4th millennia). It refers to the hypothetical understanding that 1) the 5th millennium represents a period of long-range pastoral mobility together with large-scale sepulchral land use and regional hubs which could have developed into and proto-oases after 4500 BCE; 2) in the 4th millennium we deal with short-range adaptive variability in pastoral, sepulchral and horticultural land use, increasingly based on sedentary oases of various types; while 3) the sociohydraulic, sepulchral and hydroethic dispositions of lifeways and land use remained conservative or stable.

12.4.1 The 5th Millennium BCE Sepulchral and Hydraulic Habitation Landscapes All archaeological and archaeohydrological evidence we so far have from Northern Arabia indicates that its 5th millennium environments hosted large groups of mobile pastoralists who dug wells into the wadi floors or in the proximity of seasonal or permanent lakes; fed their flocks at well-fed trough systems; managed locally surface runoff by various types of dam systems, ponds and reservoirs; constructed pens and human shelters; and gathered at burial grounds to perform their funeral practices and manifest identity by commemorating ancestral ties while negotiating social relations. Structures, including troughs, may have been built by larger or smaller ashlars once a local banked bedrock allowed it; the culture tends to express “megalithically”. The key sites so far identified are Rajajil, Rasif, Qulban Beni Murra (Figure 12.6) and possibly Rizqeh in greater Wadi Rumm area, Jordan (Kirkbride 1960, 1969). Numerous but less impressive sepulchral and hydraulic habitation sites are to be found along the routes leading off from such centres. Basically, such sites are known from all parts of the Arabian Peninsula, for example, several sites listed by al-Ghazzi (2004). These pastoral cultures are contemporaneous with the stratified Chalcolithic towns in the environmentally more favoured areas of the Fertile Crescent and might already helped cultural elements to directly cross Arabia’s steppes at that time, for example, connecting the Mesopotamian Lowlands with the Southern Levant and Egypt. For the generally different agricultural trajectories of present-day arid Arabia from those in the Fertile Crescent, cf. J. Charbonnier, this volume, Chapter 11. The sites’ general characteristics include (information supplemented by survey sites documented in both research regions): 1. Sites are related to favoured water harvesting possibilities of various kinds; in addition, they often lie at geographical “corridors” or at crossroads. Favoured locations for well sites are, for example, aquifers pressed near-surface by the buried rock topography; for dam systems: for example, natural or easily modifiable topographies allowing the collection of seasonal runoff; for other water collecting systems: for example, extensive, slightly inclining surfaces allowing a controlled collection of water by a system of deflection dams; and so on. 2. At least the central sites had both a watering and sepulchral function; their structural inventory includes graves of various types, watering places (wells with troughs) and other hydraulic installations (e.g. dams), ritual buildings, domestic structures and related installations, pens, and so on. 3. They remained aceramic as long as used only by mobile shepherds; their prominent chipped stone tool is the fan scraper. 4. Site inventories prove a modest material culture much related to grave goods; few household items are attested (mostly grinding tools, rough stone vessels and large troughs). Cupmarks and rock art are associated with structures and surrounding rock outcrops.

258

Climate Changes in the Holocene

TABLE 12.1 The Development of Hydraulic and Sepulchral Land Use in Northern Arabia’s Mid-Holocene (6th–4th Millennia BCE) Mid-Holocene Land Use

Land Use and the Attested Landscapes’ Structural Inventory without parenthesis: attested in parenthesis: expected/anticipated evidence

General Socioeconomic Trajectory (Continuing Tribal Organisation in Early Oases)

6th millennium (Late Neolithic- “Chalcolithic”)

– seasonal inselberg (isolated rock ridge or small mountain rising abruptly from surrounding plain) and open-plain settlements – often megalithic – with polygonal multi-roomed domestic structures close to permanent or seasonal water courses/lakes, high water tables, (wells with troughs and canals) – burial grounds, hunting and herding stations – seasonal? inselberg and open-plain settlements with curvilinear and circular pens and domestic structures, ritual buildings, wells with troughs and canals, platforms, isolated burials, etc. – extensive 1‒2 km2 central burial (aggregation) centres (pastoral hubs) with well/trough fields and dispersed small groups/isolated burials - often megalithic – having ashlar-lined chamber graves/chamber cairn graves on ridges/the plains (e.g. Qulban Beni Murra, Rasif) – proto-oasis settlements – often megalithic – existing at hydrologically favoured places (with natural stands of the wild date palm, e.g. Rajajil?), having ashlar-lined chamber graves/chamber cairn graves also on ridges in the second half of the 5th millennium BCE – (mshash and khabra-type of waterharvesting systems, often related to qas and wadi floors, using dams)

steppe-based mobile pastoralism and elements of pastoral sedentism/ permanency

5th millennium (“Chalcolithic”)

4th millennium (“Chalcolithic”-Early Bronze Age)

– fully flourishing oasis settlements (e.g. Rajajil) operated by wells together with – continued reducing mobile pastoralism with niche agriculture of the 5th millennium type – (mshash- and khabrat-based) water harvesting systems in the former steppes, often related to qas, may have also wells with canal-type troughs (– fields irrigated by rainwater harvesting systems)

steppe-based mobile pastoralism with permanent elements (proto-oases) establishing extensive pastoral and sepulchral landscapes with watering places, settlements, seasonal places and stations; proto oases by the second half of the 5th millennium BCE

pre-oases and oases: increasing reliance on/incipient and fully flourishing (sedentary) oasis life at hydrologically favoured spots in exchange with remaining mobile pastoral networks (the latter representing a precipitationdepending migratory seasonal and ephemeral exploitation of arid land by independent clans)

Hydraulic Cultures and Hydrology under Climatic Change

259

In terms of the culture’s social landscapes (in terms of land use, network complexity, group sizes) we expect extensive and mobile pastoral tribal chiefdoms with a hydroethic network territoriality and confined reciprocity, gradually becoming sedentary and less mobile oasis tribal chiefdoms with confined territoriality and reciprocity during the 4th millennium (for the terms used here, cf. Gebel 2014). In terms of burial practices, a collective and ancestral mind is in clear evidence: it is documented by the prominent and philopatric way the collective burials were used. For more detailed descriptions of the key sites Qulban Beni Murra, Rasif and Rajajil, please consult Gebel (2016, 2017a) and Zielhofer et al. (2018).

12.4.2 The 4th Millennium BCE Hydraulic and Sepulchral Oases Sites As said before, little direct information is so far available for 4th millennium BCE oases sites on the Arabian Peninsula; few candidates have been identified by cairn lines on nearby ridges in suitable hydrological settings, occasionally found with small amounts of potentially 4th/early 3rd millennium pottery on the slopes indicating buried Early Bronze Age oases. For larger parts of our immediate research areas, we may not find fully developed oases since these locations – most likely – could not successfully participate in such process due to developing ecological and hydrological conditions in the 4th millennium. Some of the pen sites in Wadi Sahab al-Abyad and Wadi Sahab alAsmar (Gebel and Mahasneh 2013) might be candidates for oasis onsets, especially when they were linked to previous pastoral niche agriculture. The first firm evidence for Northern Arabia we have for – presumably open-land – fig and grape cultivation from pollen records is around 4000 BCE at the oasis of Tayma (Dinies et al. 2016); date palm pollen are not in these records. The origins of date palm cultivation is yet unknown. Currently, it appears that date palms may have been domesticated or introduced quite late in countries along the Arabian-Persian Gulf and on the Omani Peninsula (3rd millennium BCE) while they possibly were cultivated much earlier in Mesopotamia (5th millennium BCE?; for further discussion, cf. Dinies 2016; Tengberg 2012; Beech and Shepherd 2001). At any rate, this research is suffering from a greater lack of records and research. Basic questions like date palm pollen preservation, the geobotanical distribution of wild date palms, and so on, aren’t answered yet. However, for us it would not be a surprise to once read that date palm cultivation was practiced in most parts of the Arabian Peninsula by 4000 BCE – if not earlier in the supposed proto-oases phase. At the same time, it should be kept in mind that oasis is not an epitome for date palm, and that oasis life already started by different types of oases. The simplified understanding that natural oases with wild date palms offered pastoral draught refugees of the late 5th millennium to cultivate them by watering and inseminating them might be promotive for the provocation of debates but it may not cover the historic complexity and reality. Certainly date palm oasis horticultures offered more gardening possibilities and food security by the shadow and micro-climate they created in their understoreys, including by imported crops and fruits supported by these greenhouse conditions.

12.5 ENGINEERING: WATER MANAGEMENT AND HYDRAULIC FEATURES 12.5.1 (Pre-) Historic Arid Land Water Management From the late 5th millennium BCE, and in terms of water resources management, we identify intensified landscaping work for modifying the natural surface runoff regimes. As stated before, rainfall became less reliable from this period. Consequently, humans needed to adapt their “hydraulic behaviour” to these changed preconditions. Ensuring daily water provision for humans and livestock meant putting more effort into increased yields once rainfall events occurred. Likewise, catchments and their watersheds as well as local depressions in topography were prepared to collect a higher yield for livestock watering, agricultural and human use once surface runoff occurred.

260

Climate Changes in the Holocene

12.5.1.1 Preparing Catchments/Modifying Landscape/Changing Watersheds Catchments were modified to withhold and thus increase available surface runoff drained to particular outlets or local depressions. This was managed by the use of shallow dams (heights less than 50 cm in average, lengths rarely more than a couple of dozen meters) which were sufficient for runoff diversion in the usually flat plains and wadis of North Arabia. By this, natural watersheds have been modified artificially to allow higher surface runoff discharges and harvests. We may regard these early manipulations of the natural hydrologic cycle to be forerunners of what today is known as rainwater harvesting systems (Evenari, Shanan, and Tadmor 1982). 12.5.1.2 Qas, Khabrats, Sabkhas, etc. as a Parts of Artificial Landscape Alteration Once drained to local depressions, surface runoff was used for direct consumption by humans and livestock, and probably horticulture, from at least the 4th millennium BCE. These seasonally filled depressions, or ponds (locally named qas, khabrats and sabkhas), acted as open landscape water storage facilities for longer periods (a couple of weeks or even months depending on the local topography and hydrology). Cisterns, as claimed to have existed already in the Neolithic (Fujii 2010; cf. also on this subject J. Charbonnier, this volume, Chapter 11), are not yet known from Mid-Holocene North Arabia. As mentioned before, retention dams supported the natural drainage regime to increase the capacity of these ponds. Dams often have been build at slightly inclined slopes to create a cascade-like succession of artificial and natural depressions. Such systems show over time quite a hydraulic complexity, since permanent measures of landscape and dam alteration was forced by altered topography in the course of ongoing sedimentation and erosion processes which had been caused by earlier measures. For example, Rasif (see below) represents a prosperous qa system, altered by dams into a more efficient artificial qa, or khabrat, system. Present-day local Bedouin in Northern Arabia crucially discriminate between the various types of ponds. In short: All are seasonally filled open land depressions characterised by an accumulation of fine sediments. Rain-fed natural depressions (qas) are almost free of vegetation, having no direct contact with lower water tables; inland sabkhas are rain-fed natural depressions with mostly halophytic vegetation, having contact with groundwater. Artificial ponds (khabrats) are created by dams in suitable topographical situations, or artificially extended qas, hosting no halophytic vegetation and having no direct contact with groundwater. Because of their groundwater contact, sabkhas do have a higher agri-/horticultural potential whenever salinity is not too high, and therefore the water management in such locations has to be expected to have fostered the evolution of early oasis horticulture, sensu sedentary gardening (see above). 12.5.1.3 Groundwater Management Aside from harvesting surface water, of course tapping groundwater by means of wells was conducted wherever possible; so far, wells have hardly been identified in Arabia’s 5th–4th millennia’s BCE open landscapes while their shepherd and oasis cultures basically must be described as well cultures. Tapping shallow groundwater is possible by digging into (temporarily) saturated sediments. If groundwater is reached by shafts in deeper layers and more permanently available, we deal with wells (bir, biyar). If groundwater is reached by open pits into near-surface layers and is less permanently available, we deal with mshash. These less sophisticated methods, when compared with those of the qas, khabrats, and sabkhas, are used in the open landscapes and allow for more restricted amounts of water; these techniques always were and are applied by nomadic groups. Once the well technologies were transferred into the water-rich locations of the later oases (see above), they became the key water source of proto-oases and oases settlements.

12.5.2 (Pre-) Historic Arid Land Hydraulic Structures 12.5.2.1 Wells and Related Techniques Concerning hydraulic engineering, many technical features related to the aforementioned strategies have been applied. For tapping groundwater, the construction of wells became mandatory, at least

Hydraulic Cultures and Hydrology under Climatic Change

261

FIGURE 12.10  Rasif Case Study: Well interior of Excavation 4 (probably 2nd half of the 5th millennium BCE; Rasif Phase 2), upper part: corbelling masonry, lower part: bedrock. (RJJ Project; photo: H.G.K. Gebel.)

in permanent settlements or wherever surface water was not reliably available. Upper well shafts commonly have been constructed by corbelling dry stone masonry with the lower shafts entering bedrock or wadi gravels (Figure 12.10). Nevertheless, the depths of such wells was at times more than 10 m, as assumed for Qulban Beni Mura (Gebel and Mahasneh 2012, 2013), for instance; some of the wells in Rasif seem to have depths of at least 5 m. Even artificial groundwater recharge in order to bridge periods without rainfall has been conducted (Gebel 2016; Zielhofer et al. 2018). 12.5.2.2 Dams, Diversion Canals and Similar Other technical features served for the retention or delineation of surface runoff. Shallow dams were applied to modify natural catchments, thus even changing watersheds. Even more, simple diversion dams, canals or ditches have been constructed in order to conduct surface water to local depressions. These hydraulic installations are built by using undressed or roughly shaped local ashlar stones. At the end, this hydraulic structural inventory allows enhancing the yield of otherwise limited water in the arid environments and promoted the evolution of artificial qas (see above and the Rasif case study below). 12.5.2.3 Troughs As for watering the livestock, rectangular or canal-shaped troughs, (semi-) circular trough basins, or systems combining both (e.g. Figures 12.4 and 12.5) leading off from well mouths were constructed by local ashlar stones; successions of troughs follow a slight inclination.

12.5.3 Rasif Case Study The site of Rasif (Gebel 2016; Zielhofer et al. 2018) is presented here in order to illustrate (Figures  12.8 through 12.11) the aforementioned features of mid-Holocene water management based on landscape alteration. Here, the application of diversion dams in the vicinity of the watersheds increased the catchment area and thus the amount of surface runoff at local depressions *. By applying such dams, sub-catchments have been connected. Another feature (Dam 10, Figure 12.8) was constructed in order to delineate surface runoff from qa06 (in the South) to the site’s core area. *

The function of the Rasif qa system is still preserved and became evident by strong and sudden rainfall in October 2013 (Figure 6).

262

Climate Changes in the Holocene

FIGURE 12.11  Rasif Case Study: Panoramic view of episodically flooded endorheic basins (qas) in October 2013. (Photo: H.G.K. Gebel.)

The centre of the (pre-) historic site was located in the vicinity of qa01, qa02, and qa03. At this spot, water was retained using shallow dams, thus allowing for a higher capacity of the artificial ponds. Rasif has five major occupational phases which demonstrate how water management strategies and techniques have been adapted to changing climatic and environmental preconditions (Figure 12.9). During the first phase, that is, in the Late Neolithic (7th millennium BCE), water was available on a more reliable/perennial basis in shallow, rather natural than artificial ponds. Once climate deteriorated and rainfall became less reliable, the groundwater level decreased and thus triggered local infiltration. Water became available only after digging shafts, which allowed for at least temporal or seasonal water acquisition. Numerous troughs and well-like features of the site are linked to this phase (Phase 2). In the course of ongoing aridisation (i.e. decreased annual rainfall rates), reduced vegetation and finally increased accumulation of fine sediments, infiltration became limited. The wells of the earlier phase have been likely used for a seasonal groundwater recharge once periodic rainfall events occurred. An aquiclude located about 1 to 2 m below the surface prevented deep percolation then. During Phase 3, therefore, temporal water availability was enhanced by means of artificial groundwater recharge. Once rainfall became even less reliable, some of the site’s wells have been deepened in order to tap local groundwater expected to have been located at some 10 m below the present-day surface. By this, water became available year-round. Of course, the digging, maintenance and safety of such wells and associated installations meant an enormous hydrosocial investment at this site, especially for the organisation of the required workforce and the implementation of social rules of water use. Hereafter, and due to its favourable hydrological location, the site was occupied further by pastoral Bedouins up to sub-recent and recent times (Phases 4–5). Surface runoff was retained by Dams 1, 4, 5 and others (Figures 12.8 and 12.9, Phase 5), thus allowing for temporal ponding of the qas once episodic rainfall occurs. The water stored in these shallow ponds was used for watering livestock and also for feeding humans. At the same time, the local depressions became spots where fodder plants and possibly vegetables have been grown.

12.6 SUSTAINABILITY: LESSONS FROM THE PAST When studying the anonymous hydraulic structures of Arabia’s past steppe-desert societies, researchers immediately are confronted with the insight about how sustainable the water management of these cultures was when compared with modern water use in Arabia’s agricultural development and housing. Wasting fossil, non-renewable groundwater on account for one’s personal comfort or for prestigious crop harvests for which the sensitive arid desert environments are not the natural habitats is a present-day daily practice and lifestyle. It is high time to now implement the lessons from the past, with the help of intensified archaeohydrological research, to counter this ongoing dramatic groundwater depletion. All researchers need to take responsibility and advocate for a re-thinking of water policy. They must demand the “translation” of the lessons from a sustainable past for today’s world. Who, if not the specialists in archaeohydrology and related fields, can understand better and translate the potentials of past hydraulic and hydrological knowledge? While prehistoric and historic water management always is a matter of reconstruction and interpretation, and thus less reliable to transfer into modern experience, fragments of the rapidly

Hydraulic Cultures and Hydrology under Climatic Change

263

vanishing water knowledge of the Bedouins is still available. It is clear that the responsibility and chances to record this knowledge and to transfer this heritage to younger local generations are being widely missed. While excavating and studying hydraulic structures of past millennia, researchers neglect to diligently save the hydrological and environmental knowledge of the old Bedouins that is leaving us day by day. All we have seen shows that their water-related ethos and environmental behaviour is similar to that of past societies, or the “Palaeo-Bedouins”, as we like to call them (Gebel 2017a; Wellbrock, Grottker, and Gebel 2017). Recording the sustainable water management of Bedouins is a painful and fruitful task at the same time, because an engineer or archaeologist does not necessarily understand the mentality by which Bedouins understand water and water deficits. Embedded engagement, bilateral inclusion, cultural advocacy, partitative concepts and community-based water heritage education are needed for such projects which are seen in the future field of an Applied Archaeohydrology (Figure 12.1). Documenting and demanding the application of past sustainable water techniques and behaviour is highly political and may easily find opposition among the stakeholders of modern life. Another threat for the survival of sustainable water management is that the lifeways of the heritage bearers are in a tremendous transformation: If the Bedouin ethos and lifeways are not preserved, we cannot expect that the related water knowledge and behaviour that does survive can be safeguarded: The great legacy of Arabia’s pastoral people is comminuting between the facilities offered by modern life and its technologies, national interests and the phlegm or disinterest and helplessness of the culture-bearers themselves. When working in our Mid-Holocene sites, we consistently came across the extensive variability of subrecent and recent Bedouin waterworks, understanding that the Bedouins are the real and better “hydrologists” and “archaeologists” who can explain to us how their systems work; to the highest degree possible, we recorded remains and remaining information. As a result, we present below examples of identified structures and water behaviour in Northern Arabia. From a technical point of view, these systems are similar to the ones operated already during (pre-) historic periods, for example, in mid-Holocene times. Their common characteristics are that 1) they modify natural drainages and topographies in a conservative way, not doing large-scale landscaping, and 2) hydraulic engineering is based on long-term observation and re-activation of older systems once hydrological conditions changed for the better. Strategies show that in the case of rainfall events, surface runoff can be saved without much input of the human workforce. Typically, pastoral Bedouins prepared several such systems at the same time, hoping that rainfall would fill at least some of them during the rainy season.

12.6.1 Mugur Systems: e.g. Bweitat 1a This type of rainwater harvesting structure consists of subterranean reservoirs (type of cisterns) fed by episodic surface runoff collected by extensive surfaces with deflection dams. The mugur, or reservoirs, are usually placed below the escarpments of higher plateaus which were prepared by a complex system of deflection dams, canals and ditches guiding the runoff to the catchment´s outlet from where reservoirs are fed. Water from mugur supplies humans and livestock as well as horticultural purposes. The mugur are linked to each other, cascade-like, by gravity flow canals (natural or artificial), thus allowing their being consecutively charged. Mugur are mostly cut into bedrock with vertical shafts of up to 2 to 4 m before the reservoir widens, bottle-shaped. Subtypes of mugur are placed in the vicinity of or directly at alluvial fans of (tributary) wadi. The site of Bweitat 1a (located some 100 km East of Sakakah) represents this type of water management. The site comprises a total of 13 mugur which are linked to each other by means of overflow canals (Figure 12.12). The site is located within a major and wide wadi’s bed (250 m). Therefore, several funnel-shaped deflection dams are engaged for concentrating the runoff in the centre of the wadi and charging a kidney-shaped mound which likely was used as an area for

264

Climate Changes in the Holocene

FIGURE 12.12  Sustainable water harvesting: Subrecent system of pastoral Bedouin water harvesting at Bweitat 1a. (Field records: K. Wellbrock, A. Suleiman; graph: K. Wellbrock.)

horticulture. Excess water is then conducted via artificial gulleys which again are created by small deflection dams and walls to subsequent mugur allocated in cascade-like manner.

12.6.2 Khabra Systems: Open Ponds for Retaining Surface Runoff Locally addressed as khabra, they are retention dams placed in tributary wadis used to retain surface runoff in open, shallow ponds, thus creating an artificial qa. The dams are constructed by using upstanding flagstones and sediments forming a barrage of less than 50 cm. After rainfall events an artificial lake for a period of two to three weeks will persist to provide water for humans and livestock (Rasif Phase 5 represents such a kahbra system.). In the traditional Bedouin markets, shebba (potassium alum) is sold; Bedouins say that khabra water can be used longer/can be “cleared” when shebba is added.

12.6.3 Mshash or Thamil’: Temporary Tapping of Shallow Groundwater These are artificial pits dug in alluvial fans of wadis or in qas to allow water harvesting. These systems are locally known as thamil’ (pl. thamayl) in Northern Saudi Arabia or mshash in Jordan. The systems provide water during summer. Neither groundwater nor surface water but interflow (infiltrated rainwater horizontally moving above aquicludes in upper soil layers; not yet groundwater) are tapped by such systems. The systems are similar to those identified in Rasif during its Phase 1 (Figure 12.9), that is, the early Mid-Holocene.

12.6.4 General Statement on Traditional Systems We may conclude that general water management strategies and techniques did not alter considerably over the millennia, while a greater structural and topographical variability can be observed. Possibilities of water provisioning are mainly depending on environmental factors and landscape (e.g. topography, geology). Thus, traditional water management in arid to semi-arid environments tends to apply similar techniques like their (pre-) historic forerunners, even if climate is changing considerably.

Hydraulic Cultures and Hydrology under Climatic Change

265

12.7 ARCHAEOHYDROLOGY: INTEGRATED AND HOLISTIC APPROACHES TO HOLOCENE ADAPTATIONS Our (H.G.K.G.) work in present-day arid lands questioned from the beginning if mere archaeological research can fully describe and understand hydraulic societies and the oasis trajectory. Hitherto, archaeology operates in such fields by using hydroarchaeological approaches, claiming that archaeology can develop or have a competence in hydraulic and hydrological matters. After we realised that our archaeological interpretations of hydraulic structures was premature, insufficient, void of hydrological and water engineering competence and even sometimes false, we were able to see the limits of archaeology, especially of dry land archaeology. The strong understanding developed that only multidisciplinary, or better integrated holistic and transdisciplinary efforts, can promote research, connecting archaeology, hydrology, geomorphology, geoarchaeology and others (cf. Figure 12.1, the Archaeohydrology Aggregate). We understood that even disciplinary frameworks were insufficient for studies in water history, and that special methodological foundations and our own disciplinary frameworks have to be established in the same way this has been done for the – today undebated – need to have a palaeoethnobotany/archaeobotany, archaeozoology, archaeomineralogy, geoarchaeology, palaeoclimatology and others: The need to establish archaeohydrology as a discipline of its own right and as a taught university subject became evident. Initial discussion shows that archaeology is quite reluctant to accept a science-guided archaeohydrology; rather, traditional archaeologies see hydroarchaeology as an option (Gebel n.d. 2; Wellbrock and Grottker n.d. 1). Both authors now are on track, together with others, to promote archaeohydrology as an academic discipline and subject with transdisciplinary “intentions”. They are currently networking in the [email protected] mailing list and at workshops and conferences. Our case studies served as initial paradigms for the propagation of archaeohydrology. Apart from the general need to promote and establish archaeohydrology by a transdisciplinary effort (Figure 12.1), we identified further related imperatives: to document and save past or vanishing traditional hydraulic competence and hydrological knowledge for the sake of sustainable future water management; to safeguard hydraulic and water-related environmental heritage/structures; and to install programmes promoting their large-scale re-activation while reducing non-natural water and land use. Studies on Arabia’s past need to become archaeohydrological works: Like it is true for all other arid regions’ histories on the globe, it is imperative to have archaeohydrological approaches for studying the water-guided history of the Arabian lands. Hydraulic structures are key features in Arabia’s archaeology; water availability and management strategies ruled all cultural development as well as the ethos of humans in Arabia’s past. It has become essential for progress in the hitherto rather neglected and underdeveloped arid land archaeology that it works archaeohydrologically. Only a limited number of projects in arid Arabia employ hydrologists or engage in building interdisciplinary frameworks for the study of their hydraulic and hydrological findings. Hydrological, palaeohydrological and palaeoclimate research has to assist the understanding of past landscapes and land use and how the ethological, social, economic and ideological meaning of water in ancient societies has to be understood. Archaeohydrology has more fields beyond sciences and engineering sciences, touching disciplines in life sciences/humanities one would not expect to be directly related to water research; among others, they are: human hydroethology, human water territoriality, hydrosociology and “water vulnerability”. As mentioned before, the study of (pre-) historic desert hydraulic features and societies has triggered the understanding that consulting, recording, understanding and protecting the hydrological and environmental knowledge of traditionally living Bedouins and oasis farmers is essential for future generations and their sustainable use of dry lands. This makes archaeohydrology an Applied Archaeohydrology which certainly not only has its natural fields in dry lands. The translation of hydraulic heritage is not only meaningful advice for sustainable water management; its application or re-activation has also become a need for survival and peace in many regions of the world.

266

Climate Changes in the Holocene

12.8 SUMMARY Hitherto new research indicates that the hydraulic cultures of North Arabia’s Mid-Holocene followed a general trajectory from mobile pastoralism (5th millennium BCE) to sedentary agri-/horticulturalism associated with reducing shares of mobile pastoral lifeways (from 4000 BCE); most probable, at hydrologically favoured spots, proto-oases already existed in the second half of the 5th Millennium. This shift, caused by a climate change towards drier conditions after 4200 BCE, led to the sustainable establishment of sedentary life on the Arabian Peninsula (cf. also J. Charbonnier, this volume, Chapter 12); it represents one of Arabia’s greatest socioeconomic achievements. Research could not reveal a major cognitive or technological – in terms of hydraulic engineering – shift related to this trajectory and expects that innovations relate more to social water organisation and related water commodification (Gebel 2010b) as well as to the introduction of new groups and gardening techniques. The hydraulic cultures and hydrological conditions in Mid-Holocene Northern Arabia are discussed from an archaeohydrological perspective: They are presented 1) as an example for interacting human social and technological adaptations reacting to changing climate and landscapes; 2) to outline the functions and interdependencies of the various elements engaged in archaeohydrological systems developing during climatic shifts; 3) to advocate for documentation, preservation/reactivation of the vanishing sustainable water systems of the traditional Bedouins and oasis people; and 4) to promote the idea of Archaeohydrology/Applied Archaeohydrology becoming a research discipline in its own right. Developments are discussed from the perspectives of vulnerability, degradation, adaptation and hydraulic engineering. (Pre-) historic climate and hydrological forcing in present-day arid Arabia had a direct impact on all sociohydraulic development, since here no stratified societies with societal, economic and territorial risk-buffering “options” were existing . The Mid-Holocene trajectory is an example for the human experience of how a severe ecological disaster affecting or terminating their lifeways and (re-) production can result in an ecological adaptation providing them sustainable food security for the following generations. The sustainable water management strategies of (pre-) historic and subrecent periods are goodpractice examples for today’s sustainable arid land water management. It is almost too late: the bearers of the hydraulic and hydrological heritage and knowledge are currently passing away without having been widely recognised as the only living source for alternative attitudes in sustainable arid land management.

ACKNOWLEDGEMENTS We owe sincere gratitude to the Saudi Commission for Tourism and Heritage (SCTH), Riyadh, and the Department of Antiquities, Amman, for the constant administrative and material support of our projects in Northern Arabia. Special thanks go to the directors-general of the Jordanian Department of Antiquities after 2001, and to HE. Prof. Dr. Ali al-Ghabban, HE Prof. Dr. Hussein Abualhassan, HE Dr. Jamal Omar, Dr. Abdallah Zahrani and many others of the SCTH. We deeply thank our devoted teams for their hard work and intellectual inputs during the field seasons in North Arabia.

REFERENCES al-Ghazzi, A. 2004 (1424). Standing stones in Saudi-Arabia. A comparative study. Journal of the Historical Society (Riyadh) 5(9): 5–46 (in Arabic). Arz, H., F. Lamy, J. Pätzold, P.J. Müller, and M. Prins. 2003. Mediterranean moisture source for an earlyHolocene humid period in the northern Red Sea. Science 300: 118–121. Bar-Matthews, M., and A. Ayalon. 2011. Mid-Holocene climate variations revealed by high-resolution speleothem records from Soreq Cave, Israel and their correlation with cultural changes. The Holocene 21: 163–171.

Hydraulic Cultures and Hydrology under Climatic Change

267

Beech, M., and E. Shepherd. 2001. Archaeobotanical evidence for early date consumption on Dalma Island, United Arab Emirates. Antiquity 75: 83–89. Benito, G., M.G. Macklin, A. Panin, S. Rossato, A. Fontana, A.F. Jones, M.J. Machado, et al. 2015. Recurring flood distribution patterns related to short-term Holocene climatic variability. Scientific Report 5: 16398. doi:10.1038/srep16398. Charbonnier, J. this volume. Chapter 11: Human adaptation in Arabia: the role of hydraulic technologies. Chiotis, E. this volume. Chapter 1. Reconstructing the environment as a scenery of human history. Clare, L. 2016. Culture Change and Continuity in the Eastern Mediterranean During Rapid Climate Change. Assessing the Vulnerability of Late Neolithic Communities to a “Little Ice Age” in the Seventh Millennium calB.C. Kölner Studien zur Prähistorischen Archäologie 7. Dinies, M., R. Neef, and H. Kürschner. 2011. Holocene vegetational development and the beginning of oasis cultivation in Tayma, northwestern Saudi Arabia. First results. Pollen 21: 9–14. Dinies, M., R. Neef, B. Plessen, and H. Kürschner. 2016. Holocene vegetation, climate, land use and plant cultivation in the Tayma region, northwestern Arabia. In The Archaeology of North Arabia. Oases and Landscapes. Proceedings of the International Congress Held at University of Vienna, 5–8th Dec., 2013, edited by M. Luciani. Oriental and European Archaeology, 4: 57–78. Vienna: Austrian Academy of Sciences. Dinies, M., B. Plessen, R. Neef, and H. Kürschner. 2015. When the desert was green: Grassland expansion during the early Holocene in northwestern Arabia. Quaternary International 382: 293–302. doi:10.1016/j. quaint.2015.03.007 Engel, M., H. Brückner, A. Pint, K. Wellbrock, A. Ginau, P. Voss, M. Grottker, et al. 2012. The early Holocene humid period in NW Saudi Arabia—Evidence from sediments, microfossils and palaeohydrological modelling. Quaternary International 266: 131–141. Engel, M., A. Matter, A.G. Parker, A. Parton, M.D. Petraglia, G.W. Preston, and F. Preusser. 2017. Lakes or wetlands? A comment on ‘The middle Holocene climatic records from Arabia: reassessing lacustrine environments, shift of ITCZ in Arabian Sea, and impacts of the southwest Indian and African monsoons’ by Enzel et al. Global Planetary Change 148: 258–267. doi:10.1016/j. gloplacha.2016.11.001. Enzel, Y., Y. Kushnir, and J. Quade. 2015. The middle Holocene climatic records from Arabia: reassessing lacustrine environments, shift of ITCZ in Arabian Sea, and impacts of the southwest Indian and African monsoons. Global Planetary Change 129: 69–91. Enzel, Y., J. Quade, and Y. Kushnir. 2017. Response to Engel et al., 2017: Lakes or wet- lands? A comment on “The middle Holocene climatic records from Arabia: reassessing lacustrine environments, shift of ITCZ in Arabian Sea, and impacts of the southwest Indian and African monsoons” by Enzel, et al., 2015. Global Planetary Change 148: 268–271. Evenari M., L. Shanan, and N. Tadmor. 1982. The Negev: The challenge of a desert. Cambridge, MA: Harvard University Press. Fleitmann, D., S.J. Burns, M. Mudelsee, U. Neff, J. Kramers, A. Mangini, and A. Matter. 2003. Holocene monsoon recorded in a stalagmite from southern Oman. Science 300: 1737–1739. Fleitmann, D., S. Burns, M. Pekala, A. Mangini, A. al-Subbary, M. al-Aowah, J. Kramers, et al. 2011. Holocene and Pleistocene pluvial periods in Yemen, southern Arabia. Quartenary Science Reviews 30: 783–787. Fuchs, M., and A. Buerkert. 2008. A 20 ka sediment record from the Hajar mountain range in N-Oman, and its implication for detecting arid-humid periods on the southeastern Arabian Peninsula. Earth Planetary Science Letters 265: 546–558. Fujii, S. 2010. Domestication of runoff surface water. Current evidence and new perspectives from the Jafr Pastoral Neolithic. Neo-Lithics 2010.2: 14–32. Berlin: ex oriente. Gebel, H.G.K. 2010a. Untergegangen im Klimawandel. Die paläo-beduinische Kultur von Qulban Beni Murra, Jordanien. Antike Welt 41: 40–44. Gebel, H.G.K. 2010b. The commodification of water. Neo-Lithics 2010.2: 4–13. Berlin: ex oriente. Gebel, H.G.K. 2013. Arabia’s fifth-millennium BCE pastoral well cultures: Hypotheses on the origins of oasis life. Proceedings of the Seminar for Arabian Studies 43: 1–16. Gebel, H.G.K. 2014. Territoriality in early Near Eastern sedentism. Neo-Lithics 2: 14–44. Berlin: ex oriente. Gebel, H.G.K. 2016. The socio-hydraulic foundations of oasis life in NW Arabia: The 5th millennium BCE shepherd environs of Rajajil, Rasif and Qulban Beni Murra. In The Archaeology of North Arabia. Oases and Landscapes. Proceedings of the International Congress Held at University of Vienna, 5–8th Dec., 2013, edited by M. Luciani. Oriental and European Archaeology, 4: 79–114. Vienna: Austrian Academy of Sciences.

268

Climate Changes in the Holocene

Gebel, H.G.K. 2017a. The origins of oasis life in NW Arabia. A model based on the Qulban Beni Murra and Rajajil Case study regions, and the need of Archaeohydrology as a discipline for studying Arabia’s past. In Water and Life in Arabia, edited by W.Y. al-Tikriti and P.A. Yule, 1–26. Abu Dhabi: Abu Dhabi Tourism and Culture Authority. Dar-Al-Ummah Publishing. Gebel, H.G.K. 2017b. Neolithic corporate identities in the Near East. In Neolithic Corporate Identities, edited by M. Benz, H.G.K. Gebel, and T. Watkins. Studies in Early Near Eastern Production, Subsistence, and Environment, SENEPSE 20: 57–80. Berlin, ex oriente. Gebel, H.G.K. n.d. 1. The bedouin legacy. Call for a rethinking. www.academia.edu/12311515. Gebel, H.G.K. n.d. 2. Archaeohydrology. The archaeological perspective. Manuscript/contribution to Archaeohydrology of Oases and Cities. Water Management, Climate, Technological Change and Social Contexts. International Conference and Workshop. German Archaeological Institute, Berlin. May 31–June 2, 2017. Gebel, H.G., C. Hannss, A. Liebau, and W. Rähle. 1989. The late Quaternary environments of ‘Ain al- Faidha/ al-‘Ain, Abu Dhabi Emirate. Archaeology in the United Arab Emirates 5: 9–48. Gebel, H.G.K. and H.M. Mahasneh. 2012. Qulban Beni Murra. Unknown Mid-Holocene sepulchral green desert landscapes, pastoral well cultures, and the origins of Arabia’s oasis economies. In Wasserwirtschaftliche Innovationen im archäologischen Kontext. Von den prähistorischen Anfängen bis zu den Metropolen der Antike, edited by F. Klimscha, R. Eichmann, C. Schuler, and H. Fahlbusch. Menschen, Kulturen, Traditionen, 5: 101–122. Rahden: Leidorf. Gebel, H.G.K., and H.M. Mahasneh. 2013. Disappeared by climate change. The shepherd cultures of Qulban Beni Murra (2nd half of the 5th millennium BC) and their aftermath. Syria 90: 127?158. Harrower, M.J. 2008. Hydrology, Ideology, and the Origins of Irrigation in Ancient Southwest Arabia,” Current Anthropology 49, no. 3 (June 2008): 497–510. DOI: 10.1086/587890. Ibouhouten, H., Zielhofer, C., Mahjoubi, R., Kamel, S., Linstädter, J., Mikdad, A., Bussmann, J., Werner, P., Fenech, K., 2010. Archives alluviales holocènes et occupation humaine en Basse Moulouya (Maroc nord-oriental). Géomorphologie: relief, processus, environnement 16, 41–56. DOI : 10.4000/ geomorphologie.781. Kirkbride, D. 1960. Khirbet Rizqeh. Revue Biblique 67: 232–235. Kirkbride, D. 1969. Ancient Arabian ancestor idols. Parts I and II: The discovery of the sanctuary at Risqeh. Archaeology 22: 116–121, 188–195. Kröpelin, S., D. Verschuren, A.M. Lézine, H. Eggermont, C. Cocquyt, P. Francus, J.P. Cazet, et al. 2008. Climate-driven ecosystem succession in the Sahara: the past 6000 years. Science 320: 765–768. di Lernia, S. this volume. Chapter 9: From ‘Green’ to ‘Brown’. The Archaeology of the Holocene Central Sahara. Lézine, AM, J.J. Tiercelin, C. Robert, J.F. Saliège, S. Cleuziou, and M.L. Inizian. 2007. Centennial to millennialscale variability of the Indian monsoon during the Early Holocene from a sediment, pollen and isotope record from the desert of Yemen. Palaeogeography, Palaeoclimatology, Palaeoecology 243: 235–249. Masry, A.H. 1974. Prehistory in northeastern Arabia: the problem of interregional interaction. Field Research Projects. Coconut Grove, Miami. Mayewski, P.A., E.E. Rohling, J.C. Stager, W. Karlen, K.A. Maascha, L.D. Meeker, E.A. Meyerson, et al. 2004. Holocene climate variability. Quaternary Research 62: 243–255. McClure, H.A. 1976. Radiocarbon chronology of late Quaternary lakes in the Arabian Desert. Nature 263: 755–756. Migowski, C., M. Stein, S. Prasad, J.F.W. Negendank, and A. Agnon. 2006. Holocene climate variability and cultural evolution in the near east from the Dead Sea sedimentary record. Quaternary Research 66: 421–431. Müller-Neuhof, B. 2014. A ‘marginal’ region with many options: the diversity of Chalcolithic/Early Bronze Age socio-economic activities in the hinterland of Jawa. Levant 46: 230–248. Pei, Q., H.F. Lee, and D.D. and Zhang. 2017. Long-term association between climate change and agriculturalists’ migration in historical China. The Holocene 28: 1–9. doi:10.1177/0959683617721325. Peltenburg E., P. Croft, A. Jackson, C. McCartney, and M.A. Murray. 2001.Well-established colonists. Mylouthkia I and the Cypro-Pre-Pottery Neolithic B. In The Earliest Prehistory of Cyprus. From Colonization to Exploitation, edited by S. Swiny, 61–93. Boston: American Schools of Prehistoric Research. Preston, G.W., D.S.G. Thomas, A. Goudie, O.A.C. Atkinson, M.J. Leng, M.J. Hodson, H. Walkington, V. Charpentier, S. Méry, F. Borgi, and A.G. Parker. 2015. A multi-proxy analysis of the Holocene humid phase from the United Arab Emirates and its implications for southeast Arabia's Neolithic populations. Quaternary International 382: 277–292.

Hydraulic Cultures and Hydrology under Climatic Change

269

Radies, D., S.T. Hasiotis, F. Preusser, E. Neubert, and A. Matter. 2005. Paleoclimatic significance of early Holocene faunal assemblages in wet interdune deposits of the Wahiba sand sea, Sultanate of Oman. Journal of Arid Environments 62: 106–125. Schultz, E, and J.W. Whitney. 1986. Upper Pleistocene and Holocene lakes in the An Nafud, Saudi Arabia. Hydrobiologia 143: 175–190. Staubwasser, M, and H. Weiss. 2006. Holocene climate and cultural evolution in late prehistoric—Early historic West Asia, Quaternary Research 66: 372–385. Tengberg, M. 2012. Beginnings and early history of date palm garden cultivation in the Middle East. Journal of Arid Environments 86: 139–147. Tierney, J.E., F.S.R. Pausata, and P.B. deMenocal. 2017. Rainfall regimes of the Green Sahara. Science Advances 3(1): e1601503. doi:10.1126/sciadv.1601503. Walker M.J.C., M. Berkelhammer, S. Björck, L.C. Cwynar, D.A. Fisher, A.J. Long, J.J. Lowe, et al. 2012. Formal subdivision of the Holocene Series/Epoch: a Discussion Paper by a Working Group of INTIMATE (Integration of ice-core, marine and terrestrial records) and the Subcommission on Quaternary Stratigraphy (International Commission on Stratigraphy). Journal of Quaternary Science 27(7): 649–659. doi:10.1002/jqs.2565. Wanner, H., L. Mercolli, M. Grosjean, and S.P. Ritz. 2015. Holocene climate variability and change; a databased review. Journal of the Geological Society London 172: 254–263. Wellbrock, K., and M. Grottker. n.d. 1. Archaeohydrology. The hydrological perspective. Manuscript/contribution to Archaeohydrology of Oases and Cities. Water Management, Climate, Technological Change and Social Contexts. International Conference and Workshop. German Archaeological Institute, Berlin. May 31–June 2, 2017. Wellbrock, K., M. Grottker, and H.G.K Gebel. 2017. Archaeohydrological investigation in NW Arabia. Potentials, problems, needs and goals. In Water and Life in Arabia, edited by W.Y. al-Tikriti and P.A. Yule, 27–43. Abu Dhabi: Abu Dhabi Tourism and Culture Authority/Dar-Al-Ummah Publishing. Wellbrock, K., P. Voß, and M. Grottker. 2011. Reconstruction of mid-Holocene climate conditions for northwestern Arabian oasis Tayma. International Journal of Water Resources and Arid Environments 1: 200–209. Wellbrock, K., P. Voß, and M. Grottker. 2012. The evolution of water management methods in north-western Arabia and the southern Levant from the Neolithic Age through Antiquity. Schriften der Deutschen Wasserhistorischen Gesellschaft 20(1): 29–48. Wellbrock, K., P. Voss, B. Heemeier, P. Keilholz, A. Patzelt, and M. Grottker. 2017. The water management of Tayma and other ancient oasis settlements on the North-western Arabian Peninsula – a synthesis. In Tayma I. Reports on Palaeoenvironment, Archaeology and History, edited by R. Eichmann and A. Hausleiter, 68–122. Oxford: Archaeopress. Weninger, B., E. Alram-Stern, E. Bauer, L. Clare, U. Danzeglocke, O. Jöris, C. Kubatzki, et al. 2006. Climate forcing due to the 8200 cal yr BP event observed at early Neolithic sites in the eastern Mediterranean. Quaternary Research 66: 401–420. Weninger, B., L. Clare, E.J. Rohling, O. Bar-Yosef, U. Böhner, M. Budja, M. Bundschuh, et al. 2009. The impact of rapid climate change on prehistoric societies during the Holocene in the Eastern Mediterranean. Documenta Praehistorica 26: 1–53. Whitney, J.W., D.J. Faulkender, and M. Rubin. 1983. The environmental history and present condition of Saudi Arabia’s northern sand seas. USGS Open-File Report 83-749. Wittfogel, K.A. 1957. Oriental despotism. A comparative study of total power. New Haven: Yale University Press. Zielhofer, C., W.J. Fletcher, M. De Batist, J.F.E. Campbell, S. Joannin, R. Tjallingii, N. El Hamouti, et al. 2017a. Atlantic forcing of Western Mediterranean winter rain minima during the last 12,000 years. Quaternary Science Review 157: 29–51. Zielhofer, C., H. von Suchodoletz, W.J. Fletcher, E. Dietze, B. Schneider, M. Schlegel, K. Schepanski, et al. 2017b. Millennial-scale fluctuations in Saharan dust supply across the decline of the African humid period. Quaternary Science Review 171: 119–135. Zielhofer, C., K. Wellbrock, A.S. al-Souliman, B. Schneider, K. Fitzsimmons, A. Stele, T. Lauer, et al. 2018. Climate forcing and shifts in water management on the Northwest Arabian Peninsula (mid-Holocene Rasif wetlands, Saudi Arabia). Quaternary International 473: S. 120–140.

13

Collapse of Bronze Age Civilizations Guy D. Middleton

CONTENTS 13.1 Introduction........................................................................................................................... 271 13.2 Defining Collapse.................................................................................................................. 272 13.3 Background............................................................................................................................ 272 13.4 The EBA Collapse in Greece................................................................................................ 273 13.5 The 4.2 ka BP Climate Event in Greece................................................................................ 274 13.6 The 4.2 ka BP Event as a Cause of Eastern Mediterranean Collapse................................... 276 13.6.1 The Levant................................................................................................................. 276 13.6.2 Egypt.......................................................................................................................... 277 13.6.3 The Akkadian Empire............................................................................................... 277 13.6.4 Summary................................................................................................................... 278 13.7 The LBA Collapses in Greece and the Eastern Mediterranean............................................ 279 13.8 Climate–Collapse–Migration as the Cause of LBA Collapses.............................................280 13.8.1 Critique......................................................................................................................280 13.8.2 Summary................................................................................................................... 282 13.9 Working Together: Opportunities and Issues........................................................................ 283 13.10 Conclusion............................................................................................................................. 286 Acknowledgments........................................................................................................................... 287 References....................................................................................................................................... 287

13.1 INTRODUCTION Archaeologists and historians, amongst others, have long been interested in climate and its potential effects on human societies (Fairbridge 2009a). Since Douglass (1867–1962) pioneered dendrochronology and dendroclimatology in the early twentieth century, archaeologists have used tree rings as proxies for past climate records (Fairbridge 2009b, 430). Huntington’s 1917 study is an early example of dendroclimatology and its application to a historical problem: the decline of Rome; he used the growth rates of Californian sequoia trees as a proxy and compiled a palaeoclimatic record for Roman history. In a chain of causality, he linked climate change with agricultural productivity, and with the economic, political and biological spheres. Bad years caused subsistence problems, which led to economic and political problems; long-term aridity was catastrophic: “thousands of people must have been driven from their homes” (Huntington 1917, 200). Climate would also affect the prevalence of diseases, such as malaria (Huntington 1917, 202–203). Now there are many more ways to gather palaeoclimatic proxy data; Dincauze (2000, 142) lists “the altitudes of tree lines, snow lines, and cirques; pollen assemblages; stable isotope ratios …; marine, lacustrine, and terrestrial microfaunal associations; and paleosols”. Much that Huntington argued remains a mainstay of climate-collapse narratives today, including for the Bronze Age collapses that are the subject of this chapter, such as combined climate and migration events. As Buntgen et al. (2011, 582), also referring the Roman Empire, note: “preindustrial societies were sensitive to famine, disease, and war, which were often driven by drought, flood, 271

272

Climate Changes in the Holocene

frost or fire events, as independently described by documentary archives”. Their study echoes that of Huntington, but is based on much more data—7,284 samples of oak and 1,546 conifers from central Europe—although while it is much more sophisticated, the principles of correlating multiple forms of evidence for climatic reconstruction are not so very different. They note apparently increased climatic variability between AD 250 and 600, a period “which coincides with some of the most severe challenges in Europe’s political, social, and economic history, during the Migration Period” (Buntgen et al. 2011, 580). However, they are careful to avoid climatic determinism, stating that, “comparison of climate variability and human history, however, prohibits any simple causal determination; other contributing factors such as sociocultural stressors must be considered in this complex interplay”. This chapter explores two episodes of collapse, the existence of which has been accepted and discussed by numerous scholars both in archaeology and in climatology and environmental science. The first is the Early Bronze Age or EBA collapse, which is dated to approximately 2200 BC, and the second is the Late Bronze Age or LBA collapse, dated to around 1200 BC. The cultural and geographical scope of each collapse is vast. The EBA collapse is associated with areas from Greece and the Aegean to the Indus Valley, including Anatolia, the Levant, Egypt and Mesopotamia; the LBA collapse has a similar scope: Greece and the Aegean, Anatolia, Cyprus and parts of the Levant, and, to some extent, Egypt. The chapter ends by discussing archaeology and climate change more generally, from disciplinary and social perspectives, including what archaeology has to offer and what responsibilities archaeologists might have at a time of environmental challenges. The hope here is to provide something of a bridge between archaeologists and, especially, paleoclimatologists. The chapter begins by briefly introducing definitions of collapse before examining the EBA and LBA collapse in turn. As befits the author’s experience, the chapter looks from Greece eastwards.

13.2 DEFINING COLLAPSE Many scholars have defined collapse (Middleton 2017a, Chapter 1 and 2017c). The neatest definition is that of Tainter (1988, 4–5), who argues that collapse takes place when a society “displays a rapid, significant loss of an established level of sociopolitical complexity”. Rapid is understood to mean a few decades. A more recent work by Storey and Storey (2017, 10) advances a theory of “slow collapse”, measured on a centennial scale. However, the word “collapse” implies rapidity and changes that happen over a longer period may be better termed “transformations” rather than “collapses”. Here a distinction between the rapid collapse of a political unit and the slower transformation of a culture may be useful, though of course political units too can transform over a longer time. A useful working list of ways in which “collapse” is used has been compiled by Schwarz (2006, 5–6), who notes that “in the archaeological literature, collapse usually entails some or all of the following: the fragmentation of states into smaller political entities; the partial abandonment or complete desertion of urban centers, along with the loss or depletion of their centralizing functions; the breakdown of regional economic systems; and the failure of civilizational ideologies”. The EBA and LBA collapses have both been seen as rapid and to include features such as sociopolitical simplification, destruction and desertion of sites—urban and rural—the end of states and empires, and ideological changes. These definitions are relevant for our case studies of both Early and Late Bronze Age collapses.

13.3 BACKGROUND Writing in 1958, Mellaart associated EBA destructions in Greece with destructions in Anatolia, arguing that they were not mere coincidence, but resulted from a series of migrations of new people into the region. Destructions and abandonments were widespread, though not universal, but culture change was clear. A few years later, Bell (1971) identified a collapse into a “dark age” beginning around 2200 BC, which affected Egypt, the Akkadian Empire, Byblos and Palestine,

Collapse of Bronze Age Civilizations

273

Anatolia—especially Troy II—Lerna and other Greek sites, and perhaps also the Indus Valley, all quite different “entities”. She argued that these events across a huge area were connected by a shared cause: “drought – widespread, severe, and prolonged – lasting for several decades and occurring more or less simultaneously over the entire eastern Mediterranean and adjacent lands”, with her reconstruction based primarily on a range of (primarily Egyptian) texts (Bell 1971, 2). Since these contributions, the idea that a 4.2 ka BP (2200 BC) climate event was a prime cause of widespread cultural, political and social collapse has become a key part of narratives of the Early Bronze Age in the old world (eg. Dalfes et al. 1997). The LBA collapse has been seen in much the same way; in Drews’ words, it “ushered in a dark age, which in Greece and Anatolia was not to lift for more than four hundred years” and was “arguably the worst disaster in ancient history, even more calamitous than the collapse of the Western Roman Empire” (Drews 1993, 3). Again, both migration and climate change, sometimes linked, have been suggested as the cause of the collapse (Carpenter 1966; Neumann 1993; Weiss 1982). Research on both “events” continues apace (eg. Cline 2014; Hoflmayer 2017; Knapp and Manning 2016; Meller et al. 2015; Middleton 2010; Weiss 2015 and 2016; Wiener 2013 and 2014), with a new volume edited by Weiss (just published) that contains chapters on the Akkadian (EBA) and Late Bronze Age collapses (2017a), but understanding of each remains problematic in a number of ways, not least in whether they should be considered as “universal” events with associated global exogenous causes, as “local” developments that happened for a variety of contingent reasons, or something in between. In part, this is a problem of chronological resolution, and in part, it reflects the particular sets of evidence and what has been read into them. The most ardent proponent of a climate-driven EBA collapse, caused by the so-called 4.2 ka BP climate event, is Weiss, who has recently (2016) restated his view that “across the Mediterranean and west Asia, the effects of the 4.2–3.9 ka BP megadrought included synchronous collapse of the Akkadian Empire in Mesopotamia, the Old Kingdom in Egypt and Early Bronze Age settlements in Anatolia, the Aegean and the Levant” (see also Weiss 2014, 2015, 2017b, 117). Discussion of the end of the LBA is haunted by the spectre of violent mass migrations in the form of the Sea Peoples, which in one recent view were a consequence of climate-driven population movement (Kaniewski et al. 2013, 2015, 2017; similar to the theory of Neumann 1993; see also Fischer and Burge 2017). Weiss and Bradley (2001) suggested that many past collapses were caused by climate change, and this has been reiterated in the popular science press by Marshall (2012), though others have voiced notes of caution and raised concerns about a new climatic determinism (Butzer 2012a and b; Middleton 2012). There is still debate about both the existence of climate change and its effects, and the reality of, and possible causes and effects of migration at the end of the EBA and LBA; theories of folk/mass migration are unfashionable although they have become attached to proposed episodes of climate change. Generally speaking, most archaeologists would set aside monocausal and deterministic exogenous explanations for collapse, which is the role often given to climate change, and accept that a collapse is usually more likely to be a quite complex affair with many interacting variables and historically specific contingencies (Cline 2014; Middleton 2017a).

13.4 THE EBA COLLAPSE IN GREECE Before attempting to explain a collapse, it is necessary first to understand what it is that the archaeology shows, i.e., how we identify that collapse. In Greece, Early Helladic (EH) II, a long period, saw development from an egalitarian society to a more complex one with signs of differentiation and ranking, possibly the appearance of small chiefdoms (Pullen 1994 and 2011). One indication of this is the development of fortifications and monumental roof-tiled corridor houses and large central buildings—some key sites being the House of Tiles at Lerna, and buildings at Akovitika, Tiryns with its circular “Rundbau”, Zygouries, Kolonna, and Thebes (Rutter 1993, 762; Shaw 2007).

274

Climate Changes in the Holocene

The corridor houses may have been the seats of chiefs and/or used for communal events, such as drinking and feasting or community meetings, with possible storage functions. Another development, not always associated with corridor or central houses, is the adoption and use of seals, 70 identified at Lerna, and many at Geraki in Lakonia, and also Petri near Nemea, which were connected with ownership and management of goods, their authenticity, and their collection and redistribution (Pullen 1994, 46; Rahmstorf 2016; Voutsaki 2010, 601; Weiberg 2010; Weingarten et al. 1999 and 2011; Wiencke 2010, 664). The ideas of seal use and common weights and measures may have spread from the Near East to Greece via Anatolia and the Cyclades (Rahmstorf 2016, 234–235). Maran and Kostoula (2014) have recently argued that some sealing from the House of Tiles were used on doors and may be connected to the use, closure, and reuse of rooms on given occasions. In Greece, the idea of an Early Bronze Age collapse between Early Helladic (EH) II and III came from excavations at Lerna, in the Argolid, in the 1950s. The monumental House of Tiles was burned and destroyed, which event marks the end of the Lerna III period, and, in Caskey’s words, “the end of an era at Lerna” (Caskey 1960, 293). A tumulus was raised over the ruins and the area was left alone. In the subsequent Lerna IV period, a settlement of a different nature was constructed, with new apsidal buildings and ceramics that were also quite different. The EH II “sauceboat” shape disappeared and new shapes such as the “tankard” and one-handled cup appeared (Caskey 1960, 296). Tumuli also started to appear later on. The archaeological evidence of destruction and the appearance of a novel culture suggested to Caskey “deliberate warlike action, and a reoccupation of the site by people of a different material culture”, and he thought that “a foreign invasion created widespread havoc in this region and brought to an end” the EH II culture (Caskey 1960, 301). Lerna was seen as the type-site for the period, and, while violent destruction was not universal, it nevertheless appeared that other EH II sites had also experienced it around the same time. Caskey envisioned widespread rapid change with the arrival of new people and culture. However, the changes of EH II–III have been investigated by Forsén (1992; 2010, 54), who argues that Caskey’s interpretation is “untenable” (see also Maran 1998). In a study of some 89 sites, she concluded that when site-wide destruction levels could be well-dated, they are not clumped together either geographically or temporally, but occur throughout EH II and EH III. In addition, some apsidal buildings are known from early EH II, including at Tiryns, Epidauros, and Thebes, existing at the same time as tile-roofed corridor houses. Rutter points out that the suddenness of site abandonment may have been overstated, perhaps beginning earlier in EH II; in Laconia, EH II may have lasted longer than in the northern Peloponnese (1993, 773). He has also argued that EH III pottery was a hybrid form using elements of EH II and “Lefkandi I” ceramics, not foreign but developed in central Greece, where the two traditions had co-existed (Rutter 1993, 763–764). Even in EH II, quite different ceramic traditions could co-exist within and at nearby sites. Thus, while there were significant changes, it now appears that these did not take place as swiftly, and were not as universal, as once thought. It could be proposed that what happened in Greece would fall into Storey and Storey’s (2017) category of “slow collapse”, taking place on a centennial scale, but this would gloss over the changes that happened within the span of individual lifetimes— the destruction of the House of Tiles, for example. It might be more accurate to abandon the term “collapse” altogether in this case.

13.5 THE 4.2 KA BP CLIMATE EVENT IN GREECE The 4.2 ka BP event is identified as a rapid climate change that affected a wide area and caused many collapses, including the EH II collapse in Greece (Booth et al. 2005; Weiss 2016 and 2017b). Finné et al. (2011) reviewed and analyzed research on the climate of the eastern Mediterranean over the last 6000 years, including 17 records from Greece and the Aegean. They concluded that while there is some evidence for climate change and a 4.2 ka

Collapse of Bronze Age Civilizations

275

BP event, the overall picture in the eastern Mediterranean is varied, with some records not indicating drought and others not indicating a drought “event”—the strongest evidence for a drought event comes from the Arabian and Red Seas and may be of relevance to the end of the Akkadian Empire. Direct climate proxy evidence from central and southern Greece is lacking, although at this time several projects are ongoing. Finné and Holmgren (2010) have cautioned against assuming that data from the Near East and elsewhere can be applied straightforwardly to Greece, because differences are apparent between this and the limited Aegean data. Evidence from Greece and Turkey does “not shed much light on the matter of a rapid climate event” and while there is evidence for cooling and drying in the Aegean Sea records, there is nothing indicating a significant climate “event” (Finné et al. 2011, 3163–3164). On the basis of their review, Finné and colleagues find that climate records indicate “a complicated picture” that does not enable them “to either confirm or deny the proposition of a widespread, rapid and dramatic climate event’ (Finné et al. 2011, 3163). As of 2010 “on the basis of the present knowledge, climatic factors seem of little or no direct importance” for the changes that were going on in the northeast Peloponnese at Lerna and elsewhere (Weiberg et al. 2010, 158; Weiberg and Finné 2013). However, one as yet unpublished study by Boyd of the Alepotrypa Cave, in the Mani, southern Greece, suggests that there may have been drying around 2200 BC (Martin Finné, personal communication 2016). More studies specific to the microregions of the Aegean are needed. Jung et al. (2015, 222) examined chronological data from the Aegean and argue, based on data from Aegina and Troy, that late EH II should be redated to between 2400 and 2300 BC, which would mean that cultural developments could not have been caused by a 4.2 ka BP climate event. They concluded that there is a “climate-conspicuous” date for the end of EH III, rather than EH II, and that while the reduction in the number of sites is most visible in EH II–III, it continued in EH III–MH I. Their main conclusion is that “the 4.2 ka cal BP climate impact coincides less with the transition from EH II–III and more with the transitions from EH III to MH I” at around 2100 BC (Jung et al. 2015, 230). Broodbank also notes for Cycladic history that “change falls decisively too early for this climate based explanation to apply”, and “the climatic crunch point would coincide instead with the nascence of the Middle Bronze Age system” rather than between EB II and III (Broodbank 2013, 539). For Weiberg and Lindblom (2014, 399) “the transition from Lerna III to IV … consisted largely of a functional shift from public and communal to private and individual that was played out over a few generations”. Changes in pottery may indicate changes in the cultures of food and drink, with a new emphasis on household rather than communal activities (Rutter 1993, 766). Weiberg (2017b, 32) sets EH III in a context of “human resourcefulness, innovation and the active reformulation of social agendas during times of change”. While these could be seen as having been influenced by climate, they could equally have resulted from the rejection of social and political systems or ideologies, which were limited in any case, represented by both the corridor-house and seal-using phenomena. Collapse, if we continue to call it this, also did not affect all parts of Greece. For Crete, Tomkins and Schoep (2010, 72) observe that the process of site growth and proto-urbanism, which began in Early Minoan (EM) II, continued in EM III–MM I, along with an increase in small rural sites. There was an increase in site size and numbers, continuity in seal use, and general evidence for increasing social complexity. At Knossos and Malia, the monumental buildings of the Middle Bronze Age overlay EM II and EM III (and even earlier) beginnings (Tomkins et al. 2010, 73). In her recent comparative study of the northeast Peloponnese and central Crete, Weiberg finds that “the Minoan case study shows no comprehensive societal transformation at the EM II–III transition” (Weiberg 2017a, 7–8). At Kolonna, Aegina, geographically proximate to the northeastern Peloponnese, there was no break between EH II and EH III, though the Weißes Haus was destroyed (Gauss 2010, 742–743; Pullen 2008, 37). The evidence appears to suggest that there were quite different trajectories before, around and after 2200 BC (Weiberg 2017a).

276

Climate Changes in the Holocene

At the moment, there is little evidence for a 4.2 ka BP climate-driven collapse event in Greece. Some of the changes that did take place, visible in the archaeology, probably did so earlier than 2200 BC, and changes continued to take place after; in other words, the situation on the ground was changing over a long period rather than as an “event”, and this took different forms in different areas. Also, later EH II, according to Jung et al. (2015), should be dated earlier, raising the date of the EH II–III transition. This suggests that change took place without climate change as a main driver; whatever the climate was 4.2–3.9 ka BP, the existence of people and communities in a range of environments around the Aegean were viable at different levels. This was not determined simply by climate or hindered by any global megadrought. The appropriateness of the term “collapse” seems questionable as applied to the EH II–III transition. The developmental trajectories of different areas varied: some areas rejected complexity, others continued at the same level of complexity, and yet others began shifting towards even greater complexity. If continued research strengthens the case for a 4.2 ka BP climate “event” in the Aegean, the variations in the archaeological evidence will still require explanation in human terms. The effects of climate changes or other environmental factors will have been mediated by local EH (or any other) societies in potentially different ways. It is tempting to see resilience as a function of social arrangements but this is reductionist—it is necessary to consider, but difficult or impossible to trace the historically unique and specific events that also shaped the courses of history in the later Early Bronze Age in Greece.

13.6 THE 4.2 KA BP EVENT AS A CAUSE OF EASTERN MEDITERRANEAN COLLAPSE The climate-collapse narrative draws in a huge area, including Anatolia, the northern and southern Levant, Egypt, and the Akkadian Empire. There is not sufficient space to go into detail on each of these cases in this chapter, and this section will be limited to the Levant, Egypt, and the Akkadian Empire.

13.6.1 The Levant Genz (2015) notes that there is little data on Holocene climate from Lebanon. However, oxygen and carbon isotope analysis of a stalagmite from the Jeita Cave show no evidence of a 4.2 ka BP climate event (Verheyden et al. 2008, 380). This result is supported by a pollen record from the Anjar wetlands, but contradicted by another from the Aamiq wetlands, both in the Bekaa Valley (Genz 2015, 98). A recent higher-definition publication of the Jeita Cave record concludes that there were “two significant drought periods, between 5.3–4.2 and 2.8–1.4 ka, straddling a relative wet period” (Cheng et al. 2015). Examining the archaeology, Genz argues that northern and southern Lebanon have different trajectories in the EBA. During EBA III (c.3000–2500 BC), there was urban settlement all along the coastal strip, north and south. From EBA III–IV settlements in the north continue, as do contacts with Egypt and Mesopotamia (e.g., at Byblos), but in southern Lebanon urbanism reduced in EBA IV, and in some cases disappeared, as it did in the southern Levant (Genz 2015, 103–105). He concludes that there is no archaeological change visible (as “collapse”) around 2200 BC. In the southern Levant, the end of the urban EBA II–III has also been blamed on the 4.2 ka BP event. Increasingly focused work on chronology using radiocarbon dating suggests that the absolute date of EBA III–IV must be raised to c. 2500 BC, in line with Lebanon, and also, importantly, that “de-urbanization” happened not rapidly, but over centuries (Hoflmayer 2015). Thus the “collapse” happened some time before 4.2 ka BP, and Hoflmayer (2015, 120; see also Regev et al. 2012a and 2012b) states that “any sudden climatic change starting around 2200 BC can therefore be excluded as a cause or contributing event”. Greenberg (2017), also looking at the southern Levant, concludes that there was no collapse, rather a long trend of de-urbanization.

Collapse of Bronze Age Civilizations

277

13.6.2 Egypt The collapse of the Old Kingdom around 2160 BC has been blamed on drought and a low Nile, which caused a period of anarchy, drought, and famine, and a dark age—the First Intermediate Period (FIP) (Bell 1971; Hassan 2007). Two studies in the Nile Delta by Krom et al. (2002) and Bernhardt et al. (2012) do identify a very dry spell and a lower Nile level at the time, although Seidlmayer (2000, 119) notes evidence of increased Nile flow at Elephantine. Literary texts such as the Dialogue of Ipuwer and tomb biographies, such as that of the nomarch Ankhtify, are taken to describe the state of chaos in Egypt during the FIP. Rather than a human catastrophe, the collapse of the Old Kingdom seems best described as only a political collapse—the end of a unified Egyptian polity ruled by a pharaoh based at Memphis (Hoflmayer 2015, 121). The view often presented of a complete social and political breakdown, with mass famine, gives far too much credence to the later literary texts (Moeller 2005; Middleton 2017a, 58–60). As Moreno Garcia (2015, 1–2) notes, many of these “pessimistic” texts are much later in date than the events they purport to describe, and their purpose was political and ideological, not historical; they helped to support the unified Middle Kingdom monarchy. Likewise, the contemporary tomb texts were designed to show the occupant of the tomb as a good leader, who, for example, was able to create stability and plenty in his own region, even to the extent of sending food elsewhere (thus implicitly denigrating other leaders). Complex society continued to exist in Egypt and functioned at least as adequately as the unified monarchy had, and culturally, the FIP was a thriving period, which influenced life in the Middle Kingdom (Morris 2006; Seidlmayer 2000). The political collapse can be traced to changes in the structure of government, with an increase in local power at the expense of the centre—it was a long process beginning in the Fifth dynasty, which eventually resulted in the redundancy of central rule (Wilkinson 2010, 116–112). It is possible that a reduction in Nile flow exacerbated the increasing autonomy of regions, perhaps because local rulers would have been better placed to react to local circumstances. However, Moreno Garcia (2015, 14) concludes that “no sudden collapse is apparent, nor any signs of catastrophic environmental events leading to political and economic chaos”. The Egyptian collapse was different in kind to the de-urbanization visible in southern Lebanon and the Levant, which also predated it by a considerable margin.

13.6.3 The Akkadian Empire The theory that the collapse of the Akkadian Empire was the result of a global climate change event, a three century megadrought, has been forcefully made by Weiss, based on extensive study of the Khabur Plains and the collection of other paleoclimatic data (e.g., 2014, 2015, 2016, 2017b). He marshals an impressive range of evidence from which he identifies a 30%–50% reduction in rainfall lasting from 2200 to 1900 BC; the onset of change was rapid—less than five years (Weiss 2015, 46). In his reconstruction, the Akkadian Empire had expanded into the north, where it became dependent on produce from the dry-farming region of the Khabur; control of the area is indicated by the public buildings at Mozan, Leilan, and Brak (Weiss 2015, 39–40; 2017, 99, 105–106). The megadrought ended the productivity of the region and massive depopulation resulted; the Akkadian Empire collapsed. The catastrophe was recorded in texts such as the Curse of Akkad (Weiss 2015, 40). However, Butzer (2012a and b) argues that there are numerous problems with the paleoclimatic data and contra-indications to Weiss’ reconstruction in the archaeology of southern Mesopotamia, such as an increase in the number of sites and canal extensions in Akkadian and post-Akkadian times. He cites isotopic and salinity data from Lake Van to suggest that there was no reduction of flow into the Euphrates. Oates (2014, 1500), working in northern Mesopotamia, finds evidence of increased rainfall. Weiss (2017b, 115–117) has responded to these and other critiques and reemphasized his view of the global nature of 4.2 ka BP climate change, and that as an event, it drove collapse. He states that “counterfactually, these synchronous West Asian and adjacent collapse and

278

Climate Changes in the Holocene

abandonment events and processes would not have occurred without the 4.2 ka BP megadrought” (Weiss 2017b, 115). Even if we accept the evidence for a 4.2 ka BP climate event in the Khabur, not everyone agrees with the narrative or that a megadrought was the direct or primary cause of Akkadian imperial collapse. Zettler, for example, citing work by Mesopotamian text specialists, questions Weiss’ use of literature—the Curse of Akkad and other traditions are not straightforward historical texts but were relevant to the time of writing, up to a century after the collapse (Zettler 2003, 18). He also notes that it is far from certain that the Akkadian dynasty relied on produce from the Khabur (Zettler 2003, 20). Another issue is that of chronology. Butzer (2012b) notes that evidence for the beginning of the megadrought has a 400 year spread, while Zettler points out uncertainties in the historical chronology of Akkad and Mesopotamia. Two suggested chronologies would place the accession of Sargon and the formation of the Akkadian Empire within the period of the proposed megadrought (Gasche et al., and Reade, cited in Zettler 2003, 20). A more plausible explanation of Akkadian collapse is that the empire was unpopular and inherently fragile; each king faced “rebellions” by states that desired a return to independence and each fought wars to reconquer his predecessors’ gains. The empire, while able to marshal considerable resources and manpower, can hardly be regarded as a stable entity; the Akkad dynasty made plenty of enemies, who sought to overthrow it, and were in a position to try. Writing of the elite rebellion faced by Rimush, Westenholz characterizes it as “a desperate, all-out effort to shake off the Akkadian yoke once and for all” (Westenholz 1999, 41). The widespread opposition to the dynasty and empire of Akkad were probably key to its collapse—it failed to successfully integrate the conquered or defeated (Van De Mieroop 2007, 69). The Gutians may also have played a role in bringing about change. Weiss’ deterministic counterfactual claim seems unlikely to be right, given the persistent instability of the Akkadian Empire—indeed, it is surprising the empire lasted so long. It is also important to remember that the Akkadian dynasty continued at a local level and there was no wider Mesopotamian collapse; the Ur III dynasty was eventually able to construct an empire of its own around 2100–2000 BC, which possibly collapsed, as did the Egyptian Old Kingdom, because central power ebbed away to provincial governors who eventually founded their own rival local dynasties (Chavalas 2006, 45). If it is correct that there was a three century megadrought between 2200 and 1900 BC, then it is also the case that complex society and empire-building remained possible even during the period of severe aridity; this required successful agriculture. Climate did not determine the degree of social and political complexity possible.

13.6.4 Summary The above discussion based on recent research shows that the reality of an EBA climate-caused collapse as a synchronous “global” Old World event is unclear; it may be more accurate to regard the EBA collapses as a set of quite different regional events and processes happening over a longer period of time than the label “4.2 ka BP event” indicates. In Greece, it is not clear that collapse is the best label for changes happening throughout EH II–III, and not as suddenly as once thought. It is possible, though not yet certain, that there was some kind of 4.2. ka BP climate event in Greece, but even if this was a three century megadrought, it may have had little effect on society—it did not prevent the increases in complexity and intensification in agriculture on Crete, for example. In Lebanon, there are divergent trajectories, with continuity in northern Lebanon and collapse in the south. In the southern Levant, de-urbanization happened earlier than the Old Kingdom collapse. The latter and the Akkadian collapse were political events in which unified states fragmented but complex society continued. If a three century megadrought happened, it did not prevent the continuation of habitation and complex society in most areas. This does not deny the ever-present threat of bad weather years that caused food shortages, famine, and political and social instability. Before trying to explain collapses by invoking global climate change as a causal factor, even when it appears coincident, it is important to go back to the archaeology and to identify what it is that we

Collapse of Bronze Age Civilizations

279

are seeking to explain and to attempt a greater degree of chronological precision. This is an ongoing project. It is also the case that more paleoclimatic data is needed so that much more localized reconstructions can be made and compared. In this regard, Manning (2017) has offered a clear caution in a careful analysis of Mediterranean-wide climate patterns. He cites evidence of clear variability in a given year across the Mediterranean, for example, wetter in Anatolia and Iberia but dryer in the Levant and Egypt, but also, importantly, of variability within individual regions (Manning 2017, 455–458). Thus “the nature of the data mitigates against generalizing metanarratives where change is simultaneous across a large area like the Mediterranean” (Manning 2017, 458).

13.7 THE LBA COLLAPSES IN GREECE AND THE EASTERN MEDITERRANEAN A number of states headed by kings developed in Greece, probably by 1400 BC (Bennet 2013; Shelton 2010, 143–146). The kings were based in a handful of palaces around Greece from Dhimini in the north to Pylos in the south. These palaces were centres of production, redistribution, and large-scale social activities such as feasting, and they imported and exported materials and goods as far afield as Egypt (Wright 2004). Selective records were kept on clay tablets using the syllabic Linear B script to write in early Greek. The Pylos tablets indicate that it was the capital of a large state, divided into two provinces (Bennet 2007; Hope Simpson 2014). Impressive tholos tombs such as the Treasuries of Atreus (Mycenae) and Minyas (Orchomenos) were built by royalty, and in the palatial period, major fortifications, roads, drainage projects, and an artificial harbor were constructed. The collapse c. 1200 BC, at the end of Late Helladic IIIB (LH IIIB–IIIC transition), meant the end of two centuries of palatial society (Dickinson 2010; Middleton 2010 and 2018a; Knapp and Manning 2016). It is indicated by the fiery destruction of the palace centres and other sites, including Mycenae, Tiryns, Midea, Thebes, Pylos, and Dhimini, as well some sites outside the palace states. Palace-based kingship ended, as did various other traditions strongly associated with the palaces, such as the use of Linear B writing to keep records, monumental building, and fresco painting. The number of sites occupied fell sharply, especially in Messenia, the former Pylos kingdom. Some explanations of the collapse have drawn on much later Greek myths, of the Dorian migration and Return of the Herakleidai, to suggest a northern invasion/migration scenario of collapse (see Middleton 2010, 41–45). Desborough (1972, 22–23) thought that Dorians led by the descendants of Mycenaean kings invaded Mycenaean Greece and then retreated—it was necessary to posit a retreat because there is no archaeological evidence for a new population. More recently, Eder (1998) and Drews (1993) have posited invasions of northern Greeks as the cause of Mycenaean collapse. There are sound criticisms, however, of using myths as garbled but essentially “true” historical sources rather than understanding them as pertaining primarily to their period of production (Hall 1997, 64), and critiques of both Eder’s and Drews’ interpretations (Middleton 2010, 47–48; Voutsaki 2000). Now a series of prominent publications have linked together climate change, collapse, and mass migration of the Sea Peoples (papers in Fischer and Burge 2017; Kaniewski et al. 2015 and Kaniewski et al. 2017). This climate–collapse–migration theory will be discussed below. In addition to the collapse of the Mycenaean palace states at c. 1200 BC, the Hittite Empire in Anatolia also collapsed, after a four century existence, its capital Hattusa was abandoned, its north Levantine vassal kingdom of Ugarit was also destroyed and there were also destructions on Cyprus. Drews (1993, Figure 13.1) mapped out the destructions around the eastern Mediterranean, counting 47, with ten in Greece, two on Crete, 13 in Anatolia and four on Cyprus. These collapses took place in the context of great conflict that marked the reign of the last Anatolian Hittite Great King Suppiluliuma II. In many parts of the east, new cultures such as the Philistines appeared, adopting traditions of Aegean-inspired pottery from Cyprus. Egypt did not collapse at this time, and some Levantine cities continued without a break.

280

Climate Changes in the Holocene

FIGURE 13.1  Map of the eastern Mediterranean region.

13.8 CLIMATE–COLLAPSE–MIGRATION AS THE CAUSE OF LBA COLLAPSES Climate change has been given as an explanation for LBA collapse in Greece by many authors (Carpenter 1966; Bryson et al. 1974; Neumann 1993). Drake (2012) has developed this argument using a range of evidence from the Soreq Cave, Israel, Lake Voulakaria in western central Greece and sediment cores in the Mediterranean (Adriatic, Aegean, eastern Mediterranean (Cyprus), and Ionian seas). He identifies long-term climate change from 1315 BC to 350 BC. This climate change was gradual and was “a continual stress put on human societies in the region for several decades” (Drake 2012, 1866). It brought about economic decline and caused instability in palatial centres, possibly leading to uprisings. Declining productivity in combination with large urban populations thus led to collapse and population movement in the form of the Sea Peoples (Drake 2012, 1867–1868). Kaniewski and colleagues (2013; 2015; 2017) posit rapid climate change and a three century “drought event” for the eastern Mediterranean from 1200 to 900 BC. This is based on samples from Cyprus and coastal Syria. They suggest that climate change led to massive land invasions from the Balkans into Greece and Anatolia and then by sea from Greece to Italy, Anatolia (west and south), North Africa, Cyprus, the Levant, and Egypt—they identify the migrants as becoming the Sea Peoples and assign specific destruction events to them. Langgut and colleagues (2013) examined a core from the Sea of Galilee and identified the driest phase at 1250–1100 BC. In their reconstruction, a shift to a cold, dry climate on northern Greece and Anatolia caused famines, which drove people south, and pushed others further south and east. Land and sea trade were disrupted, and pirates raided cities; the Sea Peoples raided in search of food. A return of wet conditions caused complex society to develop again in the Levant. These two research groups also identify drought and famine in textual sources which report grain shipments from Egypt to the Hittites.

13.8.1 Critique There is much that is puzzling about the LBA climate change collapse hypothesis as it has been presented. Firstly, the climatic reconstructions differ significantly in terms of onset and duration: Drake (2012) 1315–350 BC, gradual onset; Kaniewski et al. (2013; 2015; 2017) 1200–900 BC, rapid onset; and Langgut et al. (2013) 1250–1100 BC, rapid onset. There is also an issue with consistency of argument and potential determinism. Thus Drake (2012, 1868) suggests that while the Mycenaean collapse was caused by climate change, the development of Archaic and Classical Greek

Collapse of Bronze Age Civilizations

281

society, still in the climatic downturn, was due to human innovation. Quite why a gradual climate change would cause such a problem for the Mycenaeans, then, is unclear—in Drake’s dating, they were thriving during the first 100-odd years of the shift. Kaniewski’s reconstruction ignores the post-palatial resurgence in the Argolid and the continuities and developments elsewhere in Greece. Langgut et al. (2013, 161) link decreasing and increasing complexity with climate change. In all cases, climate change requires an agent to physically bring about the destructions and collapse— equally, agents could operate with climate change as mere background noise; assigning causality and weighting different factors is difficult. There is also the puzzle of migration: if there was a widespread drought in Greece, Anatolia and the eastern Mediterranean, it must be asked why people migrated into these regions en masse. If there were migrations, clearly people were able to feed themselves at their destinations all around the area, which suggests surpluses were available that locals could share (or that could be stolen). This in turn would suggest that any climate change had a limited impact. Archaeologically, however, there is no evidence of invaders in Greece c. 1200 BC or in Anatolia at this time (Middleton 2010, 41–45). While there is a reduction in the total number of sites, most noticeable in Messenia, southwest Greece, post-palatial LH IIIC Tiryns seems to have increased in size, and the Achaea/ Ionian region continued with much evidence of activity (Maran 2010, 729; Middleton 2010). Thus any climate change did not make habitation in the Argolid region of Greece (and other regions) impossible; there must have been other reasons why the earlier palatial culture was not reproduced after c. 1200 BC. New paleoclimatic data from southwest Greece has just been published (Finné et al. 2017). This comes from a stalagmite in the Mavri Trypa Cave, Schiza Island, located 4 km off the coast of southwestern Greece. This is an important source of evidence because it provides a high-resolution climate record from the area of a LBA Mycenaean palace and kingdom—Pylos and Messenia. It is thus likely to be much more useful than evidence from further afield. This research suggests a slight drying in the thirteenth century, followed by a wetter phase, during which the palace was destroyed. Conditions then became dryer across the twelfth century, reaching a peak of dryness during the Protogeometric. Rather than drought causing collapse, the authors of this paper suggest that drying after the collapse made it difficult to reconstruct the Pylos kingdom because agricultural output and conditions worsened. The other cultures that developed around the eastern Mediterranean, using “Aegean-looking” pottery, were also likely products of continued mobility and contact rather than mass migration (Killebrew 2014; Middleton 2015 and 2018b), and in the Levant, not all the LBA cities and states were destroyed, and there were many cultural continuities in some of these into the EIA (Killebrew 2014, 597). The development of the Neo-Hittite kingdoms in the twelfth century must also be reckoned with—a Hittite “great king” ruled from Carchemish, for example (Bryce 2012). Cyprus too has been seen as a destination for Greek emigrants, but Voskos and Knapp (2008) question this narrative, and instead note that the new elements in LC IIIA are hybridized, not simply “Mycenaean”. Cyprus remained an urbanized culture through the LBA/EIA transition, despite changes in material culture and the pattern of settlement (Steel 2010 and 2014). All in all, if there were droughts, they do not seem to have caused too much of a problem in continuing habitation and cultural continuity and/or development, which required a functioning agricultural basis. What of the Sea Peoples? The narrative of violent migration is very much a modern invention, based on interpretations of Egyptian evidence (Drews 1993 and 2000; Knapp and Manning 2016; Middleton 2015; Silberman 1998). The famous Year 8 text from Ramesses III’s mortuary temple at Medinet Habu, which describes a coalition of peoples rampaging over the Near East, destroying states and cities, cannot be regarded as historical: it is wrong in some details, such as the destruction of Carchemish (there was no destruction at this time) and Arzawa (it had ceased to be a state a century earlier), and misses out detail that might be expected, such as the destruction of Ugarit (Bryce 2005, 197; Sagona and Zimansky 2009, 299–301). Such texts were not written with literal truth in mind; they were not “historical” as we understand the term.

282

Climate Changes in the Holocene

Rather, they were intended to represent the pharaoh in an appropriate way—defeating (strong and numerous) enemies of Egypt (Roberts 2009; Wilkinson 2010, 56; Wilson 1956). Given the often-unacknowledged problems with the textual evidence for the Sea Peoples story, it is highly questionable to combine this evidence with a theory of climate change to explain widespread collapse. The written sources referred to come from Egyptian/Hittite correspondence, from Ugaritic texts, and from Assyrian and Babylonian texts, which date altogether from c. 1250 to 900 BC (Kaniewski 2015, 4–5; Langgut et al. 2013, 165; also Halayqa 2010, 301–304). A number of oft-quoted texts mention grain shortages and shipments of grain from Egypt to Hatti via Ugarit. Kaniewski et al. (2015, 5) thus claim that from around 1250 BC the Hittite capital of Hattusa and the kingdom itself were “no longer self-sustainable in food production and had to rely on imports”. Yet, while many see these texts as evidence for a great drought and famine in Anatolia, this remains speculation. We do not know the purpose of the grain shipments to Hatti. The shipments of grain could have been used as a dole at Hattusa (somewhat like the Roman dole of Egyptian grain), or have been disbursed by the king to the elite, the military, or other followers. Grain shortages may have been a destabilizing factor without any widespread drought in the Near East. Whatever the truth of the matter, it has to be recognized that the Hittite kingdom, its rulers, its military, and its administration continued to function into the reign of Suppiluliuma II (c. 1207 BC), and major building work was carried out at Hattusa (as at Mycenae and Tiryns in Greece) up until the city was abandoned. Ugarit also continued to function until its destruction—incidentally, Halayqa (2010, 304–305) has suggested that Ugarit’s food shortage was a pretense in correspondence with the Hittites. The palace states of Greece also seem to have functioned down to the point of collapse, and there is no compelling evidence of decline or anxiety in the later thirteenth century Mycenaean world (Maran 2009; Middleton 2017b). There is no evidence of food shortages in Greece at the end of the thirteenth century—“the Linear B texts of Pylos give no clear hint that there is anything wrong in the agricultural system” (Dickinson 2006, 55).

13.8.2 Summary As far as the evidence goes, it is far from certain that climate change was the cause, or even a main cause, of the LBA collapses in Greece or elsewhere. The new climatological evidence from Messenia suggests there was no megadrought that caused collapse there; the other evidence suggests different things to different people. The archaeological and historical evidence for invasions and mass migrations is highly questionable, with a lot of narrative built on very shaky foundations; the textual evidence for massive long-term droughts and famines is ambiguous at best. Evidence of continuities in the Argolid and elsewhere in Greece, and other developments, for example, the Cypriot transformations, the growth of Neo-Hittite kingdoms, and the appearance of the Philistines, also cast doubt on any issues of subsistence difficulties. Notably, Egypt did not collapse c. 1200 BC. What we do have evidence of is conflict—a stock feature of the Late Bronze Age eastern Mediterranean. Conflict would have been sufficient to bring down Mycenaean states and the Hittite Empire, and conflict is an explicit and major theme of texts from the end of the LBA. Manning’s (2017) caution about using paleoclimatic data to generalize a widespread Mediterranean or eastern Mediterranean drought, mentioned above, is equally relevant to the LBA collapse. The data used to propose an LBA climate change is too little and has been linked with difficult evidence to create a plausible-sounding narrative, but one that is really very speculative. Again, this suggests a need for the collection of a great deal more data so that much more detailed dynamic climate maps of regions can be constructed. Several projects in Greece are working towards this, and their results are eagerly awaited. A simple equation of climate change equaling collapse and mass migration also requires questioning.

Collapse of Bronze Age Civilizations

283

13.9 WORKING TOGETHER: OPPORTUNITIES AND ISSUES Understandings of collapse and historical change, and human–environment relationships, can be furthered by a better mutual understanding between archaeologists, paleoclimatologists and others, and specialists from different areas are increasingly calling for this (Izdebski et al. 2016). From the archaeology “camp”, Mitchell (2008, 1095) has pointed out that “archaeologists are in a privileged position to understand and make sense of the impact of climate change on human populations”. This comes in part from their study of communities situated within environments, and of long-term processes, which involve a range of types of evidence (Mitchell 2008, 1096). Paleoclimatic data can help in the development of long-term climate models. Holistic models of historical development that involve a range of evidence, can inform policy makers and the public. As Dawdy (2009) wrote, archaeology can be and should be “useful”. Hudson and colleagues (2012) similarly identified five ways in which archaeology can contribute to the climate change discourse through: 1. the study of historical collapse, resilience, and reorganization; 2. the study of nature/culture relationship; 3. the use of public archaeology to educate and raise awareness of past environmental impacts; 4. the study of social injustice and its relationship to responses to the environment; and 5. building “intercultural responses to climate change”. For them, the challenges archaeologists face include:

1. demonstrating “the relevance of archaeology to present and future climate change”; 2. how people can learn from the past; 3. how different archaeological traditions affect the archaeology–climate change debates; and 4. how approaches to historical human agency are affected by climate change.

The potential, and also the responsibilities, of archaeologists have been recognized by many others, including recently Kintigh et al. (2014, 879) who note already that “archaeological evidence frequently undergirds debate on contemporary issues”. In setting goals for the future of archaeology Kintigh and colleagues identified five key areas for exploration. The area of human–environment interactions included the questions “how do humans respond to abrupt environmental change?” and “how do humans perceive and react to changes in climate and the natural environment over shortand long-terms?”, amongst others. Van de Noort (2011) also argued that understanding past successful adaptations to climate change can help grow resilience in present-day communities. There is a desire, therefore, to place human–environment relations, including but also going beyond collapse, on the archaeological agenda. Both Mitchell (2008) and Riede et al. (2016) discuss the issue of whether archaeologists in fact have an ethical obligation to participate in the wider climate change and environmental discourses. Mitchell suggests that archaeologists do have this wider social responsibility and states that they “can usefully contribute to academic, popular and political debates on the prediction and the management of climate change” (Mitchell 2008, 1096). Riede et al. (2016, 467–469) identify a widening and deepening interest in environmental issues throughout the humanities and an increase in ethics thinking in disciplines such as geology. The “Geoethical Promise” includes acknowledging concern for the social implications of research, knowing researchers’ responsibilities towards society and future generations in terms of sustainable development, and suggests that researchers will “put the interest of society at large in the first place” (Matteuci et al. 2014, 191, quoted in Riede et al. 2016, 468). They argue that “current debates on climate change and climate catastrophe … make it an urgent matter to embed ethical concerns in, at the very least, environmental archaeological practice” (Riede et al. 2016, 469).

284

Climate Changes in the Holocene

Archaeologists are able to contribute data appropriate for climatological study because “the archaeological record often incorporates important, sometimes unique, proxy records of climate, of environment, and of change in both” (Sandweiss and Kelly 2012, 372). Thus Dincauze (2000, 176) points to some of the responsibilities archaeologists have with regards to climatology: they should know “(1) how to use the climatological literature responsibly, (2) how to collect climatological data at the micro-scale, and (3) how to consult with specialists for interpretation and integration of the data at all scales”. These are all issues of professional development which could easily be introduced into undergraduate courses on archaeology; contributions to such modules could be made by specialists from outside archaeology departments. It would also be helpful to have archaeologists and historians participating in classes on applied paleoclimatology. Dawdy (2009, 140) also notes that meetings with timely themes such as climate change could be convened. A recent example of a climate collapse-focused meeting of archaeologists is the 2014 Archaeological Conference of Central Germany in Halle (Meller et al. 2015), while in December 2016, University College London hosted a multidisciplinary conference entitled “Climate Change, Archaeology and History: A Multidisciplinary Approach From Archaeology, Climatology And History On Climate Change And The Possible Collapse Of Civilization”. At Princeton, there is also the Climate Change and History Research Initiative, which brings together archaeologists, historians and scientists in a co-operative environment. Inter- and multi-disciplinarity comes naturally to archaeologists, as co-operative approaches have long been an accepted and desired part of archaeological work. Indeed, a century ago, Huntington stated that: “the lessons of history cannot rightly be understood until the combined work of men in many lines gives us a clear idea of each one of the complex factors leading to such great events as the fall of Rome” (Huntington 1917, 208). If archaeologists and historians need to understand climatology better, and to integrate its data and results into their own work, and, importantly, convey their own work to others outside the discipline, the opposite is also true; those in the harder sciences need to be sensitive to how archaeology and history work as disciplines, and how they generate knowledge. One issue is to understand the dynamic and hypothetical nature of archaeology, which is always a work in progress and, when going beyond descriptions of data, involves interpretation and the constructions of narratives; these are always subject to critique, questioning, and change. They are also open to polyvocality—in an instance of collapse, for example, there may well be clear winners and losers, while for others little may change; to paint collapse simply as a “bad” or “undesirable” thing that must be explained and avoided is to oversimplify complex history peopled with conscious actors. Archaeological and historical knowledge is often not reducible to either/or. Another issue is with the nature of historical events and the demarcation of periods. “Events” once identified can take on a life of their own in academia —for example the Classic Maya collapse. Whilst recognizing the differences between the Classic and Postclassic, many archaeologists would question whether there was “a” Classic Maya collapse (Aimers 2007; Ardren 2005; Johnson 2017, 7). What happened over three centuries was the collapse of many individual polities, leading (unsurprisingly) to an overall change in culture and landscape. The wider depopulation often cited as a key feature of the Maya collapse happened over an extended period of time. Thus while we employ terms like “Classic”, “Terminal Classic”, and “Postclassic”, these are used for our convenience, and should be understood as such. The cautionary tale here is that we need to consider what it is that we are trying to explain, before we try to explain it. It is important to pay attention not only to data here but also to historiography—sometimes the object of study is a modern construct or shorthand (Middleton 2017a). Another important issue is the sometimes imprecise absolute or floating chronologies with which archaeologists work. The problem of dating the accession of Sargon was mentioned above. It is important to recognize that the different dates can mean very different histories—Sargon may have won his empire during the proposed megadrought, rather than a megadrought destroying the empire under his great-grandson Sharkalisharri (Middleton 2017a, 74). It is tendentious to argue that, because the empire collapsed and because there was climate change that seems roughly coincident,

Collapse of Bronze Age Civilizations

285

the two were simultaneous and causally connected. The proposed thirteenth century climate change and Mycenaean collapse c. 1200 BC is a similar case. For archaeologists, clearer descriptions of the rapidity of the onset of proposed climate changes, for example, the meaning in each case of “abrupt” or “gradual”, and the range of uncertainty in dating would be helpful. Other terms too, e.g., “dry”, “wet” etc. might be more helpfully defined. Finally, most archaeologists, though clearly not all, find simplistic and deterministic “explanations” of collapse as being due to climatic or environmental changes unsatisfactory in most cases. Societies and social change are complex, involving multiple actors and factors. Truthfully, we sometimes find it difficult to satisfactorily explain even contemporary events—for example, the fall of communism, the Arab Spring, the UK’s urban riots may all have been predictable in hindsight, but still few saw them coming. Recognizing this does not make archaeologists climate sceptics, or mean that they dismiss the potential contributions of paleoclimatology to the study of the human past. But the specter of environmental determinism, in which humans are passive victims of forces beyond their control, most recently represented by Weiss’ new volume (2017a), is a particular problem when the trend in archaeology is towards recognizing and crediting past people with agency and choice (Dobres and Robb 2000; Patterson 2005). Williams (2002, 372) puts it well when he concludes that “collapse and survival are social processes, affected by but not driven by their environments” and therefore that “the magnitude of a natural hazard cannot be equated with the magnitude of its social impact”. It is not just the disaster, but how people respond to it that is of interest. As he explains, it is key to consider “social elements” in narratives of collapse and resilience in the face of environmental change, whilst also admitting that change may happen predominantly for social, rather than environmental reasons. Williams suggests that “archaeology must continue to utilize both social and environmental data in order to understand the long-term impacts of ecodisasters on the landscapes humans inhabit”. In this, though, there is a difficulty in that, when a collapse (or other historical change) is seen primarily through a social lens, with no contributing environmental cause identifiable, where does the environmental data gathered then fit? When the circumstances are reversed, however, and there is a clear environmental disaster or change, such as a significant aridification, it is clear how environmental factors could have had an impact—but not always easy to demonstrate that they actually did have an impact. In other words, we are still often left with the problem of situating climatological evidence within narrative interpretations—driver, contributor or just background noise? It is almost a given now that publications on past collapse, especially those that refer to potential environmental causes, discuss the prospect of similar events causing future collapse and ask us to consider past collapses as “lessons”. Some, like Diamond’s Collapse: How Societies Choose to Fail or Succeed, and its forebears and successors, explicitly use past collapses in this way, but so do some articles in academic archaeology. A note of caution must be sounded here, that “disinterested” or objective history can sometimes fall foul of the desire to create a didactic narrative—and this is especially visible in some parts of the environmental literature and in archaeological writing too (Middleton 2017c; 2012). In functional terms, this changes the genre of writing—and of knowledge—from archaeology or history to moral, political and behavioral exhortation. Applying ethical guidelines to archaeologists along the lines of the Geoethical Promise would seem a step too far, but this in no way means that archaeologists cannot, and should not, seek co-operation and a wider audience for their relevant studies. In turn, certain sectors of the media, including the news, popular science publishing, and TV documentary commissioners, must be more accepting of the kind of knowledge and narratives that archaeologists and historians construct rather than focusing on simplistic silver-bullet explanations of collapse that come from the harder sciences and which make for good headlines. Bookstores, who also act as gatekeepers of knowledge, should also take care not to only give the public a choice of popular or theoretically fashionable texts at the expense of a variety of authoritative views. A key point from both approaches is always to question what it is that we are seeking to explain—each “collapse”—before we try to explain it. This means keeping up-to-date with what

286

Climate Changes in the Holocene

the archaeology shows and does not show. It causes confusion when explanations are proposed for what are essentially factoids or interpretations that have been superseded. It is also necessary to highlight issues of chronology and their implications for devising explanations. It is appropriate too to question the paleoclimatic data, what it does and does not show, its chronological precision, and to consider the extent to which limited data sets can be applied over larger areas. Proving the existence of climate change around times of collapse or significant change is not sufficient to make it causally connected. Both archaeologists and paleoclimatologists would benefit from greater cooperation and mutual understanding of each other’s fields and discourses, the way in order to take matters forward in a positive way.

13.10 CONCLUSION This chapter has demonstrated that there are a number of problems in explaining the EBA and LBA collapses as a result of climate change. Recent research suggests that the EBA collapses were not all contemporaneous, meaning that change in some places occurred without the suggested 4.2 ka BP climate change event. Furthermore, the collapses themselves were of different kinds—imperial or state collapse in Mesopotamia and Egypt, with continuity in complexity and population; deurbanization in parts of the Levant; clear culture change and apparent simplification in Greece that happened over time, but increasing complexity on Crete. On the face of it, it seems questionable to assume that these were connected or part of a single collapse “event”. There are other difficulties in the study of the Early Bronze Age collapse. One is that the “event” is spread across regional subfields of archaeology, in which practitioners may tend to focus on their own region more than others, or may not be overly familiar with up-to-date or reliable views on other areas. A concomitant difficulty when it comes to climatic explanations of EBA collapse is that non-archaeologists may be even less familiar with the state of archaeological knowledge in every area—and even more problematic, may place more weight on older, hypothetical or tentative interpretations and narratives as representing “historical fact” than contemporary archaeologists themselves would advise. There is a danger that constructed events take on a reality that they do not deserve. A similar problem holds with the climatological evidence. The primary paleoclimatic studies are spread over many publications and the amount of data only ever grows. It is difficult for paleoclimatoligists to stay current, let alone archaeologists, who are less familiar with the journals. It might be advisable to explore ways of storing and making accessible paleoclimatic data using new technology. Some kind of central storage would make future study much more practically possible, and more accurate. Evidence for a climate-driven LBA collapse in Greece and the eastern Mediterranean is slight and ambivalent and again what may be primarily local events (on different scales) and processes may be being wrongly conflated; it is difficult to tell (Cline 2014, 170). The new evidence from Messenia suggests climate change played no role in Mycenaean collapse. The stories of mass migrations into, and then out of, Greece and Anatolia now have little credence, at least amongst Aegean archaeologists. It would seem unlikely that desperate people would migrate towards regions plagued by severe droughts or into conflict zones. The textual references to grain shipments in the eastern Mediterranean, of relevance to the Hittite collapse, are of unclear significance, and need not relate either to climate or drought. Research on dating the EBA and its changing patterns of culture and settlement is ongoing and will clarify our picture of events and processes across the Old World and their possible relations. In order to clarify the paleoenvironmental circumstances around c. 1200 BC, much more data collection needs to be done in each region—and indeed in each microregion; there are a number of ongoing projects in this regard. It could be the case that the different trajectories visible in Greece after c. 1200 BC, with the destruction and abandonment of the Pylos palace and apparently great depopulation in the Messenia region, and a greater degree of continuity—at former palace sites such

Collapse of Bronze Age Civilizations

287

as Tiryns and Mycenae—owed something to different environmental or climatic circumstances, but this will not be picked up by data from Cyprus or Galilee. Manning’s (2017) comments on generalizing pan-Mediterranean patterns from tiny amounts of data should be widely noted. Even if there is a correlation, however, it will not be sufficient to prove causality, and this is a problem for (re) constructing history. It must be accepted that the changes we see can be explained without climate change—conflict, of which there is plenty of evidence, should surely be regarded as key, and with conflict come agricultural and labor issues, economic and political issues, and disease. Archaeologists and paleoclimatologists are all aware of the power of the environment, and that droughts have caused great misery and suffering throughout history. Besides positing significant climate changes as simple causes of collapse, we should not forget the effect of variability in weather on much shorter scales. As Dincauze reminds us, for pre-industrial populations “at interannual scales, it is weather rather than climate that is most strongly felt by people …” (Dincauze 2000, 157). There are very clear and uncontroversial examples from history of the interplay between weather, food shortages, politics and social unrest and change. This chapter has also discussed some of the ways in which archaeologists, historians, and paleoclimatologists could work together more productively, by seeking greater understanding of the different discourses of each other’s fields. Such greater understanding would facilitate the construction of better and more accurate models of human–environment dynamics, which in turn would give policy makers and the public a much more nuanced and sophisticated idea of how historical change has come about. I suggest that it is dangerous and wrong to see historical change, and then by implication future change, as primarily externally imposed. A more prominent place should be given in the wider collapse discourse to social factors rather than continuing to promote deterministic (or quasi-deterministic) environmental factors. This would be in line with several recent contributions to the collapse debate (eg. Butzer 2012a, 3638; Middleton 2012, 2017a,d; Storey and Storey 2017, 228–229; Scheidel 2017; Turchin 2010). To conclude, the evidence for climate-caused EBA and LBA collapses is ambivalent at best. But what should be clear from this chapter is that increased co-operation and also mutual understanding between archaeologists and paleoclimatologists and others, and increased synthesis of results from their respective areas of specialism, could lead to more sophisticated scenarios with regard to understanding past collapses and historical change and also could provide useful data to those seeking to promote the development of more resilient and sustainable societies.

ACKNOWLEDGMENTS I would like to thank Martin Finné for discussing the paleoclimatological evidence for Greece and Erika Weiberg for discussion of the archaeological evidence for the EBA Aegean, and for their very helpful comments on my draft; their input has improved the chapter greatly. Oliver Dickinson also kindly read the draft and saved me from numerous errors, for which I am grateful.

REFERENCES Aimers, J.J. 2007. What Maya collapse? Terminal classic variation in the maya lowlands. Journal of Archaeological Research 15: 329–377. Ardren, T. 2005. Review: The fall of the ancient Maya: Solving the mystery of the Maya collapse by David Webster. Latin American Antiquity 16: 225–226. Bell, B. 1971. The dark ages in ancient history: 1. The first dark age in Egypt. American Journal of Archaeology 75: 1–26. Bennet, J. 2007. Pylos: The expansion of a Mycenaean palatial center. In Rethinking Mycenaean Palaces II. Revised and expanded second edition, edited by M.L. Galaty and W.A. Parkinson, 29–39. Los Angeles: The Cotsen Institute of Archaeology. Bennet, J. 2013. Bronze age Greece. In The Oxford Handbook of the State in the Ancient Near East and Mediterranean, edited by P. Bang and W. Scheidel, 235–258. Oxford: Oxford University Press.

288

Climate Changes in the Holocene

Bernhardt, C.E., B.P. Horton, and J. –D. Stanley. 2012. Nile Delta vegetation response to Holocene climate variability. Geology 40(7): 615–618. Booth, R.K., S.T. Jackson, S.L. Forman, J.E. Kutzbach, E.A. Bettis, J. Kreig, and D.K. Wright. 2005. A severe centennial-scale drought in mid-continental North America 4200 years ago and apparent global linkages. The Holocene 15(3): 321–328. Broodbank, C. 2013. Minding the gap: Thinking about change in Early Cycladic island societies from a comparative perspective. American Journal of Archaeology 117(4): 535–543. Bryce, T. 2005. The Kingdom of the Hittites. New edition. Oxford: Oxford University Press. Bryce, T. 2012. The World of the Neo-Hittite Kingdoms: A Political and Military History. Oxford: Oxford University Press. Bryson, R.A., H.H. Lamb, and D.R. Donley. 1974. Drought and the decline of Mycenae. Antiquity 48: 46–50. Buntgen, U., W. Tegel, K. Nicolussi, M. McCormick, D. Frank, V. Trouet, J.O. Kaplan, et al. 2011. 2500  years of European climate variability and human susceptibility. Science 331: 578–582. Butzer, K.W. 2012a. Collapse, environment, and society. Proceedings of the National Academy of Science 109(10): 3632–3639. Butzer, K. W. 2012b. Supporting information. Available at: www.pnas.org/cgi/content/short/1114845109 Carpenter, R. 1966. Discontinuity in Greek Civilization. Cambridge: Cambridge University Press. Caskey, J.L. 1960. The early helladic period in the Argolid. Hesperia 29(3): 285–303. Chavalas, M. W., ed. 2006. The Ancient Near East: Sources in Translation. Oxford: Blackwell. Cheng, H., A. Sinha, S. Verheyden, F.H. Nader, X.L. Li, P.Z. Zhang, J.J. Yin, et al. 2015. The climate variability in northern Levant over the past 20,000 years. Geophysical Research Letters 42: 8641–8650. Cline, E H. 2014. 1177 BC: The Year Civilization Collapsed. Princeton: Princeton University Press. Dalfes, H.N., G. Kukla, and H. Weiss, eds. 1997. Third Millenium BC Climate Change and Old World Collapse. Berlin: Springer. Dawdy, S.L. 2009. Millennial archaeology. Locating the discipline in the age of insecurity. Archaeological Dialogues 16(2): 131–142. Desborough, V. 1972. The Greek Dark Ages. London: Ernest Benn Ltd. Dickinson, O.T. 2006. The Aegean from Bronze Age to Iron Age: Continuity and Change Between the Twelfth and Eighth Centuries BC. London: Routledge. Dickinson, O.T. 2010. The collapse at the end of the Bronze Age. In The Oxford Handbook of the Aegean Bronze Age, edited by E.H. Cline, 483–490. Oxford: Oxford University Press. Dincauze, D. 2000. Environmental Archaeology: Principles and Practice. Cambridge: Cambridge University Press. Dobres, M.-A., and Robb, J., eds. 2000. Agency in Archaeology. Abingdon: Routledge. Drake, B. L. 2012. The influence of climatic change on the late Bronze Age collapse and the Greek dark ages. Journal of Archaeological Science 39: 1862–1870. Drews, R. 1993. The End of the Bronze Age: Changes in Warfare and the Catastrophe ca. 1200 BC. Princeton: Princeton University Press. Drews, R. 2000. Medinet Habu: Oxcarts, ships and migration theories. Journal of Near Eastern Studies 59: 161–190. Eder, B. 1998. Argolis Lakonien Messenien. Vom Ende der mykenischen Palastzeit bis zur Einwanderung der Dorier. Vienna: Verlag der Österreichischen Akademie der Wissenschaften. Fairbridge, R.W. 2009a. History of paleoclimatology. In Encyclopedia of Paleoclimatology and Ancient Environments, edited by V. Gornitz, 414–428. New York: Springer. Fairbridge, R.W. 2009b. History of paleoclimatology—biographies. In Encyclopedia of Paleoclimatology and Ancient Environments, edited by V. Gornitz. 428–438. New York: Springer. Finné, M., and K. Holmgren. 2010. Climate variability in the eastern Mediterranean and the Middle East during the Holocene. In The Urban Mind: Cultural and Environmental Dynamics, edited by P. Sinclair, G. Nordquist, F. Herschend, and C. Isendahl, 29–60. Uppsala: Department of Archaeology and Ancient History, Uppsala University. Finné, M., K. Holmgren, H.S. Sundqvist, E. Weiberg, and M. Lindblom. 2011. Climate in the eastern Mediterranean, and adjacent regions, during the past 6000 years—A review. Journal of Archaeological Science 38: 3153–3173. Finné, M., K. Holmgren, C.-C. Shen, H.-M. Hu, M. Boyd, and S. Stocker. 2017. Late Bronze Age climate change and the destruction of the Mycenaean Palace of Nestor at Pylos. PLoS One 12(12): e0189447. https://doi.org/10.1371/journal.pone.0189447

Collapse of Bronze Age Civilizations

289

Fischer, P.M., and T. Burge., eds. 2017. Sea Peoples’ Up-To-Date: New Research on Transformations in the Eastern Mediterranean in the 13th–11th Centuries BCE. Sonderdruck: Austrian Academy of Sciences Press. Forsén, J. 1992. The Twilight of the Early Helladics: A Study of the Disturbances in East-Central and Southern Greece towards the End of the Early Bronze Age. Jonsered: Paul Astroms Forlag. Forsén, J. 2010. Mainland Greece. In The Oxford Handbook of the Aegean Bronze Age, edited by E.H. Cline, 53–65. Oxford: Oxford University Press. Gauss, W. 2010. Aegina Kolonna. In The Oxford Handbook of the Aegean Bronze Age, edited by E.H. Cline, 737–751. Oxford: Oxford University Press. Genz, H. 2015. Beware of environmental determinism: The transition from the Early to the Middle Bronze Age on the Lebanese coast and the 4.2 ka BP event. In 2200 BC – A Climatic Breakdown as a Cause for the Collapse of the Old World? edited by H. Meller, H.W. Arz, R. Jung, and R. Risch, 97–111. Halle (Saale): Landesamt fur Denkmalplflege und Archaologie Sachsen-Anhalt, Landesmuseum fur Vorgeschichte. Greenberg, R. 2017. No Collapse: Transmutations of Early Bronze Age urbanism in the Southern Levant. In The Late Third Millennium in the Ancient Near East: Chronology, C14, and Climate Change, edited by F. Hoflmayer, 33–58. Chicago: Oriental Institute Press. Halayqa, I.K. 2010. The demise of Ugarit in the light of its connections with Hatti. Ugarit-Forschungen 42: 297–330. Hall, J.M. 1997. Ethnic Identity in Greek Antiquity. Cambridge: Cambridge University Press. Hassan, F.A. 2007. Droughts, famine, and the collapse of the Old Kingdom: Re- reading Ipuwer. In The Art and Archaeology of Ancient Egypt: Essays in Honour of David B. O’Connor. Vol. 1, edited by Z. Hawass, and J. Richards, 357–377. Cairo: Conseil Supreme des Antiquites de’l Egypte. Hoflmayer, F. 2015. The southern Levant, Egypt, and the 4.2 ka BP event. In 2200 BC – A Climatic Breakdown as a Cause for the Collapse of the Old World? edited by H. Meller, H.W. Arz, R. Jung, and R. Risch, 113– 130. Halle (Saale): Landesamt fur Denkmalplflege und Archaologie Sachsen-Anhalt, Landesmuseum fur Vorgeschichte. Hoflmayer, F., ed. 2017. The Late Third Millennium in the Ancient Near East: Chronology, C14, and Climate Change. Chicago: Oriental Institute Press. Hope Simpson, R. 2014. Mycenaean Messenia and the Kingdom of Pylos. Philadelphia: INSTAP Academic Press. Hudson, M.J., M. Aoyama, K.C. Hoover, and J. Uchiyama. 2012. Prospects and challenges for an archaeology of global climate change. WIRES Climate Change. DOI:10.1002/wcc.174 Huntington, E. 1917. Climatic change and agricultural exhaustion as elements in the fall of Rome. Quarterly Journal of Economics (fall): 173–208. Izdebski, A., K. Holmgren, E. Weiberg, S.R. Stocker, U. Buentgen, A. Florenzano, A. Gogou. 2016. Realising consilience: How better communication between archaeologists, historians and natural scientists can transform the study of past climate change in the Mediterranean. Quaternary Science Reviews 136: 5–22. Johnson, S.A. 2017. Why Did Ancient Civilizations Fail? New York: Routledge. Jung, R., and B. Weninger. 2015. Archaeological and environmental impact of the 4.2 ka cal BP event in the central and eastern Mediterranean. In 2200 BC—A Climatic Breakdown as a Cause for the Collapse of the Old World? edited by H. Meller, H.W. Arz, R. Jung, and R. Risch, 205–234. Halle (Saale): Landesamt fur Denkmalplflege und Archaologie Sachsen-Anhalt, Landesmuseum fur Vorgeschichte. Kaniewski, D., E. Van Campo, J. Guiot, S. Le Burel, T. Otto, and C. Baeteman. 2013. Environmental roots of the Late Bronze Age crisis. PLOS One 8(8): e71004: 1–10. Kaniewski, D., J. Guiot, and E. Van Campo. 2015. Drought and societal collapse 3200 years ago in the eastern Mediterranean: A review. Wiley Interdisciplinary Reviews: Climate Change. DOI:10.1002/wcc.345 Kaniewski, D., and Van Campo, E. 2017. 3.2  Ka BP megadrought and the Late Bronze Age collapse. In Megadrought and Collapse: From Early Agriculture to Angkor, edited by H. Weiss, 161–182. Oxford: Oxford University Press. Killebrew, A.E. 2014. Introduction to the Levant during the transitional Late Bronze Age/Iron Age I and Iron Age I periods. In The Oxford Handbook of the Archaeology of the Levant c. 8000-323 BCE, edited by A.E. Killebrew and M.L. Steiner, 595–606. Oxford: Oxford University Press. Kintigh, K.W., J.H. Altschul, M.C. Beaudry, R.D. Drennan, A.P. Kinzig, T.A. Kohler, W.F. Limp, et al. 2014. Grand challenges for archaeology. Proceedings of the National Academy of Sciences 111: 879–880. Knapp, A.B. and S.W. Manning. 2016. Crisis in context: The end of the Late Bronze Age in the eastern Mediterranean. American Journal of Archaeology 120(1): 99–149.

290

Climate Changes in the Holocene

Krom, M.D., J.D. Stanley, R.A. Cliff, and J.C. Woodward. 2002. Nile River sediment fluctuations over the past 7000 yr and their key role in sapropel development. Geology 30(1): 71–74. Langgut, D., I. Finkelstein, and T. Litt. 2013. Climate and the Late Bronze Age collapse: New evidence from the southern Levant. Tel Aviv 40: 149–175. Manning, S.W. 2017. Comments on climate, intra-regional variations, chronology, the 2200 BC horizon of change in the eastern Mediterranean region, and socio-political change on Crete. In The Late Third Millennium in the Ancient Near East: Chronology, C14, and Climate Change, edited by F. Hoflmayer, 451–490. Chicago: Oriental Institute Press. Maran, J. 1998. Kulturwandel auf dem griechischen Festland und den Kykladen im späten 3. Jahrtausend v. Chr. Studien zu den kulturellen Verhältnissen in Südosteuropa und dem zentralen sowie östlichen Mittelmeerraum in der späten Kupfer- und frühen Bronzezeit. Bonn: Rudolf Habelt GmbH. Maran, J. 2009. The crisis years? Relections on signs of instability in the last decades of the Mycenaean palaces. Scienze dell’antichita. Storia Archeologia Antropologia 15: 241–262. Maran, J. 2010. Tiryns. In The Oxford Handbook of the Bronze Age Aegean, edited by E.H. Cline, 722–734. Oxford: Oxford University Press. Maran, J., and M. Kostoula. 2014. The spider’s web: Innovation and society in the Early Helladic period of the corridor houses. In Critical Essays on the Archaeology of the Eastern Mediterranean in Honour of E. Susan Sherratt, edited by Y. Galanakis, T. Wilkinson, and J. Bennet, 141–158. Oxford: Archaeopress. Marshall, M. 2012. Climate change: The great civilization destroyer. New Scientist 215: 32–36. Matteucci, R., G. Gosso, S. Peppoloni, S. Piacente, and J. Wasowski. 2014. The Geoethical Promise: A proposal. Episodes 37(3): 190–191. Mellaart, J. 1958. The end of the Early Bronze Age in Anatolia and the Aegean. American Journal of Archaeology 62: 9–33. Meller, H., H.W. Arz, R. Jung, and R. Risch., eds. 2015. 2200 BC—A Climatic Breakdown as a Cause for the Collapse of the Old World? Halle (Saale): Landesamt fur Denkmalplflege und Archaologie SachsenAnhalt, Landesmuseum fur Vorgeschichte. Middleton, G.D. 2010. The Collapse of Palatial Society in Late Bronze Age Greece and the Postpalatial Period. Oxford: Archaeopress. Middleton, G.D. 2012. Nothing lasts forever: Environmental discourses on the causes of past societal collapse. Journal of Archaeological Research 20(3): 257–307. Middleton, G.D. 2015. Telling stories: The Mycenaean origins of the Philistines. Oxford Journal of Archaeology 34(1): 45–65. Middleton, G.D. 2017a. Understanding Collapse: Ancient History and Modern Myths. Cambridge: Cambridge University Press. Middleton, G.D. 2017b. Reading the thirteenth century BC in Greece: Crisis, decline, or business as usual? In Crisis to Collapse: The Archaeology of Social Breakdown. Aegis 11, edited by T. Cunningham and J. Driessen, 87–97. Louvain: Louvain University Press. Middleton, G.D. 2017c. The show must go on: Collapse, resilience, and transformation in twenty-first century archaeology. Reviews in Anthropology 46(2–3): 78–105. Middleton, G.D. 2018a. Collapse of the Bronze Age Aegean. Oxford Classical Dictionary (online). Oxford: Oxford University Press. Middleton, G.D. 2018b. Should I stay or should I go? Mycenaeans, migration, and mobility in the Late Bronze Age and Early Iron Age eastern Mediterranean. Journal of Greek Archaeology, 3: 115–143. Mitchell, P. 2008. Practising archaeology at a time of climatic catastrophe. Antiquity 82: 1093–1103. Moeller, N. 2005. The First Intermediate Period: A time of famine and climate change? Agypten und Levante 15: 153–167. Moreno Garcia, J.C. 2015. Climatic change or sociopolitical transformation? Reassessing late 3rd millennium BC in Egypt. In 2200 BC—A Climatic Breakdown as a Cause for the Collapse of the Old World? edited by H. Meller, H.W. Arz, R. Jung, and R. Risch, 3–16. Halle (Saale): Landesamt fur Denkmalplflege und Archaologie Sachsen-Anhalt, Landesmuseum fur Vorgeschichte. Morris, E. 2006. Lo, nobles lament, the poor rejoice’: State formation in the wake of social flux. In After Collapse: The Regeneration of Complex Societies, edited by G.M. Schwartz and J.J. Nichols, 58–71. Tucson: The University of Arizona Press. Neumann, J. 1993. Climatic changes in Europe and the Near East in the second millennium. Climatic Change 23(3): 231–245. Oates, J. 2014. Mesopotamia: The historical periods. In The Cambridge World Prehistory, edited by C. Renfrew and P. Bahn, 1498–1507. Cambridge: Cambridge University Press.

Collapse of Bronze Age Civilizations

291

Patterson, T.C. 2005. The turn to agency: Neoliberalism, individuality, and subjectivity in late-twentiethcentury anglophone archaeology. Rethinking Marxism: A Journal of Economics, Culture and Society 17(3): 373–384. Pullen, D. 1994. A lead seal from Tsoungiza, ancient Nemea, and Early Bronze Age Aegean sealing systems. American Journal of Archaeology 98(1): 35–52. Pullen, D. 2008. The Early Bronze Age in Greece. In The Cambridge Companion to the Aegean Bronze Age, edited by C.W. Shelmerdine, 19–46. Cambridge: Cambridge University Press. Pullen, D. 2011. Redistribution in Aegean palatial societies. Before the palaces: Redistriibution and chiefdoms in mainland Greece. American Journal of Archaeology 115(2): 185–195. Rahmstorf, L. 2016. Emerging economic complexity in the Aegean and western Anatolia during earlier third millennium BC. In Of Odysseys and Oddities: Scales and Modes of Interaction Between Prehistoric Aegean Societies and Their Neighbours, edited by B. Molloy, 225–276. Oxford: Oxbow. Regev, J., P. Miroschedji, and E. Boaretto. 2012a. Early Bronze Age chronology: Radiocarbon dates and chronological models from Tel Yarmuth (Israel). Radiocarbon 54(3–4): 505–524 Regev, J., P. Miroschedji, R. Greenberg, Z. Greenhut, and E. Boaretto. 2012b. Chronology of the Early Bronze Age in the southern Levant: New analysis for a high chronology. Radiocarbon 54(3–4): 525–566. Riede, F., P., andersen, and N. Price. 2016. Does environmental archaeology need an ethical promise? World Archaeology 48(4): 466–481. Roberts, R.G. 2009. Identity, choice, and the Year 8 reliefs of Ramesses III at Medinet Habu. In Forces of Transformation: The End of the Bronze Age in the Mediterranean, edited by C. Bachhuber and R.G. Roberts, 60–68. Oxford: Oxbow. Rutter, J.B. 1993. Review of Aegean Prehistory II: The Prepalatial Bronze Age of the Southern and Central Greek Mainland. American Journal of Archaeology 97(4): 745–797. Sagona, A., and Zimansky, P. 2009. Ancient Turkey. Abingdon: Routledge. Sandweiss, D.H. and A.R. Kelley. 2012. Archaeological contributions to climate change research: The archaeological record as a paleoclimatic and paleoenvironmental archive. Annual Review of Anthropology 41: 371–391. Scheidel, W. 2017. The Great Leveler: Violence and the History of Inequality from the Stone Age to the Twenty-First Century. Princeton: Princeton University Press. Schwartz, G.M. 2006. From collapse to regeneration. In After Collapse: The Regeneration of Complex Societies, edited by G.M. Schwartz and J.J. Nichols, 3–17. Tucson: Arizona University Press. Seidlmayer, S. 2000. The First Intermediate Period (c. 2160–2055 BC). In The Oxford History of Ancient Egypt, edited by I. Shaw, 108–136. Oxford: Oxford University Press. Shaw, J.W. 2007. Sequencing the EH II ‘corridor houses. Annual of the British School at Athens 102: 137–151. Shelton, K. 2010. Mainland Greece. In The Oxford Handbook of the Bronze Age Aegean, edited by E.H. Cline, 139–148. Oxford: Oxford University Press. Silberman, N.A. 1998. The Sea Peoples, the Victorians, and us: Modern social ideology and changing archaeological interpretations of the Late Bronze Age collapse. In Mediterranean Peoples in Transition: Thirteenth to Early Tenth Centuries BCE, edited by S. Gitin, A. Mazar, and E. Stern, 268–275. Jerusalem: Israel Exploration Society. Steel, L. 2010. Cyprus. In The Oxford Handbook of the Bronze Age Aegean, edited by E.H. Cline, 804–819. Oxford: Oxford University Press. Steel, L. 2014. Cyprus during the Late Bronze Age. In The Oxford Handbook of the Archaeology of the Levant, c. 8000–323 BC, edited by M.L. Steiner and A.E. Killebrew, 577–591. Oxford: Oxford University Press. Storey, R., and G. Storey. 2017. Rome and the Classic Maya: Comparing the Slow Collapse of Civilizations. New York: Routledge. Tainter, J. 1988. The Collapse of Complex Societies. Cambridge: Cambridge University Press. Tomkins, P., and I. Schoep. 2010. Crete. In The Oxford Handbook of the Bronze Age Aegean, edited by E.H. Cline, 66–82. Oxford: Oxford University Press. Turchin, P. (2010). ‘Political instability may be a contributor in the coming decade.’ Nature 463: 608. Van De Mieroop, M. 2007. A History of the Ancient Near East, ca. 3000–323 BC. Oxford: Blackwell. Van de Noort, R. 2011. Conceptualising climate change archaeology. Antiquity 85(329): 1039–1048. Verheyden, S., F.H. Nader, H.J. Cheng, L.R. Edwards, and R. Swennen. 2008. Palaeoclimatic reconstruction in the Levant region from the geochemistry of a Holocene stalagmite from the Jeita Cave, Lebanon. Quaternary Research 70(3): 368–381. Voskos, I., and A.B. Knapp. 2008. Cyprus at the end of the Late Bronze Age: Crisis and colonization or continuity and hybridization. American Journal of Archaeology 112: 659–684. Voutsaki, S. 2000. The Dorian invasion. Classical Review 50(1): 232–233.

292

Climate Changes in the Holocene

Voutsaki, S. 2010. Argolid. In The Oxford Handbook of the Aegean Bronze Age, edited by E.H. Cline, 599– 613. Oxford: Oxford University Press. Weiberg, E. 2010. Pictures and people: Seals, figurines and Peloponnesian imagery. Opuscula 3: 185–218. Weiberg, E. 2017a. Contrasting histories in Early Bronze Age Aegean: Uniformity, regionalism and the resilience of societies in the northeast Peloponnese and Central Crete. Cambridge Archaeological Journal doi:10.1017/S095977431700018X Weiberg, E. 2017b. Early Helladic III: A non-monumental but revitalized social arena?’ In Social Change in Aegean Prehistory, edited by C. Wiersma and S. Voutsaki, 32–48. Oxford: Oxbow. Weiberg, E., and M. Finné. 2013. Mind or matter? People-environment interactions and the demise of Early Helladic II society in the northeastern Peloponnese. American Journal of Archaeology 117(1): 1–31. Weiberg, E., and Lindblom, M. 2014. The Early Helladic II-III transition at Lerna and Tiryns revisited: Chronological difference or synchronous variability. Hesperia 83(3): 383–407. Weiberg, E., M. Lindblom, B. Sjöberg, and G. Nordquist. 2010. Social and environmental dynamics in Bronze and Iron Age Greece. In The Urban Mind: Cultural and Environmental Dynamics, edited by P.J. Sinclair, G. Nordquist, F. Herschend, and C. Isendalh, 149–194. Uppsala: African and Comparative Archaeology, Department of Archaeology and Ancient History, Uppsala University. Weingarten, J., J.H. Crouwel, M. Prent, and G. Vogelsang-Eastwood. 1999. Early Helladic sealings from Geraki in Lakonia, Greece. Oxford Journal of Archaeology 18(4): 357–376. Weingarten, J., S. Macveagh Thorne, M. Prent, and J.H. Crouwel. 2011. More Early Helladic sealings from Geraki in Laconia, Greece. Oxford Journal of Archaeology 30(2): 131–163. Weiss, B. 1982. The decline of Late Bronze Age civilization as a possible response to climate change. Climatic Change 4(2): 173–198. Weiss, H. 2014. The northern Levant during the intermediate Bronze Age: Altered trajectories. In The Oxford Handbook of the Archaeology of the Levant, c. 8000–332 BCE, edited by M.L. Steiner and A.E. Killebrew, 367–387. Oxford: Oxford University Press. Weiss, H. 2015. Megadrought, collapse, and resilience in late 3rd millennium BC Mesopotamia. In 2200 BC – A Climatic Breakdown as a Cause for the Collapse of the Old World?, edited by H. Meller, H.W. Arz, R. Jung, and R. Risch, 35–52. Halle (Saale): Landesamt fur Denkmalplflege und Archaologie SachsenAnhalt, Landesmuseum fur Vorgeschichte. Weiss, H. 2016. Global megadrought, societal collapse and resilience at 4.2–3.9 ka BP across the Mediterranean and west Asia. PAGES Magazine 24(2): 62–63. Weiss, H., ed. 2017a. Megadrought and Collapse: From Early Agriculture to Angkor. Oxford: Oxford University Press. Weiss, H. 2017b. 4.2  Ka BP megadrought and the Akkadian collapse. In Megadrought and Collapse: From Early Agriculture to Angkor, edited by H. Weiss, 93–159. Oxford: Oxford University Press. Weiss, H., and R.S. Bradley. 2001. What drives societal collapse? Science 291: 609–610. Westenholz, A. 1999. The Old Akkadian Period: History and culture. In Mesopotamien. Akkade-Zeit und Ur III-Zeit. Orbis biblicus et Orientalis 160/3, edited by W. Sallaberger and A. Westenholz, 17–117. Freiburg: Universitatsverlag Freiburg Schweiz. Wiencke, M.H. 2010. Lerna. In The Oxford Handbook of the Aegean Bronze Age, edited by E.H. Cline, 660–670. Oxford: Oxford University Press. Wiener, M.H. 2013. Minding the gap: Gaps, destructions, and migrations in the Early Bronze Age Aegean. Causes and consequences. American Journal of Archaeology 117(4): 581–592. Wiener, M.H. 2014. The interaction of climate change and agency in the collapse of civilizations ca. 2300– 2000 BC. Radiocarbon 56(4): S1–S16. Wilkinson, T. 2010. The Rise and Fall of Ancient Egypt. London: Bloomsbury. Williams, R.P. 2002. Rethinking disaster- induced collapse in the demise of the Andean highland states: Wari and Tiwanaku. World Archaeology 33: 361–374. Wilson, J.A. 1956. The royal myth in ancient Egypt. Proceedings of the American Philosophical Society 100: 439–42. Wright, J.C., ed. 2004. The Mycenaean Feast. Princeton: American School of Classical Studies at Athens. Zettler, R.L. 2003. Reconstructing the world of ancient Mesopotamia: Divided beginnings and holistic history. Journal of Economic and Social History of the Orient 46: 3–45.

14

The Iranian Plateau and the Indus River Basin Cameron A. Petrie and Lloyd Weeks

CONTENTS 14.1 Introduction........................................................................................................................... 293 14.2 Spatial and Temporal Climatic Context throughout the Holocene........................................ 294 14.2.1 Modern Precipitation................................................................................................. 294 14.2.2 Paleoclimate Proxy Records...................................................................................... 297 14.2.3 Long-term Trends...................................................................................................... 298 14.2.4 Abrupt Events............................................................................................................ 299 14.3 Archaeological Evidence for Human Response to Climate..................................................300 14.3.1 Archaeological Data: Nature, Scale, Limitations and Potential................................300 14.3.2 Linking Climate Change and Human Responses: Theoretical and Practical Challenges���������������������������������������������������������������������������������������������������������������� 301 14.3.3 The Neolithic Transition of the Early Holocene........................................................302 14.3.4 Late Chalcolithic.......................................................................................................306 14.3.5 Bronze Age................................................................................................................308 14.4 Conclusions............................................................................................................................ 311 Websites.......................................................................................................................................... 312 Acknowledgment............................................................................................................................ 312 References....................................................................................................................................... 312

14.1 INTRODUCTION This chapter presents an overview of the current evidence for the Holocene climate, climate change, human adaptation and human response on the Iranian Plateau and the adjacent piedmont regions, and the extensive alluvial plains of the Indus and associated rivers in western South Asia. The topography of this extensive region is extremely varied, and today there is significant diversity in local environment and climate, influenced by the complex weather systems that operate in different regions and at varying levels of intensity in both winter and summer. It is assumed that these systems also operated during the Holocene, although available climate proxies suggest that there was variation in spatial position and strength, indicating that the climate across the region was subject to both long-term change and abrupt events. The available archaeological evidence highlights the many different ways that human populations responded to these environmental dynamics. The nature of Epi-Paleolithic occupation across the Iranian Plateau and the Indus River Basin and associated areas provides insight into the ways that early hunter foragers were able to inhabit and exploit this landscape. From the Neolithic onwards, populations moved across this landscape and adapted their subsistence practices—including traditional hunting and gathering, as well the exploitation of newly domesticated plants and animals—to enable exploitation of a range of ecological niches and the construction of new ones. By the Chalcolithic period and Bronze Age, humans occupied most of the habitable regions across this extensive zone, and communities displayed an elaborate and diverse set of adaptations that built upon and extended the subsistence production of their Neolithic precursors. The Iranian Plateau and the Indus River Basin and associated areas thus have the potential to provide fundamental insights into human adaptation and response to climate throughout the Holocene. 293

294

Climate Changes in the Holocene

14.2 SPATIAL AND TEMPORAL CLIMATIC CONTEXT THROUGHOUT THE HOLOCENE The mountains and piedmonts of the Iranian Plateau combine to form one of the dominant geographical features of Western Asia, comprising major mountain ranges, alluvial plains and fans, intermontane valleys, and desert areas (Fisher 1968) that form the key geographical entity linking Western Asia to the South Asian subcontinent and Central Asia (Petrie 2013b, 4). To the east, the Indus River Basin in South Asia comprises an imposing series of alluvial plains and associated piedmont areas which provide a distinctive example of environmental and climatic diversity and variability, in part due to the combination of topographic and climatic factors that affect the region’s precipitation. These areas are the focus of the present chapter, although surrounding regions—the alluvial plains of Mesopotamia, the inland deltas and fans of Bactria and Margiana, and the coasts and deserts of Arabia—are mentioned where they provide key paleoclimatic or archaeological evidence.

14.2.1 Modern Precipitation It is not yet possible to provide a comprehensive reconstruction of changes in the intensity and timing of precipitation across the Iranian Plateau and the Indus River Basin during the Holocene, but the extrapolation of past rainfall systems from modern weather patterns gives us indications of how we might interpret the limited number of paleoclimate proxy records that are available. Today, the Iranian Plateau receives rainfall in the winter and spring which derives from a Mediterranean moisture source in the west, with additional moisture coming from the Caspian Sea, and the whole is driven by extratropical cyclonic disturbances (Domroes et al. 1998; Pourasghar et al. 2012; Balling et al. 2016, Figure 3; Petrie et al. 2018; Jones et al. in press). Precipitation falls primarily along the Zagros ranges to the west and the Elburz to the north, with the central and eastern regions of the plateau receiving less than 100 mm of rainfall per annum (Figure 14.1; Ganji 1968, Figure 79; Domroes et  al. 1998, Figure 1; Ghasemi and Khalili 2008, Figure 1b; Pourasghar et al. 2012, Figure 1b). Alijani et al. (2008) have shown that precipitation over modern Iran tends to be irregular and intense, with a large proportion of the annual rainfall coming from a small number of high intensity to extreme rainfall events (see also Modarres and Sarhadi 2009). The nature of this winter and spring climate pattern, and its relationship to cold air intrusions from the north, results in the regular formation of an eastward moving extratropical depression that is ultimately inhibited by the Hindu Kush and the Himalayas (Dimri and Chevuturi 2016, 6–7). Known as Western Disturbances, these weather systems also appear to benefit from “secondaries”, which develop due to evaporation in the Persian Gulf and the Arabian Sea (Dimri and Chevuturi 2016, 7), and result in moderate to heavy precipitation across the northern part of the area that broadly corresponds to the Indus River Basin, which comes in the form of snow on the mountains and rain in low altitude areas (Dimri and Chevuturi 2016, 6–7). The Western Disturbances have a relatively steep winter rainfall gradient from roughly north to south, from the uplands onto the plains (Figure 14.2 [left]) Prasad and Enzel 2006; Petrie et al. 2017). The westernmost parts of South Asia are distinct from the regions further west because they also receive rain from the Indian Summer Monsoon (ISM) (Petrie et al. 2017; see also Cronin 2010, 223–8), which has a similarly steep rainfall gradient from east to west (Figure 14.2 [right]) (Prasad and Enzel 2006; Dixit et al. 2014a; Petrie et al. 2017). As for the Iranian Plateau, there is considerable inter-annual variability in the timing, intensity and distribution of rainfall across the Indus River Basin and its adjacent areas, with the added complexity of two wet seasons. It is notable that there have been many instances of extreme rainfall in this region (e.g. affecting different parts of Pakistan in 2009, 2010, 2012, and 2014 and India in 2007, 2013, 2014, and 2015; see also Possehl 1999, 286–287; Petrie 2017, 49), and there are also regular instances of drought (e.g. 1998–2002, 2004–2005, and 2009–2010) that appear to correlate to El Niño events (Kumar et al. 2006). The variability of these winter and summer climate patterns

FIGURE 14.1  Map of the region discussed in this chapter, showing the spatial distribution of the modern mean annual rainfall (dot-dashed isohyets), the location of published paleoclimate proxy records (white circles) and a selection of archaeological sites discussed in the text (black triangles, numbered as follows: 1. Arisman; 2. Bampur; 3. Chogha Mish; 4. Dholavira; 5. Ganweriwala; 6. Godin Tepe; 7. Habuba Kabira; 8. Harappa; 9. Kalibangan; 10. Keshit; 11. Konar Sandal; 12. Lothal; 13. Mehrgarh; 14. Miri Qalat; 15. Mohenjo daro; 16. Mundigak; 17. Nippur; 18. Rakhigarhi; 19. Shahdad; 20. Shahr-i Sokhta; 21. Susa; 22. Tal-e Malyan; 23. Tal-i Iblis; 24. Tappeh Sialk; 25. Tell Brak; 26. Tell Hamoukar; 27. Tepe Hissar; 28. Tepe Yahya; 29. Teppe Gawra; 30. Tol-e Nurabad; 31. Tol-e Spid; 32. Tureng Tepe; 33. Uruk).

The Iranian Plateau and the Indus River Basin 295

FIGURE 14.2  Maps of the Indus River Basin showing the distribution of modern winter (left) and summer (right) across the area occupied by the Indus Civilisation, in relation to the distribution of urban period settlements (orange circles), and urban centres (black circles).

296 Climate Changes in the Holocene

The Iranian Plateau and the Indus River Basin

297

produces considerable climatic and ecological diversity across the South Asian subcontinent as a whole, particularly where the winter and summer weather systems partly overlap in the Indus River Basin and adjacent areas (Petrie et al. 2017; see also Weber et al. 2010).

14.2.2 Paleoclimate Proxy Records Globally, a substantial number of paleoclimate proxy datasets exist to provide insight into Holocene climatic variability (e.g. https://www.ncdc.noaa.gov/data-access/paleoclimatology-data). There is, however, only a limited number of proxy datasets from the Iranian Plateau and the Indus River Basin (Figure 14.1); they show a degree of heterogeneity in precipitation patterns, but it also makes it difficult to reconstruct the chronological and spatial variability of rainfall across these regions during the Holocene. Paleoclimate proxy records from the Zagros were some of the earliest for western Asia, but they are few in number and show considerable variability. There are only four lake records from the Zagros in western Iran that make it possible to investigate long-term climate change (Figure 14.1 Zeribar, Mirabad, Parishan and Neor; Stevens et al. 2001, 2006; Jones et al. 2015; Sharifi et al. 2015; Petrie et al. 2018; Jones et al. in press). The analysis of these proxy records has not typically been undertaken to provide direct insight into archaeologically relevant questions, though there have been instances where paleoclimatic reconstruction has been undertaken in direct collaboration with archaeologists (e.g. Jones et al. 2015). Other lakes have produced pollen records that reveal evidence for changing land use (e.g. Almalou, Maharlou; Djamali et al. 2009a,b; Petrie et al. 2018), but do not provide a direct analogue for climate, as they also reflect the impact of human processes such as land management and deforestation. Speleothems have also been collected from the Zagros and analysed, but are currently unpublished (Carolin et al. 2016). Further to the east, there is a dearth of paleoclimate records from the central, eastern and southern parts of the Iranian Plateau (eastern Iran, Afghanistan, western Pakistan) and most of the alluvial plains of the Indus River and its tributaries (Possehl 1999, Figure 3.112). Nonetheless, research has now been undertaken on Lake Hamoun in the Sistan Basin in southeast Iran (Hamzeh et al. 2016a,b) and is supplemented by geomorphological studies of the region (Walker and Fattahi 2011) and a peat core recently collected adjacent to Konar Sandal in the Jiroft region (Gurjazkaite et al. 2018). Paleoclimate records from sites in southern Arabia (e.g. Fleitmann et al. 2003, 2007, 2008; Parker et al. 2006, 2016; Lézine et al. 2010, 2017; Preston et al. 2015) also contribute to understanding climate change in proximate areas of southern Iran, especially in relation to long-term trends such as the southwards regression of the Indian Ocean Monsoon over the course of the mid-Holocene. The potential importance of climate for the understanding of South Asian archaeology has long been recognized (e.g. Marshall 1931). While there has been a tradition of paleoclimate research there (Prasad and Enzel 2006; Madella and Fuller 2006; Wright et al. 2008), spatial coverage has been limited, chronological resolution has been low, and there has been limited integration with archaeological datasets (Madella and Fuller 2006; Petrie et al. 2017). Until 2012, the majority of paleoclimate records from South Asia had been obtained from the dry playas in the Thar Desert (Prasad and Enzel 2006), and archaeologists typically looked to high-resolution data from elsewhere (e.g. Qunf, Hoti in Oman; Fleitmann et al. 2003, 2007, 2008), despite these data being neither geographically proximate nor readily applicable to South Asian conditions. New high-resolution paleoclimate records from South Asia do, however, continue to appear, and attempts have been made to model climate variability (Wright et al. 2008), target paleoclimate records that have specific proximity to archaeological sites that are believed to have been affected by climate variability and change (e.g. Dixit et al. 2014a,b, 2018), and to obtain climate records directly from archaeological sites (Jones 2017). Current understanding of Holocene climate in the Indus River Basin comes from combining the data from discrete proxy records that typically lie towards the extremities of the catchment (Figure 14.1). These proxy records include river delta cores from the mouth of the Indus (63KA; Staubwasser et al. 2002, 2003), paleolakes in various locations along the edge of the Thar desert in

298

Climate Changes in the Holocene

northwest India (Nal Sarovar, Riwasa, Kotla Dahar, Karsandi; Prasad et al. 1997; Prasad and Enzel 2006; Dixit et al. 2014a,b, 2018), and speleothems from the Himalayas (Dharamjali, Tityana; Kotlia et al. 2017; Joshi et al. 2017). Several playa lakes in Rajasthan have produced pollen sequences (e.g. Didwana, Lunkaransar, Bap Malar, Sambhar; Wasson et al. 1984; Singh et al. 1990; Enzel et al. 1999; Deotare et al. 2004; Sinha et al. 2006; see Jones 2017), but only a limited range of methods have been used on these sequences, and problems with dating resolution mean that change within each is not well constrained chronologically (Madella and Fuller 2006; Prasad and Enzel 2006). There are also records from high-altitude lakes in the Himalayas (e.g. Rara, Tso-Moriri; Nakamura et al. 2016; Leipe et al. 2014; Mishra et al. 2015; Dutt et al. 2018), but as with the more distant records from Oman, it is not possible to apply the results from these records directly to an understanding of the Holocene climate of the South Asian lowlands. Taken together, the available proxies from the Iranian Plateau and the Indus River Basin provide a range of evidence for long-term trends and climatic variability. They also provide some insight into the local manifestations of the abrupt climatic events at ~8.2, 5.2, 4.2 and also 3.2 kya BP, which are roughly coincident with North Atlantic Bond events (Bond et al. 1997), though the nature and even reality of Bond events is debated (e.g. Wanner et al. 2011).

14.2.3 Long-term Trends The Zeribar and Mirabad cores show that the Zagros was a dry-steppe during the terminal Pleistocene and remained treeless during the Bölling/Allerød interstadial (see Wasylikowa and Witkowski 2008; also Hole 1996, 268, 270; Matthews 2013). Following the aridity of the late Glacial period, Zeribar shows evidence for increases in temperature and precipitation between ~10.5 and 6.5 ka BP (early Holocene), though this period was still drier than the later Holocene (Stevens et al. 2001, 2006). It is notable that oak appears at ~10 ka BP at Mirabad, which is 2000 years earlier than it appears at Zeribar, suggesting that there was more severe early Holocene aridity in the northern Zagros (Wright 1993, 464–465). The mid-Holocene (~6.5 to 4.5 ka BP) at Zeribar was a period of greater overall moisture and there is evidence for the proliferation of oak into the pistachio-oak savannah (Stevens et al. 2001), though there is some evidence for regional drying at ~5.4 ka BP (Stevens et al. 2006). The late Holocene (~4.5 to 1 ka BP) at Zeribar saw a decrease in spring rainfall, apparently to current levels (Stevens et al. 2001, 2006). Zeribar, Mirabad, and Parishan each show lower isotope values (and thus drying) between ~4.0 and 3.0 ka BP (i.e. during the second millennium BC), with extreme values appearing later further south, which may reflect several factors, including rainfall seasonality (Stevens et al. 2001), and human activity in catchment areas (Jones et al. 2015; see Petrie et al. 2018). The cores from Lake Hamoun in south-eastern Iran, although poorly constrained in dating terms, suggest a somewhat different pattern that reflects its transitional position between the Iranian Plateau and South Asia (Hamzeh et al. 2016a,b). Here, the early Holocene was relatively wet and Lake Hamoun was deep and stable, under the influence of rainfall from both the ISM and Mid-Latitude Westerlies (or MLW). The weakening of the ISM and the long-term southward migration of the Inter-Tropical Convergence Zone during the mid-Holocene created a climate in Sistan that was more arid and was affected by strong winds, with seasonal patterns of aeolian and fluvial deposition. Subsequently, rainfall in the mid-late Holocene appears to have been entirely dependent on MLW-associated winter precipitation, in a pattern similar to that observed in the present. Lake Hamoun was lower and less productive at this time, with severe fluctuations in lake levels and evidence for significant wind storms indicated by intense episodes of aeolian deposition (Hamzeh et al. 2016a,b). Paleoclimate proxy records and simulations suggest that changes in the position and intensity of the ISM reflect long-term and gradual change in the ITCZ (cf. Revel et al. 2014). The complex interrelations between rains from the ISM and MLW systems and more localized climate systems (e.g. Parker et al. 2006, 474), appear to have been a key long-term factor in water availability in the southern Iranian Plateau, the Indus River Basin, and adjacent regions such as Arabia (Lézine et al. 2010, 2017).

The Iranian Plateau and the Indus River Basin

299

Further to the east, the 63KA core from the mouth of the Indus River indicates that ~10.8 ka BP marked the end of a period of monsoon intensification (Staubwasser et al. 2002). Maximum discharge was reached after ~9.4 ka BP, but this ended abruptly ~8.4 ka BP, which Staubwasser et al. (2002) suggest was coeval with the ~8.2 ka BP event (see below). The 63KA core also has evidence for an increase in δ18O values at ~4.2 ka BP, which reflects a further reduction of river discharge (Staubwasser et al. 2003). New analysis of samples from the same core has shown that a weakening of the ISM at ~4.2 ka BP was preceded by a weakening of the Indian Winter Monsoon (IWM) at ~4.3 ka BP after a period of peak strength (Giesche et al. in press). On the eastern side of the Indus Basin at Lunkaransar and Didwana in Rajasthan, increased lake levels between ~7.2 and ~5.3 ka BP have been attributed to increased winter and strengthened summer monsoon rainfall (Prasad and Enzel 2006), which also matches the wettest period at Nal Sarovar in Gujarat (Prasad et al. 1997; Prasad and Enzel 2006). However, from ~6.0–5.9 ka BP, drier conditions begin in each of these records, attributed to weakening winter and summer rain (Prasad and Enzel 2006). Other lakes in the Thar dried out from ~5.5 ka BP (e.g. Bap Malar and Kanod; Deotare et al. 2004; Dixit et al. 2018), while Lunkaransar underwent drying from ~5.3 ka BP (Enzel et al. 1999), and Didwana from ~4.4 ka BP (Prasad and Enzel 2006). These records thus suggest a west to east pattern of drying following the modern precipitation gradient (Roy and Singhvi 2016; Dixit et al. 2018). The Riwasa lake record (southern Haryana) from the northern edge of the Thar Desert has evidence of a deep lake between ~10.4 and 8.3 ka BP, though ostracods and gastropods decrease after ~9.4 ka BP (Dixit et al. 2014a). There is, however, evidence for an abrupt drying of the lake at ~8.2 ka BP, which produced a cemented hardground and suggests a dramatically weakened monsoon at that point, with desiccation continuing until ~7.9 ka BP (Dixit et al. 2014a). Increased precipitation and lower evaporation is attested to at Riwasa after ~7.5 ka BP, which has been attributed to increased winter precipitation (Dixit et al. 2014a; also Singh et al. 1990). The end of the sequence has not been dated but may well be close in time to the ~5.3 ka BP drying at Lunkaransar. The Kotla Dahar lake record (eastern Haryana) from the eastern edge of the Thar has evidence of a deep lake up to ~5.8 ka BP, when there was a transition to a shoaling phase, which continued until ~4.2–4.1 ka BP at which point δ18O values increased abruptly, coinciding with a drop in CaCO3 and the disappearance of ostracods (Dixit et al. 2014b). Taken together, this evidence suggests a sharp reduction in monsoon intensity at a time roughly coincident with the evidence from 63KA, and also that from the Mawmluh speleothem in north-east India (Dixit et al. 2014b; also Staubwasser et al. 2003; Berkelhammer et al. 2012). The Karsandi lake record, which comes from an area of northern Rajasthan that lies in the Thar Desert margins, provides evidence for monsoon intensification between ~5.0 and 4.4 ka BP after which drier conditions set in and massive gypsum deposits formed, and the lake dried permanently by ~3.2 ka BP (Dixit et al. 2018). The results from Karsandi suggest that although the monsoon and winter precipitation may have reduced in much of Rajasthan by ~5.3 ka BP, there was a period of relatively increased precipitation between ~5.0 and 4.4 ka BP that affected the desert margins to the north, before monsoon weakening from ~4.4 and 4.2–4.1 ka BP. It is notable that the lakes/playas in these margins are considerably closer to the Himalayas and the impact of both Western Disturbances and the north western curve of the Indian Summer Monsoon, so the difference from the evidence from the Thar records is not surprising (Figures 14.1 and 14.2).

14.2.4 Abrupt Events Overlaid on these long-term trends, a number of major abrupt climate events can be documented for the Holocene, of which the ~8.2, 5.2, 4.2 and 3.2 ka BP events appear to be most significant on the Iranian Plateau and the Indus River Basin and surrounding regions. The ~8.2 ka BP event appears to have been related to deglaciation into the north Atlantic (Staubwasser and Weiss 2006; Cronin 2010, 267). It is expressed as an isotopic shift to more positive values in Hoti and Qunf caves in Oman (Fleitmann et al. 2007) and is also recorded at Riwasa (Dixit et al. 2014a). At ~5.2 ka BP, Europe was affected by an abrupt cold and wet period, called the Piora Oscillation, which spanned from ~5.2 to 4.9 ka BP (Lamb 1982 [2005]). Interestingly, this was a dry period seen in lakes and

300

Climate Changes in the Holocene

wetlands in southeast Arabia (Parker et al. 2006) and Turkey (Kuzucuoğlu et al. 2011). The ~5.2 ka BP date also coincides with the decreased rainfall in the central Rajasthani lakes (e.g. Lunkaransar; Enzel et al. 1999), and the shoaling phases at Kotla Dahar (Dixit et al. 2014b), though it appears to have been followed by an increase in precipitation at Karsandi (Dixit et al. 2018). Another period of drought at ~4.2 ka BP is recorded in lakes in Turkey (Eastwood et al. 2007; Dean et al. 2015), southeast Arabia (Parker et al. 2006), the Dead Sea (Litt et al. 2012) and the Gulf of Oman (Cullen et al. 2000), and low δ18O values occur at this time at Zeribar and Mirabad (Stevens et al. 2006). It has also been identified at Kotla Dahar in South Asia where it is seen to mark a weakening of the monsoon (Dixit et al. 2014b), and is likely to have contributed to the post ~4.4 ka BP drying at Karsandi (Dixit et al. 2018). The Kotla Dahar record suggests that the ~4.2 ka BP event might have lasted for up to 200 years on the plains of northwest India (Dixit et al. 2014b). Inspection of the δ18O results from the Dharamjali speleothem (Kotlia et al. 2017, Figure 5), however, suggests that there may have been two periods of weak precipitation between ~4.1 and 3.9 ka BP, interrupted by a period of increased precipitation, highlighting that speleothems have the potential to capture fine-grained nuances that are missed in lake records, although this is not a given. Overall, it has been argued that the ~4.2 ka BP event would have resulted in less late summer water run-off through the Indus hydrological system, reducing the water that was available for the start of the winter growth seasons in areas that did not receive direct summer rain (e.g. Berkelhammer et al. 2012; Dixit et al. 2014b; Miller 2006, 2015; Petrie 2017). There was also a dry period from ~3.2 ka BP recorded at Lake Zeribar (Stevens et al. 2001), Lake Parishan (Jones et al. 2015) and in a peat core near Konar Sandal in southeastern Iran (Gurjazkaite et al. 2018). High-resolution analysis suggests that this drier period comprised several drought episodes interspersed within decades of wetter climate (Kuzucuoğlu, 2009). It is notable that this event is not recorded in the lake records from the Thar or its margins, although it is evident in the Dharamjali speleothem (Kotlia et al. 2017, 6, Figure 6). The Thar lakes may well have been completely ephemeral by this point, and potentially affected by human activity. Although significant, the ~3.2 ka BP event will not be discussed further here.

14.3 ARCHAEOLOGICAL EVIDENCE FOR HUMAN RESPONSE TO CLIMATE 14.3.1 Archaeological Data: Nature, Scale, Limitations and Potential Archaeological sites add thousands of data points to maps of ancient climate/environment proxies from lake or sea cores and from speleothem studies (Sandweiss and Kelley 2012), and have the potential to provide information—albeit attenuated by anthropogenic selection—at higher chronological and spatial resolution than is typical for paleoclimate datasets (d’Alpoim Guedes et al. 2016). An expanding range of techniques, including both remote sensing and ground-based approaches, facilitate high-resolution landscape-level studies of ancient settlement systems (e.g. Wilkinson 2003; Orengo and Petrie 2017, 2018; Green and Petrie 2018). These allow the reconstruction of site types, numbers, location, size, hierarchy, etc., including the identification of water-management systems (e.g. irrigation canals, dams, barrages, etc.) and components of agricultural systems (e.g. field systems, animal enclosures), as well as other elements of production that supported human occupation. At the site level, studies of plant and animal remains from excavations support reconstructions of species representation, site environment, subsistence strategies and the organization of productive activities. Further scientific analyses of these bioarchaeological assemblages—e.g. with stable and radiogenic isotopes (C, O, N, Sr)—can support the reconstruction of aspects as diverse as diet, foddering, seasonal mobility, and water stress. Similarly, a wide variety of scientific analyses of human remains provide insights on demography, health/disease and inter-personal violence, as well as information on diet, mobility, and genetic affiliations, all of which is relevant to an understanding of the relationships between humans and their environment. The archaeological datasets from the Iranian Plateau and the Indus River Basin show considerable variations in spatial and chronological coverage and resolution, and the nature of the archaeological

The Iranian Plateau and the Indus River Basin

301

evidence. There has been a long tradition of archaeological survey that provides insight into changing settlement systems, fluctuations in population size and distribution, and the development of urbanism on parts of the Iranian Plateau and its associated piedmonts (e.g. Sumner 1990; Johnson 1973; Alden 2013; Prickett 1986; Hole 1987; Moghaddam and Miri 2003, 2007), the regions around the Persian Gulf (Cleuziuo and Tosi 2007; Højlund 2007; Magee 2014; Laursen 2017), and the plains and piedmonts of the Indus River Basin (Figure 14.2; Joshi et al. 1984; Mughal 1997; Possehl 1999; Kumar 2009; Green and Petrie 2018). While there is an ever-increasing number of excavated sites and surveys in different parts of the region (e.g. Azarnoush and Helwing 2005; Chakrabarti 1999, 2007), these archaeological surveys have typically had varying extent, intensity and methods, and there remain considerable gaps in our knowledge of specific regions and individual periods (e.g. Singh et al. 2008; Petrie et al. 2017; Green et al. 2018). There have also been numerous excavations of significant archaeological sites in each of these areas (e.g. from west to east: Susa, Tal-e Malyan, Tepe Yahya, Konal Sandal South, Shahr-i Sokhta [Shahr-e Sukhteh], Mundigak, Mehrgarh, Mohenjo-daro, Harappa; Voigt and Dyson 1992; Shaffer 1992; see Figure 14.1). However, the number of excavated archaeological sites occupied in any one period is generally very limited, and our information remains patchy and is almost certainly unrepresentative (e.g. Singh et al. 2008; Petrie 2013a, 404–5; Petrie et al. 2017). Moreover, there has been relatively limited use of new technologies to investigate the remains from archaeological sites throughout much of the Iranian Plateau and the Indus River Basin. Although archaeobotany and archaeozoology have been applied widely, new excavations are providing opportunities to implement the full spectrum of archaeological approaches. For example, isotopic analysis is being increasingly utilized to investigate questions of mobility, diet, and the impact of climate change on water availability and use (e.g. Gregorika 2013; Kenoyer et al. 2013; Chase et al. 2014, 2018; Valentine et al. 2015; Jones 2017). These approaches are being complemented by methods such as residue analysis of ceramics—including material from new excavations and museum collections—and the analysis of dental calculus, which both provide insight into diversity and change in human diet (e.g. Warinner 2016; Hendy et al. 2018). High-throughput (“next-generation”) sequencing (e.g. Narasimhan et al. in press) is also revealing entire new vistas on population mobility and interaction.

14.3.2 Linking Climate Change and Human Responses: Theoretical and Practical Challenges Numerous papers have presented climate as a driver of cultural transformation at different points during the Holocene, including consideration of both long-term climate trends and abrupt events, where it has been common to link abrupt events to instances of cultural change observed in the archaeological record (e.g. Prasad and Enzel 2006; Staubwasser and Weiss 2006; Weiss 2015, 2016). For example, it has been proposed that the ~8.2 ka BP event contributed to the dispersal of Neolithic settlement into the Aegean and Europe, though it has also been suggested that such migration processes could well be caused by over-population and over-stress on local ecological habitats (Weninger et al. 2006, 2014). It has also been noted that the ~8.2 ka BP event had no significant impact upon local societies in those regions (Flohr et al. 2016), so its impact might have been localized to certain regions, including the Levant, north Syria, and southeast Anatolia, and also central Anatolia and Cyprus (Weninger et al. 2006; see below). The ~5.2 ka BP event has been linked to the end of the late Uruk period societies in Mesopotamia (Weiss 2003; Staubwasser and Weiss 2006; see below), though its potential impact upon local populations in many areas has not been fully investigated. The period of drying around ~4.2 ka BP has perhaps seen the most interest, being linked to the “collapse” of the Akkadian civilization in Mesopotamia, the possible disintegration of urban communities in the southern Levant (Weiss et al. 1993; Staubwasser and Weiss 2006; Weiss 2016; see below), as well as the dramatic decline in sedentary settlement at the end of the third millennium BC in southeast Iran and southeast Arabia (Fouache et al. 2015; Preston et al. 2015) and also the decline of Indus urbanism (Staubwasser and Weiss 2006; Kotlia et al. 2017; see below).

302

Climate Changes in the Holocene

It is increasingly recognized, however, that the relationships between humans and their environments were diverse and variable, and there is no simple correlation between climate change and human response (e.g. Aimers and Hodell 2011; Miller et al. 2011; Middleton 2017). This is particularly the case for the geographically and culturally diverse Iranian Plateau and Indus River Basin (Petrie et al. 2017; Jones et al. in press), where the mosaic nature of long-term cultural development and change provides insight into the fundamentally local and historically contingent nature of human responses to changing climate. At a practical level, the available climate records from the Iranian Plateau and the Indus River Basin are unevenly distributed and have variable proximity to archaeological regions of interest, which means that it is difficult to draw reliable correlations between evidence for climate change and cultural changes observed in the archaeological record (Wright et al. 2008; Wright 2010; Petrie 2017; Petrie et al. 2017; Jones 2017). The major issue for gaining an understanding of the relationship between humans and climate on the Iranian Plateau is the limited number and restricted distribution of paleoclimate records in the region, which lie in relatively close proximity in the west and northwest parts of an extensive geographical area, though there is now a record from Hamoun Lake (see above). While there are various paleoclimate records from the edges of the Indus River Basin, most of the early climate records from South Asia have poor chronological control (Madella and Fuller 2006), and while this has been improved by more recent studies (Berkelhammer et al. 2012; Dixit et al. 2014a,b, 2018), the spatial distribution of these records remains limited. At a more fundamental level, archaeological approaches to identifying climate impact on past societies are largely correlative: as a result, they suffer not only from low resolution and poor spatial association between archaeological sites and paleoclimate proxy records, but also from the fact that they are based on untested assumptions. If correlation does not imply causation, then understanding human responses to climate change requires accurately modelling the link between climate change, and factors such as resource variation, and human demographic and socio-economic change (d’Alpoim Guedes et al. 2016). We are now in a position where emerging approaches utilising computational modelling have the potential to move debates “well beyond merely noting coincidences of social and climatic change” (d’Alpoim Guedes et al. 2016, 14489). It is not possible within the confines of this contribution to fully explore the relationships between human cultural development and a variable and changing Holocene climate across an area as vast as the Iranian Plateau and the Indus River Basin. Thus, we will focus discussion on three key chronological periods—the Neolithic, the Late Chalcolithic, and the Early Bronze Age—and review evidence from various sub-regions of the western and eastern Iranian Plateau and the Indus River Basin. Consideration of the archaeology of these periods will enable us to discuss the impact of both long-term warming in the wake of the Late Glacial Maximum, and the effects of the abrupt ~8.2, 5.2 and 4.2 ka BP events.

14.3.3 The Neolithic Transition of the Early Holocene The transition to farming was one of the major innovations of human history and, amongst a range of other social developments, involved a significant transformation of the ways that human populations interacted with their environment (Cauvin 1994, 2000; Marchiniak and Czerniak 2007). Climate change during the late Pleistocene and early Holocene has been regarded as key to Neolithic origins since the time of R. Pumpelly and V.G. Childe (Harris 2010, 229; Weisdorf 2005), and is still seen as broadly necessary to the development of agricultural communities in southwest Asia (e.g. Bar-Yosef 2011; Richerson et al. 2001). These variations included not only long-term and abrupt changes in temperature and rainfall patterns, but also broad scale changes in climatic stability and atmospheric CO2 levels that, it is argued, limited opportunities for cereal domestication prior to the Holocene (e.g. Willcox 2013; Riehl et al. 2015; Cunniff et al. 2017). Nevertheless, it is clear that correlations between cultural and climate change are not always easily identified (e.g. Maher et al. 2011) and that environmental change alone is not sufficient to explain the transition to agricultural

The Iranian Plateau and the Indus River Basin

303

production in Asia. The archaeological record of the Iranian Plateau and South Asia is a testament to the complexity and variability of human adaptations to the changing early Holocene climate, and to the development of productive agricultural economies. Various parameters affected the ways that people carried out farming and sedentary life, with rainfall and water availability being both critically important and extremely variable across the Iranian Plateau and the Indus River Basin. Water availability in some areas is reasonably straightforward, particularly in areas that have access to karst springs. Some regions receive sufficient direct winter rainfall to support dry-farming, while various piedmont areas receive broadly predictable water supplies via run-off onto alluvial fans that are likely to have been reliable and largely predictable sources of water for early farmers (Petrie and Thomas 2012). The presence of occupation in upland and highland areas of the Zagros during both the Upper and Epi-Paleolithic suggests that the region was habitable during the Last Glacial Maximum, and occupation continued into the early Holocene as attested by Epi-Paleolithic or “proto-Neolithic” sites in the highlands of the Elburz and central and southern Zagros (e.g. Belt, Ghar-i Khar, Gar Arjeneh, Hotu, Izeh sites, Marv Dasht sites, Palegawra, Pa Sangar, Shanidar, Zarzi; Coon 1951, 1952; Smith 1986; Rosenberg 2003; Conard et al. 2006; Tsuneki et al. 2007; Hole 2008; Tsuneki and Zeidi (eds) 2008). To the south, the progressive infilling of the Persian Gulf from the end of the LGM (Lambeck 1996) submerged a large area of former river valley that may have been a key locus of habitation for late Pleistocene and early Holocene human groups (Rose 2010; Rose et al. 2013; Cuttler 2013), and any related archaeological sites of the Epi-Paleolithic and early Neolithic periods are now deep under the waters of the Persian Gulf. For South Asia, this period is relatively under-studied in comparison to earlier and later occupation. Nevertheless, surveying has identified numerous terminal Pleistocene and early Holocene open-air microlithic sites in Sindh, Rajasthan and Gujarat (Allchin et al. 1978; Chakrabarti 1999; Blinkhorn et al. 2017; Biagi and Starnini 2018). It has been suggested that these hunter–forager groups proliferated in the favorable conditions of the early Holocene (e.g. Allchin et al. 1978), and there appears to be a chronological hiatus between these occupations and the subsequent settlement by sedentary agriculturalists. The earliest, and some of the best, information regarding the transition from mobile hunting and foraging groups to sedentary farming communities in the region comes from the central Zagros Mountains, in Luristan and Kermanshah. This region is within the natural distribution of the wild ancestors of sheep, goats, cattle and pigs, and wild cereal crops, including wheat and barley, and had sufficient rainfall to support agriculture without the need for irrigation (Harlan and Zohary 1966; Hole 1987, Figure 5; Clutton-Brock 1999; Uerpmann 1987; Zohary and Hopf 2000; Weeks 2013a). Sites such as Ganj Dareh, Tepe Abdul Hosein, Tepe Guran, Sheikh-e Abad and Jani, which are situated between 1200 and 1900 m asl (Matthews 2013) suggest that moderate altitude mountain valleys were suitable for settlement in the early Holocene, from ~12.0–10.0 ka BP/c.10,000–8,000 BC (Weeks 2013a). These sites have provided critical evidence that processes of animal and plant domestication in the Zagros began as early as in other parts of the Fertile Crescent, including evidence for the long-term exploitation and management of the wild precursors to plant and animal domesticates, eventually leading to a diverse Neolithic subsistence base built on domesticated ungulates and cereals, alongside a continued tradition of the exploitation of diverse wild plant and animal resources (e.g. Zeder and Hesse 2000; van Zeist et al. 1984; Weeks 2013a). The earliest aceramic Neolithic site recorded in the lower ranges of the central Zagros Mountains is Chogha Golan, 485 masl, where excavations have revealed more than 8 m of occupational deposits spanning two millennia from ~11.8–9.6 ka BP/c.9,800–7,600 BC (Conard et al. 2013; Riehl et al. 2015). Detailed analyses of plant and animal remains document a slow and uneven transition to agriculture—incorporating a long period of cultivation of wild plant precursors and the exploitation of other wild species that were never ultimately domesticated (Riehl et al. 2015; Weide et al. 2017)—that is argued to reflect a pattern common across ancient southwest Asia, in which human adaptations were strongly determined by variable local environmental conditions (Riehl 2016).

304

Climate Changes in the Holocene

Comparisons can be made with the partly contemporary aceramic site of East Chia Sabz, in the Seimarreh Valley at a slightly higher elevation of c.660 masl (Darabi et al. 2013), where preliminary analyses also support the pre-domestication cultivation of wild forms of founder crops (barley, hulled wheats, lentil, grass pea, bitter vetch) and possibly the exploitation of wild ungulates. From ~9.6–9.2 ka BP/c.7,600–7,200 BC, other aceramic Neolithic settlements are recorded in lowland areas, including sites such as Ali Kosh and Chagha Sefid in Deh Luran (Hole et al. 1969; Hole 1977) and Chogha Bonut in Susiana (Alizadeh 2003). The occupants of these settlements made use of domestic goats, sheep and cattle, and supplemented their diet with hunting, as well as growing barley and wheat, and making use of weeds and dung for fuel (Weeks 2013a). Slightly later aceramic Neolithic settlements are also known in other highland areas, including Sang-e Chakhmaq on the north Central Plateau (c.7,140–6,825 cal BC; Thornton 2013; Roustaei et al. 2015), and Tappeh Rahmatabad in the high Zagros of Fars province (c.7,070–6,670 cal BC; Azizi Kharanagi et al. 2013). At the site of Qaleh Rostam, situated at 1,900 masl in the Bakhtiari region of the Zagros Mountains and occupied at the transition from the aceramic to the ceramic Neolithic, an integrated program of bioarchaeological analyses has been used to suggest that the site was occupied for only a part of the year during the warmer months from spring to autumn (Daujat et al. 2016). Its occupants appear to have been focused on the herding of domesticated goats and subsidiary cultivation of emmer wheat, the latter of which is regarded as not inconsistent with use of the site by a mobile group (Daujat et al. 2016). Seasonality and sedentarization are significant issues for the earliest Neolithic groups in the central Zagros (Thomalsky 2016; Darabi et al. 2013; Matthews et al. 2010, 2013; Pullar 1990; Smith 1990; Mortensen 1972, Figure 1). Mobility—by individuals, small groups and perhaps whole communities, undertaken at multiple chronological (daily to seasonal) and geographical (local to long-distance) scales—appears to have been a key mechanism that facilitated the resilience of Neolithic communities, allowing them to adapt to climates and landscapes that varied considerably across time and space. Further to the south and east, there is aceramic Neolithic occupation at Tall-e Atashi and contemporary sites, which is situated adjacent to the (now dry) Poshtrood River on the southern part of the plateau close to Bam (Garazhian 2016; Garazhian and Shakooie 2013) and Mehrgarh at the far south-eastern edge of the Iranian Plateau on the Kacchi plain, close to the Indus River in Pakistani Baluchistan (e.g. Jarrige 2008; Jarrige et al. 2013; also Petrie et al. 2010; Petrie 2015). Early crop use at Mehrgarh is known through impressions of naked six-row barley (Hordeum vulgare), which makes up more than 90 per cent of the seeds observed, though domestic hulled six-row and wild and domestic hulled two-row barley were present, along with very low proportions of domestic emmer, domestic einkorn, and a free threshing wheat (Costantini 1984; Meadow 1996; Jarrige et al. 2013; Petrie 2015). Significant numbers of goat kids in human burials and the remains of relatively small sub-adult or adult animals in trash deposits indicate that behaviorally domesticated goats were exploited from the earliest levels, though the remainder of the faunal assemblage is dominated by wild species, including gazelle, goats, sheep, deer, buffalo, and cattle (Meadow 1981, 1996; Jarrige et al. 2013; Petrie 2015). The dating of the earliest occupation at Mehrgarh is debated and it is not clear that it was established before c.6,000 BC (Petrie 2015), while the excavators suggest that Tall-e Atashi dates to c.5,500 BC (Gharazian and Shakooie 2013; Garazhian 2016, Figure 16). These more easterly settlements thus appear to have been somewhat later than the aceramic Neolithic sites in the upland areas further to the west and the north of the plateau (Petrie 2015). It is notable that many of these settlements are situated on or adjacent to alluvial fans (Petrie and Thomas 2012), which suggests that their occupants were able to identify favorable ecological contexts in new areas that were suitable for the use of their existing farming practices. In the seventh millennium BC, Neolithic settlements on both the western and eastern sides of the plateau provide evidence for the invention and then relatively widespread adoption of soft ware pottery, which was initially undecorated, but was soon after slipped and decorated with increasingly sophisticated motifs (e.g. Iran: Ganj Dareh, Chogha Bonut, Chogha Mish, Qaleh Rostam; Smith 1990; Alizadeh 2003; Petrie 2011, 2012; Weeks 2013a; Pakistan: Mehrgarh, Kili Gul Mohammad;

The Iranian Plateau and the Indus River Basin

305

Vandiver 1995; Petrie et al. 2010). Neolithic sites occupied by pottery users engaging in hunting, herding, farming and gathering were soon distributed widely, though there was a chronological separation in the developments at the western and eastern sides of the Iranian Plateau. For example, from ~8.3–8.0 ka BP/c.6,300–6,000 BC onwards, ceramic Neolithic settlements were spread throughout the Zagros Mountains and around the northern and southern edges of the Iranian Plateau (e.g. north-west: Yanik Tape, Hajji Firuz Tepe; Central Plateau: Tepe Sialk, Cheshme Ali, Tappe Pardis, Tepe Zagheh; north-east: Sang-e Chakhmaq East, Tappeh Deh Kheir, Kalateh Khan; south: Tepe Rahmatabad, Tal-e Mushki, Tal-e Bashi, Tal-e Jari, Tal-e Nurabad; Weeks 2013a,b; Rezvani and Roustaei 2016). In the north-east, this expansion continued into Central Asia with the appearance of Jeitun period Neolithic sites in Turkmenistan from c. 6,100 BC, which has been linked in part with the climatic deterioration of the ~8.2 ka BP event (Harris 2010, 236). There is some evidence that these north-eastern Neolithic communities exploited a narrower range of domesticates, with a particular focus on caprines and glume wheats, and limited evidence for the use of pulses, though they supplemented this diet with the continued hunting of wild animals (including gazelle, caprines, onager and wild boar) (Tengberg and David 2016; Mashkour et al. 2016; Charles and Bogaard 2010; Dobney and Jaques 2010). While it has been tentatively suggested that the narrow spectrum of plant domesticates at these sites may reflect selection of species adapted to low soil fertility and low water availability (Charles and Bogaard 2010, 162–3), it is clear that access to a wide range of ecological zones was a key strategy in their subsistence adaptations, which needed to be resilient to changing environmental conditions (Harris 2010, 191). Ceramic Neolithic sites in Kerman on the southern part of the plateau, appear to be slightly later than those to the west, dating to the mid-sixth millennium BC (Tepe Yahya, Tal-e Iblis; Petrie 2011; Weeks 2013a), which also matches the well-dated ceramic Neolithic levels at Mehrgarh (Shaffer 1992; Jarrige 2000; Petrie et al. 2010). It is unclear at present how extensive the earliest ceramicusing population in the Indus River Basin was, as such early material has only been identified through excavations at Mehrgarh and Kili Gul Mohammad, and found on the surface of a small number of other sites (Possehl 1999). As with the aceramic Neolithic settlements, a significant number of these settlements occupied by early ceramic users were also situated on or near alluvial fans (Petrie and Thomas 2012). Weeks (2013a, 69) has noted that it seems likely that the variability seen in early Iranian Neolithic culture and economy reflects not only adaptations by farmers to the new environments they were experiencing, but also to their interactions with foraging groups. Viewed as a whole, the evidence suggests that there was a rapid expansion of the Neolithic population during the period of early ceramic use, both in terms of the number of settlements within individual areas, and the extent of their distribution into and across highland areas of the plateau. The increase in settlement density and dispersal of population appears to have primarily occurred after the ~8.2 ka BP event, although the degree to which that event impacted upon these processes is unclear. For instance, there was a notable increase in the number of settlements in the intermontane and high valleys of the southern Zagros from ~8.0 ka BP onwards (Nishiaki 2010; Weeks 2013a). There then appears to have been a delay before a phase of dispersal, 600–700 years later, around ~7.5 ka BP, when ceramic-using Neolithic populations appeared in Kerman and areas further to the east. It is unclear whether this overarching process had separate phases of a rapid increase in settlement density followed later by a rapid dispersal, or whether it was all a more gradual process. Furthermore, these processes show no obvious correlation with long-term or abrupt climatic changes, so perhaps demographic pressures had necessitated further dispersals. The ~8.2 ka BP event potentially affected certain areas and environments, and may have encouraged population displacement, as it appears to have done in Anatolia (see above; Weninger et al. 2006), but the degree to which the intensification and expansion of Neolithic settlement in the highlands of the Iranian Plateau can be seen as a response to that process needs to be further explored. It appears that there was a dispersal of Neolithic populations throughout the highlands and piedmonts at the western edge of the Indus River Basin from ~6.7 to 5.5 ka BP, with settlements

306

Climate Changes in the Holocene

ultimately appearing out on the Indus floodplain (Petrie et al. 2010; also Possehl 1999). Where data are available, it appears that these populations were making use of a crop species and animals similar to those seen at Mehrgarh and the sites in Iran to the west (Weber 2003; Fuller 2006; Thomas and Cartwright 2010). As with processes on the Iranian Plateau, this more eastern dispersal does not appear to have been motivated by a climatic driver, and demographic pressure may have been a more significant factor (Petrie and Thomas 2012). As a further contrast to the developments that took place on the western side of the Iranian Plateau, it is now clear that there were at least three distinct Neolithic trajectories within South Asia: “western”, dominated by winter cereals like wheat and barley; “northern”, dominated by summer cereals like rice and millet; and “southern”, dominated by summer pulses, predominantly beans (Fuller 2006, 2011; Kingwell Banham et al. 2015). Each of these developments was clearly a local manifestation of Neolithic practices where the particular crops that were being exploited were suited to the local environmental conditions. As will be discussed below, subsequent inhabitants of the Indus plains employed a range of specific subsistence adaptations (in terms of crops and associated farming practices) derived from each of these distinctive South Asian Neolithic traditions (Petrie et al. 2016, 2017; Bates et al. 2017a,b; also Weber et al. 2010).

14.3.4 Late Chalcolithic The fourth millennium BC was yet another a critical period of innovation and transformation for early human societies, as it marked the appearance of the first cities and the first evidence for elaborate administrative technologies, in the form of cylinder seals being used in combination with clay bullae, numerical tablets and ultimately proto-writing, which all appear at Uruk in southern Mesopotamia during the Late Uruk period c.3,300–3,100 BC/~5.3–5.1 ka BP (Nissen 2001; Englund 1998; Pollock 1999; Algaze 2008). There were major urban centres elsewhere in southern Mesopotamia (e.g. Nippur, Abu Salabikh, Umma), northern Mesopotamia (e.g. Tell Brak, Tell Hamoukar, Tepe Gawra), and Iran (Susa and Chogha Mish), which all appear to have been involved in a medium- to long-range exchange and trading network that linked populations in many of the surrounding regions, including settlements in the Taurus Mountains (e.g. Haçinebi, Arslan Tepe), and on the Iranian Plateau (e.g. Godin Tepe, Tappeh Sialk, Mahtoutabad; Algaze 2005; Rothman (ed.) 2001; Postgate (ed.) 2002; Butterlin 2003; Petrie 2013a,b). The pattern of distribution of distinctive material, including bevel-rim bowls and shouldered storage jars, has been taken to indicate the existence of a Mesopotamia-centred “world-system” (Algaze 2005, 2008), though other interactive models such as “trade diasporas” and “distance parity” have been used to interpret the same evidence (Stein 1999). Here we will simply refer to it as the “Uruk phenomenon” (Petrie 2013b). At least partly contemporaneous with these developments (c.3,300–2,900 BC/~5.3–4.9 ka BP), a set of similar administrative technologies with distinctive seal iconography and a related yet distinct style of proto-writing (“Proto-Elamite” or Susa III) were in use at settlements of various sizes across an extensive area of the Iranian Plateau, including the lowland plain of Susiana (Susa) and Ram Hormuz (Tal-e Gezer), and highland areas at various points on the plateau, including Godin Tepe, Tappeh Sialk, Tal-e Malyan, and Tepe Yahya (Dahl et al. 2013). Although the Iranian proto-writing is not always attested, sites with distinctive material culture are found in many other areas of the Iranian Plateau, including other parts of the Central Plateau (Arisman), Fars (Mamasani and the Kur River Basin), Kerman (Mahtoutabad), and even Pakistani Makran (Miri Qalat), while protowriting without the other material has been documented in the earliest levels at Shahr-i Sokhta (see Petrie [ed.] 2013). Precisely what the distribution of this material represents has been much debated; although there is no consensus, it is clear is that this “Proto-Elamite Phenomenon” was much more spatially extensive than its Uruk counterpart (Petrie 2013a,b). During this period, there were clear shifts in the nature and intensity of cultural interaction taking place at the eastern edge of the Iranian Plateau and the Indus River Basin, involving increases in settlement sizes, the development of nascent administrative technologies including stamp seals, and

The Iranian Plateau and the Indus River Basin

307

the exploitation and distribution of raw material resources at short, medium and long-range scales (Chakrabarti 1995; Kenoyer 1997; Possehl 1999; Wright 2010; Law 2011). Nothing directly comparable to the Uruk or Proto-Elamite Phenomena developed in this region at this time. The site of Mundigak in southern Afghanistan was potentially occupied throughout this period, with Phase III starting in the late fourth millennium BC, and occupation continuing throughout much of the third millennium BC in Phases III5–6 –IV1, which were characterized by the building and rebuilding of major structures, platforms and retaining walls, and a “palace” and a “temple” in Phase IV1, (Casal 1961; Shaffer 1978; Petrie and Shaffer in press). As noted above, the ~5.2 ka BP event has been linked to the end of the late Uruk period societies in Mesopotamia (Weiss 2003; Staubwasser and Weiss 2006). Although there is clear evidence for socio-economic and even political disruption during the Late Uruk period (c.3,300–3,100; Johnson 1973, 1989), there is no neat “break-point” at or after ~5.2 ka BP that marks the end of the Uruk Phenomenon chronologically (see Wright and Rupley 2001; Petrie 2014). In fact, a number of erstwhile Uruk period sites have radiocarbon dates suggesting that occupation continued for some time after that date, particularly sites in Syria such as Habuba Kabira and Jebel Aruda (Wright and Rupley 2001), which, it has been suggested, were established by populations fleeing southern Mesopotamia (Johnson 1989). The precise date for the end of the Uruk period is unclear, but it is usually argued that it continued down to c.3,100 BC, and was then followed by the Jemdet Nasr period on the plains of Mesopotamia (Pollock 1999). However, some of the Uruk period radiocarbon dates extend to c.2,900 BC (e.g. Wright and Rupley 2001, Figures 3.1 through 3.3) and they overlap statistically with virtually all of the known radiocarbon dates for the Proto-Elamite period, meaning that it is impossible to differentiate the two phases chronologically (Dahl et al. 2013; Petrie 2014, in press). The Proto-Elamite period occupation does, however, appear to end around c.2,900 BC, after which there is an apparent gap in the occupation sequences at many sites across the Iranian Plateau (e.g. Arisman, Tepe Yahya, Tol-e Spid, Tol-e Nurabad). However, there is also evidence for continuity at some sites (e.g. Susa, Carter 1980; Tal-e Malyan, Miller and Sumner 2004; Alden et al. 2005; Tappeh Hisar, Dyson and Howard [eds] 1989; Shahr-i Sokhta, Tosi [ed.] 1983; Salvatore and Vidale 1997), and occupation at major settlements in the east, including Konar Sandal South in the Halil Rud region (Madjidzadeh 2008), and Mundigak in Afghanistan (Casal 1961; Shaffer 1978; Petrie and Shaffer in press). Several of these Proto-Elamite settlements were situated on alluvial fans or piedmont areas, but others were situated on extensive alluvial plains. Unfortunately, we know very little about subsistence practices at these settlements, and it is only at Tal-e Malyan that detailed analysis of bioarchaeological remains has been carried out (e.g. Miller 1991; Zeder 1991). The potential impact of climate and/or climate change at ~5.2 ka BP on any of these processes is very dependent upon the nature and duration of the event in these regions, and also the degree to which it had similar effects across the diverse and varied landscapes of the Iranian Plateau and the Indus River Basin. The Uruk Phenomenon was largely constrained to the alluvial plains of Mesopotamia, and to a limited extent the upland areas of southern Anatolia and western Iran, which is a substantial area with diversity in environment and marked differences in the distribution of rainfall and surface water availability (Petrie 2013b). Agriculture in southern Mesopotamia was highly dependent on the use of irrigation systems of various scales, and the city-states and early empires of the alluvium grew and developed within or on the edges of an anastomosed and deltaic riverine system characterized by abundant salt- and freshwater marshes (Wilkinson 2013; Pournelle and Algaze 2014). Following Pournelle (2003, 5), Wilkinson (2013) has emphasized the importance of “littoral” resources as a critical third pillar to the Mesopotamian economy alongside cereal cultivation and animal husbandry, and Hritz and Pournelle (in press; also Pournelle and Algaze 2014) have proposed that the southern Mesopotamian system displayed a “deltaic resilience”. In this context, littoral resources and context were one aspect of long-term, sustainable, human-environment interaction that can usefully be considered as a process of niche construction. These niches allowed early southern Mesopotamian societies not only access to varied and abundant resources and productive strategies (and ultimately to develop unevenly distributed wealth and social hierarchies).

308

Climate Changes in the Holocene

In contrast to the Uruk Phenomenon, the Proto-Elamite Phenomenon occurred across a much larger area, and included settlements in piedmont zones, intermontane valleys and basins, desert margins, alluvial fans and inland lake zones (Petrie 2013b). Although it has been argued that there was a chronological correlation between the ~5.2 ka BP event and the Uruk decline (Weiss 2003; Staubwasser and Weiss 2006), we argue that this link is likely overstretched and tenuous, as it is largely unsupported at a local level in any of the relevant areas. It thus does not come close to adequately explaining the interrelationships between different populations and categories of material evidence across the Iranian Plateau let alone across the various neighboring areas.

14.3.5 Bronze Age The mid-late third millennium BC was a period of dramatically increased socio-economic and political complexity across the Iranian Plateau, the Indus River Basin, and adjacent regions. In ancient Mesopotamia, this period saw the development and proliferation of the power of city-states that had been developing since the late fourth millennium BC, and the appearance of the first large-scale empires under Sargon of Agade, his descendants, and their political successors (Pollock 1999). On the Iranian Plateau, a number of political entities and/or confederations are documented in Mesopotamian royal inscriptions (e.g. Elam, Anshan, Awan, Pashime, Marhashi, Shimashki; Steinkeller 2012, 2016; Potts 1999), and other entities in the Persian Gulf and the Indus River Basin are also mentioned in these sources (e.g. Dilmun, Makan, Meluhha; Potts 1997; Laursen and Steinkeller 2017). It was noted above that the construction of ecological niches in southern Mesopotamia made it possible for allowed early societies to access varied and abundant resources and productive strategies, and it is also likely that these niches also provided a buffer against significant climate fluctuations, such as the ~4.2. ka BP event, that are argued to have dramatically affected rain-fed agricultural systems in northern Mesopotamia (Weiss 2015, 2016). The archaeology of the Iranian Plateau in the third millennium BC is less well resolved. In the western part of the Iranian Plateau, there appears to have been continued occupation from the fourth millennium BC onwards at major sites including Susa (Carter 1980) and Tal-e Malyan (Miller and Sumner 2004; Alden et al. 2005) in the south, and Tappeh Hissar (Dyson and Howard [eds] 1989) and Tureng Tepe (Deschayes 1970, 1975) in the north (Voigt and Dyson 1992). A number of other major sites appear to have been occupied or reoccupied at some point during the third millennium BC, including Godin Tepe (Gopnik and Rothman 2011), but many settlements with major fourth millennium BC occupation were abandoned in the third millennium BC, including Tappeh Sialk (Ghirshman 1938) and Arisman (Vatandoust et al. [eds] 2011). As with the Late Chalcolithic, several of these settlements were situated on alluvial fans or piedmont areas, but more diverse environments were also exploited, including inland lakes and extensive alluvial plains, where some type of irrigation is likely. There is relatively limited information about the subsistence practices of these populations, but where data are available, it is clear that they made extensive use of winter crops, including wheat and barley (e.g. Miller 1982, 1991, 2003), and animal domesticates that had been exploited since the Neolithic, particularly sheep, goat and cattle (e.g. Meadow 1986; Zeder 1991). Grape seeds have been attested in some abundance at Kaftari period Tal-e Malyan, and grape wood was also attested, suggesting wine-making was practiced (Miller 1982, 2003). Wood charcoals suggest that Tal-e Malyan lay in an area of pistachio and almond forest that contained juniper, and during the Kaftari period this thinned, and more distant oak forests began to be exploited, though this is seen to be the result of demographic pressure on resources (Miller 1982, 1985, 2003). Bronze Age settlement in south-eastern Iran at this time was most certainly enabled by the coalescence of an oasis agricultural system (Tengberg 2012) built around irrigated gardens of domesticated date palms that provided not only their fruit, but also shade and thus a suitable micro-climate for successfully growing other crops including fruit trees, vegetables and cereals, in an otherwise arid landscape. Major sites (re-)occupied during the third millennium BC include Shahdad (Hakemi 1997) and Keshit (Eskandari et al. 2014) on the edges of the Lut Desert, Konar Sandal South in

The Iranian Plateau and the Indus River Basin

309

the Halil Rud/Jiroft region (Madjidzadeh 2008), Tepe Yahya in Daulatabad (Potts 2001; Mutin and Lamberg-Karlovsky 2014), Bampur in the eastern Jazmourian Basin (de Cardi 1970; Mutin 2015) and Shahr-i Sokhta in Sistan (Tosi [ed.] 1983; Salvatore and Vidale 1997). The significance of the date palm to these Bronze Age societies is amply demonstrated not only by its presence in archaeological assemblages from Konar Sandal South, Tepe Yahya and Shahr-i Sokhta (Mashkour et al. 2013, 236), but also by artistic representations of date palms on “Intercultural Style” soft-stone vessels from southeastern Iran (Madjidzadeh 2003, 2008; Perrot 2012 [2008]; Tengberg 2012). Archaeobotanical and zooarchaeological analyses from archaeological sites in the Jiroft region (Mashkour et al. 2013) indicate a subsistence focus on the alluvial plains for cereal and fruit production (wheat, barley, dates, grapes), and the husbandry of domesticated animals, particularly goats. These activities were supplemented by the hunting and gathering of local plant and animal species from an environment that was probably wetter and more floristically dense and diverse than at present. Plant and animal remains from beyond the immediate environs of the site indicate connections with colder and higher elevation regions (for wood-cutting, fruit-gathering, and possibly seasonal herding), and the remains of marine fish indicate connections with the Persian Gulf more than 250 km to the south. At Shahr-i Sokhta, the typical Bronze Age subsistence pattern combining the cultivation of cereals (wheat, barley) and fruits (dates, grape) alongside the use of domestic sheep, goat and cattle, was supplemented with a strong focus on hunting (gazelle, onager, and a great variety of marsh bird species) and fishing, attesting to the significance of Lake Hamoun and its associated swamp lands at this time (Gala and Tagliacozzo 2014). During the third millennium BC, the Indus River Basin witnessed the development of the Indus Civilization, which incorporated populations living across much of the greater watershed for the Indus River and saw the appearance of the first cities and complex administration in South Asia (Lal 1997; Kenoyer 1997; Possehl 2002; Agrawal 2007; Wright 2010; Coningham and Young 2015). The extensive zone occupied by Indus populations, who made use of what has been described as a veneer of similar types of material culture (Meadow and Kenoyer 1997; Petrie 2013c; Chase et al. 2014; Petrie et al. in press) was characterized by considerable ecological variability, and different iterations of marginality (Petrie 2017; Petrie et al. 2017). This variability in turn required a range of specific adaptations in terms of subsistence practices, including the use of crops, and their associated farming practices, that originated within each of the distinctive South Asian Neolithic traditions (see above; Petrie et al. 2016, 2017; Bates et al. 2017a,b; also Weber et al. 2010). Prior to and during the urban phase of the Indus Civilization, populations in some areas appear to have engaged in variations of mono-cropping involving almost exclusive use of winter or summer crops, while in other areas winter and summer crops were both used, and either one set of crops was dominant, or the crops were used in a flexible way that appears to have been down to the agency of village farmers (Petrie and Bates 2017). The ~4.2 ka BP event has been given considerable attention due to its perceived impact upon complex societies of the late third millennium BC from the Mediterranean to the Indus (e.g. Staubwasser and Weiss 2006; Weiss 2015, 2016). However, this argument fails to differentiate the complexities of the climate systems in operation across these regions, particularly the degree to which the winter and/or summer systems were affected, and also the degree of cultural and environmental diversity that already existed (Petrie 2017; Petrie et al. 2017). The diversity of the environments and rainfall patterns across the Indus River floodplain suggest that the ~4.2 ka BP event is unlikely to have had consistent effects across the region, and although factors like tectonics and river shifts have also been considered, at present we lack definitive dates for most of these processes (Wright 2010: Petrie et al. 2017). Thus, while the ~4.2 ka BP event appears to have dramatically affected the winter rainfed agricultural systems of northern Mesopotamia (Weiss 1986), the contemporaneous evidence from the Iranian Plateau and the Indus River Basin demonstrates that the impact of the ~4.2 ka BP event was not consistent. It is notable that the ~4.2 ka BP event has been recognized in zones that receive winter rain and zones that receive monsoon rain, but the interconnection between these two weather systems and the event remains poorly understood and must be the subject of future research.

310

Climate Changes in the Holocene

In archaeological terms, we know that a wide variety of water access and control regimes was employed across this substantial and environmentally diverse zone, and while some areas may have suffered direct impact due to a lack of rainfall, particularly those where farming was adapted to direct rainfall or predictable rainfall runoff, farmers in other areas were already adapted to variable and unpredictable rainfall and potentially resilient to climate change (Petrie et al. 2017; Petrie 2017). In the Kur River Basin in highland southwest Iran, for example, the period after ~4.2 ka BP (c.2,200 BC) witnessed a significant expansion in the number of known settlements and the reappearance of a large-scale urban centre at Tal-e Malyan (Sumner 1989, 1990), which had been established as a major centre in the later fourth millennium BC and then seen a dramatic reduction in the extent of occupation in the early third millennium BC (Miller and Sumner 2004; Alden et al. 2005; Petrie et al. 2005; Alden 2013). There is also clear evidence for the revival of long-range links between Tal-e Malyan and Susa in Khuzestan, and it is particularly notable that distinctive Kaftari type pottery originating in the Kur River Basin has been recovered from sites across highland southwestern Iran and outside the region on Bahrain (Qala’at al-Bahrain) and in the United Arab Emirates (Tell Abraq and UNAR2; Petrie et al. 2005). A very different pattern is recorded in south-eastern Iran, where the late fourth and third millennia BC witness dramatic settlement growth, followed by a rapid decline in sedentary settlement in the late third or early second millennium BC, although regional variations are seen (e.g. Prickett 1986, Figure 7.7; Madjidzadeh 2008, 74; Fouache 2012 [2008]; Fouache and Garçon 2009; Lawler 2011; Pittman 2013; Moradi et al. 2014). Although numerous scholars have noted the correlation between settlement decline and the abrupt climatic deterioration of the ~4.2 ka BP event (e.g. Madjizadeh 2008; Pittman 2013), interpretations of the role of climate change in the decline of Bronze Age civilizations in south-eastern Iran vary considerably. Pittman (2013, 304, 318), although allowing for regional variations and the resilience and adaptation of specific communities, has stated that “it is generally assumed that increasing aridity in the late Holocene intensified toward the end of the third millennium, contributing to the gradual depopulation of the area in the early second millennium BC.” Madjidzadeh (2008, 74), drawing on the survey data from the Halil Rud/ Jiroft, likewise has suggested that the reduction in sedentary settlement was caused by extended drought and related salinization of agricultural land. Similar arguments giving climate change a primary causal role in the reduction of archaeologically visible settlement have also been postulated for the region at the transition from the Bronze Age to the Iron Age, in the late second millennium BC (Magee 2013, 494; Gurjazkaite et al. 2018). Others have noted the variable climatic sensitivity and hydrological instability of sub-regions of south-eastern Iran, suggesting that abrupt climate change is more likely to have impacted areas such as Sistan than the comparatively well-watered Halil Rud/Jiroft region (e.g. Fouache and Garçon 2009). Such considerations allow for explanations in which climate change is one component in regionally variable trajectories of human adaptation that also incorporate contemporary changes to economic and social systems and water access technologies. At Shahr-i Sokhta, local environmental changes, including the drying up of the Helmand delta, have been directly linked to the gradual abandonment of southern Sistan at the end of the third millennium BC (Gala and Tagliacozzo 2014, 320), with consequent increasing need to store and stockpile agricultural process being linked to increasing risks from food spoilage (Milanesi et al. 2015) and associated subsistence pressure. Similar, multi-causal explanations have also been put forward, including the local impacts of much larger contemporary changes to economic systems across the ancient Near East (e.g. Moradi et al. 2014, 286–287; Lawler 2011). Particularly critical for continued human occupation in south-eastern Iran in such periods of climatic and environmental unpredictability was the flexibility of groups to transition between variable modes of subsistence and social organization. Lamberg-Karlovsky and Tosi (1973, 53) have regarded this resilience as the product of a “continuous dialectic … between pastoralists, agriculturalists, nomads, villagers and city dwellers”. Overall, it is clear that episodes of “collapse” documented in the region’s archaeological record can in many instances—and from a different

The Iranian Plateau and the Indus River Basin

311

perspective—be regarded as successful adaptations by communities that became more dispersed, mobile, and differently organized, but archaeologically much less visible (cf. Prickett 1986). Given the diversity of Indus farming practices, the 4.2 ka BP event is unlikely to have had consistent effects across the entire Indus River Basin. Careful re-analysis of the settlement distribution data from this period suggests that Indus populations displaced towards the areas that received monsoon rain, as settlements appear to have proliferated in the areas that today receive more than 300 mm of summer rain per annum (Petrie et al. 2017; Petrie 2017). Although it has been suggested that Indus urbanism declined in the wake of the 4.2 ka BP event (e.g. Staubwasser and Weiss 2006), the de-urbanization process appears to have been particularly protracted, and it was only after c.1900 BC that the cities appear to have been largely abandoned, suggesting that there was no Indus “collapse” in the strictest sense (Petrie 2017). It was thus a process of socio-economic transformation where there were changes to long- and medium-range exchange and trading operations, gradual processes of de-urbanization, and the continuity of a protracted period of population mobility and settlement displacement (Petrie et al. 2017; Petrie 2017). It is also interesting to consider areas immediately adjacent to the Iranian Plateau. For example, although the Early Bronze Age population of south-eastern Arabia grew rapidly on a newly consolidated mixed agro-pastoral subsistence base (Al-Jahwari 2009), significant climate deterioration in the late third millennium BC (i.e. at the end of the MHHP; Parker et al. 2016), contemporary with the 4.2 ka BP event, has been linked to a dramatic reduction in known settlements in the subsequent Middle and Late Bronze Ages (c.2000–1200 BC) (Parker and Goudie 2007). Significantly, largescale local metallurgical production and its associated fuel requirements may also have impacted and exacerbated the de-vegetation/forestation of this semi-arid region at this time (Weisgerber 1991; Weeks 2003, 35). While many sites were abandoned by c.2,000–1,900 BC— as in the Indus region some 200–300 years after the onset of the ~4.2. ka BP abrupt climate event—some substantial settlements show continued sedentary occupation across the Early–Middle Bronze Age boundary and there is evidence for a sizeable continuing human presence in the form of large collective and individual tombs (Magee 2014) and persistently occupied temporary sites with substantial residues of food gathering activities (Weeks et al. 2017). In addition to the difficulties of recognizing what may be a more dispersed system of sedentary agricultural settlement at this time (Velde 2009), this pattern can also be regarded as evidence for segments of local communities responding to changed environmental circumstances by exploiting their capacity to move between mobile and more sedentary systems of adaptation and to balance continued agro-pastoral production with hunting, gathering, and fishing (Weeks et al. in press). Therefore, despite significant climate deterioration in the late third millennium BC and increasing desertification and episodes of more intense climate deterioration in more recent millennia (Parker and Goudie 2008), occupation of Arabia remained continuous throughout the late Holocene due to the resilience and adaptability of local communities. A counterpoint to these changes in south-eastern Arabia is provided by the contemporary Early Dilmun civilization on Bahrain, which displayed rapid growth in scale and social complexity in the four centuries after the ~4.2 ka BP event (Laursen 2008, 2009, 2017; Højlund 2007). With an integrated agro-pastoral system built around date palm oasis agriculture that drew predominantly upon springs and artesian wells tapping the vast Eastern Arabian aquifer (Larsen 1983), the Early Dilmun communities established a subsistence base that allowed them to flourish in the period of climatic deterioration in the four centuries after the ~4.2 kyr BP event (Laursen 2008, 2009; Højlund 2007).

14.4 CONCLUSIONS This review of the evidence for climate, environment and human occupation across the Iranian Plateau and the Indus River Basin amply demonstrates the diversity and complexity of each of these parameters during the early–mid Holocene. The breadth of possible human responses to a changing climate widened dramatically at the beginning of this period, with the transition to productive

312

Climate Changes in the Holocene

plant and animal economies and the associated changes in social organization and demography (e.g. Bocquet-Appel 2011). This slowly assembled “package” of resources and behaviors—perhaps initially achievable only in the comparatively stable, warmer, wetter and more CO2-rich climate of the early Holocene—accumulated over the course of the terminal Pleistocene and Holocene and facilitated the creation of new niches for subsistence, although paradoxically, it has been argued that early Neolithic societies were less resilient than their Paleolithic forebears (Bar-Yosef 2015). Regardless, the complexity of subsequent mid-Holocene adaptations reflects the ability of human groups to reconstitute these elements in many ways, in articulation with changes to socio-economic organization, technologies, and mobility. There are certainly limitations to our knowledge of each category of evidence, but it is clear that despite a number of chronological correlations between abrupt events of climate change and instances of cultural transformation, there is no simple relationship between the two. Cultural transformation is thus not contingent on climate change, but it clearly has the potential to be influenced by it. However, the variability of the climate and environment across the Iranian Plateau and the Indus River Basin, and the behavioral adaptations that were required to live in those environments, intersect with considerable variation in the ways that human societies organized themselves and interacted with others, which were independent of climate or climate change. A review such as that presented here hopefully serves to highlight difficulties in the exploration of past human–climate interactions. Looking to the future of research, it appears that addressing these difficulties will require not only an expansion of archaeological and paleoclimatic field and laboratory research, but also the development of new practical and theoretical approaches to the exploration of causality. Whereas previous research has focused on abrupt events and the identification of chronological correlations between instances of climate and cultural change, it is becoming increasingly important to make use of methods that allow modelling of the links between climate change, resource variation, and processes of human demographic and socio-economic change (e.g. d’Alpoim Guedes et al. 2016). Computational approaches including remote sensing and agentbased and network modelling are revolutionizing the way that these questions are approached and addressed, and their systematic implementation will ultimately support a much clearer understanding of the scale and scope of human responses to long-term and abrupt climate change across the Iranian Plateau and the Indus River Basin.

WEBSITES https://www.ncdc.noaa.gov/data-access/paleoclimatology-data

ACKNOWLEDGMENT The contribution made by C.A. Petrie was supported by funding from the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation program (grant agreement no. 648609).

REFERENCES Agrawal, D.P. 2007. The Indus Civilization. An Interdisciplinary Perspective. Delhi: Aryan Books. Aimers, J., and D. Hodell. 2011. Societal collapse: Drought and the Maya. Nature 479: 44–45. Alden, J.R. 2013. The Kur River Basin in the Proto-Elamite Era—surface survey, settlement patterns, and the appearance of full-time transhumant pastoral nomadism. In Ancient Iran and Its Neighbours: Local Developments and Long-Range Interactions in the Fourth Millennium BC, edited by C.A. Petrie, 207– 232. Oxford and Oakville: Oxbow Books. Alden, J.R., K. Abdi, A. Azadi, G. Beckman, and H. Pittman. 2005. Fars archaeological project 2004: Excavation at Tal-e Malyan. Iran 43: 39–47.

The Iranian Plateau and the Indus River Basin

313

Algaze, G. 2005. The Uruk World System: The Dynamics of Expansion of Early Mesopotamian Civilisation. 2nd revised edition. Chicago: University of Chicago Press. Algaze, G. 2008. Ancient Mesopotamia at the Dawn of Civilisation: The Evolution of an Urban Landscape. Chicago: University of Chicago Press. Alijani, B., J. O’Brien, and B. Yarnal. 2008. Spatial analysis of precipitation intensity and concentration in Iran. Theoretical and Applied Climatology 94 (1–2): 107–124. Alizadeh, A. 2003. Excavations at the Prehistoric Mound of Chogha Bonut, Khuzestan, Iran, Seasons 1976/77, 1977/78, and 1996. OIP 120, Chicago: Oriental Institute. Al-Jahwari, N.S. 2009. The agricultural basis of Umm an-Nar society in the northern Oman peninsula (2500– 2000 BC). Arabian Archaeology and Epigraphy 20: 122–133. DOI:10.1111/j.1600–0471.2009.00315.x Allchin, B., A. Goudie, and K. Hegde. 1978. The Prehistory and Palaeogeography of the Great Indian Desert. London: Academic Press. Azarnoush, M., and B. Helwing 2005. Recent archaeological research in Iran—Prehistory to Iron Age. Archaeologische Mitteilungen aus Iran und Turan 37: 189–246. Azizi Kharanagi, H., H. Fazeli, and Y. Nishiaki. 2013. Tepe Rahmatabad: A pre-pottery and pottery Neolithic site in Fars Province. In The Neolithisation of Iran: the Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 108–123. Oxford: BANEA and Oxbow Books. Balling, R.C., M.S. K. Kiany, S.S. Roy, and J. Khoshhal. 2016. Trends in extreme precipitation indices in Iran: 1951–2007. Advances in Meteorology 2456809: DOI:http://dx.doi.org/10.1155/2016/2456809 Bar-Yosef, O. 2011. Climatic fluctuations and early farming in West and East Asia. Current Anthropology 52: S175–S193. Bar-Yosef, O. 2015. Facing climatic hazards: Paleolithic foragers and Neolithic farmers. Quaternary International 428: 64–72. Bates, J., C.A. Petrie, and R.N. Singh. 2017. Approaching rice domestication in South Asia: New evidence from Indus settlements in northern India. Journal of Archaeological Science 78: 193–201. doi:http:// dx.doi.org/10.1016/j.jas.2016.04.018 Bates, J., R.N. Singh, and C.A. Petrie. 2017. Exploring Indus crop processing: Combining phytoliths and macrobotanical analysis to consider the organisation of agriculture in northwest India c.3200–1500BC. Vegetation History and Archaeobotany (special issue) 26: 25–41 [published online 21 May 2016; doi:http://dx.doi.org/10.1007/s00334-016-0576-9]. Berkelhammer, M., A. Sinha, L. Stott, H. Cheng, F.S.R. Pausata, and K. Yoshimura. 2012. An abrupt shift in the Indian monsoon 4000 years ago. In Climates, Landscapes, and Civilizations, edited by L. Giosan, D.Q. Fuller, K. Nicoll, R.K. Flad, and P.D. Clift, 75–87. Geophysical Monograph Series 198. Washington D.C.: AGU. Biagi, P., and E. Starnini. 2018. Neanderthals and modern humans in the Indus valley? The middle and late (Upper) palaeolithic settlement of Sindh, a forgotten region of the Indian subcontinent. In The Middle and Upper Paleolithic Archeology of the Levant and Beyond, Replacement of Neanderthals by Modern Humans Series, edited by Y. Nishiaki and T. Akazawa. Springer. DOI:10.1007/978-981-10-6826-3_12 Blinkhorn, J., H. Achyuthan, P. Ditchfield, and M. Petraglia. 2017. Palaeoenvironmental dynamics and Palaeolithic occupation at Katoati, Thar Desert, India. Quaternary Research 87: 298–313. Bocquet-Appel, J.P. 2011. The agricultural demographic transition during and after the agriculture inventions. Current Anthropology 52/S4: S497–S510. Bond, G., W. Showers, M. Cheseby, R. Lotti, P. Almasi, P. Priore, H. Cullen, et al. 1997. A pervasive millennial-scale cycle in North Atlantic holocene and glacial climates. Science, 278(5341): 1257–1266. Butterlin, P. 2003. Les Temps Proto-Urbains de Mésopotamie. Paris: CNRS Éditions. Carolin, S.A., J. Morgan, E. Peckover, R. Walker, G. Henderson, P. Rowe, J., andrews, et al. 2016. Iranian speleothems: Investigating quaternary climate variability in semi-arid Western Asia. Geophysical Research Abstracts 18: EGU2016–923. Carter, E. 1980. Excavations in Ville-Royale-I at Susa: The third millennium B.C. Cahiers de la Délégation Archéologie Française en Iran 11: 11–134. Casal, J.-M. 1961. Fouilles de Mundigak. Mémoires Délégation archéologique française en Afghanistan. Paris: C. Klincksieck Cauvin, J. 1994. Naissance des divinités, naissance de l’agriculture: La révolution des symboles au Néolithique. Paris: CNRS Éditions. Cauvin, J. 2000. The Birth of the Gods and the Origins of Agriculture. Translated by Trevor Watkins, with updated postscript. Cambridge: Cambridge University Press. Chakrabarti, D.K. 1995. The Archaeology of Ancient Indian Cities. Delhi: Oxford University Press, Delhi. Chakrabarti, D.K. 1999. India: An Archaeological History. Delhi: OUP.

314

Climate Changes in the Holocene

Chakrabarti, D.K. 2007. The Oxford Companion To Indian Archaeology: The Archaeological Foundations of Ancient India: Stone Age to AD 13th Century. Delhi. Oxford University Press. Charles, M., and A. Bogaard. 2010. Charred plant macro-remains from Jeitun: implications for early cultivation and herding practices in western Central Asia. In Origins of Agriculture in Western Central Asia: An Environmental-Archaeological Study, edited by D.R. Harris, 150–165. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Chase, B., D. Meiggs, P. Ajithprasad, and P.A. Slater. 2014. Pastoral land-use of the Indus Civilization in Gujarat: Faunal analyses and biogenic isotopes at Bagasra. Journal of Archaeological Science 50: 1–15. DOI:http://dx.doi.org/10.1016/j.jas.2014.06.013 Chase, B., D. Meiggs, P. Ajithprasad, and P.A. Slater. 2018. What is left behind: Advancing interpretation of pastoral land-use in Harappan Gujarat using herbivore dung to examine biosphere strontium isotope (87Sr/86Sr) variation. Journal of Archaeological Science 92: 1–12. DOI:https://doi.org/10.1016/j. jas.2018.01.007 Cleuziou, S., and M. Tosi. 2007. In the Shadow of the Ancestors. Muscat: Ministry of Culture and Heritage of Oman. Clutton-Brock, J. 1999. A Natural History of Domesticated Mammals. Cambridge: CUP. Conard, N.J., E. Ghasidian, S. Heydari and M. Zeidi 2006. Report on the 2005 survey of the TübingenIranian stone age research project in the provinces of Esfahan, Fars and Kohgiluyeh-Boyerahmad. Archaeological Reports 5: 9–34. Conard, N.J., S. Riehl and M. Zeidi. 2013. Revisiting neolithisation in the Zagros foothills: excavations at Chogha Golan, an aceramic Neolithic site in Ilam Province, western Iran. In The Neolithisation of Iran: The Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 76–83. Oxford: BANEA and Oxbow Books. Coon, C.S. 1951. Cave Explorations in Iran 1949. Museum Monographs, University Museum, University of Pennsylvania Press. Coon, C.S. 1952. Excavations in Hotu Cave Iran, 1951. Philadelphia: American Philosophical Society. Coningham, R.A. E., and R. Young. 2015. The Archaeology of South Asia: From the Indus to Ashoka, c.6500 BCE-200 CE. Cambridge: Cambridge University Press. Costantini, L. 1984. The beginning of agriculture in the Kacchi Plain, the Evidence of Mehrgarh. In South Asian Archaeology 1981, edited by B. Allchin, 29–33. Cambridge: Cambridge University Press. Cronin, T.M. 2010. Paleoclimates: Understanding Climate Change Past and Present. New York: Columbia University Press. Cullen, H.M., P.B. deMenocal, S. Hemming, F.H. Brown, T. Guilderson, and F. Sirocko. 2000. Climate change and the collapse of the Akkadian empire: Evidence from the deep sea. Geology 28: 379–382. doi:10.1130/0091-7613 Cunniff, J., G. Jones, M. Charles and C.P. Osborne. 2017. Yield responses of wild C3 and C4 crop progenitors to subambient CO2: A test for the role of CO2 limitation in the origin of agriculture. Global Change Biology 23: 380–393. Cuttler, R. 2013. Considering marine transgression as a mechanism for enforced migration and the littoral Gulf Ubaid phenomenon. Arabian Archaeology and Epigraphy 24: 37–43. d’Alpoim Guedes, J.A., S.A. Crabtree, R.K. Bocinsky, and T.A. Kohler. 2016. Twenty-first century approaches to ancient problems: Climate and society. PNAS 113/51: 14483–14491. Dahl, J., C.A. Petrie, and D.T. Potts. 2013. Chronological parameters of the earliest writing system in Iran. In Ancient Iran and its Neighbours: Local Developments and Long-Range Interactions in the Fourth Millennium BC, British Institute of Persian Studies Archaeological Monographs Series III, edited by C.A. Petrie, 353–378. Oxford: Oxbow Books. Darabi, H., H. Fazeli Nashli, R. Naseri, S. Riehl and R. Young. 2013. The neolithisation process in the Seimarreh Valley: Excavations at east Chia Sabz, Central Zagros. In The Neolithisation of Iran: The Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 55–74. Oxford: BANEA and Oxbow Books. Daujat, J., M. Mashkour, A. Emery-Barbier, R. Neef, and R. Bernbeck. 2016. Qale Rostam: Reconsidering the rise of a highland way of life: An integrated bioarchaeological approach. In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 107–136. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. Dean, J.R., M.D. Jones, M.J. Leng, S.R. Noble, S.E. Metcalfe, H.J. Sloane, D. Sahy, W.J. Eastwood, C.N. Roberts. 2015. Eastern Mediterranean hydroclimate over the late glacial and Holocene, reconstructed from the sediments of Nar lake, central Turkey, using stable isotopes and carbonate mineralogy. Quaternary Science Reviews 124: 162–174.

The Iranian Plateau and the Indus River Basin

315

Deotare, B.C., M.D. Kajale, S.N. Rajaguru, S. Kusumgar, A.J. T. Jull, and J.D. Donahue. 2004. Palaeoenvironmental history of Bap-Malar and Kanod playas of western Rajasthan, Thar desert. Journal of Earth System Science 113: 403–425. de Cardi, B. 1970. Excavations at Bampur, a third millennium settlement in Persian Baluchistan, 1966. Anthropological papers of the AMNH 51.3. New York: American Museum of Natural History. Deschayes, J. 1970. Tureng Tépé. Iran 6: 165–166. Deschayes, J. 1975. Les fouilles récent de Tureng Tépé: La terrasse haute de la fin du IIIe millénnaire. Academic des Inscriptions et Belles-Lettres, Comptes Rendus 1975: 522–530. Dimri, A. P., and A. Chevuturi. 2016. Western Disturbances—An Indian Meteorological Perspective. Springer Switzerland. Dixit, Y., D.A. Hodell, A. Giesche, S.K. Tandon, S.F. Gázquez, H.S. Saini, L. Skinner, et al. 2018. Intensified Indian summer monsoon and the urbanization of the Indus Civilization in northwest India. Scientific Reports 8:4225. doi:10.1038/s41598-018-22504-5. Dixit, Y., D.A. Hodell, and C.A. Petrie. 2014b. Abrupt weakening of the summer monsoon in northwest India ~4100 year ago. Geology 42: 339–342. doi:10.1130/G35236.1. Dixit, Y., D.A. Hodell, R. Sinha, and C.A. Petrie. 2014a. Abrupt weakening of the Indian summer monsoon at 8.2 kyr B.P. EPSL 391: 16–23. doi:10.1016/j.epsl.2014.01.026. Djamali, M., J.-L. de Beaulieu, V., andrieu-Ponel, M. Berberian, N.F. Miller, E. Gandouin, H. Lahijani, et al. 2009. A late Holocene pollen record from Lake Almalou in NW Iran: Evidence for changing land-use in relation to some historical events during the last 3700 years. Journal of Archaeological Science 36: 1364–1375. doi:10.1016/j.jas.2009.01.022 Djamali, M., J.-L. de Beaulieu, N.F. Miller, V., andrieu-Ponel P. Ponel, R. Lak, N. Sadeddin, et al. 2009. Vegetation history of the SE section of the Zagros Mountains during the last five millennia: A pollen record from the Maharlou Lake, Fars Province, Iran. Vegetation, History and Archaeobotany 18: 123–136. Dobney, K., and D. Jaques 2010. The vertebrate assemblage from excavations at Jeitun, 1993 and 1994. In Origins of Agriculture in Western Central Asia: An Environmental-Archaeological Study, edited by D.R. Harris, 174–179. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Domroes, M., M. Kaviani, and D. Schaefer. 1998. An analysis of regional and intra-annual preicpitation variability over Iran using multivariate statistical methods. Theoretical and Applied Climatology 61: 151–159. Dutt, S., A.K. Gupta, B. Wünnemann, and D. Yan. 2018. A long arid interlude in the Indian summer monsoon during δ4,350 to 3,450 cal. yr BP contemporaneous to displacement of the Indus valley civilization. Quaternary International, online first 17 April 2018. DOI:10.1016/j.quaint.2018.04.005 Dyson Jr., R. H., and S.M. Howard, eds. 1989. Tappeh Hesar: Reports of the Restudy Project, 1976. Florence: Case Editrice le Lettere. Eastwood, W.J., M.J. Leng, N. Roberts and B. Davis. 2007. Holocene climate change in the eastern Mediterranean region: a comparison of stable isotope and pollen data from a lake record in southwest Turkey. Journal of Quaternary Science 22(4): 327–341. Englund, R.K. 1998. Texts from the Late Uruk Period. In Späturuk–Zeit und frühdynastische Zeit, OBO 160/1, edited by P. Attinger and M. Wäfler, 15–233. Göttingen. Enzel, Y., L.L. Ely, S. Mishra, R. Ramesh, R. Amit, B. Lazar, S.N. Rajguru, V.R. Baker, and A. Sandler. 1999. High-resolution Holocene environmental changes in the Thar Desert, Northwestern India. Science 284: 125–128. Eskandari, N., A. Aberdi, M. Shafie, and M. Javadi. 2014. Keshit: an early Bronze Age urban centre on the western edge of the Lut Desert, south-eastern Iran. Antiquity 341/88 Project Gallery. Fisher, W.B. 1968. Physical geography. In The Cambridge History of Iran, Volume 1: The Land of Iran, edited by W.B. Fisher, 3–110. Cambridge: Cambridge University Press. Fleitmann, D., S.J. Burns, A. Mangini, M. Mudelsee, J. Kramers, I. Villa, U. Neff, et al. 2007. Holocene ITCZ and Indian monsoon dynamics recorded in stalagmites from Oman and Yemen (Socotra). Quaternary Science Reviews 26: 170–188. Fleitmann, D., S.J. Burns, M. Mudelsee, U. Neff, J. Kramers, A. Mangini, and A. Matter. 2003. Holocene forcing of the Indian monsoon recorded in a stalagmite from Southern Oman. Science 300: 1737–1739. Fleitmann, D., M. Mudelsee, S.J. Burns, R.S. Bradley, J. Kramers, and A. Matter. 2008. Evidence for a widespread climatic anomaly at around 9.2 ka before present. Paleoceanography 23. Flohr, P., D. Fleitmann, R. Matthews, W. Matthews, and S. Black. 2016. Evidence of resilience to past climate change in Southwest Asia: Early farming communities and the 9.2 and 8.2 ka events. Quaternary Science Reviews 136: 23–39. doi:10.1016/j.quascirev.2015.06.022

316

Climate Changes in the Holocene

Fouache, E.2012 [2008]. JIROFT ii. Human Geography and Environment. Encyclopaedia Iranica XIV/6:648–653. Fouache, E., and D. Garçon. 2009. Holocene geomorphological dynamics and Bronze Age settlements along the Halil Roud River (Jiroft, SE of Iran). In Ol’ Man River: Geo-Archaeological Aspects of Rivers and River Plains, edited by M. De Dapper, 363–373.Ghent: Academia Press. Archaeological Reports Ghent University 5. Fouache, E., H.-P. Francfort, C. Cosandey, and C. Adle. 2015. La crise d’aridité climatique de la fin du 3ème millénaire av. J.-C., à la lumière des contextes géomorphologique de 3 sites d’Iran Oriental (Bam, Tepe Damghani, Jiroft). In Greater Khorasan: History, Geography, Archaeology and Material Culture,edited by R. Rante. Paris: Walter de Gruyter GmbH. Fuller, D.Q. 2006. Agricultural origins and frontiers in South Asia: A working synthesis. Journal of World Prehistory 20: 1–86. Fuller, D.Q. 2011. Finding plant domestication in the Indian subcontinent. Current Anthropology 52(suppl.4): S347–S362. Gala, M., and A. Tagliacozzo 2014. The role of birds in the settlement of Shahr-i Sokhta (Sistan, Iran) during the 3rd millennium BC. International Journal of Osteoarchaeology 24: 319–330. Ganji, M.H. 1968. Climate. In The Cambridge History of Iran, Volume 1: The Land of Iran, edited by W.B. Fisher, 212–249. Cambridge: Cambridge University Press. Garazhian, O. 2016. Stratigraphic excavation at Tal-e δtashi, Dδrestδn, Bam: living floors of a pre-pottery Neolithic site, southeastern Iran. In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 49–74. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. [In Persian] Garazhian, O., and M. Shakooie. 2013. Tell-e Atashi (Bam, southeastern Iran) and the Neolithic of the eastern Near East. In The Neolithisation of Iran: The Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 284–296. Oxford: BANEA and Oxbow Books. Ghasemi, A. R., and D. Khalili. 2008. The association between regional and global atmospheric patterns and winter precipitation in Iran. Atmospheric Research 88: 116–133. Ghirshman, R. 1938. Fouilles de Sialk, Musee du Louvre, Department des antiquites orientales, Serie Archeologique, 4, Paris, Geuthner. Giesche, A., M. Staubwasser, C.A. Petrie, and D.A. Hodell in press. Re-examining the 4.2 ka BP event in foraminifer isotope records from the Indus River delta in the Arabian Sea, Climate of the Past [preprint online first September 3, 2018; doi: https://doi.org/10.5194/cp-2018-104]. Gopnik, H., and M. Rothman, eds. 2011. On the High Road: The History of Godin Tepe, Iran, Toronto: Royal Ontario Museum/Mazda Press. Green, A. S., and C A. Petrie. 2018. Landscapes of urbanisation and de-urbanization: Integrating site location datasets from northwest India to investigate changes in the Indus Civilization’s settlement distribution. Journal of Field Archaeology. DOI:10.1080/00934690.2018.1464332. Gregoricka, L.A. 2013. Geographic origins and dietary transitions during the Bronze Age in the Oman Peninsula. American Journal of Physical Anthropology 152(3): 353–369. Gurjazkaite, K., J. Routh, M. Djamali, A. Vaezi, Y. Poher, A. Naderi Beni, V. Tavakoli, and H. Kylin. 2018. Vegetation history and human-environment interactions through the late Holocene in Konar Sandal, SE Iran. Quaternary Science Reviews 194: 143–155. Hakemi, A. 1997. Shahdad: Archaeological Excavations of a Bronze Age Center in Iran. Rome: ISMEO. Hamzeh, M.A., M.H. M. Gharaie, H.A. K. Lahijani, M. Djamali, R.M. Harami, and A.N. Beni. 2016a. Holocene hydrological changes in SE Iran, a key region between Indian summer monsoon and Mediterranean winter precipitation zones, as revealed from a lacustrine sequence from Lake Hamoun. Quaternary International 408: 25–39. Hamzeh, M.A., M.H.M. Gharaie, H.A. K. Lahijani, R.M. Harami, M. Djamali, and A.N. Beni. 2016b. Paleolimnology of Lake Hamoun (E Iran): Implication for past climate changes and possible impacts on human settlements. Palaios 31: 616–629. Harlan, J. R., and D. Zohary 1966. Distribution of wild wheats and barley. Science 153 (3740): 1074–1080. Harris, D.R. 2010. Origins of Agriculture in Western Central Asia: An Environmental-Archaeological Study. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Hendy, J., F. Welker, B. Demarchi, C. Speller, C. Warinner, and M.J. Collins. 2018. A guide to ancient protein studies. Nature Ecology and Evolution. DOI:10.1038/s41559-018-0510-x Højlund, F. 2007. The Burial Mounds of Bahrain: Social Complexity in early Dilmun. Moesgaard: Jutland Archaeological Society. Hole, F. 1977. Studies in the Archaeological History of the Deh Luran Plain: the Excavation of Chagha Sefid. Memoirs of the Museum of Anthropology Number 9. Ann Arbor: University of Michigan Museum of Anthropology.

The Iranian Plateau and the Indus River Basin

317

Hole, F. 1987. Archaeology of the village period. In The Archaeology of Western Iran: Settlement and Society from Prehistory to the Islamic Conquest, edited by F. Hole, 29–78. Smithsonian Series in Archaeological Inquiry. Washington D.C.: Smithsonian Institution Press. Hole, F., ed. 1987. The Archaeology of Western Iran: Settlement and Society from Prehistory to the Islamic Conquest, Smithsonian Series in Archaeological Inquiry. Washington D.C.: Smithsonian Institution Press. Hole, F. 1996. The context of caprine domestication in the Zagros region. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, edited by D.R. Harris, 263–281. London: Institute of Archaeology, University College London. Hole, F. 2008. Palaeolithic Age in Iran. Encyclopaedia Iranica http://www.iranicaonline.org/articles/paleolithicage Hole, F., K.V. Flannery, and J.A. Neely. 1969. Prehistory and Human Ecology of the Deh Luran Plain: An Early Village Sequence from Khuzistan, Iran. Memoirs of the Museum of Anthropology Number 1. Ann Arbor: University of Michigan Museum of Anthropology. Hritz, C., and J. Pournelle. in press. Feeding history: deltaic resilience, inherited practice, and millennialscale sustainability in an urbanized landscape. In From Field to Table: Historical Ecology of Regional Subsistence Strategies, edited by D. Goldstein. Columbia: University of South Carolina Press. Jarrige, J.-F. 2000. Mehrgarh Neolithic: new excavations. In South Asian Archaeology 1997, edited by M. Taddei and G. De Marco, 259–283. Rome: Instituto italiano per l’Africa e l’Oriente. Jarrige J.-F. 2008. Mehrgarh Neolithic. Pragdhara, 18: 135–154. Jarrige, J.F. et al. 2013. Mehrgarh: Neolithic Period – Seasons 1997 – 2000. Paris: Éditions de Boccard. Johnson, G.A. 1973. Local Exchange and Early State Development in Southwestern Iran, Archaeological Papers, 51, Museum of Anthropology. Ann Arbor: University of Michigan. Johnson, G.A. 1989. Late Uruk in greater Mesopotamia: expansion or collapse? Origini 14: 595–611. Jones, P.J. 2017. Climate change, water stress and agriculture in the Indus Civilisation, 3000–1500 BC, Unpublished PhD Dissertation, University of Cambridge. Jones, M.D., N. Abu Jaber, A. Al Shdaifat, D. Baird, B. Cook, M. Cuthbert, J. Dean, et al. in press. 20,000 years of societal vulnerability and adaptation to climate change in southwest Asia. WIREs Water. Jones, M.D., M. Djamali, J. Holmes, L. Weeks, M.J. Leng, A. Lashkari, K. Alamdari, et al. 2015. Human impact on the hydroenvironment of Lake Parishan, SW Iran, through the late-Holocene. The Holocene, 25(10): 1651–1661. Joshi, J.P., M. Bala, and J. Ram. 1984. The Indus Civilisation: a reconsideration on the basis of distribution maps. In Frontiers of the Indus Civilisation: Sir Mortimer Wheeler Commemoration Volume, edited by B.B. Lal and P. Swaraj, 511–530. Delhi: Books & Books. Joshi, L.M., B.S. Kotlia, S.M. Ahmad, C.-C. Wu, J. Sanwal, W. Raza, A.K. Singh, et al. 2017. Reconstruction of Indian monsoon precipitation variability between 4.0 and 1.6 ka BP using speleothem δ18O records from the Central Lesser Himalaya, India. Arabian Journal of Geoscience 10.356: 1–16. Kenoyer, J. M. 1997. Early city-states in South Asia: comparing the Harappan phase and early historic period. In The Archaeology of City-States: Cross-Cultural Approaches, edited by D.L. Nichols and H.C. Thomas, 51–70. Washington D.C.: Smithsonian. Kenoyer, J.M., D.T. Price, and J.H. Burton. 2013. A new approach to tracking connections between the Indus valley and mesopotamia: Initial results of strontium isotope analyses from Harappa and Ur. Journal of Archaeological Science 40: 2286–2297. Kingwell-Banham, E., C.A. Petrie, and D.Q. Fuller. 2015. Early agriculture in South Asia. In Cambridge World History, Volume 2: A World with Agriculture, 12,000 BCE–500 CE, edited by G. Barker and C. Goucher, 261–288. Cambridge: Cambridge University Press. Kotlia, B.S., A.K. Singh, L.M. Joshi, and K. Bisht. 2017. Precipitation variability over Northwest Himalaya from δ4.0 to 1.9 ka BP with likely impact on civilization in the foreland areas. Journal of Asian Earth Sciences online first Nov 14, 2017. DOI:10.1016/j.jseaes.2017.11.025 Krishna Kumar, K., B. Rajagopalan, M. Hoerling, G. Bates, and M. Cane. 2006. Unravelling the mystery of Indian monsoon failure during El Niño. Science 314.5796: 115–119. DOI:10.1126/science.1131152. Kumar, M. 2009. Harappan settlements in the Ghaggar-Yamuna divide. Linguistics. Archaeology and the Human Past 7: 1–75. Kuzucuoδlu, C. 2009. Climate and environment in times of cultural changes from the 4th to the 1st mill. BC in the near and middle East. Scienze dell’Antichità, 15(15): 193–216. Kuzucuoδlu, C., W. Dörfler, S. Kunesch, and F. Goupille. 2011. Mid- to late-holocene climate change in central Turkey: The Tecer Lake record. The Holocene 21.1: 173–188. Lal, B.B. 1997. The Earliest Civilisation of South Asia. Delhi: Aryan Books. Lamb, H.H. 1982 [2005]. Climate, History and the Modern World. London and New York: Routledge.

318

Climate Changes in the Holocene

Lambeck, K. 1996. Shoreline reconstructions for the Persian Gulf since the last glacial maximum. Earth and Planetary Science Letters 142: 43–57. Lamberg-Karlovsky, C. C., and M. Tosi. 1973. Shahr-i Sokhta and Tepe Yahya: Tracks on the earliest history of the Iranian Plateau. East and West 23: 21–57. Larsen, C.E. 1983. Life and Land Use on the Bahrain Islands: the Geoarchaeology of an Ancient Society. Chicago: University of Chicago. Laursen, S.T. 2008. Early Dilmun and its rulers: New evidence of the burial mounds of the elite and the development of social complexity, c. 2200–1750 BC. Arabian Archaeology and Epigraphy 19: 156–167. Laursen, S.T. 2009. The decline of Magan and the rise of Dilmun: Umm an-Nar ceramics from the burial mounds of Bahrain, c.2250–2000 BC. Arabian Archaeology and Epigraphy 20: 134–155. Laursen, S.T. 2017. The Royal Mounds of A’ali in Bahrain: The Emergence of Kingship in Early Dilmun. Aarhus: Aarhus University Press. Laursen, S. T., and P. Steinkeller 2017. Babylonia, the Gulf Region, and the Indus: Archaeological and Textual Evidence for Contact in the Third and Early Second Millennia B.C. Winona Lake: Eisenbrauns. Law, R.W. 2011. Inter-regional interaction and urbanism in the ancient Indus valley: A geological provenience study of Harappa’s rock and mineral assemblage. Linguistics, Archaeology and the Human Past, Occasional Paper 11:1–800. Lawler, A. 2011. The world in between. Archaeology 64/6 November/December. https://archive.archaeology. org/1111/features/dasht-e_lut_iran_shahr-i-sokta-shahad_tepe_yaha.html Leipe, C., D. Demske, P.E. Tarasov, and HIMPAC Project Members, 2014. A holocene pollen record from the northwestern Himalayan lake Tso Moriri: Implications for palaeoclimatic and archaeological research. Quaternary International 348: 93–112. Lézine A.M., C. Robert, S. Cleuziou, M.L. Inizan, F. Braemer, J.F. Saliège, F. Sylvestre, et al. 2010. Climate change and human occupation in the Southern Arabian lowlands during the last deglaciation and the Holocene. Global and Planetary Change 72: 412–428. Lézine, A.M., S.J. Ivory, P. Braconnot, and O. Marti. 2017. Timing of the southward retreat of the ITCZ at the end of the holocene humid period in Southern Arabia: Data-model comparison. Quaternary Science Reviews 164: 68–76. Litt, T., C. Ohlwein, F.H. Neumann, A. Hense, and M. Stein. 2012. Holocene climate variability in the Levant from the dead sea pollen record. Quaternary Science Reviews 49: 95–105. doi:10.1016/j. quascirev.2012.06.012 Madella, M., and D.Q. Fuller. 2006. Palaeoecology and the Harappan civilisation of South Asia: A reconsideration. Quaternary Science Review 25: 1283–1301. Madjidzadeh, Y. 2003. Jiroft. The Earliest Oriental Civilization. Tehran: Research Center, Cultural Heritage Organization. Madjidzadeh, Y. 2008. Excavations at Konar Sandal in the region of Jiroft in the Halil Basin: First preliminary report (2002–2008). Iran 46: 69–103. Magee, P. 2013. Iron age southeastern Iran. In The Oxford Handbook of Ancient Iran, edited by D.T. Potts, 493–499. Oxford: OUP. Magee, P. 2014. The Archaeology of Prehistoric Arabia: Adaptation and Social Formation from the Neolithic to the Iron Age. Cambridge: Cambridge University Press. Maher, L.A., E.B. Banning, and M. Chazan. 2011. Oasis or Mirage? Assessing the role of abrupt climate change in the prehistory of the Southern Levant. Cambridge Archaeological Journal 21: 1–29. Marciniak, A., and L. Czerniak. 2007. Social transformations in the late neolithic and the early chalcolithic periods in Central Anatolia. Anatolian Studies 57: 115–130. Marshall, J. 1931. Mohenjo-Daro and the Indus Civilisation (3 volumes). London: Arthur Probsthain. Mashkour, M., F. Azadeh Mohaseb, and H. Rezvani. 2016. Faunal remains from Deh Kheir and Kalδteh Khδn: Two Neolithic sites in the Shahroud and Bastam Plains (northeast Iran). In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 85–100. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. Mashkour, M., M. Tengberg, Z. Shirazi, and Y. Madjidzadeh. 2013. Bio-archaeological studies at Konar Sandal, Halil Rud basin, southeastern Iran. Journal of Environmental Archaeology 18: 222–246. Matthews, W. 2013. Contexts of Neolithic interaction: geography, palaeoclimate, and palaeoenvironment of the Central Zagros. In The Earliest Neolithic of Iran: 2008 Excavations at Sheikh-e Abad and Jani, edited by R. Matthews, W. Matthews and Y. Mohammadifar, 13–19. Oxford: Oxbow and BIPS. Matthews, R., W. Matthews, and Y. Mohammadifar. 2013. Networks of neolithic engagement. In The Earliest Neolithic of Iran: 2008 Excavations at Sheikh-e Abad and Jani, edited by R. Matthews, W. Matthews and Y. Mohammadifar, 225–233. Oxford: Oxbow and BIPS.

The Iranian Plateau and the Indus River Basin

319

Matthews, R., Y. Mohammadifar, W. Matthews, and A. Motarjem. 2010. Investigating the early neolithic of western Iran: The Central Zagros Archaeological Project (CZAP). Antiquity 84/323 Project Gallery. Meadow, R.H. 1981. Early animal domestication in South Asia: a first report of the faunal remains from Mehrgarh, Pakistan. In South Asian Archaeology 1979, edited by H. Härtel, 143–179. Berlin: Dietrich Reimer Verlag. Meadow, R.H. 1986. Faunal exploitation in the greater Indus Valley: a review of recent work to 1980. In Studies in the Archaeology of India and Pakistan, edited by J. Jacobsen, 46–64. Delhi: Oxford & IBH Publishing/The American Institute of Indian Studies. Meadow, R.H. 1996. The origins and spread of agriculture and pastoralism in northwest South Asia. In The Origins and Spread of Agriculture and Pastoralism in Eurasia, edited by D.R. Harris, 390–412. London: UCL Press. Meadow, R. H., and J. Mark Kenoyer. 1997. Excavations at Harappa 1994–1995: New perspectives on the Indus script, craft activities, and city organization. In South Asian Archaeology 1995, edited by F.R. Allchin and B. Allchin, 139–172. New Delhi: Oxford & IBH. Middleton, G.D. 2017. Understanding Collapse: Ancient History and Modern Myths, Cambridge: Cambridge University Press. Milanesi, C., M. Cresti, L. Costantini, M. Gallo, G. Gallo, S. Crognale, C. Faleri, et al. 2015. Spoilage of oat bran by sporogenic microorganisms revived from soil buried 4000 years ago in Iranian archaeological site. International Biodeterioration and Biodegradation 104:83–91. Miller, N. 1982. Economy and Environment of Malyan, a Third Millennium B.C. Urban Center in Southern Iran. PhD Dissertation, Department of Anthropology, University of Michigan, Ann Arbor. Miller, N. 1985. Palaeoethnobotanical evidence for deforestation in ancient Iran: A case study of urban Malyan. Journal of Ethnobiology 5: 1–19. Miller, N.F. 1991. Archaeobotanical Perspectives on the Rural-Urban Connection. In Economy and Settlement in the Near East: Analyses of Ancient Sites and Materials, edited by N.F. Miller, 7: 79–83. Pennsylvania: MASCA, The University Museum of Archaeology and Anthropology, University of Pennsylvania. Miller, N. 2003. Archaeobotany in Iran, Past and Future. In Yeki bud, yeki nabud: Essays on the Archaeology of Iran in Honor of William M. Sumner, edited by N. F. Miller and K. Abdi, 48: 9–15. Los Angeles: Cotsen Institute of Archaeology, University of California. Miller, H.M-L. 2006. Water supply, labor requirements, and land ownership in Indus floodplain agricultural systems. In Agricultural Strategies, edited by J. Marcus and C. Stannish, 92–128, Los Angeles: Cotsen Institute. Miller, H.M-L. 2015. Surplus in the Indus Civilisation, agricultural choices, social relations, political effects. In Surplus: The Politics of Production and the Strategies of Everyday Life, edited by C.T. Morehart and K. De Lucia, 97–119. Boulder: University Press of Colorado. Miller, N., K.M. Moore, and K. Ryan, eds. 2011. Sustainable Lifeways: Cultural Persistence in an EverChanging Environment. Philadelphia: University of Pennsylvania Museum of Archaeology and Anthropology. Miller, N.F., and W.M. Sumner, 2004. The Banesh-Kaftari interface: The view from operation H5, Malyan. Iran XVII: 77–89. Mishra, P.K., A. Anoop, G. Schettler, S. Prasad, A. Jehangi, P. Menzel, R. Naumann, et al. 2015. Reconstructed late Quaternary hydrological changes from Lake Tso Moriri, NW Himalaya. Quaternary International 371: 76–86. Modarres, R., and A. Sarhadi, 2009. Rainfall trends analysis of Iran in the last half of the twentieth century. Journal of Geophysical Research: Atmospheres 114(3). doi: IDD03101, 2009. Moghaddam, A., and N. Miri. 2003. Archaeological surveys in the ‘Eastern Corridor’, south-western Iran. Iran XLV: 23–55. Moghaddam, A., and N. Miri. 2007. Archaeological research in the Mianab Plain of lowland Susiana, southwestern Iran. Iran XLI: 99–137. Moradi, H., H. Sarhaddi-Dadian, N. Hassan Shuhami, and N. Abdul Rahman. 2014. Development and decline of the Bampur Valley, based on the new archaeological evidence from prehistoric period. Iranian Studies 47:263–287. doi:10.1080/00210862.2013.860332 Mortensen, P. 1972. Seasonal camps and early villages in the Zagros. In Man, Settlement and Urbanism, edited by P.J. Ucko, R. Tringham, and G.W. Dimbleby, 293–297. London: Duckworth. Mughal, M.R. 1997. Ancient Cholistan: Archaeology and Architecture. Ferozsons: Lahore. Mutin, B. 2015. Chrono-cultural sequence of the Bampur Valley (southeastern Iran) during the Chalcolithic Period: a re-evaluation of the Stein Collection at the Peabody Museum, Harvard University. Iran 53: 1–28.

320

Climate Changes in the Holocene

Mutin, B., and C.C. Lamberg-Karlovsky. 2014. The Proto-Elamite Settlement and its Neighbors: Tepe Yahya Period IVC. American School of Prehistoric Research (ASPR) Monograph Series, Harvard University and Oxbow Books, Oxford, UK and Oakville, CT, USA. Nakamura, A., Y. Yokoyama, H. Maemoku, H. Yagi, M. Okamura, H. Matsuoka, N. Miyake, et al. 2016. Weak monsoon event at 4.2 ka recorded in sediment from Lake Rara, Himalayas. Quaternary International 397: 349–359. doi:10.1016/j.quaint.2015.05.053. Narasimhan, V.M., N. Patterson, P. Moorjani, I. Lazaridis, M. Lipson, S. Mallick, N. Rohland, et al. in press. The genomic formation of south and central Asia. bioRχiv pre-print server, doi:10.1101/292581. Nishiaki, Y. 2010. A radiocarbon chronology for the Neolithic settlement of Tall-i Mushki, Marv Dasht plain, Fars, Iran. Iran 48: 1–10. Nissen, H.J. 2001. Cultural and political networks in the ancient Near East during the fourth and third millennia B.C. In Uruk Mesopotamia and Its Neighbours: Cross-Cultural Interactions in the Era of State Formation, edited by M. Rothman, 149–179. Santa Fe: SAR Press. Orengo, H.A., and C.A. Petrie. 2017.  Large-scale, multi-temporal remote sensing of palaeo-river networks: A case study from northwest India and its implications for the Indus Civilisation. Remote Sensing 9:735. doi:10.3390/rs9070735. Orengo, H.A., and C.A. Petrie. 2018. Multi-Scale Relief Model (MSRM): A new algorithm for the analysis of subtle topographic change in digital elevation models. Earth Surface Processes and Landforms [online first 14 December 2017. doi:10.1002/esp.4317]. Parker, A.G., and A.S. Goudie. 2007. Geomorphological and palaeoenvironmental investigations in the southeastern Arabian Gulf region and the implication for the archaeology of the region. Geomorphology 101:458–470. Parker, A.G., A.S. Goudie, S. Stokes, K. White, M.J. Hodson, M. Manning, and D. Kennet. 2006. A record of Holocene climate change from lake geochemical analyses in southeastern Arabia. Quaternary Research 66(3): 465–476. Parker, A.G., G.W. Preston, A. Parton, H. Walkington, P.E. Jardine, M.J. Leng, and M.J. Hodson. 2016. Low-latitude Holocene hydroclimate derived from lake sediment flux and geochemistry. Journal of Quaternary Science 31(4): 286–299. doi:10.1002/jqs.2859 Perrot, J. 2012 [2008]. Jiroft iv. Iconography of Chlorite Artifacts. Encyclopaedia Iranica XIV/6:656–664. Petrie, C.A. 2011. ‘Culture’, innovation and interaction across southern Iran from the Neolithic to the Bronze Age (6500–3000 BC). In Investigating Archaeological Cultures: Material Culture, Variability and Transmission, edited by B. Roberts and M. Vander Linden, 151–182. Springer. Petrie C.A. 2012. Ceramic production. In A Companion to the Archaeology of the Ancient Near East, edited by D.T. Potts, 279–294. Oxford: Wiley-Blackwells. Petrie, C.A. 2013a. Ancient Iran and Its Neighbours: Emerging paradigms and future directions. In Ancient Iran and its Neighbours: Local Developments and Long-Range Interactions in the Fourth Millennium BC, edited by C.A. Petrie, 385–410. British Institute of Persian Studies Archaeological Monographs Series III, Oxford: Oxbow Books. Petrie, C.A. 2013b. Ancient Iran and its neighbours: State of play. In Ancient Iran and Its Neighbours: Local Developments and Long-Range Interactions in the Fourth Millennium BC, edited by C.A. Petrie, 1–24. British Institute of Persian Studies Archaeological Monographs Series III. Oxford: Oxbow Books. Petrie C.A. 2013c. South Asia. In The Oxford Handbook of Cities in World History (hardback), edited by P. Clark, 83–104. Oxford: OUP. Petrie, C.A., ed. 2013. Ancient Iran and Its Neighbours: Local Developments and Long-range Interactions in the Fourth Millennium BC. British Institute of Persian Studies Archaeological Monographs Series III. Oxford: Oxbow Books. Petrie, C.A. 2014. Iran and Uruk Mesopotamia: Chronologies and connections in the 4th millennium BC. In Preludes to Urbanism: Studies in the Late Chalcolithic of Mesopotamia in Honour of Joan Oates, edited by A. McMahon and H. Crawford, 137–155. Cambridge: McDonald Institute Monographs. Petrie, C.A. 2015. Case study: Mehrgarh. In Cambridge World History, Volume 2: A World with Agriculture, 12,000 BCE-500 CE, edited by G. Barker and C. Goucher, 289–309. Cambridge: CUP. Petrie, C.A. 2017. Crisis, what crisis? Adaptation, resilience and transformation in the Indus Civilisation. In Crisis to Collapse: The Archaeology of Social Breakdown, edited by T. Cunningham and J. Driessen, 43–64. UC Louvain: Aegis Publications. Petrie, C.A., A. Asgari Chaverdi, and M. Seyedin. 2005. From Anshan to Dilmun and Magan: The spatial and temporal distribution of Kaftari and Kaftari-related ceramic vessels. Iran 43: 49–86. Petrie, C.A., and J. Bates. 2017. ‘Multi-cropping’, intercropping and adaptation to variable environments in the Indus Civilisation. Journal of World Prehistory 30: 81–130. doi:10.1007/s10963–017-9101-z.

The Iranian Plateau and the Indus River Basin

321

Petrie, C.A., J. Bates, T. Higham, and R.N. Singh. 2016. Feeding ancient cities in South Asia: dating the adoption of rice, millet and tropical pulses in the Indus Civilisation. Antiquity 90.354: 1489–1504. doi:10.15184/aqy.2016.210. Petrie, C.A., M. Djamali, and M.D. Jones. 2018. Physical geography and environment of elam. In The Elamite World, edited by J. Álvarez-Mon, G.P. Basello, and Y. Wicks, 99–117. Oxford: Routledge Worlds Series. Petrie, C.A., F. Khan, J.R. Knox, K.D. Thomas, and J.C. Morris. 2010. The investigation of early villages in the hills and on the plains of western South Asia. In Sheri Khan Tarakai and Early Village Life in the Borderlands of North-West Pakistan, edited by C.A. Petrie, 7–28. Bannu Archaeological Project Monographs—Volume 1. Oxford: Oxbow Books. Petrie, C.A., D. Parikh, A.S. Green, and J. Bates, in press. Looking beneath the veneer: Thoughts about environmental and cultural diversity in the Indus Civilisation. In Walking with the Unicorn. Social Organization and Material Culture in Ancient South Asia, edited by D. Frenez, G. Jamison, R. Law, M. Vidale, and R.H. Meadow. Oxford: Archaeopress. Petrie, C.A., and J.G. Shaffer, in press. The development of stratified societies south of the Hindu Kush. In The Archaeology of Afghanistan, edited by W. Ball, and N. Hammond. Edinburgh: Edinburgh University Press. Petrie, C.A., R.N. Singh, J. Bates, Y. Dixit, C.A.I. French, D. Hodell, P.J. Jones, et al. 2017. Adaptation to variable environments, resilience to climate change: investigating Land, Water and Settlement in northwest India. Current Anthropology 58.1: 1–30. doi:10.1086/690112. Petrie, C.A., and K.D. Thomas. 2012. The topographic and environmental context of the earliest village sites in western South Asia. Antiquity 86.334: 1055–1067. Pittman, H. 2013. Eastern Iran in the Early Bronze Age. In The Oxford Handbook of Ancient Iran, edited by D.T. Potts, 304–324. Oxford: OUP. Pollock, S.M. 1999. Ancient Mesopotamia: The Eden that Never Was. Cambridge: Cambridge University Press. Possehl, G.L. 1999. Indus Age: The Beginnings. Philadelphia: University of Pennsylvania Museum. Possehl, G.L. 2002. The Indus Civilization. A Contemporary Perspective. Walnut Creek: Altamira Press. Postgate, J.N., ed. 2002. Artefacts of Complexity: Tracking the Uruk in the Ancient Near East. Cambridge: British School of Archaeology in Iraq, Cambridge University Press. Potts, D.T. 1997. Mesopotamian Civilisation: The Material Foundations., Ithaca, NY: Cornell University Press. Potts, D.T. 1999. The Archaeology of Elam: Formation and Transformation of an Ancient Iranian State. Cambridge: Cambridge World Archaeology, Cambridge University Press. Potts, D.T. 2001. Excavations at Tepe Yahya, Iran 1967–1975: The Third Millennium. Cambridge: American School of Prehistoric Research Bulletin, 45, Peabody Museum of Archaeology and Ethnology, Harvard University. Pourasghar, F., T. Tozuka, S. Jahanbakhsh, B.S. Sarraf, H. Ghaemi, and T. Yamagata. 2012. The interannual precipitation variability in the southern part of Iran as linked to large-scale climate modes. Climate Dynamics 39: 2329–2341. Pournelle, J. 2003. Marshland of Cities: Deltaic Landscapes and the Evolution of Early Mesopotamian Civilisation, Ph.D. dissertation, University of California, San Diego. Pournelle, J.R., and G. Algaze 2014. Travels in Edin: Deltaic resilience and early urbanism in greater Mesopotamia. In Preludes to Urbanism: Studies in the Late Chalcolithic of Mesopotamia in Honour of Joan Oates, edited by A. McMahon and H. Crawford, 7–34. Cambridge: McDonald Institute Monographs. Prasad, S., and Y. Enzel. 2006. Holocene paleoclimates of India. Quaternary Research 66(3): 442–453. Prasad, S., S. Kusumgar, and S.K. Gupta. 1997. A mid to late Holocene record of palaeoclimatic changes from Nal Sarovar: a palaeodesert margin lake in western India. Journal of Quaternary Science 12(2): 153–159. Preston, G.W., D.S. Thomas, A.S. Goudie, O.A. Atkinson, M.J. Leng, M.J. Hodson, H. Walkington, et al. 2015. A multi-proxy analysis of the Holocene humid phase from the United Arab Emirates and its implications for southeast Arabia’s Neolithic populations. Quaternary International 382: 277–292. doi:10.1016/j.quaint.2015.01.054. Prickett, M. 1986. Man, Land, and Water: Settlement Distribution and the Development of Irrigation Agriculture in the Upper Rud-i Gushk Drainage, Southeastern Iran. PhD dissertation, Harvard University. Pullar, J. 1990. Tepe Abdul Hosein: A Neolithic site in western Iran, excavations 1978. Oxford: BAR International Series 563.

322

Climate Changes in the Holocene

Revel, M., C. Colin, S. Bernasconi, N. Combourieu-Nebout, E. Ducassou, F.E. Grousset, Y. Rolland. 2014. 21,000 years of Ethiopian African monsoon variability recorded in sediments of the western Nile deepsea fan. Regional Environmental Change 14: 1685–1696. doi:10.1007/s10113-014-0588-x. Rezvani, H., and K. Roustaei. 2016. Preliminary report on two seasons of excavations at Tappeh Deh Kheir, Bastam Plain, northeast Iran. In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 15–52. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. Richerson P.J., R. Boyd and R.L. Bettinger. 2001. Was agriculture impossible during the Pleistocene but mandatory during the Holocene? A climate change hypothesis. American Antiquity 66: 387–411. Riehl, S. 2016. The role of the local environment in the slow pace of emerging agriculture in the Fertile Crescent. Journal of Ethnobiology 36: 512–534. doi:10.2993/0278-0771-36.3.512. Riehl, S., E. Asouti, D. Karakaya, B.M. Starkovich, M. Zeidi, and N.J. Conard. 2015. Resilience at the transition to agriculture: the long-term landscape and resource development at the aceramic Neolithic tell site of Chogha Golan (Iran). BioMed Research International. doi:10.1155/2015/532481. Rose, J.I. 2010. New light on human prehistory in the Arabo-Persian Gulf Oasis. Current Anthropology 51: 849–883. Rose, J.I., V. δerný, and R. Bayoumi. 2013. Tabula rasa or refugia? Using genetic data to assess the peopling of Arabia. Arabian Archaeology and Epigraphy 24: 95–101. Rosenberg, M. 2003. The Epipaleolithic in the Marvdasht. In Yeki bud, yeki nabud: Essays on the Archaeology of Iran in Honor of William M. Sumner, edited by N. F. Miller and K. Abdi, 98–108. Los Angeles: The Cotsen Institute of Archaeology, University of California. Rothman, M.S., ed. 2001. Uruk Mesopotamia and Its Neighbours: Cross-Cultural Interactions in the Era of State Formation. Santa Fe: SAR Press. Roustaei, K., M. Mashkour, and M. Tengberg. 2015. Tappeh Sang-e Chakhmaq and the beginning of the Neolithic in north-east Iran. Antiquity 89(345): 573–595. Roy, P.D., and A.K. Singhvi. 2016. Climate variation in the Thar Desert since the Last Glacial Maximum and evaluation of the Indian Monsoon. TIP 19(1): 32–44. Salvatore, S., and M. Vidale 1997. Shahr-i Sokhta 1975–1978: Central Quarters Excavations. Preliminary Report. IsIAO Reports and Memoirs, Series Minor I, Rome. Sandweiss, D.H., and A.R. Kelley. 2012. Archaeological contributions to climate change research: The archaeological record as a paleoclimatic and paleoenvironmental archive. Annual Review of Anthropology 41: 371–391. Shaffer, J.G. 1978. The later prehistoric periods. In The Archaeology of Afghanistan: From Earliest Times to the Timurid Period, edited by F.R. Allchin and N. Hammond. London: Academic Press. Shaffer, J.G. 1992. The Indus Valley, Baluchistan, and Helmand Traditions: Neolithic through Bronze Age. In Chronologies in Old World Archaeology, edited by R.W. Ehrich, 441–464 (Vol. I) and 425–446 (Vol. II). Chicago: University of Chicago Press. Sharifi, A., A. Pourmand, E.A. Canuel, E. Ferer-Tyler, L.C. Peterson, B. Aichner, S.J. Feakins, et al. 2015. Abrupt climate variability since the last deglaciation based on a high-resolution, multi-proxy peat record from NW Iran: The hand that rocked the Cradle of Civilization? Quaternary Science Reviews, 123, 215–230. doi:10.1016/j.quascirev.2015.07.006. Singh, R.N., C.A. Petrie, C.A.I. French, A.S. Goudie, S. Gupta, R. Tewari, A.K. Singh, et al. 2008. Settlements in Context: Reconnaissance in western Uttar Pradesh and Haryana, April and May 2008. Man and Environment 33(2): 71–87. Singh, G., R.J. Wasson, and D.P. Agrawal. 1990. Vegetational and seasonal climatic changes since the last full glacial in the Thar Desert, north-west India. Revue of Palaeobotany and Palynology 64: 351–358. Sinha, R., W. Smykatz-Kloss, D. Stüben, S.P. Harrison, Z. Berner, and U. Kramarb. 2006. Late Quaternary palaeoclimatic reconstruction from the lacustrine sediments of the Sambhar playa core, Thar Desert margin, India. Palaeogeography, Palaeoclimatology, Palaeoecology 233: 252–270. Smith, P.E.L. 1986. Palaeolithic Archaeology in Iran. The American Institute of Iranian Studies Monographs. Philadelphia: University Museum, University of Pennsylvania. Smith, P.E.L. 1990. Architectural innovation and experimentation at Ganj Dareh, Iran. World Archaeology 21: 323–335. Staubwasser, M., F. Sirocko, P.M. Grootes, and H. Erlenkeuser. 2002. South Asian monsoon climate change and radiocarbon in the Arabian Sea during early and middle Holocene. Palaeoceanography 17(4): 15/1–12. Staubwasser, M., F. Sirocko, P.M. Grootes, and M. Segl. 2003. Climate change at the 4.2 ka BP termination of the Indus Valley Civilization and Holocene South Asian monsoon variability. Geophysical Research Letters 30(8): 1425.

The Iranian Plateau and the Indus River Basin

323

Staubwasser, M., and H. Weiss. 2006. Holocene climate and cultural evolution in late prehistoric-early historic West Asia-introduction. Quaternary Research 66: 372–387. Stein, G.R. 1999. Rethinking World-Systems: Diasporas, Colonies, and Interaction in Uruk Mesopotamia. Tuscon: University of Arizona Press. Steinkeller, P. 2012. New light on Marhaši and its contact with Makkan and Babylonia. In Aux marges de l’archéologie: Hommage à Serge Cleziou, edited by J. Girard and G. Gernez, 261–274. Travaux de la Maison Rene-Ginouves 16. Paris: de Boccard. Steinkeller, P. 2016. The role of Iran in the inter-regional exchange of metals: tin, copper, silver and gold in the second half of the third millennium BC. In Ancient Iran: New Perspectives from Archaeology and Cuneiform Studies: Proceedings of the Internatioanl Colloquium Held at the Center for Eurasian Cultural Studies, Kyoto University, December 6–7, 2014, edited by K. Maekawa. Ancient Text Studies in the National Museum, vol. 2: 127–150, Kyoto. Stevens, L.R., E. Ito, A. Schwalb, and H.E. Wright Jr. 2006. Timing of atmospheric precipitation in the Zagros Mountains inferred from a multi-proxy record from Lake Mirabad, Iran. Quaternary Research 66: 494–500. Stevens, L.R., H.E. Wright, and E. Ito. 2001. Proposed changes in seasonality of climate during the Lateglacial and Holocene at Lake Zeribar, Iran. The Holocene 11(6): 747–755. Sumner, W.M. 1989. Anshan in the Kaftari Phase: Patterns of Settlement and Land Use. In Archaeologia Iranica et Orientalis: Miscellanea in Honorem Louis Vanden Berghe, edited by L. de Meyer and E. Haerinck, 135–161. Gent. Sumner, W.M. 1990. Full-coverage regional archaeological survey in the near east: An example from Iran. In The Archaeology of Regions: A Case for Full-Coverage Survey, edited by S.K. Fish and S.A. Kowalewski, 87–115. Washington D.C.: Smithsonian Institution Press. Tengberg, M. 2012. Beginnings and early history of date palm garden cultivation in the Middle East. Journal of Arid Environments 86: 139–47. Tengberg, M., and B. David. 2016. First results of the archaeobotanical analysis at Kalδteh Khδn in the Shahroud Plain, NE Iran. In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 169–188. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. Thomalsky, J. 2016. The development of lithic industries in Iran in the light of the processes of neolithisation. In The Neolithic of the Iranian Plateau: Recent Research, edited by K. Roustaei and M. Mashkour, 169–188. Berlin: ex oriente. Studies in Early Near Eastern Production, Subsistence, and Environment 18. Thomas, K.D and C. Cartwright. 2010. The biological remains from Sheri Khan Tarakai. In Sheri Khan Tarakai and Early Village Life in the Borderlands of North-West Pakistan, edited by C.A. Petrie. Bannu Archaeological Project Monographs—Volume 1, 305–342. Oxford: Oxbow Books. Thornton, C.P. 2013. Tappeh Sang-e Chakhmaq. In The Neolithisation of Iran: the Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 241–255. Oxford: BANEA and Oxbow Books. Tosi, M., ed. 1983. Prehistoric Sistan 1. Rome: IsMEO (IsMEO Reports and Memoirs XIX.1). Tsuneki, A., and M. Zeidi. 2008. Tang-e Bolaghi: The Iran-Japan Archaeological Project for the Sivand Dam Salvage Area. AL-SHARK 3 (University of Tsukuba, Studies for West Asian Archaeology).Tsukuba, Japan: University of Tsukuba. Tsuneki, A., M. Zeidi, and K. Ohnuma. 2007. Proto-Neolithic caves in the Bolaghi Valley, south Iran. Iran 45: 1–22. Uerpmann, H.-P. 1987. The Ancient Distribution of Ungulate Mammals in the Middle East. Beihefter zum Tübinger Atlas des Vorderen Orients. Wiesbaden: Dr. L. Reichert. Vandiver, P.B. 1995. The production technology of early pottery from Mehrgarh. In Mehrgarh: Field Reports 1974–1985 from Neolithic Times to the Indus Civilization, edited by C. Jarrige et al., 648–661. Karachi: Department of Culture and Tourism, Government of Sindh, Pakistan, in collaboration with the French Ministry of Foreign Affairs. Van Zeist, W., P.E.L Smith, R.M. Palfenier-Wegter, M. Suwijn, and W.A. Casparie. 1984. An archaeobotanical study of Ganj Dareh Tepe, Iran. Palaeohistoria 26: 201–224. Vatandoust, A., H. Parzinger, and B. Helwing, eds. 2011. Early mining and metallurgy on the Central Iranian Plateau. Report on the first five years of research of the Joint Iranian-German Research Project, Archäologie in Iran und Turan 9, Philip von Zabern, Mainz-am-Rhein. Velde, C. 2009. The Landscape of the Middle Bronze Age in the UAE—Where did people live? In Proceedings of the International History Conference on New Perspectives on Recording UAE History, 61–74. Abu Dhabi: National Center for Documentation and Research.

324

Climate Changes in the Holocene

Valentine, B., G.D. Kamenov, J.M. Kenoyer, V. Shinde, V. Mushrif-Tripathy, E. Otarola-Castillo, and J. Krigbaum. 2015. Evidence for Patterns of Selective Urban Migration in the Greater Indus Valley (2600–1900 BC): A Lead and Strontium Isotope Mortuary Analysis. PlosONE. doi:10.1371/journal. pone.0123103 Voigt, M., and R.H. Dyson Jr., 1992. Chronology of Iran, ca. 8000–2000 B.C. In Chronologies of Old World Archaeology, edited by Ehrich, R.W., Vols. I–II: 122–178; 125–153, Chicago: Chicago University Press. Walker, R.T., and M. Fattahi 2011. A framework of Holocene and Late Pleistocene environmental change in eastern Iran inferred from the dating of periods of alluvial fan abandonment, river terracing, and lake deposition. Quaternary Science Reviews 30: 1256–1271. Doi: 10.1016/j.quascirev.2011.03.004 Wanner, H.O. Solomina, M. Grosjean, S.P. Ritz, M. Jetel. 2011. Structure and origin of Holocene cold events. Quaternary Science Reviews 30: 3109–3123. Warinner C. 2016. Dental calculus and the evolution of the human oral microbiome. California Dental Association Journal 44(7): 411–420. Wasson, R.J., J.I. Smith, and D.P. Agrawal. 1984. Late Quaternary sediments, minerals and inferred geochemical history of Didwana lake, Thar Desert, India. Palaeogeography, Palaeoclimatology, Palaeoecology 46: 345–372. Wasylikova, K., and A. Witkowski, eds. 2008. The Palaeoecology of Lake Zeribar and surrounding areas, Western Iran, during the last 48,000 years. Diatom Monographs Volume 08. Weber, S.A. 2003. Archaeobotany at Harappa: indications of change. In Indus Ethnobiology, edited by S.A. Weber, and W.R. Belcher, 175–198. Lanham: Lexington Books. Weber, S.A., T. Barela, and H. Lehman. 2010. Ecological continuity: an explanation for agricultural diversity in the Indus Civilisation and beyond. Man and Environment 35(1): 62–75. Weeks, L.R. 2003. Early Metallurgy of the Persian Gulf: Technology, Trade, and the Bronze Age World. Boston: American School of Prehistoric Research and Brill Academic Publishers. Weeks, L.R. 2013a. The development and expansion of a Neolithic way of life. In The Oxford Handbook of Ancient Iran, edited by D.T. Potts, 49–75. Oxford: OUP. Weeks, L.R. 2013b. The Neolithisation of Fars. In The Neolithisation of Iran: The Formation of New Societies, edited by R. Matthews and H. Fazeli Nashli, 97–107. Oxford: BANEA and Oxbow Books. Weeks, L., C. Cable, K. Franke, C. Newton, S. Karacic, J. Roberts, I. Stepanov, et al. 2017. Recent archaeological research at Saruq al-Hadid, Dubai, UAE. Arabian Archaeology and Epigraphy 28: 31–60. Weeks, L., C.M. Cable, K. Franke et al., in press. Saruq al-Hadid: A persistent temporary place in late prehistoric Arabia. World Archaeology 50/2. Weide, A., S. Riehl, M. Zeidi and N.J. Conard. 2017. Reconstructing subsistence practices: taphonomic constraints and the interpretation of wild plant remains at aceramic Neolithic Chogha Golan, Iran. Vegetation History and Archaeobotany 26: 487–504. Weisdorf, J.L. 2005. From foraging to farming: explaining the Neolithic revolution. Journal of Economic Surveys 19: 561–586. Weisgerber, G. 1991. Archäologisches Fundgut des 2. Jahrtausends v. Chr. in Oman. Möglichkeiten zur chronologischen Gliederung. In Golf-Archäologie: Mesopotamien, Iran, Kuwait, Bahrain, Vereinigte Arabische Emirate und Oman, edited by K. Schippmann, A. Herling and J.-F. Salles, 321–330. Internationale Archäologie 6. Weiss, H., ed. 1986. The Origins of Cities in Dry-Farming Syria and Mesopotamia in the Third Millennium B.C. Guilford: Four Quarters Publishing, 71–108. Weiss, H. 2003. Ninevite 5 periods and processes. In The Origins of North Mesopotamian Civilization. Subartu IX, 593624, Turnhout: Brepols. Weiss, H. 2015. Megadrought, collapse, and resilience in late 3rd millennium BC Mesopotamia. 2200 BC-Ein Klimasturz als Ursachefür den Zerfall der Alten Welt: 35–52. Weiss, H. 2016. Global megadrought, societal collapse and resilience at 4.2–3.9 ka BP across the Mediterranean and west Asia. PAGES 24: 62–63. Weiss, H., M.A. Courty, W. Wetterstrom, F. Guichard, L. Senior, R. Meadow, and A. Curnow. 1993. The genesis and collapse of 3rd millennium north Mesopotamian civilization. Science 261: 995–1004. Weninger, B., E. Alram-Stern, E. Bauer, L. Clare, U. Danzeglocke, O. Jöris, L. Kubatzki, et al. 2006. Climate Forcing due to the 8200 cal BP event observed at Early Neolithic sites in the Eastern Mediterranean. Quaternary Research 66: 401–420. Weninger, B., L. Clare, F. Gerritsen, B. Horejs, R. Krauss, J. Lind-städter, R. Ozbal, et al. 2014. Neolithisation of the Aegean and Southeast Europe during the 6600–6000 cal. BC period of rapid climate change. Document Praehististroy 41: 1–31. Wilkinson, T.J. 2003. Archaeological Landscapes of the Near East. Tucson: University of Arizona Press.

The Iranian Plateau and the Indus River Basin

325

Wilkinson, T.J. 2013. Hydraulic landscapes and irrigation systems of Sumer. In The Sumerian World, edited by H. Crawford, 33–54. Routledge: Abingdon. Willcox, G. 2013. The beginnings of cereal cultivation and domestication in southwest Asia. In A Companion to the Archaeology of the Ancient Near East, edited by D.T. Potts, 163–180. New York: Blackwells. Wright, H.E.J. 1993. Environmental determinism in near eastern prehistory. Current Anthropology 34(4): 458–469. Wright, R.P., R.A. Bryson, and J. Schuldenrein. 2008. Water supply and history: Harappa and the Beas regional survey. Antiquity 82: 37–48. Wright, H.T., and E.S.A. Rupley. 2001. Calibrated radiocarbon age determinations of Uruk-related assemblages. In Uruk Mesopotamia and Its Neighbours: Cross-Cultural Interactions in the Era of State Formation, edited by M.S. Rothman, 85–122. Santa Fe: SAR Press. Wright, R.P. 2010. The Ancient Indus: Urbanism, Economy and Society. Case Studies in Early Societies. Cambridge: CUP. Zeder, M.A. 1991. Feeding Cities: Specialized Animal Economy in the Ancient Near East, Smithsonian Series in Archaeological Inquiry. Washington, D.C.: Smithsonian Institution Press. Zeder, M. A., and B. Hesse. 2000. The initial domestication of goats (capra hircus) in the Zagros Mountains 10,000 years ago. Science 287: 2254–2257. Zohary, D., and M. Hopf. 2000. Domestication of Plants in the Old World. (3rd edition). Oxford: Oxford University Press.

15

Interaction of Climate, Environment and Humans in North and Central Asia during the Late Glacial and Holocene Renato Sala

CONTENTS 15.1 Introduction........................................................................................................................... 327 15.2 Modern Climate and Environment of Siberia and North Central Asia................................. 329 15.2.1 Geography................................................................................................................. 329 15.2.2 Climate...................................................................................................................... 330 15.2.3 Vegetation Cover........................................................................................................ 331 15.2.3.1 Biomes........................................................................................................ 331 15.2.3.2 Regime Shift............................................................................................... 333 15.3 Climate-Environmental Changes and Human Responses During the Last 20 ka................ 334 15.3.1 Global Considerations................................................................................................ 334 15.3.2 Evolution of Climate and Environment..................................................................... 335 15.3.2.1 General Patterns.......................................................................................... 335 15.3.2.2 LGM, Interstadial and Younger Dryas (22–10.7 ka BP)............................. 336 15.3.2.3 The Holocene (11.7–0 ka BP)...................................................................... 338 15.3.3 Cultural Evolution Through Adaptation....................................................................340 15.4 Case Studies........................................................................................................................... 342 15.4.1 The Siberian Paleolithic Refuge During the Last Glacial Period (45–12 ka BP)............................................................................................................. 342 15.4.2 Cultural Dynamics and Emergence of Early Pottery in the Lower Amur Basin (14 ka BP).............................................................................................. 345 15.4.3 The Northern Far East: Colonizers of America and Pioneers of the Circumpolar Cultural Region (14.3–2.2 ka BP).............................................. 347 15.4.4 Shifting of the East Asian Summer Monsoon and Cultural Collapse in the Cis-Baikal During the Early Atlantic Period (7.2–6 ka BP)........................... 349 15.4.5 Climate Change, Vegetation Regime Shift and Cultural Dynamics in the Forest-Steppe Belt of the Trans-Urals (7.0–3.0 ka BP).................................... 351 15.5 Conclusions............................................................................................................................ 355 References....................................................................................................................................... 356

15.1 INTRODUCTION The study of the interaction between climate changes and human cultures falls within the field of bioclimatology, an ecological science dealing with the relation between climate and distribution of living species on Earth. Climatically induced temperature (T) and precipitation (P) changes 327

328

Climate Changes in the Holocene

affect the distribution of plant and animals in particular regions (biomes), compelling human communities to adaptations of different kinds. Human cultures are an integral part of a specific biome and will respond to environmental changes in different ways, depending on their mode of sustenance. A simple behavioral model can successfully explain the long-term evolution patterns of communities of mobile foragers and stockbreeders, and even of early farmers: these categories will respond by shrinking population numbers, moving with the shifting ecological system, or adapting to the incoming one. More difficult is modeling the behavior of large, productive, stratified societies of the metal ages, economically based on immovable complex infrastructures (fields, irrigation systems, towns, industrial plants) and endowed with means of tactic resistance to climate reversals (technical innovation, commerce, governmental structures, territorial expansion). Particular social strata would introduce partial regulations delaying a sustainable response and increasing the size and complexity of the productive system. In these cases the bioclimatological model could be still valid in a condition of developing its temporal resolution and sophistication, taking into consideration the accumulation effect of tactic short-term adaptations, the establishment of unsustainable thresholds, and the possibility of abrupt socio-economic catastrophes resetting the game between man and nature. The aim of the present chapter is to use the bioclimatological approach to interpret the Holocene interactive evolution of climate, biomes, and human cultures in the northern part of the Asian continent, that is, the boreal-arctic expanses of North Asia (Siberia) and the steppe and desert of North Central Asia (Kazakhstan, Dzungaria, Mongolia), covering, all together, more than 13 + 5=18 million km2. The task is made difficult by the vastness of the territory and eased by the continental autochthonous character of its climate: the capricious stimuli from the surrounding oceans are tempered, and climate anomalies are spatially distributed as a regular series of parallel latitudinal isotherms. Due to the distance from maritime air masses, the territory has scanty precipitation: anyway moist in the north where precipitation always exceeds potential evaporation but seasonally and cyclically arid in the southernmost band. Being that the forcing of climate variables on human cultures is not direct but mediated by their effect on the environment, that is, on the botanical and animal systems sustaining human communities, an absolute priority is the analysis, together with climate, of the correspondent biotas: their modern arrangement, the main patterns of regime shifts, their historical evolution and, finally, their cultural impact. The biotic systems of Siberia, due to the high latitude, cold temperature and low evaporation rates, are very sensitive to T gradients and much less to variations of P, which instead at lower latitude is seasonally exceeded by evaporation rates and becomes the main factor of aridization and environmental change. In any case, the spatial distribution of biomes happens in the form of parallel latitudinal bands following isotherms, from north (N) to south (S): polar barren desert, tundra and cold steppe, boreal forest (coniferous taiga and mixed forest), temperate forest-steppe and steppe, semidesert and desert. Tundra and boreal forest are characteristic of Siberia, semidesert and desert of Central Asia, forest-steppe and steppe develop in between. Similarly, climate changes induce to these vegetation zones quite regular transformations, displacing them latitudinally a few degrees to the north or to the south and widening or shrinking some bands at the expense or gain of others. Concerning human adaptations, their historical patterns in Siberia are relatively simple, being that foragers always represented here the main category and in the Arctic region the only one until a couple of centuries ago. To environmental anomalies, fishermen, hunters and collectors always responded by moving together with their ecological system, unless their biome was not just shifting but contracting, in which case or they risked extinction by adhering to the fading environment or responded by adapting to the incoming one through cultural change.

Interaction of Climate, Environment and Humans in North and Central Asia

329

The bioclimatic model provides the theoretical framework of the chapter. The evolution of climate, environment, and cultural adaptation deserves the same attention: the three subjects are treated in three different sections, with the aspects of mutual interaction getting progressively focused along the text. Section 15.2 describes the modern climate and environment of Siberia and North Central Asia, and Section 15.3 describes their evolution during the last 20 ka. Section 15.4 analyzes five different cases of climate, environment, and cultural interaction, chosen because they provide the best evidence of the interactive process and high significance of the cultural outcome. They are ordered geographically from east to west and also chronologically, following the spatial and temporal progression of postglacial climate amelioration: correlation between glacial landscapes and glacial refuges of Anatomically Modern Humans (15.4.1); innovative (15.4.2) and conservative (15.4.3) responses, of late Upper Paleolithic groups to Late Glacial transformation of climate and vegetation cover; the cultural collapse in the Baikal region during the early Atlantic period, attributable to the southern redirection of the East Pacific monsoon (15.4.4); and finally, the role of the Atlantic climate in opening in the forest-steppe between the Urals and the Altai a door between the Siberian expanses and the “outer world”, which, after millennia of independent evolution, triggered cultural processes that changed the course of Eurasian history (15.4.5).

15.2 MODERN CLIMATE AND ENVIRONMENT OF SIBERIA AND NORTH CENTRAL ASIA 15.2.1 Geography Siberia extends between 60°–190°E and 50°–75°N, defined by clear geomorphological borders: from the Ural Mountains to the Pacific shore, from the Ural-Altai-Sayan-Amur regions to the Arctic seashore. The entire territory is characterized by annual and seasonal temperature values decreasing from SWW to NEE and continentality indexes growing in the same direction, mainly attributable to decreasing winter temperature. Climate-environmental considerations suggest the partition of the entire area in few zones: in three latitudinal bands – South, Middle, North (SIB-S, SIB-M, SIB-N), and three longitudinal sectors – West, East, Far East (SIB-W, SIB-E, SIB-F); that is, in nine zones (SIB-SW, SIB-MW, SIB-NW, SIB-SE, etc.). As shown in Figure 15.1, the longitudinal partition (West: 60°–90°E; East: 90°–135°; Far East: 135°–170°) matches the succession of alternating relieves, from west to east: the flat plain of the Ob and Yenisei rivers (West Siberian Plain); the Central Siberian plateau and the flat depression of the Lena river basin (Yakutia); the Far East mountain ranges and the coastal provinces (on the Okhotskian coast the Amur, Primorsky Krai and Khabarovsk, and on the Beringian coast the Magadan, Kamchatka, Chukotka). The latitudinal tripartite division (South: 50°–60°; Medium: 60°–68°; North: >68°N) corresponds to three different zones of modern permafrost, which covers almost the totality of the Siberia plains (it is only absent in SIB-SW) and even overflows to N-Mongolia: it decreases N to S from continuous, discontinuous and sporadic, varying in depth from 1500 to 0.5 m (average 5 m), and widens W to E along the −2°T isotherm from 60°N in the Urals to 54°N on the Pacific coast. North Central Asia is located below Siberia between 55°–35°N (from the Ural-Altai-Sayan mountains to the Syrdarya-Tienshan-Kunlun) and 45°–120°E (from the Caspian Sea to the GreaterKhingan range). It is divided in four parts by climatic considerations and clear geomorphological markers: on the West N-Kazakhstan and the Syrdarya region, on the East Dzungaria-Mongolia and Tarim.

330

Climate Changes in the Holocene

FIGURE 15.1  Physical map of Siberia. White dots = Siberian borders.

15.2.2 Climate The description of modern climate and environmental conditions is of prior importance for paleo-reconstructions: it defines the nomenclature and character of the natural elements in question; it fixes the modern reference for paleo-data, which are most often quoted as deviation from modern values (∆p) *. The climate of Siberia, due to the high latitude and the distance from temperate oceans, is boreal continental and dry, with short, warm summers and long, cold or very cold winters (Dwb, Dwc, Dwd in terms of Köppen climate classification). Temperature (T) and precipitation (P) values depend on latitude and from atmospheric interaction with three oceans: Atlantic in the west, Pacific in the east, Arctic in the north. Atlantic air masses weaken after the Urals and don’t reach SIB-E; Pacific air masses penetrate the Amur valley but don’t proceed beyond the Baikal and the longitudinal mountain chains of the Far East. This situation confers to the climate of Sib-E (Yakutia) an autochthonous character, exposed just to the north as part of the Arctic climate zone: extremely continental, cold and dry, never arid. Mean annual T is −5°C (Jan −25°, July +16°), varying between +2.5°C in SIB-SW and −9° in SIB-NE. As shown in Figure 15.2, climate and environmental differences are gradual in both the south-to-north and west-to-east directions, without pronounced thresholds. The continentality index (amplitude between January and July values) increases from SWW to NEE, mainly due to decreasing winter gradients. The average growing season with daily T > +5°C shortens in the same direction from 140 to 80 days. Mean annual P averages 400 mm: higher in the W and in the SE regions exposed to oceanic air masses, and lower in the most continental central and northeastern regions: decreasing to 200 mm in Yakutia and to less than 50 mm in Beringia. Aridity is absent because precipitation always exceeds evapotranspiration (~20–40 mm year). *

On the Siberian territory, the analysis of the latitudinal and longitudinal distribution of T and P deviation from present (1980 AD) local values (∆p) is a significant index of the intensity and diffusion of climate impact from the Atlantic or Pacific or Arctic sectors.

Interaction of Climate, Environment and Humans in North and Central Asia

331

FIGURE 15.2  Asia: Geographical distribution of modern average annual T and P. (From the Atlas of the Biosphere.)

The climate of North Central Asia is dry continental with hot summer, cool to cold winter and year-round semi-arid to arid, under the influence of three kinds of air masses, from the west, from the east and from Siberia, which defines its partition in three climate zones. The northern zones bordering Siberia shares the same climate of SIB-S; Western Central Asia is reached by Mediterranean and Atlantic air masses bringing precipitation during spring in the south and mid summer in the north; Eastern Central Asia is drier, slightly moistened by the East Asian Summer Monsoon (EASM). Precipitation values vary regularly by altitude (mainly) and by latitude, from above 1000 mm/year in the mountain zone to 330 mm in the southern Urals, 130 mm in the Aral depression and less than 60 mm in the Tarim desert.

15.2.3 Vegetation Cover 15.2.3.1 Biomes Almost the entire vegetal cover of Siberia lies on shallow permafrost soil (gelisol, histosol, podzol) characterized by a black acidic topsoil rich in half-processed organic matter, and an unfertile subsoil that, because of the waterlogged context, is heavily leached and deprived of basic elements. P doesn’t represent a limiting factor, and the selection of the botanical assemblages strictly depends on T gradients. The vegetal cover is distributed by latitude as a series of parallel bands, as shown in Figure 15.3. Wooden boreal species are dominant, growing between the July isothermal lines of +20°–+15° and +12°C (today established between 45°–52° and 70°N), bordered on the N by arctic tundra and cold steppe and on the S by temperate forest-steppe, steppe and desert. Tundra and cold steppe. The tundra-steppe is established between polar desert in the north and boreal forest in the south. Tundra is an assemblage of mosses, small herbs and shrubs developing in polar regions above 70°N or in high mountains, wherever warmer month T is less than +12°, precipitation and evaporation less than 40 mm/year and moisture just provided at mid summer (T=+4°–+6°C) by the thaw of the permafrost surficial layer. The reduction of soil moisture (which happens in better drained slopes deprived of permafrost or where the establishment of a colder climate decreases thaw rates) favors herbs against mosses and the shift from tundra (today dominant) to cold steppe.* Quite significant and mutable is the transitional area (ecotone, see Section 15.2.3.2) between tundra and forest that, depending from T values and latitude, manifests sub-types like shrubs-tundra and forest-tundra. *

Cold steppe is differentiated from temperate steppe (‘real steppe’) by being drier and colder, in spatial proximity to the tundra biome.

332

Climate Changes in the Holocene

FIGURE 15.3  Modern vegetation zones of NE-Europe, Siberia and North Central Asia.

Typical herbivores are reindeer and muskox that, by trampling and grazing, favor cold steppe grasses and tundra mosses against shrubs. Boreal forest. The boreal forest (taiga) is the largest biome on Earth and covers around 60% of the entire Siberian territory. It consists of coniferous taxa: evergreen Scots pine (Pinus sylvestris), spruce (Picea) and fir (Abies) in SIB-W; deciduous larch in the coldest SIB-ME and SIB-FE.* It grows between July isotherm +18° and +10°C (between 55°–58° and 68°–70° N), where mean annual temperature varies between +5° and −5°C and precipitation (basically as summer rain) between 200–750 mm. Summer is lasting 1–3 months and the growing season 4–5 months. The forest canopy diversifies by latitude and moisture: close canopy in the warm-moist southern-middle band (forest), open canopy (parkland) under colder drier conditions and all plants getting an ice-pruned dwarf form at the forest-tundra ecotone. Typical mammals are deer and moose, wild boar and beaver. Temperate mixed forest. Boreal forest grades to temperate mixed forest (mixing coniferous and broadleaf species) at lower latitude escaping extreme winter temperatures, that is, south of the +18° July isotherm (i.e. below 58°N) with mean annual T >–2°C and P >400 mm. Its southern band is dependent also from P gradients, so that it is not very consistent in the central continental SIB-E but well developed in the moist SIB-SW (where pine and birch dominate) and in the very humid SIB-F (with spruce, poplar, helm and cedar taxa). In all forests with close canopy, boreal or mixed, fires are frequent, with cycles of 50–200 years, bringing rejuvenation and shaping the forest mosaic. Steppe. The Eurasian temperate steppe is established between temperate forest in the north and semidesert and desert in the south. It is located where a semi-arid continental climate is limiting the growth of trees and favoring grasses: a 400–800 km wide latitudinal belt developing from Ukraine to Manchuria (from 30° to 120°E), between +26° and +20°C July isotherms (50° and 57°N) where annual P is 44000 BP …and the earliest occupation of Upper Paeolithic to c.43200 BP (layer 6)…and the Upper Paleolithic appears to have developed directly from the Middle Paeolithic” (Derevianko 2007, 61). † In Siberia, “prehistoric cultural complexes developed gradually, without sudden replacement, compared with more dynamic cultural changes in Europe” (Kuzmin 2007a, 763). ‡ The earliest evidence of microblade manufacture in Northern Eurasia is found in Upper Paleolithic complexes of the Altai Mountains dated to 28 ka BP. *

344

Climate Changes in the Holocene

In the context of the Siberian landscape (tundra and boreal forest) lacking food production, the start and periodization of the following Neolithic period is not connected with the appearance of plant cultivation or animal husbandry but, like in the Paleolithic period, with changes of technological-typological characters in the tools assemblage. Characteristic of Neolithic is the standardized lamellar technology and the mass-grinding of some tools, predominantly axes and adzes. Pottery as a tool is a secondary marker of archaeological periodization, fixing stages within a period, so that we can have a ceramic Epipaleolithic or a pre-ceramic Neolithic. In the Far East, pottery appeared as early as 14 ka BP and gradually reached all Siberian corners by 2.5 ka BP:* it is generally accompanying or following closely the start of Neolithic, but in Northeast and East Asia is sometimes even anticipating it.† The general trend of UP population dynamics has been reconstructed by numerical analysis of a set of 240 sites provided of 14C chronological dating. Siberia was continuously occupied all along the UP from 43 to 12 ka BP, across the coldest glacial maximum at 22–18 ka: the number of occupied sites increased gradually, with a peak at 32–30 ka, until 19 ka BP; then the growth became almost exponential until 12 ka‡ (Kuzmin and Keates 2005; Fiedel and Kuzmin 2007). A hasty attempt to explore the possibility of direct correlation between evolution of the number of Paleolithic sites and GISP2 ice-core records (global T trend) had a negative result. This relation has a higher complexity that cannot be ignored: the direct factor inducing human responses is not temperature but environment, and the latter is responding not to global T but to local climate.§ It is difficult determining whether temporal changes of sites’ density are better attributed to internal population dynamics or to migratory waves. In Russian archaeology, both causes are suspected, being that the Siberian Paleolithic is considered part of the Siberian-Mongolian Paleolithic province (inclusive of the West Central Asian region). This large territory has been marked by dramatic latitudinal shifts of inhabitable areas¶ evidencing the ecological complementarity of southern and northern latitudes and explaining the overall similar industrial characteristic and conservatism of the stone tool inventory.** Concerning the question of migratory

The technique of ceramic manufacture has been known from the early Upper Paleolithic when it was used for making objects of portable art. In Siberia its production for domestic ware appears at the end of the Paleolithic and the start of the Neolithic. The earliest domestic pottery known is from the lower Amur region in the Far East (15–14 ka BP) and in the Japanese islands (about 13.8–13.5 ka BP). It appears in the Baikal-Angara at 8–7 ka, in the westernmost S-Siberia regions in 7–6 ka and in the northern Far East territories round 4.5–2.5 (Kuzmin 2010) (see Section 15.4.4). † In Russian archaeology, the transitional stages Epipaleolithic and/or Pre-ceramic Neolithic correspond to the Mesolithic of the western schools. In the case of Siberia, the term Mesolithic is not used for periodization because it is considered not of technological-typological but of the economic-ecological type, and it refers particularly to the transition from the Paleolithic to a productive economy, which didn’t happen in Siberia. ‡ This final demographic explosion happened in times of quasi-devastation of the Paleolithic core areas and of intensive migrations, and its reasons have been probably complex (Vereshchagin and Baryshnikov 1984). § Fiedel and Kuzmin (2007) are well aware of the complexity of the interaction between climate and cultural systems: “In the case of the Paleolithic human occupation of northern Eurasia, we must consider the possible complexities of adaptive responses of both ecosystems and human socio-cultural systems to climatic changes. We should not simply assume that humans responded directly to colder temperatures as a negative stress factor by means of population reduction, contraction of settlement area, or outright abandonment of the entire region.” ¶ “At this time the glacial fauna that inhabited the steppe and riverine corridors of Central Kazakhstan moved northward to Siberia, following the displacement of its ecological environment. And so did the hunting Paleolithic tribes of Sary-Arka (Northern Pre-Balkhash), like the ones of the Ordos and Mongolia, determining all together in Siberia what is called the Siberian-Mongolian (or Siberian-Chinese) Late Paleolithic formation” (Medoev 1982). ** All over the Siberian-Mongolian Paleolithic province, Middle Paleolithic traits are common well into the late Upper Paleolithic and even later (Ranov and Nesmeyanov 1973). *

Interaction of Climate, Environment and Humans in North and Central Asia

345

waves, some hints are provided by typological correlation between stone tools assemblages* and by ancient DNA analyses.†

15.4.2 Cultural Dynamics and Emergence of Early Pottery in the Lower Amur Basin (14 ka BP) The climate warming of the Late Glacial mega-interstadial provoked paludification and disintegration of the cold steppe: tundra from the north and wooden vegetation from southern refuges got supremacy over the cold steppe, dismembering the wide homogeneity of the glacial landscape into a mosaic of ecological systems of reduced extent. The climate-environmental switch extinguished the megafauna (see Section 15.3.2.2) and forced Upper Paleolithic cultures to deep transformation. Foraging communities were compelled to rely on new species, in particular on aquatic resources enhanced by the postglacial rehabilitation of the river network. The new foraging strategies required just the introduction of new tools within the existing microliths blade industry, and helped preserving semi-sedentary habits and the high social status of women. Population numbers increased, foraging groups reduced in size and so did the dimensions of dwellings. The forms and results of these natural and cultural changes share the same traits all over Siberia, but each region presents peculiar characteristics, particularly in the Far East district in both its southern and northernmost parts. The climate and environmental amelioration of the mega-interstadial had its earliest and most accelerated manifestation in the southern Far East, more precisely in the Lower Amur region. Between 14–9 ka BP, this transitional zone between Siberian and East Asian landscapes, most exposed to the influence of warming SST and monsoonal activity of the North Pacific Ocean, manifested the greatest and most accelerated T and P changes of all Siberia: July T rose above modern values from −6° to +1°, January T from −14° to −1° and P as present, favoring regime shift from forest-tundra to larch forest with admixture of broadleaved species. Briefly, the region saw the introduction of southern botanical and cultural elements, and entered in permanent economic and cultural contact with the neighboring territories of NE China, Korea and Japan. Around 14 ka BP in the middle-low Amur basin are coexisting two distinct cultural traditions, which is a common trait not only of the Siberian Far East but of the entire East Asia: a persistent microblade final Upper Paleolithic tradition and an incipient Neolithic tradition inclusive of early ceramic (Kuzmin et al. 1994). Traces of ancient ceramics are very old. They are found in few Upper Paleolithic sites of Europe (the most famous being the clay-fired Venus figurine of Vestonice, 24 ka BP) and in Paleolithic China (the thick, coarse-pasted, round-bottomed, bag-shaped vessels of the Xianrendong Cave, Jianxi, South China, 18 ka BP). The Jianxi ceramic manufacture persisted continuously at low scale until its methods and models, around 14 ka BP, spread northwards to Japan and the Lower Amur, where they acquired the form of domestic artifacts (pots). In the Lower Amur several sites provide evidence of early pottery, foragers’ pottery, the earliest being the site of Khummi, radiocarbon dated at 14470 cal BP and pertaining to the so-called In the “post-Mousterian” UP, manufacture of points, knives and wedge-shaped cores are reflected in 2 basic traditions, both rooted in the Mousterian substratum. The first, established in two waves between 43–26 and 14–7 ka BP and characterized by unifacial stone tools, goes back to the “Mousterian-Levallois cultures” of Middle and Central Asia and concerns the southern zones of Siberia, from Altai to Transbaikal. The second and later (26–5.5 ka BP) but more primitive with bifacial stone tools, comes from the bifacial “Mousterian cultures with Acheulian tradition” of the southern Urals, Central Kazakhstan and other territories bordering Siberia on the south, and concerns Central Yakutia and further to the NE: it successively bifurcated in two sub-traditions with wedge-shaped cores (Dyuktai) or without (Abramova 1975, Mochanov 1977). † Ancient DNA analyses of a 24 ka old individual from Mal’ta (Cis-Baikal) and of a 17 ka individual from Afontova gora (east of the Yenisei) revealed the presence of mitochondrial genome belonging to haplogroup U, which has possibly place of origin in Western Asia and is found at high frequency among UP western Eurasians, modern Europeans (11%) and modern native Americans (14%–37%). This suggests that populations related to contemporary western Eurasians had a more north-easterly distribution into Siberia 24,000 years ago, and that from here departed the colonization of America. (Raghavan et al. 2014a). *

346

Climate Changes in the Holocene

Osipovka culture (14–10.3 ka BP). Archaeological findings testify the continuity of lithic techniques made of knife blades and prismatic cores, scrapers and adzes, and the presence of quite distinctive ceramic products, made by specific technology and with typical shapes (point or knob base, flared rims). As a whole the site represents the earliest monument of a long-standing pottery Neolithic period that in the Amur region will endure until 3 ka BP. “Spore-pollen spectra of the layers with early pottery reflect a warming and wetting of climate, which caused the spread of broad-leaved trees” (Tsydenova 2015) enhancing the interbiotic* character of the region and favoring a lifestyle within a reduced space range, based on hunting, gathering and foraging of new plants and game as well as intensive use of local aquatic resources, which all over Siberia will represent the basic mean of sustenance of the Neolithic cultures. The catching of spawning salmon was predominant and played the main role in promoting a semi-settled lifestyle and the manufacture of pottery.† We can say that, like farming in the Middle East, fishing in Siberia induced the sedentary habits normally connected with the start of the neolithization process. Improvements in ceramic technology (kilns type, firing temperatures, smoothed surfaces, ornamentation, cord-mark decorations, etc.) suggest the distinction of three Neolithic stages (initial, early, late), the last being chronologically fixed between the appearance of millet agriculture around 4.2–3.7 ka BP and the transition to the Early Iron Age at 3.3–3.1 ka BP. Between 14–10 ka BP the process of neolithization and the use of pottery spread beyond its early sites, at first in the neighboring areas of the Amur river basin and Transbaikal region and then gradually westward, following the steps of climate amelioration. Its occurrence around 7.5 ka BP in the Cis-Baikal, with different characteristics (net-impressed design) and possibly by independent discovery, signals a definite step towards its accelerated diffusion to the west and the north: by the end of the 8th millennium it reached the middle Yenisei, Ob and Irtysh basins, and around 6 ka BP the Aldan river in Central Yakutia, from which it diffused across the northern Far East up to the Arctic coast around 5 ka BP and the remote Beringia at 2.3 ka (Kuzmin 2014, Figure 15.1). The basic features of the early Amur pottery characterize all the subsequent Siberian pottery types, supporting the hypothesis of a common point of origin, even if in some cases regional differences are relevant enough to make presumable independent genesis. Constant is the fact that the early foragers’ pottery always appeared in the context of fishing communities, as if not know-how but necessity has been the determining factor of its introduction (Barnett 2009). As shown in Figure 15.6, by the end of the 8th millennium BP the presence of the Siberian foragers’ pottery extended from the Pre-Urals to the Baltic. It appeared at 7.4 ka BP in Northern Russia within the widespread and long-standing hunter gatherers’ cultures of Narva and Comb Ceramic (Pit-Comb Ware), and at 5.3 ka in the Eneolithic Yamnaya and Corded Ware cultures of the Eastern European plains. Being that the presence of pottery in the Urals and Eastern Europe antedates the first kilns in the Middle East by 1000 years, and that nuclear DNA analyses indicate for the Yamnaya people an admixture of Siberian and Caucasian haplogroups (Haak et al. 2015), it is legitimate suspecting the establishment during 7–5 ka BP of grandiose cultural exchanges between the forest-steppe foragers of the Urals and the farming societies of the Middle East, resulting in the diffusion of the art of pottery to the south and the art of stockbreeding to the north (see Section 15.4.5). The successive appearance of pottery in the east-to-west direction, independently from the forms and causes of such kind of progression, witnesses the existence across southern Siberia of a kind of ecological corridor favoring human interaction and cultural diffusion. Its activity will become more evident with the start of the Subboreal period when this latitudinal band, invaded by the steppe * †

About the definition of interbiotic zone, see Section 9.2.3.2. “It seems that the combination of environmental changes and the necessity to process freshwater fish and mollusks and terrestrial plants (including acorns and nuts) resulted in the introduction of pottery-making in East Asia” (Kuzmin et al. 2016). In fact, the cook-ware shape of the majority of the Khummi pots suggests their use as boilers for fish soup.

Interaction of Climate, Environment and Humans in North and Central Asia

347

biome, favored the displacements of pastoralist groups in the opposite direction, from the Eastern European plains to the Baikal region, and definitely changed into a commercial road between the Baltic and the Baikal at the end of the 5th millennium BP with the establishment of the Urals and Altai metallurgic districts.

15.4.3 The Northern Far East: Colonizers of America and Pioneers of the Circumpolar Cultural Region (14.3–2.2 ka BP) The interactive evolution of climate, environment and cultures has been quite different (gradual and conservative) in the northern Far East, like the Kamchatka peninsula or upper Kolyma or Chukotka, which belong to a very special geographical and cultural region: Beringia. Beringia is located at 55°–70°N between the watershed of the Lena river in Siberia and the Mackenzie river in Canada. It has been the first segment of the Circumpolar Zone to be occupied, the route of the human colonization of the Americas, and the nest of all Eskimo cultures. The presence of pre-UP early hominids is not excluded. UP tools are found on both sides of the Bering strait starting from 14 ka BP, and their homogeneous relict character (including bifacial, in lesser measure unifacial, and also cobble traditions from East Asia, often combined in one and the same site) testifies to the existence of a cultural zone with its own personal technological tradition.* The Bering land bridge at −49 m asl and at its max extent more than 500 km wide, constituted an arctic route for mammoth and human movements between the N-Asian and Alaskan sides of Beringia, not just at 14 ka BP (as witnessed by stone tools similarities) but well before and after, from 28 to 10 ka BP. The fact that the passage between the North American Laurentide and Canadian ice sheets opened between 28–25 and 23 ka and then again only after 13–10 ka, and that the tool assemblage of the Clovis culture of Western N-America predates the 13 ka BP and presents significant differences from Beringian tools, all that makes suspecting a still undetected migration from N-Asia into Alaska between 28–23 ka. (Kotlyakov et al. 2017). The Kamchatka peninsula hosts a very important stratified site, the Ushki lake site, where a sequence of seven cultural layers covers all together the entire time span between 14 ka BP and now (Dikov 1977). 14.3–10.7 ka BP (Layer-VII, late UP). The lower Layer-VII embodies a stone tool inventory characterized by the relict characters spoken of above. This makes of the Ushki culture a better candidate in playing a certain role in the colonization of America before and after 23 ka BP than the famous Dyuktai site on the Alden river (Central Yakutia), which has similar earliest dates to 14–12 ka but presents a different, more developed, late UP stone tool inventory including a microblade complex. Moreover, the Ushki specific geographical settings between the persisting glaciers of the Lena watershed on the west and the Pacific coast on the east would have provided a direct meridional corridor for movements along the coastal zone of the Okhotsk sea toward the north, without any obstacle, to Beringia and Alaska. At 14 ka, climate was still cold, valley glaciers covered the mountain range, rivers will begin flowing and shaping terraces only at the start of the Holocene, and winter sea ice was 22 m thick against the present 4–5 m (Chizhov 1970: 74). But the fertile mammoth steppe of high grass type was still the prevailing biome. From the excavations of Layer-VII can be reconstructed the economic-ecological features of the early UP Ushki culture: life strategies semi-settled, based on fishing from a lake that never freezes and on hunting of reindeer, horses, bison and few pachyderms with the use of bow and spear; large houses for 6–10 families; two human burials with grave and surrounding area sprinkled with red ochre, and one burial of a domesticated wolf.† The archaeological

* †

Diagnostic are stemmed points of Levallois-Mousterian technique, possibly of mixed Asian-American origin. In the northern Far East, the dog has been the only animal to be domesticated until recent times. Even today, following the Soviet colonization, the reindeer of the Chukchi and the Koryak are only slightly tamed and may be regarded as in a primitive state of domestication (Mirov 1945).

348

Climate Changes in the Holocene

complex of neighbor sites suggests that the inhabitant group was probably already part of some vast ethnic commonality marked in the west by the Dyuktai culture and in Alaska by the Denali culture. What is more significant from the point of view of our research is the interactive evolution of environment and culture recorded in the succession from Layer-VII to Layer-I. The main feature of the whole process is conservatism coupled with cultural longevity, that is, a series of slow, gradual but successful adaptations, in deep contrast with the cultural dynamism of the southern Far East. Conservatism matches the relatively low amplitude of temporal climate and environmental deviations and the geographical isolation; successful adaptation came from the growing intensity of arctic maritime fishing adaptation. Conservatism and maritime adaptation will promote during the 4–3 ka BP a specific social integration resulting in the definition of an Arctic Circumpolar Cultural Zone (ACCZ). Everything evolved very slowly. The Paleolithic was prolonged until 8.0 ka BP, the Neolithic until modern times. Land transports didn’t develop (domesticated dogs or semi-domesticated reindeer will never be harnessed), which reinforced stationary habits; maritime transport, well known from most ancient times, increased significantly only in the last 4,000 years. Cultural connections with northwestern America can be observed in all stages of the history being examined, most intensively during the late Paleolithic and Epipaleolithic, and then again after 3 ka BP with the formation and spread of the Eskimo culture (Fitzhugh 1975). 10.7–8.0 ka BP (Layer VI-V, late-final UP): T reached present values at 9.9 ka BP and a Boreal thermal optimum (summer ∆pT + 1°) at 8.5 ka. The dry cold steppe progressively disappeared, substituted by wet, swampy tundra and shrub tundra, which means more severe conditions for grazers. Mammoths went immediately extinct; bison and horses decreased in number until disappearing at the end of the period. Only reindeer was left and, as the most adapted to the new environment, became the primary object of land hunting in the entire northern Siberia, preserving Paleolithic traits in these regions for millennia. The lithic industry shifted to late UP characters, introducing microliths from wedge-shaped core and reaching in that way, after more than 4 ka, similarity with the Dyuktai inventory: the innovation spread to Alaska. Hunting groups reduced in size and so did their semi-interred houses, acquiring a shape that will be preserved until a century ago. 8.0–4.2 ka BP (Layer IV, pre-ceramic Neolithic-1). Relevant and continuous T rise started at 7.4 ka BP culminating in HTM (summer ∆pT + 2°) at 5.5–4.5 ka, accompanied by regime shift to open forest and the appearance of a Neolithic industry with thinner microliths from conical core as diagnostic tools. Some relict cultures, like the Siberdik, which was using uniface choppers and cleavers weighing a few kilos, were still based on just hunting. The majority of communities were now using Neolithic tools, relying on mixed hunting and fishing strategies and, favored by the retreat of sea ice, increasingly turning to marine mammals (seals and, later, whales). The process favored sedentarization, population growth and the establishment of matrilinear clans. 4.2–3.2 ka BP (Layer III, pre-ceramic Neolithic-2). The period is climatically similar to the preceding, slightly cooler, but marks a deep cultural turn, supporting the apogee of the primitive economy based on sea mammal procurement, increasing contacts with neighboring groups and territorial expansion.* The enhancement of interregional relations is testified by growing analogies with the Yakutian Neolithic, immigration of fishing communities from the Transbaikal and possibly establishment of distant contacts with the Japan archipelago and initial activation of an Arctic TranSiberian road to Scandinavia. 3.2–2.2 ka BP (Layer II, ceramic developed and late Neolithic, with undeveloped Bronze). Temperature drops by 1°C, reducing the open forest to birch patches and finally to the present conditions of shrub tundra, making the entire Arctic as a peculiar well-defined circumpolar landscape (Stefanovich et al. 1986). The Neolithic stone-working technique rose to perfection, grinding tools *

Muskoxen and humans from the Canadian archipelago colonized W and NE Greenland by the end of the 5th millennium BP (Kotlyakov et al. 2014).

Interaction of Climate, Environment and Humans in North and Central Asia

349

and ceramics appeared and, around 3 ka, even bronze, which at any rate didn’t substitute lithic tools but remained undeveloped. Latitudinal coastal and marine circumpolar interactions increased until the emergence of an autochthonous circumpolar cultural region enlarged to the northwestern and northern coast of North America, which the landscape specificity kept isolated from southern influences. Social aggregations preserved equalitarian traits, and matrilinear and patrilinear clans coexisted in the same villages. As a whole, the period signals the formation of the ethnogenetic roots of Eskimos, Yukagir, Chukchi, Koryak and of their mythology and arts: totemic signs, petroglyphs, obsidian blade figurines of men and animals, and bone-sculpted images of original female ancestors often with a half-animal, half-human appearance. 2.2–0.2 ka BP (Layer-I, remnant Neolithic, with undeveloped Iron). Iron appeared less than 2 ka ago and, like bronze, remained undeveloped. All the former tendencies increased, in particular the strengthening of circumpolar interactions and detachment from southern latitudes, cutting Siberia and the entire Northern Hemisphere into a polar and a non-polar zone.* This went on until the Russian colonization of the 18th century when, with the imposition of paternal clans and private ownership, the primitive society entered the phase of disintegration.

15.4.4 Shifting of the East Asian Summer Monsoon and Cultural Collapse in the Cis-Baikal During the Early Atlantic Period (7.2–6 ka BP) The three cases quoted above are about transformations of Paleolithic cultures through successful adaptation to new climate trends and biotic conditions. The Cis-Baikal case is about the collapse of a long-standing culture as consequence of an abrupt climatic change disrupting its economicecological background. During all historical periods, the Cis-Baikal constituted one of the most favorable human habitats of Siberia, as witnessed by the largest concentration of Upper Paleolithic sites, the early transition to the microlithic phase (22 ka BP) and, as explained in note † at p. 344, its role as genetic niche of native Americans and NW Europeans. Here, the amelioration of climate started with the Holocene, reaching at 10–7 ka BP average annual ∆pT of –2°–3°C and P similar to present. Average annual T values similar to present are established at 6.0 ka, and the HTM of ∆pT + 1° quite late at 3.5–3 ka. These average values depend from the persistence of continentality and hash winters, because the trend of summer T values is quite different, with values similar to present anticipated at 9 ka, and HTM of ∆pT + 2° at 6.5–3 ka. Precipitation values are similar to present from 12 to 8 ka, reach the Holocene max. at 8–7 ka, fall abruptly between 7–6 ka and after 3.5 ka gradually increase to a second peak at 2.5 ka. Aridity is very high at 7–6 ka due to the coupling of max summer T and min P; then it fluctuates across slightly moist (5.3–4.5 ka), arid (4.5–3.5 ka) and slightly moist (3.5–2.5 ka) phases until 2.5 ka, when wet conditions are reestablished (Figure 15.7) (White and Bush 2010). The vegetal cover evolved matching climate change: during the Late Glacial and Early Holocene, spruce- and fir-dominated forest increased across much of the region; then, at the onset of the Middle Holocene, under enhanced aridity, it was basically substituted by pine forest (Prokopenko et al. 2007). These climate and environmental fluctuations make suspecting the occurrence in the Baikal region at the start of the Atlantic period (7 ka BP) the sudden weakening of the influence of EASM, followed, after a hiatus of one millennium, by increasing (cyclical) influence of the Atlantic atmospheric circulation. The EASM between 7–5 ka was apparently redirected to Central China and reached again periodically the Baikal region at 5.1–3.6 ka and, with lesser intensity, after 2.1 ka, as quoted in Section 15.3.1 (An et al. 2000; Chen 2015).

*

“Nuclear and mitochondrial DNA data unequivocally show that the Paleo-Eskimos are closer to each other than to any other present-day population” (Raghavan et al. 2014b).

350

Climate Changes in the Holocene

The aridization event of 7–6 ka BP apparently affected the base of subsistence of the local foragers, because the sudden collapse of a long-standing culture occurred at the same time. Around 8 ka BP, in the upper Angara (Irkut river) and on the western coast of the Baikal, appeared a quite peculiar culture, the Kitoi culture, made up of hunting but mainly fishing mobile equalitarian communities with ancestral ties extending to the local Epipaleolithic period. They entered the millennium using a pre-pottery Early Neolithic toolkit* and then started producing ceramics characterized by net-impressed and comb designs – it is unclear if out of cultural diffusion from the Transbaikal or local invention.† They were relying on a wide range of subsistence strategies (hunting red deer, boar, and bear, and fishing) that progressively switched to the large use of aquatic resources (predominantly salmon but also fresh water seal from the Baikal lake), promoting some kind of sedentarization, demographic expansion and higher social complexity (Lieverse et al. 2009; Waters-Rist et al. 2011). Settlements are very rare and located at the mouth of rivers, and cemeteries are instead very numerous (Lokomotiv, Shamanka, etc.), which witnesses very mobile living habits within a circumscribed territory. Each cemetery has its own specific spatial and mt-DNA configuration and evidently belonging to different clans (Mooder et  al. 2004).‡ The funerary assemblage consisted of ochre, prismatic blades and animal burial, among which are fishes, dogs (a kind of husky) and wolves. Dog burials are many, located in the same cemetery and often in the same tomb together with humans, accompanied with stone tools and spoons, and even adorned with the same type of necklace the Kitoi wore (with four red deer–teeth pendants). Evidently “dogs were considered not impure but equaled as humans, a friends near-human status. Most probably they were hunting companions” and even “shared the same diet” (Losey et al. 2011; Bazaliiskii 2010).§ Around 7.2 ka BP the burial complexes of the Baikal region were totally and abruptly abandoned, and new cemeteries belonging to different cultures reappeared only after 1,000 years. Being that some rare settlings (without cemeteries) continued existing, the reason for such population turnover could not have been, as initially suspected, the complete displacement of a tribe followed by demographic hiatus. It had been a real cultural collapse: the deep, sudden economic conversion of the majority of the Kitoi population to different food resources, accompanied by an accelerated degradation of social links and disordered dispersion (Kuzmin 2007b).¶ The Late Neolithic cultures that after 6 ka repopulated the region (Isakovo, Serovo, Glazkovo) had different toolkits associated mostly with land hunting (Weber et al. 2002) and different subsistence and diet, mobility patterns, social and political relations and genetic affiliation.** They were possibly coming from the Yenisei region in the context of the increasing mobility and domino The toolkit includes microliths for composite tools: shanks for fishhooks and prismatic blades for harpoons, daggers, spears and knives (Weber et al. 2002). † To the idea of some authors, the pottery-making of the initial Neolithic didn’t appear in the Cis-Baikal region through a migration carrying a complex of new technologies but was incorporated as local innovation into a cultural sphere in technological continuity with the lithic complexes of the aceramic Late Upper Paleolithic of the region (Tsydenova 2015). ‡ mt and Y-chromosome DNA tests of samples from different Kitoi cemeteries found total maternal heterogeneity but some cases of paternal similarity, providing information about the prehistoric population structure: “Combining the maternal and the paternal results from the prehistoric populations of Lake Baikal suggested a patrilocal post-marital residence pattern, where females moved to their husbands’ birthplace after marriage” (Moussa 2015), which matches the mobile patterns suggested by the archaeological complex. § Losey (2013) explains the special status of some dogs as follows: “Dogs could sniff out the seals’ pupping dens. Dogs with this talent were likely rare…and local people today claims that only one in 25 dogs has the skill”. ¶ Migratory waves following the collapse of the Kitoi culture could have accelerated the westward diffusion of cultural elements. Arrow-shaped points with an asymmetric lateral notch typical of Transand Cis-Baikal are found in the Altai, Ural and Pre-Caspian regions in layers dated to the 8th-6th millennium BP and in the tool inventory of the Kelteminar culture of the Pre-Aral (7.5–5.5 ka BP). mt-DNA frequencies in samples from EN Kitoi cemeteries show similarities with samples from a couple of LN cemeteries of the Cis-Baikal, Late Bronze tombs of Kazakhstan, and Siberian and Xiongnu kurgans (Der Sarkissian et al. 2013). ** Analyses of mtDNA and Y-chromosome samples from cemeteries before and after the hiatus found in all cases different origin, with the exception of two cemeteries showing maternal continuity (Moussa 2015). *

Interaction of Climate, Environment and Humans in North and Central Asia

351

effects of the Late Holocene. They show genetic discontinuity with the preceding Kitoi culture and affinities with the subsequent Bronze age cultures.‡ The disappearance of the Kitoi culture coincides with one of the most significant periods of climate change in the Baikal region during the Holocene. Given the relatively low population densities, over-harvesting alone is excluded as determinant factor. It must be instead suspected that a natural climatically induced disruption of the aquatic ecosystems occurred. But what had been the environmental response to the climate reversal, and how did it affect the fishery base of subsistence of the Kitoi groups? The establishment of new precipitation patterns can induce changes of water levels and temperature, surface runoff and nutrient input into rivers and lakes, influencing “primary production, community structure and food chains, habitat diversity and availability, fish abundance and dispersal patterns, seasonality of spawning behavior, and larval survival and growth rates”, transforming in few decennia a fishing ground into a desert (White et al. 2010, 17; Beamish ed. 1995). Contemporaneously, shifts in forest taxa could have induced changes in the abundance and distribution of land game. The question, still unsolved, constitutes the so-called “Baikal conundrum”. An intriguing fact is that a second mortuary hiatus is documented in the Baikal region at 3.6–2.7 ka: it again follows precisely the end of a pluvial phase by AESM diversion and accompanied by the diffusion of steppe and the arrival at 3.2 ka of the first pastoralist groups (Losey et al. 2017).

15.4.5 Climate Change, Vegetation Regime Shift and Cultural Dynamics in the Forest-Steppe Belt of the Trans-Urals (7.0–3.0 ka BP) The eastern and western corners of the meridional band of Siberia, the lower Amur and the TransUrals, are the Siberian gates to the “outer” world. The southern Far East has been the first region concerned by postglacial environmental and cultural changes and also the first gate to open; the plains between the Urals and the Altai, located in a long-standing periglacial context, have been the last. Here, cultural processes were postponed until the Boreal period, but as soon as climate and environment conditions similar to present were restored, the recovery was much accelerated, and in less than one millennium made of the region a cultural and geopolitical center of primary importance for the subsequent historical development of Siberia and of the entire Eurasia. At the approaches of the cold hyper-arid conditions of the LGM, the few UP inhabitants of the Trans-Urals deserted the region, finding refuge in the southernmost areas of the foothills. Human groups reappeared only around 12 ka, equipped with late UP microblade technology. Ceramic Neolithic is documented starting around 8.5 ka, and Eneolithic between 6.3 and 4.5 ka (7 ka later than in the Far East), followed by the Middle Bronze epoch (Chairkina et  al. 2017; Matyushin and Zvelebil 1986) (Figure 15.7). The prehistoric cultural complexes in the Trans-Urals from the Mesolithic to the Bronze Age span, in calibrated years BC, approximately as the following after Chairkina et al. (2017, Figure 2): Mesolithic: 10,000 to 6,500; Neolithic: 7,000 to 3,800; Eneolithic (Chalcolithic): 4,300 to 2,800; Bronze Age: 2,500 to 1.000 cal BC. During the 7th millennium BP, at the establishment of the Atlantic period, the Ural region entered a phase of exceptional cultural dynamism. Neolithic communities diffused from the premountain zone and colonized the Tobol-Ishim and Irtysh-Ob interfluves, where is documented the occurrence of a very high concentration of settlements (more than 300, half of which were excavated) all together defining the Atbasar early Neolithic culture (7.5–6 ka BP). This cultural wave spread west and south of the Urals, opening Siberia to the outer world and inducing a chain of large-scale. The reasons for such phenomenal cultural blossoming are various and interrelated: the southern Ural region is the last frontier of the Neolithic revolution of the Siberian foragers that through seven millennia diffused to the west from the Pacific sector; the region is energized by climate changes

352

Climate Changes in the Holocene

and specific regime shifts (see below); it is located in a strategic node of potential migratory routes where movements are facilitated by a continuous steppe belt spanning from the Yenisei to the Pontic; and it is blessed by such abundance of copper and tin ore deposits that, starting from 4.5 until 3.5–3 ka BP, it represented the most powerful metallurgic district of Eurasia: the EurAsian Metallurgical Province (EAMP) (Chernyk 2008). Through the “Uralian gate” (across central and southern Urals) during the 7th millennium BP ceramic styles, stone toolkits, living strategies and genomes of Siberian origin literally poured to the west, over NE Europe and the Baltic, where they appeared within a series of cultures (Figure 15.8). The earliest is the Pit-Comb Ware (or Comb Ceramic) culture (6.2–4.0 ka BP) found in NE Europe between the Baltic and the Trans-Urals, which, by presenting ceramic types of the Siberian tradition and a genetic makeup as admixture of East European and Siberian components,* is considered intrusive and influencial of the preexisting neighboring Narva culture (SE Baltic, 7.3–3.7 ka). The Pit-Comb Ware is followed on the south by two widespread Neolithic-Chalcolithic cultures of very mobile stockbreeders and (only later and in the Cis-Urals) also mix-farmers: the Yamnaya and the Corded Ware. The Yamnaya culture (Yamna or Pit Grave, 5.3–4.6 ka) succeeded the preexisting Dnieper-Donets, Samara and Khvalinsk (6.7–5.8 ka) cultures in the forest-steppe (as Northern Yamnaya) and steppe (as Southern Yamnaya) from the Danube to the Urals and probably further east in the Yenisei as Afanasevo culture (5.3–4.5 ka). North of the Yamna, between the Rhine and the Volga, is found the Corded Ware culture (4.9–4.3 ka), in reality a series of independent cultures succeeding the Northern Yamnaya and sharing similar traits. Yamna and Corded Ware are both

FIGURE 15.8  Neolithic and Eneolithic cultures of the CircumCaspian and East European regions: Kelteminar, Pit-Comb, Yamnaya, Corded Ware, and some preexisting cultures. Their origin coincides with the western spread of Siberian cultures and genotypes. Color refer to paleo-ethnic affiliation: greenish = Uralic (Y-DNA: N1c); yellow-orange = Proto-Indoeuropean (Y-DNA: R1b, R1a); reddish = Caucasian (Y-DNA: G, F’, J2, R1b). Number = ka BP. *

Ancient DNA analyses of samples from sites of Pit-Comb Ware culture at Smolensk (Russia) reveal the presence of Y-DNA haplogroup N1a1. It probably arose in North China around 14 ka BP, and subsequently experienced serial bottlenecks in Siberia and secondary expansions in eastern Europe (Rootsi 2006). N1a1 is dominant in male individuals of the Liao culture in the Chinese Far East and of the Kelteminar culture in W-Central Asia (Mazurkevich and Dolbunova 2012); its subclade N1c characterizes individuals from a LN Baikal cemetery and is dominant among modern Uralic people.

Interaction of Climate, Environment and Humans in North and Central Asia

353

a result of the blending of local populations with massive migrations from the Urals, Siberia* and the Caucasus†, and together constitute an integral part of the original niche (“urheimat”) of the proto-Indo-European genetic makeup and languages. Through the Uralian gate the Siberian cultures also established contacts with southern farming societies. The existence of a meridional corridor of cultural exchanges between the Urals, the Caucasus and the Middle East along both sides of the Caspian Sea is suggested by some synchronicities: the sudden development during the 7th millennium BP of pottery technology in Northern Mesopotamia‡; the appearance during the 6th millennium BP of domesticated species of southern origin (sheep and cow) in the Eurasian steppes; and the fact that the genome of the 43% of Yamnaya culture individuals submitted to ancient DNA testing includes components typical of Caucasian farmers. The human interactions that occurred along this meridional corridor worked like an ethnogenetic spark: on the west of the Caspian Sea, in the Russian plains, they ended up by structuring Proto-Indoeuropean societies and languages; on the east, in the Aral basin, they generated the Kelteminar culture (7.5–5.5), a Neolithic culture of Uralian ancestry with phenotypes and DNA characteristic of Finno-Ugric people (Yablonsky 1985; Kozintsev 2009) and diagnostic microlithic tools of South Siberian origin.§ In that way, during the 7th–6th millennia BP, a powerful “CircumCaspian Cultural Zone” made its historical appearance.¶ The subsequent cultural development of the southern Urals no longer depended on just its Siberian substratum and its protagonists were not just foragers but became involved in two largescale processes that emerged out of the interaction between Siberia and the outer world. • The first large-scale process has been the pastoralist colonization of the steppe belt. It started with the import of few domesticated taxa (cattle, sheep, goat) and then was enhanced by independent domestication of the horse: in Mullino II (Southern Urals) at 6.2–6.0 ka BP (debatable); in Botai (North Kazakhstan) at 5.5–4.8 ka BP (Matyushin 2003; Levine 1996). Domestication reached the Yenisei at 5.3 ka, the Baikal at 3.2 and Yakutia only at 0.6 ka, together representing the most significant return effect into the Siberian territory of the opening of the Ural gate. It never reached the northern Far East. Haak et al. (2015) note that tests of autosomal DNA from individuals of the Yamna culture revealed an “Ancient North Eurasian” genetic component representing descent from the people of the Mal’ta-Buret culture. Dominant paternal lineages R1b and R1a are detected: the first is the most common haplogroup among Yamna samples and modern West Europeans; the second is the most common among modern North Europeans. The authors also note that about 75% of the autosomal DNA of Corded Ware skeletons found in Germany matches DNA from individuals of the Yamna. † Jones et al. (2015) and Haak et al. (2015) found that the genome of 43% of Late Neolithic or Bronze Age individuals of the Yamna culture submitted to DNA analyses contains a Caucasian component (Y-DNA J2), possibly introduced by people together with domestication and farming techniques. ‡ In the Middle East, the transition from pre-pottery to pottery Neolithic occurred in Early Neolithic semi-nomadic camps of Northern Iraq, where are also documented some of the earliest cases of domestication (goat). In Ganj Dareh (Kurdistan), the first ceramic products are statuettes and geometric objects (7.3–6.9 ka BP). mtDNA analysis of bone from a local woman shows that she belonged to Haplogroup X, which is dominant among the ancient Caucasian huntergatherers and present among people of the Yamna and the Afanasevo cultures (Gallego-Llorente et al. 2016; Jones et al. 2015). But ceramic independent invention in complex interaction with external influences is not excluded. An earliest exceptional presence of pottery is found at the site of Tell Sabi Abyad (N-Syria) where pottery appears at 9 ka BP fully developed with mineral temper and painted stripes, apparently as prestige object, before to be substituted within 300 years by the production of coarse ceramics for everyday use (Nieuwenhuyse et al. 2010). § The late “Kelteminar” arrowhead with typical asymmetrical side notch (also called Khina or Daurian point) has a very large distribution, of which the reasons could have been diffusion or independent emergence. Its earliest presence is found in the Transbaikal Daurian steppe and in the Kitoi culture of the Cis-Baikal (8–7 ka BP) from where it diffused to Mongolia, Trans-Urals and the whole of Central Asia to the Volga and possibly the Zagros, until the Chalcolithic period (Kiryushin 2011; Brunet 2005). ¶ It is not possible to evaluate precisely the number of people involved in those processes, but some information can be sorted out about its scale. There is a large accord among authors in estimating for the 6 ka BP a world global population of seven million, which means 1/1000 of the present population number. The same ratio, when applied to particular regions, would suggest for the 6 ka the presence of 30–40 thousand people in the entirety of Siberia, 13–15,000 in the Urals and 150–200,000 in the CircumCaspian region. *

354

Climate Changes in the Holocene

• The second process has been the establishment, during the Middle-Late Bronze epoch, of two Metallurgical Provinces based on the exploitation of the copper mines in the Urals (EAMP) and the Altai (SEAMP), plus the tin mines of the Chingyz range and, more in general, the ore deposits of West Central Asia. As shown in Table 15.1, the metallurgic district conferred to the region absolute control of the whole Eurasian bronze production from 4.2 to 3.5 ka BP and of the East Asian sector from 3.5 to 2.8 BP. Both provinces and phases were ruled by elite military groups. During the first phase, the Urals production was controlled by proto-Indo-Aryan tribes oriented towards southern markets, and the Altai production by very mobile Finno-Ugrian Seima-Turbino clans moving along the forest-steppe zone between the Baltic and the Yenisei (of which only tombs are found or, better, mainly memorials in the form of cenotaphs with metal and flint as grave goods). During the second phase, the Ural mines were exhausted and the western Eurasian markets passed into the hands of the resurging Carpathian production (EuMP); on the east, the Seima-Turbino people merged within Karasuk tribes that, centered in the Altai, kept furnishing the Baikal and Amur regions in Siberia and the Ordos in China (Chernyk 2008, 2012). TABLE 15.1 Development of the Metallurgic Provinces of Eurasia during the Bronze Age (7.5–2.8 ka BP) Periods

Metallurgic Provinces of Eurasia*

Years BP

Age

Europe

3200–2800

FB

EuMP

3500–3200

LB

Caucasus

EAMP-3

EAMP-2

MB

EAMP-1

4900–4200

CMP-2 CMP-1

5800–4900

EB

Proto-CMP CBMP

7500–5800

CH

Altai SEAMP-2

3900–3500

4200–3900

Urals

EMA

SEAMP-1

West Central Asia Steppe and Forest Steppe Cultures Final Bronze: Karasuk-2 (Altai, C-KZ) Late Bronze II: Andronovo-2 (C-KZ); Karasuk-1 (Altai) Late Bronze I Andronovo-1 (Urals, C-KZ, W-Altai); SeimaTurbino (N-Altai) Middle Bronze II Sintashta-Petrovka (Urals); SeimaTurbino (N-Altai) Eneolithic - Middle Bronze I Poltovka, Abashevo (Urals); Afanasevo, Okunevo, Elunino (Altai) Eneolithic Botai (N-KZ); Yamnaya (Urals), Afanasevo (S-Siberia); Kelteminar (W-Central Asia) Neolithic Pit-Comb (Cis-Urals); Atbasar (Trans-Urals) Kelteminar (W-Central Asia);

*  Acronyms. CH: Chalcolithic; EB, MB, LB, FB: Early, Middle, Late, Final Bronze; EMA: Early Metal Age (Chalcolithic); CBMP, CMP, EAMP, SEAMP, EuMP: Carpatho-Balkan, Circumpontic, Eur-Asian, Steppe-East-Asian, European Metallurgic Province. N-KZ, C-KZ: North, Central Kazakhstan. Highlighted cells: light gray = start or disintegration; dark gray = inactivity. Based on data from Chernyk (2008).

Interaction of Climate, Environment and Humans in North and Central Asia

355

We will not develop further these two most interesting subjects, about which we recommend an introductory bibliography. We conclude instead reporting how climate and environmental changes could have supported this unique cultural development of the Southern Urals, referring to the climate-environmental reconstruction provided by Velichko et al. (1997), the cultural responses to climate reversals in the south Urals region provided by Vybornov (2010), and the analysis of the interaction between alternating arid-pluvial phases and migratory waves in the Baraba steppe and the Tobol-Irtysh interfluve provided by Zakh et al. (2010). Periglacial conditions with tun dra and cold steppe elements prevailed in the Trans-Urals until the start of the Holocene, when large areas were finally taken out by parkland and steppe. The Boreal climate, with ∆pT slightly lower than now, saw the penetration of patches of spruce-birch forest and the establishment of a forest-steppe belt more or less similar to present. As explained in Section 15.2.3.2, the interbiotic character of the forest-steppe ecotone provided the very favorable background of all the subsequent cultures, from the Neolithic Atbasar to the Early Iron communities of mounted pastoralists (Tasmola, Sauromates, etc.), conferring to them mosaic-like structure, mobility and adaptability. The warming trend approaching the Boreal-Atlantic transition was in the north the most favorable time span for the development of forest; but in the south, that is, in the Irtysh-Ob interfluve, it increased aridity and provoked the first of a series of cyclical regime shifts between forest, foreststeppe and steppe, accompanied by correspondent migratory waves. • Between 7.7 and 6.3 ka BP the forest-steppe moved north, arid steppe became dominant, and people expelled by desertification immigrated from the south to the region, contributing to the start of the Atbasar culture. • Between 6.3–6.1 ka, at the approaches of the thermal optimum of the Atlantic period, mixed forest had an exceptional development that supported the blossoming of the Atbasar culture. Then climate conditions gradually worsened until a short, abrupt, arid phase at 5.3–5.0 ka when the region was crossed again by southern people moving with domesticated cattle, sheep and goats. Most probably this migratory wave is related also to the establishment of the Afanasevo culture in the Altai and in the Minusinsk basin. • A long cooling arid trend phase occurred between 4.5–3.2 ka: its final stage saw the immigration from the south of pastoralist groups who colonized the Trans-Ural floodplain and in few centuries would be counted among the first protagonists of the Early Iron Scythian confederations. • Following the 3.2 ka BP, climate changed to cool-moist and the vegetation of the region acquired the present configuration, with taiga in the north and forest-steppe in the south, attracting, around 2.7 ka, Scythian immigrant groups coming this time from the NE (broader Altai area) and fostering the first demographic peak of the Late Holocene.

15.5 CONCLUSIONS The periodization of the multi-millennial postglacial evolution of the foraging communities of Siberia is strictly based on technological considerations and presents a fundamental character: every cultural period starts and ends as a response to the establishment of new climate-environmental conditions. In the absence of challenging stimuli from climate, internal social dynamics could just introduce technical ameliorations within the period and well-defined cultural-bio-geographic realms. This pattern changed only at the start of the 7th millennium BP with the opening of the Ural region to the “outer world” of productive pastoralist and farming economies. The outpouring of Siberian foragers’ into the Circum-Caspian region intermingled with pastoralist and metallurgist groups, inducing an ethnogenetic process that made the “Circum-Caspian Cultural Zone” and its steppe extension the geopolitical center of Eurasia all throughout the following millennia, until the advent of artillery.

356

Climate Changes in the Holocene

The return-effect into Siberia of such pivotal phenomenon only affected the southern part of the territory, from the Urals to the Baikal, where during the 3rd–2nd millennium BP stockbreeding became well established and, starting from the 1st millennium BP, was nurtured a long series of expansive mounted military confederations. The central and northern regions of Siberia preserved the former cultural rhythm, punctuated by just climate change. The foragers of the extreme Arctic north, well defined by the tundra biome, arrived even to sever southern links and to gather within an autochthonous “Arctic Circumpolar Cultural Zone”.

REFERENCES Abramova, Z.A. 1975. Arkheologicheskie kul’tury v verkhnem paleolite Severnoi Azii i yuzhnosibirskaia kul’turnaia oblast [Archaeological Cultures in the Upper Paleolithic of Northern Asia and the Southern Siberian Cultural Region]. In Sootnoshenie drevnikh kul’tur Sibiri skul’turami sopredel’nykh territorii [Correlation of Early Cultures of Siberia with the Cultures of Adjoining Territories]. Novosibirsk. [in Russian]. An, Z., S.C. Porter, J.E. Kutzbach, W. Xihao, W. Suming, L. Xiaodong, L. Xiaoqiang, et al. 2000. Asynchronous Holocene optimum of the East Asian monsoon. Quaternary Science Reviews 19(8): 743–762. Andreev, A.A. and P.E. Tarasov. 2013. Northern Asia. In The Encyclopedia of Quaternary Science, edited by S.A. Elias, 4: 164–172. Amsterdam: Elsevier. Baclanov, P.Y. and A.A. Velichko. 2011. Environmental development of Northeast Asia in the Pleistocene and Holocene. Quaternary International 237: 1–3. Baker, J.L., M.S. Lachniet, O. Chervyatsova, Y. Asmerom, and V.J. Polyak. 2017. Holocene warming in western continental Eurasia driven by glacial retreat and greenhouse forcing. Nature Geoscience 10(6): 430. Barnett, W.K. 2009. Discussion: Living at the edge. In Ceramics Before Farming: The Dispersal of Pottery Among Prehistoric Eurasian Hunter-Gatherers, edited by P. Jordan and M. Zvelebil, 53. Oakland: Left Coast Press. Bauch, H.A., T. Mueller-Lupp, E. Taldenkova, R.F. Spielhagen, H. Kassens, P.M. Grootes, J. Thiede, et al. 2001. Chronology of the Holocene transgression at the North Siberian margin. Global and Planetary Change 31(1): 125–139. Bazaliiskii, V.I. 2010. Mesolithic and Neolithic mortuary complexes in the Baikal region. In Prehistoric Hunter-Gatherers of the Baikal Region, Siberia: Bioarchaeological Studies of Past Life Ways, edited by A.W. Weber et al., 51–86. Edited by A.W. Weber, M. A. Katzenberg, and T. G. Schurr. Philadelphia: University of Pennsylvania Press. Beamish, R.J., ed. 1995. Climate Change and Northern Fish Populations, 121. Ottawa: NRC Research Press. Binney, H., M. Edwards, M. Macias-Fauria, A. Lozhkin, P. Anderson, J.O. Kaplan, A. Andreev, et al. 2017. Vegetation of Eurasia from the last glacial maximum to present: Key biogeographic patterns. Quaternary Science Reviews 157: 80–97. Brunet, F. 2005. Pour une nouvelle etude de la culture neolithique de Kel’teminar, Ouzbekistan. Paleorient 31(2): 100–102. Campbell, J. 1959. The Masks of God: Primitive Mythology. London: Secker & Warburg. Chairkina, N.M., Y.V. Kuzmin, and G. Hodgins. 2017. Radiocarbon chronology of the Mesolithic, Neolithic, Aeneolithic and Bronze Age sites in the Trans-Urals (Russia: a general framework). Radiocarbon 59(2): 505–518. Chen, R., J. Shen, C. Li, E. Zhang, W. Sun, and M. Ji. 2015. Mid-to late-Holocene East Asian summer monsoon variability recorded in lacustrine sediments from Jingpo Lake, Northeastern China. The Holocene 25(3): 454–468. Chernykh, E.N. 2008. Formation of the Eurasian “steppe belt” of stockbreeding cultures: Viewed through the prism of archaeometallurgy and radiocarbon dating. Archaeology, Ethnology and Anthropology of Eurasia 35(3): 36–53. Chernykh, E.N. 2012. The northern hunters, fishers and gatherers and their southern neighbours in Eurasia in the early metal age. In Networks, Interaction and Emerging Identities in Fennoscandia and Beyond, Papers from the conference held in Tromsø, Norway, October 13–16, 2009. Mémoires de la Société Finno-Ougrienne 265C, edited by Damm and J. Saarikivi, 105–122. Helsinki: Damm and J. Saarikivi. Chizhov, O.P. 1970. Ob izmeneniiakh sostoianiia arkticheskogo basseina so bremeni maksimuma poslednego oledeneniia [On Changes of the Condition of the Arctic Basin from the Time of the Maximum of the Last Glaciation]. In Severnyi Ledovityi okean i ego poberezh’e v kainozoe [The Arctic Ocean and Its Coast in the Cenozoic]. Edited by O. A. Ivanov. Leningrad: Gidrometeozidat. [In Russian]

Interaction of Climate, Environment and Humans in North and Central Asia

357

Dallmeyer, A., M. Claussen, J. Ni, X. Cao, Y. Wang, N. Fischer, M. Pfeiffer, et al. 2017. Biome changes in Asia since the mid-Holocene-an analysis of different transient Earth system model simulations. Climate of the Past 13(2): 107–134. Der Sarkissian, C., O. Balanovsky, G. Brandt, V. Khartanovich, A. Buzhilova, S. Koshel, V. Zaporozhchenko, et al. 2013. Ancient DNA reveals prehistoric gene-flow from Siberia in the complex human population history of North East Europe. PLoS genetics 9(2), p.e1003296. Derevianko, A.P. 2010. Three scenarios of the middle to upper paleolithic transition: Scenario 1: the Middle to Upper Paleolithic Transition in Northern Asia. Archaeology, Ethnology and Anthropology of Eurasia 38(3): 2–32. Derevianko, A.P., A.V. Postnov, E.P. Rybin, Y.V. Kuzmin, and S.G. Keates. 2007. The Pleistocene peopling of Siberia: a review of environmental and behavioural aspects. Indo-Pacific Prehistory Association 3. Dikov, N.N. 1977. Archaeological sites in Kamchatka, Chukotka, and the Kolyma River headwaters.Moscow: Nauka. [in Russian]. Dolitsky, A.B. 1985. A critical review of ‘the Mesolithic’ in relation to Siberian archaeology. Arctic, Sept. 1: 178–187. Dolukhanov, P.M., A.M. Shukurov, P.E. Tarasov, and G.I. Zaitseva. 2002. Colonization of Northern Eurasia by modern humans: Radiocarbon chronology and environment. Journal of Archaeological Science 29(6): 593–606. Fiedel, S.J., and Y.V. Kuzmin. 2007. Radiocarbon date frequency as an index of intensity of Paleolithic occupation of Siberia: Did humans react predictably to climate oscillations? Radiocarbon 49(2): 741–756. Fitzhugh, W., ed. 1975. Prehistoric Maritime Adaptations of the Circumpolar Zone. Berlin: Walter de Gruyter. Frenzel, B., M. Pecsi, and A.A. Velichko eds. 1992. Atlas of Paleoclimates and Paleoenvironments of the Northern Hemisphere. Geographical Research Institute, Hungarian Academy of Sciences, Budapest, Gustav Fischer Verlag, Stuttgart, Jena, New York. Gallego-Llorente, M., S. Connell, E.R. Jones, D.C. Merrett, Y. Jeon, A. Eriksson, V. Siska, et al. 2016. The genetics of an early Neolithic pastoralist from the Zagros, Iran. Scientific Reports 6: 31326. Jin, L., F. Chen, C. Morrill, B.L. Otto-Bliesner, and Rosenbloom N. 2012. Causes of early Holocene desertification in arid central Asia. Climate Dynamics 38(7–8): 1577–1591. Jones, E.R., G. Gonzalez-Fortes, S. Connell, V. Siska, A. Eriksson, R. Martiniano, R.L. McLaughlin, et al. 2015. Upper Paleolithic genomes reveal deep roots of modern Eurasians. Nature Communications 6: 8912. Jordan, P. and Z. Marek, eds. 2009. Ceramics Before Farming: The Dispersal of Pottery Among Prehistoric Eurasian Hunter-Gatherers, 53. Oakland: Left Coast Press. Haak, W., I. Lazaridis, N. Patterson, N. Rohland, S. Mallick, B. Llamas, G. Brandt, et al. 2015. Massive migration from the steppe was a source for Indo-European languages in Europe. Nature 522(7555): 207. Kim, J.H., N. Rimbu, S.J. Lorenz, G. Lohmann, S.I. Nam, S. Schouten, C. Rühlemann, et al. 2004. North Pacific and North Atlantic sea-surface temperature variability during the Holocene. Quaternary Science Reviews 23(20): 2141–2154. Kiryushin, Yu F., K. Yu Kiryushin, and V.P. Semibratov. 2011. Kelteminar arrowheads from the Altai. Archaeology, Ethnology and Anthropology of Eurasia 39(1): 56–64. Kotlyakov, V.M., A.A. Velichko, and S.A. Vasil’ev. 2014. Pervonachal’noe zaselenie Arktiki chelovekom v usloviyah menyayuscheisia prirodnoi sredy [Initial human colonization of the Arctic in changing paleoenvironments]. Moscow: GEOS. Kotlyakov V.M., A.A. Velichko, and S.A. Vasil’ev. 2017. Human Colonization of the Arctic: The Interaction Between Early Migration and the Paleoenvironment. Cambridge: Academic Press. Kozintsev, A.G. 2009. Craniometric evidence of the early Caucasoid migrations to Siberia and eastern Central Asia, with reference to the Indo-European problem. Archaeology, Ethnology and Anthropology of Eurasia 37(4): 125–136. Kuzmin, Y.V. 2007a. Chronological framework of the Siberian Paleolithic: Recent achievements and future directions. Radiocarbon 49(2): 757–766. Kuzmin, Y.V. 2007b. Hiatus in prehistoric chronology of the Cis-Baikal Region, Siberia: Pattern or artifact?. Radiocarbon 49(1): 123–129. Kuzmin, Y.V. 2010. The origin of pottery in East Asia and its relationship to environmental changes in the Late Glacial. Radiocarbon 52(2): 415–420. Kuzmin, Y.V. 2014. The Neolithization of Siberia and the Russian Far East: major spatiotemporal trends (the 2013 state-of-the-art). Radiocarbon 56(2): 717–722. Kuzmin, Y.V., G.S. Burr, J.M. O’Malley, and A.J. Jull. 1994. Radiocarbon dating of climatic and cultural changes on the Russian Far East during the Late Glacial and Holocene. Pacific Institute of Geography. Vladivostok. https://www.lpi.usra.edu/meetings/gold/pdf/2198.pdf

358

Climate Changes in the Holocene

Kuzmin, Y.V., and S.G. Keates. 2005. Dates are not just data: Paleolithic settlement patterns in Siberia derived from radiocarbon records. American Antiquity 70(4): 773–789. Kuzmin Y.V., and L.A. Orlova. 1998. Radiocarbon Chronology of the Siberian Paleolithic. J. of World Prehistory 12: 1–53. Levine, M.A. (1996). Domestication of the horse. The Oxford Companion to Archaeology, 315–317. Lieverse, A.R., V.I. Bazaliiskii, O.I. Goriunova, and A.W. Weber. 2009. Upper limb musculoskeletal stress markers among middle Holocene foragers of Siberia’s Cis-Baikal region. American Journal of Physical Anthropology 138(4): 458–472. Losey, R.J., V.I. Bazaliiskii, S. Garvie-Lok, M. Germonpré, J.A. Leonard, A.L. Allen, M.A. Katzenberg, et al. 2011. Canids as persons: Early Neolithic dog and wolf burials, Cis-Baikal, Siberia. Journal of Anthropological Archaeology 30(2): 174–189. Losey, R.J., A.L. Waters-Rist, T. Nomokonova, and A.A. Kharinskii. 2017. A second mortuary hiatus on Lake Baikal in Siberia and the arrival of small-scale pastoralism. Scientific Reports 7(1): 2319. MacDonald, G.M., A.A. Velichko, C.V. Kremenetski, O.K. Borisova, A.A. Goleva, A.A. Andreev, L.C. Cwynar, et al. 2000. Holocene treeline history and climate change across northern Eurasia. Quaternary Research 53(3): 302–311. Mann, D.H., P. Groves, R.E. Reanier, B.V. Gaglioti, M.L. Kunz, and B. Shapiro. 2015. Life and extinction of megafauna in the ice-age Arctic. Proceedings of the National Academy of Sciences 112(46): 14301–14306. Marsicek, J., B.N. Shuman, P.J. Bartlein, S.L. Shafer, and S. Brewer. 2018. Reconciling divergent trends and millennial variations in Holocene temperatures. Nature 554(7690): 92. Matyushin, G. 2003. Problems of inhabiting central Eurasia: Mesolithic–Eneolithic exploitation of the central Eurasian steppes. In Prehistoric Steppe Adaptation and the Horse, 367–393. Cambridge: McDonald Institute for Archaeological Research. Matyushin, G., and M. Zvelebil. 1986. The Mesolithic and Neolithic in the Southern Urals and Central Asia. New York: Cambridge University Press. Mazurkevich A. and E. Dolbunova. 2012. The oldest pottery in hunter-gatherer communities and models of Neolithisation of Eastern Europe. Documenta Praehistorica 42: 13–66. Medoev, A. 1982. Geokhronologia paleolita Kazakhstana [Geochronology of the Paleolithic of Kazakhstan]. Alma-Ata: Nauka [in Russian]. Mirov, N.T. 1945. Notes on the domestication of reindeer. American Anthropologist 47(3): 393–408. Mochanov, I.A. 1977. Drevneishie etapy zaseleniia chelovekom severo-vostochnoi Azii [The Most Ancient Stages in the Settlement by Man of Northeast Asia]. Novosibirsk: Nauka. [in Russian]. Mooder, K.P., T.G. Schurr, F.J. Bamforth, V.I. Bazaliiski, and N.A. Savel’ev. 2004. Population affinities of Neolithic Siberians. A snapshot from prehistoric Lake Baikal. American Journal of Human Biology 16(2): 215. Moussa, N.M. 2015. Maternal and paternal polymorphisms in prehistoric Siberian populations of Lake Baikal. PhD diss., University of Alberta. Muratova, M.V., L.R. Serebryanny, and O.V. Denissenko. 1993. Climate and vegetation changes in the Northern Eurasia during the Younger Dryas. Geographica Helvetica 48(3): 111–119. Nieuwenhuyse, O.P., P. Akkermans, and J. Van der Plicht. 2010. Not so coarse, nor always plain—The earliest pottery of Syria. Antiquity 84(323): 71–85. Pei, Q. and D. Zhang. 2014. Long-term relationship between climate change and nomadic migration in historical China. Ecology and Society 119(2). Prokopenko A.A., G.K. Khursevich, E.V. Bezrukova, M.I. Kuzmin, X. Boes, D.F. Williams, S.A. Fedenya, et al. 2007. Paleoenvironmental proxy records from Lake Hovsgol, Mongolia, and a synthesis of Holocene climate change in the Lake Baikal watershed. Quaternary Research 68(1): 2–17. Raghavan, M., M. DeGiorgio, A. Albrechtsen, I. Moltke, P. Skoglund, T.S. Korneliussen, B. Grønnow, et al. 2014b. The genetic prehistory of the New World Arctic. Science 345(6200). Raghavan, M., P. Skoglund, K. Graf, M. Metspalu, A. Albrechtsen, I. Moltke, S. Rasmussen, et al. 2014a. Upper Paleolithic Siberian genome reveals dual ancestry of Native Americans. Nature 505(7481): 87–91. Ranov, V.A., and S.A. Nesmeyanov. 1973. Paleolit i stratigrafia antropogena Srednei Azii [The Paleolithic and the Anthropogene Stratigraphy in Central Asia]. Dushanbe: Donish [in Russian]. Ricankova, V., J. Robovský, J. Riegert, and J. Zrzavý. 2015. Regional patterns of postglacial changes in the Palearctic mammalian diversity indicate retreat to Siberian steppes rather than extinction. Scientific Reports 5: 12682.

Interaction of Climate, Environment and Humans in North and Central Asia

359

Stefanovich, E.H., V.A. Klimanov, Z.K. Borisova, and S.N. Vinogradova. 1986. Paleogeographic conditions during the Holocene on the northern coast of the Penzhina Bay. Bulletin of the Commission for the Study of the Quaternary period 55. [in Russian] Tsydenova, N., and H. Piezonka. 2015. The transition from the Late Paleolithic to the Initial Neolithic in the Baikal region: Technological aspects of the stone industries. Quaternary International 355: 101–113. Velichko, A.A., ed. 2009. Paleoclimates and Paleoenvironments of Extra-Tropical Regions of the Northern Hemisphere. Late Pleistocene-Holocene. Moscow: GEOS Publishers. [in Russian]. Velichko, A.A., A.A. Andreev, and V.A. Klimanov. 1997. Climate and vegetation dynamics in the tundra and forest zone during the Late Glacial and Holocene. Quaternary International 41: 71–96. Vereshagin, N.K., and I.M. Gromov. 1977. The formation of the teriofauna of USSR during the Quaternary period. In Achievements of Modern Theriology, 37. Moscow: Nauka. Vybornov, A.A. 2010. On the correlation of natural and cultural processes in the Neolithic-Volga-Kama area. Documenta Praehistorica 37: 293–298. Waters-Rist, A.L., V.I. Bazaliiskii, A.W. Weber, and M.A. Katzenberg. 2011. Infant and child diet in Neolithic hunter-fisher-gatherers from cis-baikal, Siberia: Intra-long bone stable nitrogen and carbon isotope ratios. American Journal of Physical Anthropology 146(2): 225–241. Weber, A.W., D.W. Link, and M.A. Katzenberg. 2002. Hunter-gatherer culture change and continuity in the Middle Holocene of the Cis-Baikal, Siberia. Journal of Anthropological Archaeology 21(2): 230–299. White, D., and A. Bush. 2010. Holocene climate, environmental change, and Neolithic biocultural discontinuity in the Baikal region. In Prehistoric Hunter-Gatherers of the Baikal Region, Siberia. Bioarchaeological Studies of Past Life Ways, edited by Weber et al., 1–26. Edited by A.W. Weber, M. A. Katzenberg, and T. G. Schurr. Philadelphia: University of Pennsylvania Press. Xia, D., J. Jia, G. Li, S. Zhao, H. Wei, and F. Chen. 2014. Out-of-phase evolution between summer and winter East Asian monsoons during the Holocene as recorded by Chinese loess deposits. Quaternary Research 81(3): 500–507. Yablonsky L.T. 1985. Kelteminar Culture in Neolithic steppe zone of Middle Asia. Soviet Ethnography 2: 127–140. Zakh, V.A., N.E. Ryabogina, and J. Chlachula. 2010. Climate and environmental dynamics of the mid-to late Holocene settlement in the Tobol–Ishim forest-steppe region, West Siberia. Quaternary International 220(1–2): 95–101. Zimov, S.A., N.S. Zimov, A.N. Tikhonov, F.S. Chapin. 2012. Mammoth steppe: A high-productivity phenomenon. Quaternary Science Reviews 57: 26–45.

Section IV Challenges Ahead

16

Perspectives of Climate Monitoring in the Satellite Era Mika G. Tosca

CONTENTS 16.1 A “Pre-History” of Climate-Monitoring Satellites............................................................... 363 16.2 The Contemporary Era of Climate-Monitoring Satellites..................................................... 366 16.3 Satellite-Based Observational Investigations of Climate Change......................................... 367 16.3.1 Combining Terra and Aqua to Reconcile Equatorial Cloud Dynamics.................... 368 16.3.2 Using Lidar to Observe Vertically Resolved Features of the Global Atmosphere....... 368 16.3.3 Monitoring Changes to Carbon Dioxide Concentrations During El Niño................ 369 16.4 The Future of Satellite Observations in Climate Research................................................... 370 References....................................................................................................................................... 371

16.1 A “PRE-HISTORY” OF CLIMATE-MONITORING SATELLITES On July 23, 1972, the “Earth Resources Technology Satellite” (ERTS-A, seen in Figure 16.1) was launched from Vandenberg Air Force Base in Southern California, and, once in orbit, became the first-ever satellite with the express purpose of studying and monitoring the planet remotely (Colvocoresses 1970). In 1975, ERTS-A was renamed “Landsat 1” and, while Landsat 1 has since been decommissioned, six similarly purposed Landsat satellites have been launched into orbit as part of a multi-decadal effort to preserve the continuity of the climate data record begun by ERTS-A. The newest of these satellites (Landsat 7 and 8) continue to observe the Earth from a nearpolar, sun-synchronous orbit, 438 miles above the planet, providing valuable moderate-resolution imagery of Earth’s surface. The instruments on each satellite cover the entire Earth (from 81°S to 81°N) every 16 days, and are offset in orbit by eight days. The Landsat 8 satellite (launched in 2013 and originally called the Landsat Data Continuity Mission) is the newest in the series, and not only extends the 40 year Landsat observation record but also includes new capabilities to measure cirrus clouds and aerosols, and to investigate coastal zones and water management (Roy et al. 2014; https://landsat.usgs.gov/landsat-8). The consistent, long-term terrestrial data record that began with Landsat 1 over 45 years ago, and extends to the present with the continued operation of both Landsat 7 and 8, plays an ongoing and outsized role in the monitoring of anthropogenic climate change, specifically changes to the Earth’s surface (Roy et al. 2014; https://www.space. com/19716-nasa-landsat-satellites-earth-space-infographic.html). Since Landsat 1 was launched over 45 years ago, dozens of other climate-monitoring satellites have entered orbit and continue to be maintained by agencies from the United States, Europe, India, Russia, and Japan. One of the most important long-term climate datasets in the contemporary satellite era began in 1978 with the launch of the scanning multichannel microwave sensor (SMMR) on board the NIMBUS-7 satellite (Gloersen 1984). Between 1978 and 1987, the SMMR instrument used passive microwave sensing to detect changes to Arctic sea ice extent (even when the Arctic was completely shrouded in darkness from October through March). The archive of Arctic sea ice extent data was extended to the present with the launch, in 1987, of the Special Sensor Microwave/Imager (SSM/I) instrument flown on the United States Air Force Defense Meteorological Satellite Program (DMSP) Block 5D-2 satellites (Flights F8-F15, with the exception of F-9). 363

364

Climate Changes in the Holocene

FIGURE 16.1  Workers in lab coats around the Earth Resources Technology Satellite (ERTS-A, Landsat 1) in flight configuration with solar panels deployed. (Image courtesy of Smithsonian Institution, National Air and Space Museum.)

Data from both the SMMR and SSM/I have been combined to create a 40-year continuous time series of Arctic sea ice extent. As of 2018, this is the most complete and scientifically significant sea-ice record available to climate scientists. During this period, the average late summer (September) extent of Arctic sea ice has declined from nearly eight million square kilometers in 1979 to under five million square kilometers in 2017—a staggering loss of over 13% per decade (Serreze and Stroeve 2015). Viewed in context with future projections of ice loss due to climate change, this satellite-observed decline is faster than expected and accelerating (see Figure 16.2). This and other similar datasets exist because of the continuous operation and maintenance of Earth-observing satellites since the 1970s. These datasets have greatly improved our understanding of the Arctic sea ice system and the representation of the sea-ice—albedo feedback in advanced Earth system models. In addition to the Landsat and DMSP satellites launched in the 1970s, the United States National Oceanic and Atmospheric Administration (NOAA) launched a series of polar-orbiting satellites—beginning in 1979—which were equipped with microwave sounding units (MSUs) used to derive the temperature of the troposphere via passive microwave measurements. Initially, these satellites remained in orbit for 2–4 years before being replaced; more recent satellites containing MSU instruments have remained in orbit for up to a decade, with the most recent satellite—the NOAA Metop-B—entering orbit in 2012. Though the explicit goal of these satellites was to improve weather forecasting, the continuity of the data record beginning in 1979 has provided climate scientists with an abundance of important data. The MSU instrument, in particular, was originally designed to fill in geographical gaps between ground-based atmospheric sounding measurements (began in 1958) and constituted a major advancement in weather forecasting. In the late 1990s, NOAA began launching a series of updated advanced microwave sounding units (AMSUs), which are now used extensively in weather prediction and include the capability to derive temperature and water vapor atmospheric profiles, snow and ice coverage, cloud properties, and stratospheric ozone. The data from these instruments have

Perspectives of Climate Monitoring in the Satellite Era

365

FIGURE 16.2  Model simulations of Arctic sea ice extent for September (1900–2100) based on four twentyfirst century climate-change scenarios, compared to observations (black line) for 1953–2012. (Adapted from Stroeve et al. 2012 and obtained from the National Climate Assessment.)

been synthesized by two groups (the University of Alabama, Huntsville [UAH] and Remote Sensing Systems [RSS]) to derive a 40-year time series of tropospheric temperature. Both the RSS and UAH datasets indicate a tropospheric warming trend of about 0.14°C per decade (Santer et al. 2017). Data from Earth-orbiting microwave sounding units has also been synthesized with temperature data from ground stations to estimate global surface temperature over the last century and a half. The combination of MSU/AMSU data and surface station temperature observations has allowed for the creation of long-term global temperature data records, the most well known of which include the records from NASA (https://data.giss.nasa.gov/gistemp/), NOAA (https://www.ncdc.noaa.gov/ data-access/marineocean-data/noaa-global-surface-temperature-noaaglobaltemp), and the United Kingdom Meteorological Office’s Hadley Center (HadCRUT4; https://www.metoffice.gov.uk/ research/monitoring/climate/surface-temperature). The NASA, NOAA, and HadCRUT4 temperature records indicate a surface temperature warming trend of 0.17°, 0.16°, and 0.18°C per decade, respectively (Simmons et al. 2017). Accompanying the MSU instrument on board NOAA satellites was the Advanced Very High Resolution Radiometer (AVHRR), first launched in 1978. An improved radiometer (AVHRR/2) was launched in 1981 on the NOAA-7 satellite, and a third, more sophisticated version (AVHRR/3) was launched in 1998 on the NOAA-15 satellite. The most recent iteration of AVHRR/3 was launched in 2012 on board the Metop-B satellite; there are currently four AVHRR/3 instruments orbiting the Earth–offset by several hours–thereby ensuring no gaps in data coverage. The AVHRR instrument is capable of indirectly measuring cloud cover, “surface” temperature, and the concentration of particulates in the atmosphere (a product known as Aerosol Optical Depth [AOD]), and has primarily been used by NOAA to monitor weather systems and ocean temperatures. The AVHRR instrument (and, later on, other radiometers with higher spatial resolution designed specifically for climate monitoring) uses passive sensing of visible and near-infrared light (0.58– 3.93 µm) to derive daytime and nighttime cloud coverage, land-water boundaries, and snow and ice coverage. AVHRR also uses long-wavelength infrared (10.3–12.5 µm) passive sensing to derive surface temperature—specifically sea surface temperature—and has allowed for the development of a long-term time series of ocean temperatures (http://www.ospo.noaa.gov/Products/ocean/sst/conto ur/). Data from the suite of AVHRR instruments has also been used to construct a long-term record

366

Climate Changes in the Holocene

of global AOD over the oceans (Chan et al. 2013), providing insight on how human activity has affected the emission and dispersal of small particles.

16.2 THE CONTEMPORARY ERA OF CLIMATE-MONITORING SATELLITES Prior to 1997, satellites and instruments that monitored climate from space were not specifically designed to do so. Landsat, AMSU, AVHRR, and others were designed and launched as operational weather-monitoring satellites, yet, as described above, their continued maintenance and longevity in orbit have provided for the creation of several very important long-running climate data sets. Their contribution to climate science is unrivaled. However, between 1997 and 1999, NASA launched three satellites—the Quick Scatterometer (QuikSCAT), the Sea-Viewing Wide Field-of-View Sensor (SeaWiFS), and the Tropical Rainfall Measuring Mission (TRMM)—with the express purpose of monitoring the climate. All three satellites have since been decommissioned, but their legacy of monitoring wind speeds (QuikSCAT), tropical rainfall (TRMM), and ocean color (SeaWiFS) greatly expanded our understanding of the Earth system and contemporary climate change, and, subsequently, inspired a new era of climatemonitoring satellites that continue to orbit the Earth today. On December 18, 1999, NASA launched the 4,800 kg, bus-sized Terra satellite into a sun-synchronous orbit with the express purpose of researching various changes to the Earth’s climate. Terra flies at 705 km above the Earth’s surface, takes 98.8 minutes to make a full orbit around the planet, and passes over the equator around 10:30 a.m. local time. On February 24, 2000, Terra began collecting data and, as of 2018, is still in orbit with no immediate indications that it will lose altitude or be decommissioned anytime soon. Data collected by Terra and the many climate research satellites that preceded it, have forever changed the field of climate science. In fact, the successful launch and operation of Terra spawned the creation of NASA’s Earth Observing System (EOS) program, which continues today. There are five instruments on board the Terra satellite: The Advanced Spaceborne Thermal Emissions and Reflection Radiometer (ASTER), the Clouds and the Earth’s Radiant Energy System (CERES), the Multi-angle Imaging SpectroRadiometer (MISR), the Moderate-resolution Imaging Spectroradiometer (MODIS), and the Measurements of Pollution in the Troposphere (MOPITT). Data from these instruments have been used in nearly 15,000 publications, and have been cited nearly 300,000 times, demonstrating their far-reaching utility. Just one of these instruments— MODIS, which uses passive radiometric sensing in the UV, visible, and infrared wavelengths—has been used to retrieve measurements of surface albedo, NDVI, aerosol optical depth, cloud fraction, leaf area, snow cover, and cloud physical properties (e.g. Justice et al. 2002a,b; Christopher et al. 2002; Hall et al. 2002; Huete et al. 2002). Similarly, the MISR instrument, equipped with nine cameras fixed at varying angles which passively sense in three visible and one near-infrared wavelength(s), has produced data which has been used to derive cloud and aerosol properties as well as changes to land cover (Diner et al. 1998). Additionally, MISR’s nine cameras allow for the retrieval of cloud and aerosol plume heights using stereographic techniques (Kahn et al., 2008; Tosca et al. 2011). Shortly after Terra attained orbit and began collecting data, NASA and its international counterparts initiated a program that would install a constellation of satellites into a sun-synchronous orbit with an equatorial crossing at 1:30 p.m. local time. The satellites, many of which are still in commission, are closely spaced (just a few minutes apart from each other), ensuring that their collective observations, when combined, help construct a complete, three-dimensional understanding of the Earth system. This constellation of satellites is referred to as the “A-Train” and began in 2002 with the launch of the Aqua satellite. The same MODIS instrument that flies on Terra is also included on Aqua, allowing for multi-platform studies of clouds, as described later. Also on board the Aqua satellite is the AMSU instrument (described above), the Atmospheric

Perspectives of Climate Monitoring in the Satellite Era

367

Infrared Sounder (AIRS), which measures atmospheric temperature and humidity, land, and sea surface, and the Clouds and the Earth’s Radiant Energy System (CERES), which measures both reflected sunlight and Earth-emitted infrared radiation in an ongoing effort to record changes in the radiative flux at the top of the atmosphere (with direct applications to measuring and understanding climate change). Two years after Aqua began collecting data, in 2004, NASA added the Aura satellite to the A-Train. Aura, which will remain in orbit until at least 2022 (and likely beyond) follows 15 minutes behind Aqua, and its primary mission is to observe the ozone layer, air quality, and climate. In 2006, NASA (in partnership with the French-based Centre national d’etudes spatiales, CNES) added the CALIPSO satellite to the A-Train. CALIPSO, which is still in orbit, represents a significant advancement in the way we measure and understand the atmosphere’s vertical structure; the primary instrument on CALIPSO is the Cloud-Aerosol Lidar with Orthogonal Polarization (CALIOP), which retrieves high-resolution vertical profiles of aerosols and clouds. Using lidar, the CALIOP instrument employs active remote sensing and can detect clouds and aerosols at night, representing a significant advancement in our understanding of the diurnal cycle—and how it is changing—of these important atmospheric features. Also launched in 2006, and flying nearly in tandem with CALIPSO, CloudSat uses radar to measure the altitude and vertical properties of clouds. In 2018, CloudSat was removed from the A-Train and lowered to a different orbit. More recently, NASA added the Orbiting Carbon Observatory 2 (OCO-2) satellite, and the Japanese Aerospace Exploration Agency (JAXA) added the Global Change Observation Mission (GCOM-W1) satellite to the A-Train formation. OCO-2 passively measures atmospheric concentrations of carbon dioxide by what amount of sunlight reflected by Earth’s surface is absorbed by carbon dioxide in the troposphere. This technique is employed through three spectral channels—one in the visible spectrum which measures oxygen, and two in the near-infrared (≤ 3µm) which measure carbon dioxide—and the retrieval algorithm calculates a dry-air mole fraction of carbon dioxide which is used to express the more commonly understood concentration (Frankenberg et al. 2015). Finally, not included in the A-Train, but vitally important to our understanding of global environmental change, is the Gravity Recovery and Climate Experiment (GRACE; launched in 2002; retired in 2017), and its recent follow-on, GRACE II (launched in 2016). GRACE measured the distance between a pair of satellites (GRACE-1 and GRACE-2), and used these measurements to calculate changes to Earth’s gravitational field. In addition to providing valuable information on the solid Earth, GRACE also measured how water and ice moved on Earth’s surface in response to changing climate. GRACE measurements have provided insight on how large-scale underground aquifers are changing (Famiglietti 2014; Richey et al. 2015), the dynamics of the melting Greenland and Antarctic ice sheets (Rodell et al. 2015), and the rate at which the oceans are warming and rising (Reager et al. 2016). The contributions of the GRACE and GRACE-II satellites to our understanding of the global water cycle and how it is changing in the twenty-first century is unrivaled in the field of climate science.

16.3 SATELLITE-BASED OBSERVATIONAL INVESTIGATIONS OF CLIMATE CHANGE Satellites in the contemporary era have proven to be critical instruments for studying and understanding global and regional climate change. As humans continue to dramatically change Earth’s climate, our ability to observe and understand these changes becomes increasingly important. In this section, a few selected experiments are discussed in an effort to provide a broad look at the various intersecting ways that satellites can substantially improve our understanding and representation of the planet’s complex environment processes.

368

Climate Changes in the Holocene

16.3.1 Combining Terra and Aqua to Reconcile Equatorial Cloud Dynamics The Terra satellite, the inaugural member of NASA’s Earth Observing System (EOS) and in orbit since late 1999, has an equatorial crossing time of 10:30 a.m. local time, and crosses from north to south (descending node). Three hours later, at 1:30 p.m. local time, the Aqua satellite, the original member of NASA’s A-Train constellation and the third EOS satellite, crosses the equator from south to north (ascending node). There are two versions of the same radiometric instrument (MODIS), on both Terra and Aqua, and between about 30°S and 30°N, the measurement swaths of MODIS-Terra and MODIS-Aqua overlap. Atmospheric features that MODIS observes at 10:30 a.m. can then be compared to atmospheric features at 1:30 p.m. from the same instrument in the same geographical location. Specifically, MODIS passively detects cloud fraction, and this unique configuration provides scientists a way to observe changing diurnal cloud dynamics in an oft-understudied part of the world. Tosca et al. (2014) exploited the known Terra/Aqua overlap in the low latitudes and researched the impact that human-initiated burning of the African savanna and tropical forest—and subsequent release of fire-generated smoke particles (aerosols)—has on dry season cloud growth. The authors conclude that fire aerosols stunt the growth of afternoon cumulus clouds (and rainfall) in the north African savanna region. The study notes that the change in cloud fraction from 10:30 a.m. and 1:30 p.m., as observed by MODIS, cannot be fully attributed to fire aerosol-modified cloud dynamics, and apply a comprehensive adjustment to the observed change in cloud fraction to account for mesoscale meteorology. In a follow-up manuscript, Tosca et al. (2015) combine the results from Tosca et al. (2014) with aerosol data from the MISR instrument on Terra and meteorological data assimilated from the European Organisation for the Exploitation of Meteorological Satellites’ (EUMETSAT) Meteosat-8, 9, 10, and 11 satellites, and deduce that fire-driven changes to cloud coverage in north Africa are a result of atmospheric heating and increased subsidence. A representation of this process is shown in Figure 16.3. This result suggests the existence of a feedback loop between landscape fire and rainfall that could help explain extreme burning events in tropical regions.

16.3.2 Using Lidar to Observe Vertically Resolved Features of the Global Atmosphere Monitoring Earth’s climate from space was limited to two dimensions until the launch of the CALIPSO satellite—equipped with the CALIOP lidar instrument—in 2006. CALIPSO provided the first vertically resolved look at the global atmosphere and has since expanded our understanding of the vertical distribution of aerosols, the existence of sub-visible cirrus clouds, and the formation and dissipation of hurricanes, among many other processes. For example, Marquis et al. (2017) use CALIPSO to show that radiometric observations of sea surface temperatures were biased due to the presence of unscreened optically thin cirrus clouds (seen only using lidar). Stubenrauch et al. (2010) used CALIPSO to validate a six-year cloud climatology and to explore new dynamics (such as sub-visible cirrus cloud-tops) that prior satellites were unable to see. Tosca et al. (2017) used CALIPSO data to evaluate the climate response to the decline of industrial aerosols in the Southeast United States in the twenty-first century. They concluded that, while cleaning the air (reducing the aerosol optical depth) improved air quality, it also allowed for rapid global warming at the surface. The inclusion of vertically resolved aerosol information (from CALIPSO) provided the authors with the opportunity to validate their hypothesis using a radiative transfer model; the inclusion of CALIPSO data improved the robustness of the results. The conclusions of this particular study, and the implication that better air quality may accelerate surface global warming, is relevant for other polluted regions of the world, like eastern China and the Indo-Gangetic plain, that are currently working to improve air quality.

Perspectives of Climate Monitoring in the Satellite Era

369

FIGURE 16.3  A simplified depiction of the process described in Tosca et al. (2014) and Tosca et al. (2015) showing how fire aerosols inhibit cloud growth (and subsequently, rainfall) in the north African savanna burning region during the dry season. (Image created by M.G. Tosca.)

16.3.3 Monitoring Changes to Carbon Dioxide Concentrations During El Niño It is widely known that El Niño-La Niña events are correlated with changes to the carbon dioxide growth rate in the atmosphere, but, prior to 2015, less was known about the causal relationship between the two. For example, during the 2011 La Niña, only 34% of anthropogenic emissions remained in the atmosphere, whereas during the 2015–2016 El Niño, 56% did, suggesting that tropical processes during El Niño events outweighed strong land uptake of carbon dioxide in the extratropics (Poulter et al. 2014; Liu et al. 2017). Fortunately, the OCO-2 satellite, which measures carbon dioxide concentrations in the atmosphere, attained orbit shortly before the 2015–2016 El Niño—the largest El Niño since 1997–1998 and the second strongest since the 1950s—and offered climate scientists an unprecedented opportunity to diagnose how and why fluctuations in the carbon cycle respond to El Niño. Liu et al. (2017) examined the carbon cycle response to the 2015–2016 El Niño by combining satellite and climate model observations of precipitation, temperature, and carbon dioxide (from OCO-2) over the three tropical continents (Africa, South America, Asia). Their results show that no single dominant process determines carbon cycle interannual variability (Figure 16.4). On all three continents, El Niño conditions produce a net flux of carbon dioxide into the atmosphere. However, in tropical Asia, increased landscape fire is the primary driver of this increased flux, while in Africa, elevated plant respiration appears to be the dominant process, and in South America a decline in gross primary production in tropical forests results in less absorption of anthropogenic carbon dioxide. The authors warn that the 2015–2016 El Niño is a proxy for projected future climate

370

Climate Changes in the Holocene

FIGURE 16.4  Schematic showing climate anomaly patterns of the three tropical continents for the 2015 El Niño showing the relative increase or decrease in net carbon flux relative to the 2011 La Niña, in gigatons of carbon. (Figure from Liu et al. 2017.)

conditions in the tropical oceans and that the role of tropical forests as a net sink of atmospheric carbon dioxide and a buffer for industrial carbon emissions may be greatly diminished, or even eliminated, in the latter half of this century.

16.4 THE FUTURE OF SATELLITE OBSERVATIONS IN CLIMATE RESEARCH The continued funding and maintenance of Earth-observing satellite missions is paramount for weather and climate science research. The foresight of scientists in the 1970s, who proposed and launched the first Earth-monitoring satellites, and the creation of many Earth-observing programs (including NASA’s EOS) in the 1990s that funded satellites designed specifically to monitor the climate system, provided for the very important long-term satellite records (like the Arctic sea ice extent record described above) that are now threatened by funding lapses. The uninterrupted continuation of satellite-derived climate data records begun in previous decades is vital to understanding the ways our planet will change in the next century. The long-term record of sea ice extent is most at risk; the satellite program that has maintained and operated the SSM/I instrument for over three decades is nearing the end of its lifetime, and a replacement mission is not scheduled to launch until at least 2022. This gap in satellite coverage presents not only a short-term time-series continuation problem but would also compromise future research. Measurements from new instrument missions are calibrated to measurements from the mission they replace, and this calibration is crucial for continuity in the record. Currently there are Arctic ice-observing instruments aboard three aging U.S. satellites. If all three die before the replacement satellite is launched, takes its first measurements, and is calibrated, a lapse in data will threaten the integrity of the long-term record—the longest available. American scientists are currently preparing for the worst-case scenario by exploring ways to incorporate data from Japan’s microwave sensor and, potentially, from China’s Fengyun satellite series (though the United States Congress restricts NASA scientists from directly collaborating with Chinese scientists) into the current record (https://www.scientificamerican.com/article/ ageing-satellites-put-crucial-sea-ice-climate-record-at-risk/). Climate-monitoring satellites have substantially improved our understanding of Earth’s changing climate. As we’ve discussed, satellites have elucidated a deeper understanding of the dynamics of decreasing Arctic sea ice, changing patterns of landscape fire, interannual variability of the atmospheric carbon dioxide growth rate, the vertical structure of clouds, global temperature warming, and much more. The once-emerging field of climate science was transformed in the last several decades by the wealth of data derived from the satellites and instruments that have orbited—and continue to orbit—the planet. With appropriate foresight and political commitment, we can continue

Perspectives of Climate Monitoring in the Satellite Era

371

to fund and maintain climate-monitoring satellites and promise future scientists the data they will need to observe our rapidly changing climate in the twenty-first century.

REFERENCES Chan, P.K., X.-P. Zhao, and A.K. Heidinger. 2013. Long-term aerosol climate data record derived from operational AVHRR satellite observations. Dataset Papers in Geosciences 140791, doi:10.7167/2013/140791. Christopher, S.A., and J.L. Zhang. 2002, Shortwave aerosol radiative forcing from MODIS and CERES observations over the oceans, Geophysical Research Letters 29(18): 1859. Colvocoresses, A.P. 1970. ERTS-A satellite imagery. Photogrammetric Engineering 36: 555–561. Diner, D.J., J.C. Beckert, T.H. Reilly, C.J. Bruegge, J.E. Conel, R. Kahn, J.V. Martonchik, et al. 1998. Multiangle Imaging SpectroRadiometer (MISR) description and experiment overview, IEEE Transactions on Geoscience and Remote Sensing 36(4): 1072–1087. doi:10.1109/36.700992. Famiglietti, J. 2014. The global groundwater crisis, Nature Climate Change 4(11): 945–948, doi:10.1038/ nclimate2425. Frankenberg, C., R. Pollock, R.A.M. Lee, R. Rosenberg, J.-F. Blavier, D. Crisp, C.W. O’Dell, et al. 2015. The Orbiting Carbon Observatory (OCO-2): Spectrometer performance evaluation using pre-launch direct sun measurements, Atmospheric Measurement Techniques 8: 301–313. doi:10.5194/amt-8–301-2015. Gloersen, P, D.J. Cavalieri, A.T.C. Chang, T.T. Wilheit, W.J. Campbell, O.M. Johannessen, K.B. Katsaros, et al. 1984. A summary of results from the first NUMBUS 7 SMMR observations, Journal of Geophysical Research 89: 5335–5344. doi:10.1029/JD089iD04p05335. Hall, D.K., G.A. Riggs, V.V. Salomonson, N.E. DiGirolamo, and K.J. Bayr. 2002. MODIS snow-cover products, Remote Sensing of Environment 83: 181–194. Huete, A, K. Didan, T. Miura, E.P. Rodriguez, X. Gao, and L.G. Ferreira. 2002. Overview of the radiometric and biophysical performance of the MODIS vegetation indices, Remote Sensing of Environment 83 (2-Jan): 195–213. Justice, C.O., L. Giglio, S. Korontzi, J. Owens, J.T. Morisette, D. Roy, J. Descloitres, et al. 2002b. The MODIS fire products. Remote Sensing of Environment 83: 244–262. Justice, C.O., J.R.G. Townshend, E.F. Vermote, E. Masuoka, R.E. Wolfe, N. Saleous, D.P. Roy, et al. 2002a. An overview of MODIS Land data processing and product status. Remote Sensing of Environment 83: 3–15. Kahn, R.A., Y. Chen, D.L. Nelson, F.-K. Leung, Q. Li, D.J. Diner, and J.A. Logan. 2008. Wildfire smoke injection heights: Two perspectives from space. Geophysical Research Letters 35. doi: 10.1029/2007GL032165. Liu, J., K.W. Bowman, D.S. Schimel, N.C. Parazoo, Z. Jiang, M. Lee, A.A. Bloom, et al. 2017. Contrasting carbon cycle responses of the tropical continents to the 2015–2016 El Niño. Science 358(6360). doi:10.1126/ science.aam5690. Marquis, J.W., A.S. Bogdanoff, J.R. Campbell, J.A. Cummings, D.L. Westphal, N.J. Smith, and J. Zhang. 2017. Estimating infrared radiometric sea surface temperature retrieval cold biases in the tropics due to unscreened optically thin cirrus clouds. Journal of Atmospheric and Oceanic Technology 34: 355–373. doi:10.1175/JTECH-D-15-0226.1. Poulter, B., D. Frank, P. Ciais, R.B. Mynemi, N. Andela, J. Bi, G. Broquet, et al. 2014. Contribution of semiarid ecosystems to interannual variability of the global carbon cycle, Nature 509: 600–603. doi:10.1038/ nature13376. Reager, J.T.G., A.S. Gardner, J.S. Famiglietti, D.N. Weiss, A. Eicker, M.H. Lo. 2016. A decade of sea level rise slowed by climate-driven hydrology. Science 351(6274): 699–703. doi:10.1126/science. aad8386. Richey, A.S, B.F. Thomas, M.-H. Lo, J.T. Reager, J.S. Famiglietti, K. Voss, S. Swenson, et al. 2015. Quantifying renewable groundwater stress with GRACE. Water Resources Research 51(7): 5217–5238. doi:10.1002/2015WR017349. Rodell, M., H.K. Beaudoing, T.S. L’Ecuyer, W.S. Olson, J.S. Famiglietti, P.R. Houser, R. Adler, et al. 2015. The observed state of the water cycle in the early twenty-first century. Journal of Climate 28: 8289– 8318. doi:10.1175/JCLI-D-14-00555.1. Roy, D.P., M.A. Wulder, T.R. Loveland, C.E. Woodcock, R.G. Allen, M.C. Anderson, D. Helder, et al. 2014, Landsat-8: Science and product vision for terrestrial global change research. Remote Sensing of Environment 145: 154–172, doi:10.1016/j.rse.2014.02.001. Santer, B.D., S. Solomon, F.J. Wentz, Q. Fu, S. Po-Chedley, C. Mears, J.F. Painter, et al. 2017. Tropospheric warming over the past two decades. Scientific Reports 7: 2336. doi:10.1038/s41598–017-02520-7.

372

Climate Changes in the Holocene

Serreze, M.C., and J. Stroeve (2015), Arctic sea ice trends, variability, and implications for seasonal forecasting, Philosophical Transactions of the Royal Society A 373. doi:10.1098/rsta.2014.0159. Simmons, A.J., P. Berrisford, D.P. Dee, H. Hersbach, S. Hirahara, and J.-N. Thepaut. 2017. A reassessment of temperature variations and trends from global reanalyses and monthly surface climatological datasets, Quarterly Journal of the Royal Meteorological Society 143: 101–119. doi:10.1002/qj.2949. Stroeve, J. C., V. Kattsov, A. Barrett, M. Serreze, T. Pavlova, M. Holland, and W. N. Meier. 2012. Trends in Arctic sea ice extent from CMIP5, CMIP3 and observations. Geophysical Research Letters 39: L16502. doi:10.1029/2012GL052676. Stubenrauch, C.J., S. Cros, A. Guignard, N. Lamquin. 2010, A 6-year global cloud climatology from the Atmospheric InfraRed Sounder AIRS and a statistical analysis in synergy with CALIPSO and CloudSat, Atmospheric Chemistry and Physics 10: 7197–7214. doi:10.5194/acp-10–7197-2010. Tosca, M.G., J.R. Campbell, M.J. Garay, S. Lolli, F.C. Seidel, J. Marquis, and O.V. Kalashnikova. 2017. Attributing accelerated summertime warming in the Southeast United States to recent reductions in aerosol burden: Indications from vertically-resolved observations. Remote Sensing 9(7): 674. doi:10.3390/ rs9070674. Tosca, M.G., D.J. Diner, M.J. Garay, and O.V. Kalashnikova. 2014. Observational evidence of fire-driven reduction of cloud fraction in tropical Africa. Journal of Geophysical Research 119: 8418–8432. doi:10.1002/2014JD021759. Tosca, M.G., O.V. Kalashnikova, M.J. Garay, D.J. Diner, and J.T. Randerson. 2015. Human caused fires limit convection in tropical Africa: First temporal observations and attribution. Geophysical Research Letters 42. doi:10.1002/2015GL065063. Tosca, M.G., J.T. Randerson, C.S. Zender, D.L. Nelson, D.J. Diner, and J.A. Logan. 2011. Dynamics of fire plumes and smoke clouds associated with peat and deforestation fires in Indonesia. Journal of Geophysical Research 116. doi:10.1029/2010JD015148.

17

Perspectives of Clean Energy and Carbon Dioxide Capture, Storage and Utilization Nikolaos Koukouzas, Vasiliki Gemeni, and Nikolaos Tsoukalas

CONTENTS 17.1 Clean Energy......................................................................................................................... 373 17.2 Carbon Dioxide Capture Storage and Utilization.................................................................. 375 17.2.1 CO2 Capture Technologies......................................................................................... 375 17.2.1.1 Pre-Combustion Capture............................................................................ 376 17.2.1.2 Post-Combustion Capture........................................................................... 377 17.2.1.3 Oxy-Fuel Combustion Capture................................................................... 377 17.2.2 CO2 Geological Storage............................................................................................. 378 17.2.2.1 Storage in Deep Saline Formations............................................................ 378 17.2.3 CO2 Utilization.......................................................................................................... 379 17.2.3.1 Direct Utilization of CO2............................................................................ 379 17.2.3.2 Enhanced Oil and Coal-Bed Methane Recovery........................................ 379 17.2.3.3 Conversion of CO2 into Chemicals and Fuels............................................. 379 17.2.3.4 Mineral Carbonation................................................................................... 380 17.2.3.5 Biofuels from Microalgae........................................................................... 380 17.3 Implementation Scale............................................................................................................ 380 17.3.1 CO2 Geological Storage............................................................................................. 380 17.3.1.1 Safety of Geological Storage...................................................................... 381 17.3.1.2 CO2-Enhanced Oil Recovery (CO2-EOR).................................................. 381 17.3.2 CO2 Utilization.......................................................................................................... 381 17.4 Implications and Barriers to the Implementation of CCUS.................................................. 382 17.5 Conclusions............................................................................................................................ 383 References....................................................................................................................................... 383

17.1 CLEAN ENERGY The rapid socioeconomic development that took place in the last few decades is responsible for considerable changes that impact our society. Moreover, the number of people with access to electricity has increased by 1.7 billion since the 1990s, and as the global population continues to rise, so will the demand for cheap energy. A global economy reliant on fossil fuels, and the increase of greenhouse gas (GHG) emissions is creating drastic changes to our climate system. Based on this, the International Energy Agency (IEA 2017) states that the global demand for energy will increase tremendously in the following years. Unfortunately, this phenomenon has its negative implications on a global scale, affecting air quality degradation and global warming, which it is thought will rise to levels which Earth last faced thousands of years ago, albeit without the man’s influence then, and thus without us having seen the effects on mankind’s history.

373

374

Climate Changes in the Holocene

Efforts to encourage clean energy has resulted in more than 20% of global power being generated by renewable sources as of 2011. However, according to the United Nations Development Programme (Goal 7: Affordable and clean energy, http://www.undp.org/content/undp/en/home/ sustainable-development-goals/goal-7-affordable-and-clean-energy.html), one in seven people still lack access to electricity, and as the demand continues to rise, there needs to be a substantial increase in the production of renewable energy across the world. While impressive progress has been made in developing clean energy technologies in recent years, the success stories are overshadowed by surging demand for fossil fuels, which are outstripping deployment of clean energy technologies. Coal has met 47% of the global new electricity demand since the turn of the century, eclipsing clean energy efforts made over the same period of time, which include improved implementation of energy efficiency measures and rapid growth in the use of renewable energy sources. The 2020 climate and energy package, a set of measures agreed by the EU in 2008 to implement the so-called “20–20–20” targets, aims at transforming Europe into a highly energy-efficient, low carbon economy (EC 2008). Binding targets to be met by 2020 are 1. a reduction in EU GHG emissions of at least 20% below 1990 levels; 2. 20% of EU energy consumption coming from renewable resources; and 3. a 20% reduction in primary energy use compared with projected levels, achieved by improving energy efficiency (Czernichowski-Lauriola and Stead 2014). In the last few decades, many alternative forms of energy have emerged as advances in clean energy. These clean energy systems should have the following capabilities: • • • •

emissions reduction by taking the advantage of renewable and cleaner sources; lesser energy consumption supplies; increased efficiency based on useful outputs (i.e., multigeneration); emissions and waste reduction by recovering energy (Dincer and Acar 2015).

The International Energy Association supports the description of energy derived from natural processes using continually replenished sources as “renewable energy” (IEA 2014). This energy may derive from sources as a consequence of solar radiation (e.g., hydro, wind, wave and biomass) or it can be non-solar (e.g., geothermal, tidal, and oceanic). Types of renewable energies along with their output types are shown in Figure 17.1.

FIGURE 17.1  Renewable energy sources and their associated outputs. (Modified from Dincer et al. 2015.)

Perspectives of Clean Energy and Carbon Dioxide Capture

375

FIGURE 17.2  Carbon capture, storage and utilization. (Modified from Cuellar-Franca et al. 2015.)

17.2 CARBON DIOXIDE CAPTURE STORAGE AND UTILIZATION Carbon dioxide capture storage (CCS) and carbon dioxide utilization (CCU) aim to capture CO2 emissions from point sources such as power plants and industrial processes, thus contributing to the mitigation measures related to the reduction of GHG emissions (Cuellar-Franca and Azapagic 2015). The difference between CCS and CCU is in the final destination of the captured CO2. In CCS, CO2 is captured and transferred to a selected site for long-term geological storage (Metz et al. 2005; Meylan et al. 2015), while in CCU, captured CO2 is used to produce products with the objective of an economic benefit (Markewitz et al. 2012; EIT Climate-KIC 2015). CCU is aligned with the industrial ecology strategies, taking into consideration material and energy constraints, enabling the industrial utilization of the carbon dioxide to provide various final products. On the other hand, the geological sequestration is an important option that can be economically interesting when using CO2 for Enhanced Oil Recovery (EOR), a technique between sequestration and utilization. Currently, the proportion of CO2 used by industrial activities (approximately 0.2 Gt/yr) is considerably smaller compared to the anthropogenic emissions (more than 30 Gt/yr). However, new processes are being developed to increase this proportion and reduce the environmental impact. Different CCS and CCU options are summarized and described in the following Figure 17.2.

17.2.1 CO2 Capture Technologies The carbon dioxide capture procedure is a valuable means of reducing greenhouse gases in the atmosphere and restricting the planet’s global warming. Therefore, in 2005, the IPCC report for Carbon Capture and Storage (CCS) indicated that the procedure is applicable in large point sources,

376

Climate Changes in the Holocene

such as electrical power generation constitutes (Metz 2005). Other main industrial sources are oil refineries, biogas sweetening, and petrochemical, cement and iron plants (Styring et al. 2011; Markewitz et al. 2012). In the US, greenhouse gas emissions from steam power plants makes up about 29% of all GHG emissions over the course of a year, mostly arising from the combustion of fossil fuels such as coal and natural gas (US Environmental Protection Agency 2017). Capture is the first critical step of CCS activities, storing CO2 in geological structures, both onshore and offshore, which would be geologically secure without any CO2 leaks from the reservoir to the overlying formations and at a later stage to the sea or atmosphere (Bachu et al. 2003; Bentham and Kirby 2005). The captured CO2 can also be used as feedstock for industrial processes and enhanced oil recovery. The great variety of plants and thus the diversity of gas emissions indicates that one-size-fits-all technology is not practically applicable. For this reason, there is a wide variety of CO2 capturing systems, to ensure compatibility with the specific industry (UNIDO 2011). However, the level of maturity among different capturing systems varies across industries (Cuellar-Franca and Azapagic 2015). The CO2 capturing process can follow three main procedures and thus is classified as precombustion, post-combustion, and post oxy-fuel combustion (Figueroa et al. 2008). These three technologies are discussed below and in Table 17.1. The combustion process is what directly affects the applicable CO2 capture route. CO2 capture technologies are nowadays available but they still remain costly, as they require 70%–80% of the total cost of a full CCS system, including capture, transport and storage (Blomen et al. 2009). Consequently, significant efforts are focusing on the reduction of operating costs and energy consumptions. 17.2.1.1 Pre-Combustion Capture Pre-combustion capture refers to CO2 capturing generated as an undesired co-product of an intermediate reaction of a conversion process (UNIDO 2011). Examples where this technique is applicable is the production of ammonia and coal gasification in power plants (Pehnt and Henkel 2009). Prior to ammonia synthesis, CO2 that is co-produced with hydrogen must be removed. Organic framework membranes are also applicable for CO2 capture, due to their high CO2 selectivity and uptake; however, no applications have been reported lately (Gao et al. 2014).

TABLE 17.1 CO2 Capture Options, Technologies, Methods and Applications Capture Option Post-conversion

Pre-conversion

Oxy-fuel combustion

Separation Technology

Applications

Absorption by chemical solventsa

Power plants; iron and steel industry; cement industry; oil refineries

Adsorption by solid sorbents Membrane separation Cryogenic separation Absorption by physical solvents Absorption by chemical solvents Adsorption by porous organic frameworks Separation of oxygen from air

No application reported Power plants; natural gas sweetening Power plants; iron and steel industry Power plants (IGCC)

(Data from Cuellar-Franca et al. 2015) a Mature technology b May become available in the long term (>2030)

Ammonia production Gas separations Power plants; iron and steel industry; cement industryb; syngas production and upgrading

Perspectives of Clean Energy and Carbon Dioxide Capture

377

A noteworthy drawback of the pre-combustion capture option is the energy consumption for regeneration of chemical solvents (e.g., amine-based solvents); however, this is lower for the physical solvents, as they are regenerated by reducing pressure rather than by heat. It is understood that physical solvents are more suitable for applications with high operating pressure. One of the most significant and widespread technical methods for post-combustion capture is the IGCC process, in which the O2 is separated from the air. The mixture of coal and oxygen at high temperatures (as high as 1350˚C), causes a partial combustion thereby producing a combination of gases containing CO, CO2, H2, COS, and H2O which is known as syngas. A clean-up method is then followed, producing syngas primarily containing pure H2 and CO2, as water, mercury and sulfur have been successfully removed. The CO2 removal is achieved by a solvent-based process. The significance of this solvent is that has the ability to remove from syngas both H2S and CO2, using two separate absorbers in series. It is remarkable that the pressure of the recovered CO2 gas from the flash drums is moderately high, and therefore the expected costs of CO2 compression are much lower than for SCPC and NGCC post-combustion approaches that capture the CO2 at nearatmospheric flue gas pressures (Adams et al. 2017). In 2009, Hoffmann et al. (2009) conducted a pilot natural gas project followed by cost analysis on advanced combined cycle gas turbine plants, with a pre-combustion CO2 capture system, obtaining a CO2 capture efficiency of 80%. The cost of CO2 reached $29/t CO2 for an advanced design concept (Leung et al. 2014). 17.2.1.2 Post-Combustion Capture Post-combustion technologies are the preferred option for modifying existing equipment and structures in existing power plants (Leung et al. 2014). In this process, CO2 is sequestrated from flue gas after fuels have completely burned, and after combustion has taken place (Wang et al. 2017). The challenge for post-combustion CO2 capture is its large parasitic load and this is because the CO2 level in combustion flue gas is normally low (i.e., 7%–14% for coal-fired and as low as 4% for gasfired). The energy consumption and the associated costs for the capture unit to reach the concentration of CO2 needed for transport and storage are high (de Visser et al. 2008; ICF 2009; Olajire 2010). The U.S. National Energy Technology Laboratory holds that CO2 post-combustion capture would drive a tremendous increase regarding the cost of electricity production that could reach up to 70% (Elwell and Grant 2006). Post-combustion capture as an “end-of-pipe” technology is a viable option for existing coal-fired power plants, because all conventional coal-fired power plants combust fuels directly in a boiler for power generation (Merkel et al. 2010). During CO2 capture, the gas leaving the final cleaning stage is sent to a solvent-based CO2 capture process. At this stage, CO2 is isolated for later compression, transport, and storage. The rest of the gases are released to the atmosphere. Many post-combustion capture technologies have been investigated the last few years; however, in most of the cases these are not applicable for CO2 capture in power plants. Besides the low partial pressure of CO2 in the flue gas, an obstacle to the implementation of post-combustion capture technology in coal-fired power plants is that the flow rate of flue gas is frequently 5–10 times higher than streams usually treated in natural gas and chemical industries (Merkel et al. 2010). 17.2.1.3 Oxy-Fuel Combustion Capture The oxy-fuel combustion is a procedure that can only be functional at power generation related to fossil-fueled plants, cement production and the iron and steel industry (Cuellar-Franca and Azapagic 2015). The nitrogen component is removed from the air and therefore the combustible fuel is burnt with pure oxygen, affecting the subsequent separation process. The product is flue gas of high CO2 concentration, water, particulates and SO2. After a specific procedure, the residual gases contain a high concentration of CO2—80%–98%, depending on the fuel used (Zero 2013). This method is technically feasible, as it does not require chemicals for CO2 separation, but suffers from drawbacks. A significant problem is the large amounts of oxygen that are required to come

378

Climate Changes in the Holocene

from an energy intensive air separation unit (Robeson 1991), as oxygen is expensive and the environmental impacts related to its production are high, due to the high energy consumption required (Azapagic et al. 2005, 57). The additional cost may reach over 7% compared with a plant without CCS (Low et al. 2013; Powell and Qiao 2006). At present, there is no full-scale oxy-fuel project developed at a large scale, such as 1000–2000 MWth. However, a few—mostly in pilot-scale—commercial demonstration plants are under development worldwide, such as the 25 MWe and 250 MWe oxy-coal units planned by CS Energy and Vattenfall (Powell and Qiao 2006).

17.2.2 CO2 Geological Storage Geological storage is, at present, considered the most viable option for the storage of the large CO2 quantities needed to effectively reduce global warming and related climate change (Metz et al. 2005; Fang et al. 2010; van der Zwaan and Smekens 2009; Doughty et al. 2008; Leung et al. 2014). A typical geological storage site, such as deep saline aquifers, can hold several tens of million tonnes of CO2 trapped by different physical and chemical mechanisms (Doughty et al. 2008). It has been shown that CO2 storage potential can reach 400–10,000 GT for deep saline aquifers (IEA 2004). Suitable geological sites for CO2 storage need to be carefully selected (Figure 17.3). General requirements for geological storage of CO2 include: appropriate porosity, thickness, and permeability of the reservoir rock; a cap rock with good sealing capability; and a stable geological environment (Metz et al. 2005; Doughty et al. 2008; Gibson-Poole et al. 2008). Requirements such as distance from the source of CO2, effective storage capacity, pathways for potential leakage, and, in general, economic constraints may limit the feasibility of a place being a storage site. In addition, economic aspects related to infrastructure and socio-political conditions will also affect the site selection. 17.2.2.1 Storage in Deep Saline Formations Saline formations are deep sedimentary rocks saturated with formation waters or brines containing high concentrations of dissolved salts. These formations are widespread and contain enormous quantities of water, but are unsuitable for agriculture or human consumption (Eke et al. 2011; Celia et al. 2015). Saline aquifers are used locally by the chemical industry and formation waters of varying salinity are used in health spas and for producing low-enthalpy geothermal energy (Metz et al. 2005).

FIGURE 17.3  Overview of geological storage and enhanced oil and gas recovery cases.

Perspectives of Clean Energy and Carbon Dioxide Capture

379

17.2.3 CO2 Utilization Though significant amounts of CO2 are being currently utilized, the potential is much higher. CO2 is used as a carbon source to produce new, marketable products. It is, in essence, CO2 reuse. Carbon dioxide utilization technologies (CDU) can either give products that sequester the CO2 for a lengthy period (such as polymers or mineralization); or only for a matter of weeks or days but also perhaps between seasons, as in the case of fuels and methanol. There are many methods for CDU available, which include catalytic reduction and direct addition (Armstrong and Styring 2015). These products are used commercially, either directly or after conversion. Examples of direct utilization include its use in the food and drink industry and for EOR; CO2 can be converted into chemicals or fuels (Cuellar-Franca and Azapagic 2015). Carbon dioxide utilization is not a new technology. Carbon dioxide has been used to produce urea for many decades. Currently, CO2 utilization processes such as urea and methanol production use 122 Mt of CO2 annually (Aresta et al. 2013). This by far exceeds the current amount of CO2 captured by CCS which is 26.6 Mt/year. 17.2.3.1 Direct Utilization of CO2 Several industries utilize CO2 directly. For example, in the food and drink industry, CO2 is commonly used as a carbonating agent, a preservative, a packaging gas and as a solvent for the extraction of flavors and in the decaffeination process. Other applications can be found in the pharmaceutical industry where CO2 can be used as a respiratory stimulant, or as an intermediate in the synthesis of drugs (Yu et al. 2008). However, these applications are restricted to sources producing CO2 waste streams of high purity, such as ammonia production (Markewitz et al. 2012; Metz et al. 2005). 17.2.3.2 Enhanced Oil and Coal-Bed Methane Recovery EOR and ECBM are other examples of direct utilization of CO2 where it is used to extract crude oil from an oil field or natural gas from unmineable coal deposits, respectively. While the latter is not commercially available yet, the former has been widely practiced for over 40 years in several oilproducing countries, including Norway, Canada and the United States (Metz et al. 2005). 17.2.3.2.2 Coal Seams Many of the world’s coal reserves are too deep to exploit by conventional methods. Underground coal gasification provides access to coal deposits that would otherwise remain unused, and is also an attractive route to carbon capture and utilization (Yang et al. 2016). On the other hand, there is the enhanced coalbed methane (ECBM), a promising clean coal technology for methane recovery, as well as CO2 sequestration. Coal contains fractures (cleats) that impart some permeability to the system. Gaseous CO2 injected through wells will flow through the cleat system of the coal, diffuse into the coal matrix and be adsorbed onto the coal micropore surfaces, freeing up gases with lower affinity to coal (e.g., methane) (Metz et al. 2005). The process of CO2 trapping in coal for temperatures and pressures above the critical point is not well understood (Larsen 2003). It seems that adsorption is gradually replaced by absorption and the CO2 diffuses or “dissolves” in coal. 17.2.3.3 Conversion of CO2 into Chemicals and Fuels CO2 can also be utilized by processing and converting it into chemicals and fuels. This can be achieved through carboxylation reactions where the CO2 molecule is used as a precursor for organic compounds such as carbonates, acrylates and polymers, or reduction reactions where the C=O bonds are broken to produce chemicals such as methane, methanol, syngas, urea and formic acid (Styring et al. 2011; Yu et al. 2008; Markewitz et al. 2012). Furthermore, CO2 can be used as a feedstock to produce fuels.

380

Climate Changes in the Holocene

However, although CO2 can replace petrochemical feedstocks for production of chemicals and fuels (Styring et al. 2011; Armstrong and Styring 2015), a disadvantage is that its conversion is energy intensive and it requires high-selectivity catalysts since CO2 is thermodynamically highly stable. Furthermore, chemicals and fuels offer limited storage periods for captured CO2 because of their short life span (typically less than six months). Consequently, CO2 is released into the atmosphere before the benefits of the capture can be realized. For that reason, future research efforts should focus on the synthesis of materials and products with longer lifespans. An example is mineral carbonates that can be used in construction (Metz et al. 2005). 17.2.3.4 Mineral Carbonation Mineral carbonation (MC) is an accelerated form of weathering of naturally occurring silicate rocks and has been proposed as an alternative approach for CO2 sequestration since the 1990s (Seifritz 1990). Some MC technologies have recently approached the commercial stage. In mineral carbonation, CO2 reacts with a metal oxide, such as magnesium or calcium, to form carbonates in a chemical process. Magnesium and calcium are normally found in nature in the form of silicate minerals such as serpentine, olivine and wollastonite (Metz et al. 2005; Koukouzas et al. 2009). Moreover, MC can take advantage of different starting materials, which include Mg-silicate minerals and Caor Fe-rich silicates (Sanna et al. 2014).

Metal oxide + CO2 → Metal carbonate + Heat

The use of pure CO2 is not essential for mineral carbonation as the presence of impurities such as NOx in flue gas will not interfere with the carbonation reaction (Metz et al. 2005). Therefore, the separation and capture step that produces a pure stream of CO2 can be omitted, as waste emissions containing CO2 can be used directly. Compared to these other utilization/storage options, mineral carbonation has some advantages, such as: theoretical power generation, as it is an exothermic reaction; the large storage capacity, since there is an abundance of magnesium and calcium silicate deposits worldwide; and the thermodynamic stability of the resulting solid products. As a result, the storage method is permanent, safe, and does not require monitoring of the storage site (Goff and Lackner 1998; Lackner et al. 1997; Metz et al. 2005; Sanna et al. 2014). 17.2.3.5 Biofuels from Microalgae CO2 can be used to cultivate microalgae used for the production of biofuels (Styring et al. 2011; Brennan and Owende 2010). Microalgae have the ability to fix CO2 directly from waste streams such as flue gas, as well as using nitrogen from the gas as a nutrient (Styring et al. 2011). Biochemical conversion relies on biological and chemical processes, such as anaerobic digestion, fermentation and esterification (Brennan and Owende 2010).

17.3 IMPLEMENTATION SCALE 17.3.1 CO2 Geological Storage While large-scale CCS has yet to be implemented, there are a few industrial-scale injection operations and a number of demonstration or pilot-scale injections. The Sleipner Project in the North Sea is the best available example of a CO2 storage project in a saline formation. It was the first commercial-scale project dedicated to geological CO2 storage. Approximately 1 Mt CO2 is removed annually from the produced natural gas and injected underground at Sleipner. The operation started in October 1996, and over the lifetime of the project, a total of 20 Mt CO2 is expected to be stored (Metz et al. 2005; Celia et al. 2015). Other injection operations of note include the In Salah injection in Algeria, which was operational from 2004 to 2011 and injected approximately 0.5 Mt CO2/yr;

Perspectives of Clean Energy and Carbon Dioxide Capture

381

and the more recent injection associated with the Snohvit gas project, with injection initially into the Tubaen Formation and now into the Stø Formation, both of which are under the North Sea. Injection began in 2008 and the injection rate is approximately 0.5 Mt CO2/yr (Hansen et al. 2013). 17.3.1.1 Safety of Geological Storage CO2 is likely to remain stored for millions of years; however, in any given project, the safety of storage must be ensured by monitoring the CO2 in the subsurface. Although CO2 is a non-combustible low-toxicity gas, in high concentrations, it can be dangerous to human health and the environment. Therefore, safety and risk assessments are always necessary to evaluate potential sources of CO2 leakages or seepages away from the storage complex, in order to plan remediation options. The safety of subsurface storage can be supported by the several known natural accumulations of CO2, and can be used to gather further knowledge on containment conditions (Ziogou et al. 2013). 17.3.1.2 CO2-Enhanced Oil Recovery (CO2-EOR) CO2-enhanced oil recovery (CO2-EOR) has emerged as a major option for productively utilizing CO2 emissions captured from industrial plants. Oil fields can provide secure, well-characterized sites for storing CO2, while at the same time provide revenues to offset the costs of capturing CO2. Although utilization of captured CO2 emissions for enhanced oil recovery has been underway for some time, further advances in CO2-EOR technology could significantly improve the technology’s applicability as a revenue generator for CO2 capture and a large-scale CO2 storage option (Kuuskraa et al. 2013). CO2-EOR operations have the potential to extend the economic life of individual fields by a decade or more, as proven by a number of projects where several tens of percent of additional oil reserves have been produced. The ultimate storage capacity provided by any CO2-EOR project will depend on a number of technical and economic factors (CEPAC 2014) (Figure 17.4).

17.3.2 CO2 Utilization Armstrong and Styring (2015) have produced a scenario for CO2 utilization which incorporates current uses such as urea production, and replaces fossil oils in other processes to produce a small range of organic chemicals, diesel and aviation fuel, methane (synthetic natural gas), and some polymers.

FIGURE 17.4  Schematic illustration of enhanced oil recovery and CO2 storage (https://www.globalccsinstitute.com/understanding-ccs/information-resource).

382

Climate Changes in the Holocene

Moreover, they proposed the quantity of CO2 that could be utilized at different market shares based on current levels of production and compared this against CO2 reduction targets for the EU and the world in CCS, and EU and USA overall CO2 reduction targets. It was observed that producing only 10% of each product would make significant inroads into the EU CCS target or exceed it. Although the majority of these products are produced to provide energy via combustion—hence re-releasing the CO2—the net reduction in CO2 emitted due to switching from fossil sources will be significant. Worldwide there are a number of commercial- and pilot-scale CDU projects. Carbon Recycling International in Iceland is producing 5 million liters (950 t) of renewable methanol per annum from CO2 accounting for 1.5% of world production. Bayer Material Science has recently invested 15 million in the construction of the world’s first commercial plant to produce polyols from CO2 as a precursor for CO2-based polyurethane foams. Most of CO2-EOR projects are located in the United States, and a large proportion of the CO2 used has been obtained from natural accumulations, with a smaller part coming from anthropogenic sources.

17.4 IMPLICATIONS AND BARRIERS TO THE IMPLEMENTATION OF CCUS The deployment of CCS projects worldwide is facing many challenges, including financial issues, public acceptance and the establishment of regulatory frameworks. Different legal approaches are under development in most countries that have significant potential CO2 storage resources and CCS activities. While progress is being made, the legal framework is still immature and often insufficient to assure a successful and effective permitting process. To better understand that, in 2009 the European Commission launched the Directive on the geological storage of carbon dioxide, which aimed to establish the legal framework for the environmentally safe geological storage of CO2 (EC 2009). By 2012, most European Member States with CCS projects underway had implemented the Directive (EC 2012). However, the best candidate to implement a CCS demonstration project, Germany, failed to fully transpose the European Directive before the EC deadline. In September 2011, the government rejected the budget allocated that would allow the underground storage of CO2. This caused Vattenfall to abandon its CCS demonstration project in Jänschwalde, Brandenburg, and stop the planned €1.5 billion investment (Vattenfall 2011). The project had been awarded €180 million from the EEPR and submitted an application for the European NER300 funding program. Another example is the ROAD project in the Netherlands, which filed the storage permit application in 2010. The EU Directive was implemented in the Dutch legislation in its original format without any amendment adding national provisions. However, the project still does not have the final storage permit (ROAD 2013). The high capital intensity as well as regulatory constraints seem to be the recurring element. Demonstrators, prototypes, and pilot plants are essential to convince investors and customers via proof of technology. These facilities are often very capital intensive. Additionally, there is missing access to technical expertise, as well as lack of management knowledge (EIT 2015). Conversion into CO2 utilization products requires more energy than conversion from conventional feedstocks, because of the thermodynamic stability of carbon dioxide. Research and development is focusing on catalysis and other conversion processes to reduce the amount of required energy. Moreover, the main focus for climate action policies is the reduction of CO2 emissions in the EU. The contribution to abatement is implicated by important factors like permanence of storage (permanent, semi-permanent, short cycle), possible leakage, and the overall efficiency improvement. To what extent the various CCU technologies may (potentially) contribute to emission reduction is not yet determined, at least for most technologies and in detail. This is mainly due to the fact that there is no developed, accepted and approved methodology to determine the contribution of each technology to emission reduction.

Perspectives of Clean Energy and Carbon Dioxide Capture

383

17.5 CONCLUSIONS The International Energy Agency (IEA) agrees that the global demand for energy will increase tremendously over the following years which will therefore dramatically affect CO2 emissions. In order to avoid this situation, scientists and companies work together to move towards cleaner energy sources and procedures. Clean energy is nowadays collected through efficient and environmentally sustainable technologies, such as exploiting sun power and geothermal energy directly or indirectly. CO2 capturing methods and technologies are a crucial weapon in the arsenal of clean energy. There are three main technologies that are mostly used. The first one is the pre-combustion method which works with coal-gasification combined cycle power plants. The coal is gasified to produce a synthetic gas (syngas) and thus reacting with water, yields CO2 which is captured as well as hydrogen. The post-combustion technology is the most preferred, working with liquid solvents that isolate CO2 phase from the flue gas. The third capturing method is the oxy-fuel combustion which is applicable in power and cement plants. A specific combustion environment leads to emissions of mostly CO2 and water vapor, and the CO2 can be isolated. The IEA target for CCS for 2020 is around 60 Mt CO2/year (Energy Technology Perspectives 2014). In comparison CO2 utilization projects are growing in deployment and are providing a net reduction in CO2, both by utilizing CO2 in production and by providing a new fossil-free source for these products. CDU capacity is currently higher (180 Mt/year) than operational CCS capacities (26.6 Mt/year). Carbon dioxide utilization will provide much-needed additional capacity, and profit, in the move toward a low-carbon economy. CO2 is used as a resource, not as waste. Like CCS, it should be regarded as one of the key emissions mitigation technologies in the fight against climate change (Armstrong and Styring 2015). Since CO2 utilization has become a trending topic in recent years, it can be expected that the funding support in CO2 utilization as a clean technology will continue to stay strong. The Paris Agreement signed in 2015 includes technology and finance mechanisms with a budget of at least US$100 billion per year that also supports clean tech innovations for climate change mitigation (UNFCCC Secretariat 2015). Additionally, the G7 countries agreed on the decarbonization of the world economy until 2100. Overall, both declarations shape a strong future demand of low emission technologies and circular economy concepts worldwide. CO2 utilization will only help meet climate goals if CO2 products are widely deployed. The prospects for that are good in some market segments, although in others, high costs, well-established alternatives and entrenched incumbents create barriers to market entry. Sound market strategies, targeted technological development and supportive policies all have a role in accelerating deployment and supporting CO2 utilization technologies in the market.

REFERENCES Adams II, A.A., L. Hoseinzade, P. Bhaswanth Madabhushi, and I.J. Okeke. 2017. Comparison of CO2 capture approaches for fossil-based power generation: Review and meta-study. Processes 5: 44. Aresta, M., A. Dibenedetto, and A. Angelini. 2013. The changing paradigm in CO2 utilization. Journal of CO2 Utilization 3–4: 65–73. Armstrong, K., and P. Styring. 2015. Assessing the potential of utilization and storage strategies for postcombustion CO2 emissions reduction. Frontiers in Energy Research 3(8): 1–9. Azapagic, A., R. Clift, and C. Duff. 2005. Sustainable Development in Practice: Case Studies for Engineers and Scientists, edited by A. Azapagic, S. Perdan, and R. Clift. Chichester: John Wiley & Sons. Bachu, S., and J.J. Adams. 2003. Sequestration of CO2 in geological media in response to climate change: capacity of deep saline aquifers to sequester CO2 in solution. Energy Conversion and Management 44(20): 3151–3175. Bentham, M., and G. Kirby. 2005. CO2 storage in saline aquifers, oil & gas science and Technology. IFP Energies Nouvelles 60(3): 559–567. Blomen, E., C. Hendriks, and F. Neele. 2009. Capture technologies: improvements and promising developments. Energy Procedia 1: 1505–1512.

384

Climate Changes in the Holocene

Brennan, L., and P. Owende. 2010. Biofuels from microalgae—A review of technologies for production, processing, and extractions of biofuels and co-products. Renewable and Sustainable Energy Reviews 14(2): 557–577. Celia, M.A., S. Bachu, J.M. Nordbotten, and K.W. Bandilla. 2015. Status of CO2 storage in deep saline aquifers with emphasis on modeling approaches and practical simulations. Water Resources Research 51: 6846–6892. CEPAC. 2014. Center of Excellence in Research and Innovation in Petroleum, Mineral Resources and Carbon Storage, Brazilian Atlas of CO2 Capture and Geological Storage. Cuellar-Franca, R.M., and A. Azapagic. 2015. Carbon capture, storage and utilization technologies: A critical analysis and comparison of their life cycle environmental impacts. Journal of CO2 Utilization 9: 82–102. Czernichowski-Lauriola, I., and R. Steada. 2014. Developments of CO2 geological storage in Europe and the role of CO2GeoNet. Energy Procedia 63: 8107–8115. de Visser, C. Hendriks, M. Barrio, M.J. Molnvik, G. Koeijer, S. Liljemark, Y. Le Gallo. 2008. Dynamis CO2 quality recommendations. International Journal of Greenhouse Gas Control 2(4): 478–484. Dincer, I., and C. Acar. 2015. A review on clean energy solutions for better sustainability. International Journal of Energy Research 39: 585–606. Doughty, C., B.M. Freifeld, and R.C. Trautz. 2008. Site characterization for CO2 geologic storage and vice versa: the Frio brine pilot, Texas, USA as a case study. Environmental Geology 54: 1635–1656. EC. 2008. Communication’20 20 by 2020, Europe’s climate change opportunity. COM. EC. 2009. European Commission. Directive of the European Parliament and of the Council of 23 April 2009 amending Directive 2003/87/EC. 2009/29/EC. EC. 2012. European Commission. Implementation of the CCS Directive - state of play. EIT. 2015. European Institute of Innovation and Technology Climate – KIC, The Business Side of Innovative CO2 Utilisation, Report for the wider public. Eke, P.E., M. Naylor, S. Haszeldine, and A. Curtis. 2011. CO2-Brine Surface Dissolution and Injection: CO2 Storage Enhancement. Society of Petroleum Engineers 6(1). Elwell, L.C., and W.S. Grant. 2006. Technology options for capturing CO2—Special reports. Power 150: 60–65. Fang, Y., B. Baojun, T. Dazhen, S. Dunn-Norman, and D. Wronkiewicz. 2010. Characteristics of CO2 sequestration in saline aquifers. Petroleum Science 7: 83–92. Figueroa, J. D., et al. 2008. Advances in CO2 capture technology - The U.S. Department of Energy’s Carbon Sequestration Program. International Journal of Greenhouse Gas Control 2: 9–20. Gao, X., X. Zou, H. Ma, S. Meng, and G. Zhu. 2014. Highly selective and permeable porous organic framework membrane for CO2 capture. Advanced Materials 26(22): 3644–3648. Gibson-Poole, C.M., L. Svendsen, J. Underschultz, M.N. Watson, J. Ennis-King, P.J. Van Ruth, E.J. Nelson, et  al. 2008. Site characterisation of a basin-scale CO2 geological storage system: Gippsland Basin, southeast Australia. Environmental Geology 54: 1583–1606. Goff, F., and K.S. Lackner. 1998. Carbon dioxide sequestering using ultramafic rocks. Environmental Geosciences 5(3): 89–101. Hansen, O., D. Gilding, B. Nazarian, B. Osdal, P. Ringrose, J.B. Kristoffersen, O. Eiken, et al. 2013. Snohvit: The history of injecting and storing 1 Mt CO2 in the fluvial Tubaen Fm. Energy Procedia 37: 3565–3573. Hoffmann, S., M. Bartlett, M. Finkenrath, A. Evulet, and T.P. Ursin. 2009. Performance and cost analysis of advanced combined cycles with precombustion CO2 capture. Journal of Engineering for Gas Turbines and Power 131: 021701–1 to 7. ICF. 2009. Developing a pipeline infrastructure for CO2 capture and storage; issues and challenges. Technical report prepared for INGAA Foundation (106 pp.). IEA. 2004. International Energy Agency Report – Improvements in power generation with post-combustion capture of CO2. IEA Greenhouse Gas R&D Programmes, PH4/33. IEA. 2017. International Energy Agency Technical Report. Key World Energy Statistics. http://www.iea.org/ publications/freepublications/publication/KeyWorld2014.pdf Koukouzas, N., V. Gemeni, and H.J. Ziock. 2009. Sequestration of CO2 in magnesium silicates, in Western Macedonia, Greece. International Journal of Mineral Processing 93(2): 179–186. Kuuskraa, V.A., M.L. Godec, and P. Dipietro. 2013. CO2 utilization from “next generation” CO2 enhanced oil recovery technology energy. Procedia 37: 6854–6866. Lackner, K.S., D.P. Butt, C.H. Wendt, F. Goff, and G. Guthrie. 1997. Carbon Dioxide Disposal in Mineral Form: Keeping Coal Competitive. Tech. Report No. LA-UR-97-2094 (Los Alamos National Laboratory).

Perspectives of Clean Energy and Carbon Dioxide Capture

385

Larsen, J.W. 2003. The effects of dissolved CO2 on coal structure and properties. International Journal of Coal Geology 57: 63–70. Leung, D.Y.C., G. Caramanna, and Maroto-Valer, M. 2014. An overview of current status of carbon dioxide capture and storage technologies. Renewable and Sustainable Energy Reviews 39: 426–443. Low, T.B., L. Zhao, T. Merkel, M. Weber, and D. Stolten. 2013. A parametric study of the impact of membrane materials and process operating conditions on carbon capture from humidified flue gas. Journal of Membrane Science 431: 139–155. Markewitz, P., W. Kuckshinrichs, W. Leitner, J. Linssen, P. Zapp, R. Bongartz, A. Schreiber, et al. 2012. Worldwide innovations in the development of carbon capture technologies and the utilization of CO2. Energy Environmental Science 5(6): 7281–7305. Merkel, T. C., H. Lin, X. Wei, and R. Baker. 2010. Power plant post-combustion carbon dioxide capture: An opportunity for membranes. Journal of Membrane Science 359: 126–139. Metz, B., O. Davidson, de Coninck, H., M. Loos, and L. Meyer. 2005. IPCC Special Report on Carbon Dioxide Capture and Storage, Intergovernmental Panel on Climate Change. http://www.ipcc.ch/pdf/ special-reports/srccs/srccs_wholereport.pdf. Meylan, F., V. Moreau, and S. Erkman. 2015. CO2 utilization in the perspective of industrial ecology, an overview. Journal of CO2 Utilization 12: 101–108. Olajire, A.A. 2010. CO2 capture and separation technologies for end-of-pipe application—A review. Energy 35: 2610–2628. Pehnt, M., and J. Henkel. 2009 Life cycle assessment of carbon dioxide capture and storage from lignite power plants. International Journal of Greenhouse Gas Control 3(2): 49–66. Powell, C. E., and G.G. Qiao. 2006. Polymeric CO2/N2 gas separation membranes for the capture of carbon dioxide from power plant flue gases. Journal of Membrane Science 279: 1–49. ROAD. 2013. European CCS Network: Case study of the ROAD storage permit. A report by the ROAD project, part of the European CCS Demonstration Project Network. Robeson, L. M. 1991. Correlation of separation factor versus permeability for polymeric membranes. Journal of Membrane Science 62: 165–185. Rochelle, G. T. 2009. Amine scrubbing for CO2 capture. Science 325: 1652–1654. Sanna, A., M. Uibu, G. Caramanna, R. Kuusik, and Maroto-Valerac, M.M. 2014. A review of mineral carbonation technologies to sequester CO2. Chemical Society Reviews 43 (23): 8049–8080. Seifritz, W. 1990. CO2 disposal by means of silicates. Nature 345 (6275). Styring, P., D. Jansen, H. de Coninck, H. Reith, and K. Armstrong. 2011. Carbon Capture and Utilisation in the Green Economy. Centre for Low Carbon Futures. http://co2chem.co.uk/wp-content/uploads/2012/06/ CCU%20in%20the%20green%20economy%20report.pdf. UNFCCC Secretariat. 2015. Synthesis report on the aggregate effect of the intended nationally determined contributions, FCCC/CP/2015/7. UNIDO. 2011. Technology Roadmap: Carbon Capture and Storage in Industrial Applications. United Nations Industrial Development Organization, Organisation for Economic Development & International Energy Agency. http://www.unido.org/fileadmin/user_media/News/2011/CCS_Industry_Roadmap_WEB.pdf US Environmental Protection Agency. 2017. Inventory of U.S. Greenhouse Gas Emissions and Sinks 1990– 2015; EPA 430-P-17-001; Washington, DC: EPA. van der Zwaan, B., and K. Smekens. 2009. CO2 capture and storage with leakage in an energy-climate model. Environmental Modeling and Assessment 14: 135–148. Vattenfall, 2011. Uncertainties with CCS Law Stop Vattenfall Investment in Demo Plant. https://corporate. vattenfall.com/press-and-media/press-releases/press-releases-imported/uncertainties-with-ccs-lawstop-vattenfall-investment-in-demo-plant/. Wang, Y., L. Zhao, A. Otto, M. Robinius, and D. Stolten. 2017. A review of post-combustion CO2 capture technologies from coal-fired power plants. Energy Procedia 114: 650–665. Yang, D., N. Koukouzas, M. Green, and Y. Sheng. 2016. Recent development on underground coal gasification and subsequent CO2 storage. Journal of the Energy Institute 89(4): 469–484. Yu, K., I. Curcic, J. Gabriel, and S. Tsang. 2008. Recent advances in CO2 capture and utilization. ChemSusChem 1(11): 893–899. Zero. 2013. Zero Emissions Resource Organization. http://www.zeroCO2.no. Ziogou, F., V. Gemeni, N. Koukouzas, D. De Angelis, S. Libertini, S.E. Beaubien, S. Lombardi, et al. 2013. Potential environmental impacts of CO2 leakage from the study of natural analogue sites in Europe. Energy Procedia 37: 3521–3528.

18 The Future of the Earth and What Lies Ahead?

Society as an Adaptive System Timothy Karpouzoglou and Feng Mao CONTENTS 18.1 Introduction........................................................................................................................... 387 18.2 Current Risks and Future Uncertainties for Earth and Society............................................ 388 18.3 Towards a New Understanding of “Earth and Society” as a Coupled System...................... 389 18.3.1 Resilience-Based Approaches................................................................................... 389 18.3.2 Absorptive, Adaptive and Transformative Capacities............................................... 390 18.3.3 Three Perspectives of Human–Nature Coupling and the Resilience Cube............... 390 18.4 Future Development Pathways for Improved Human–Nature Relations.............................. 392 18.4.1 Governance Implications of Future Development Pathways..................................... 392 18.4.2 Recent Innovations for Connecting Social and Natural Systems.............................. 393 18.5 Conclusions............................................................................................................................ 393 References....................................................................................................................................... 394

18.1 INTRODUCTION The current era has been described as the Anthropocene due to the powerful force that humanity is now exerting over the Earth’s environment. As the critically acclaimed historian Harari states, Homo sapiens are in a moment in history of becoming Homo deus (Harari 2015); that is, godlike creatures that have the power to alter the very core of the biological fabric of the Earth as well as stocks and flows of major elements in the planetary machinery, such as carbon (Steffen et al. 2007). From an archaeological and historical perspective this is not surprising, as humans have evolved from small communities of hunter-gatherers living in valleys, forests and mountains, to large populations of urban dwellers that dominate the Earth’s surface. Indeed, as per the latest figures, the urban share of the world population will reach 6.419 billion (66%) by 2050 (United Nations 2014). As our cities expand, so will the pressure on vital ecosystems such as wetlands, forests, lakes and oceans. Agricultural production will have to increase to achieve food security, placing enormous pressures on limited availability of land and freshwater in order to feed the expanding population. Human activities have thus become so pervasive and profound that they rival the forces of nature and are pushing the Earth into unchartered waters (Steffen et al. 2007). Rather than viewing humanity as a powerful force over the Earth, an increasing scholarly movement from the social and natural sciences increasingly positions humanity and the Earth as interacting forces that are coupled both in space and time, given that much of our human history entails efforts of decoupling from and controlling the biophysical world. This chapter therefore departs from the point of view that as societies continue to evolve there is an inherent human responsibility towards redressing the historical power imbalance between humans and nature.

387

388

Climate Changes in the Holocene

The chapter examines different bodies of literature that seek to address this challenge. First, the chapter will broadly cover some of the current risks and future uncertainties for the Earth and society. We then explore some of the recent thinking on human–environment interactions, and describe some concepts including the concept of “resilience” that we believe will become useful in establishing a new kind of shared human knowledge (gnosis) about the Earth and society. Subsequently, we explore different ways in which gnosis may be translated into practice (praxis) through placing particular attention on the governance and innovation dimensions of how to shape future pathways that place more emphasis on Earth and society interactions. We conclude the chapter with some key reflections to take forward.

18.2 CURRENT RISKS AND FUTURE UNCERTAINTIES FOR EARTH AND SOCIETY The alteration of the Earth begun with the dawn of human civilizations but has rapidly intensified since the industrial revolution. Humanity has today achieved comparatively high levels of human well-being compared to pre-industrial times, but at a high cost for the Earth. This creates a complex set of uncertainties for the future of ecosystems and the ability of the Earth to sustain human civilizations. Scientific understanding of the risks and uncertainties for Earth and society has certainly grown in recent decades. Policy and science are converging around a growing consensus on the need to drastically reduce greenhouse gas emissions in order to limit the increase in global average temperatures. This will be one of the biggest global challenges and will pose particular challenges for developed and developing countries. According to some projections, a reduction of greenhouse gases close to 70%–95% over the next 40 years will need to occur (Söderholm et al. 2011) Humanity is certainly not passive to these changes. International awareness about the risks of human activities on the Earth has been growing in recent years and is linked with unprecedented policy and global governance efforts to decarbonize economies. The Sustainable Development Goals (SDGs) and the Paris Agreement on Climate put an important emphasis on reducing humanrelated disturbances to the climate system (i.e., limiting further increases in global mean temperatures) through technological measures and new policy measures. In some sectors, these efforts are showing results. For example, globally the cost of renewable energy technologies is decreasing which makes them more competitive to fossil-based energy technologies and may mean that these will become adopted faster in certain sectors (IRENA 2018). However, there are concerns surrounding the prospects of achieving a transition in energy systems before dangerous interference with the climate system occurs (Rockström et al. 2017). We have not yet fully understood where new energy supplies will come from in order to achieve future development targets, particularly in the developing world which is now growing fast in terms of both human population and ecological footprints. Although global trends are important, it is useful to identify the particular geographies where some of these global changes will have long-lasting impacts. Deltaic regions are good examples of some of these critically vulnerable geographies. Deltas are essentially landforms that have historically offered a wider variety of benefits to society; they are, in fact, a good representation of Earth and society interaction. The deltas provide two very important resources for societies to develop—an abundant water source and fertile soils for agricultural production. These resources, water and land, have been critical for many countries to produce food. In some places, deltas have been crucial for human societies to transition from village settlements to complex city networks (Meyer and Peters 2016). However, in order for these processes to occur, coastal deltas have suffered numerous disturbances over the centuries. For example, increases in water pollution and loss of soil fertility and biodiversity as well as land subsidence are considered to lead to an increased likelihood of passing critical thresholds, otherwise referred to as “tipping points” (Renaud et al. 2013).

What Lies Ahead?

389

As in the case of deltas, similar interactions with potential for reaching tipping points have been observed in other fragile ecosystems, including wetlands and tropical forests, as well as arid and semi-arid regions where critical Earth–society interactions occur (Scheffer et al. 2015; Riehl 2017). We are only beginning to decipher the management implications associated with tipping points, particularly because they are inherently unpredictable and dynamic. But we are beginning to observe them in different types of ecosystems. For example, increasing land desertification is predicted as a result of increases in air temperature, which in turn will cause acute droughts and loss of vegetative growth in large parts of the world that are already struggling with desertification (Reynolds et al. 2007). In parts of the African continent, changes in the climate are likely to generate challenges for food security and safeguarding agricultural activities (Karpouzoglou and Barron, 2014). These effects will in turn hamper societal development and will pose substantial risks to human well-being in many countries. Similar feedbacks loops can be observed in the oceans, for example. In particular, post-industrial levels of carbon dioxide have increased, and as a result much more carbon dioxide finds its way into the oceans causing ocean acidification. This will have devastating consequences on marine ecosystems as well as for the billions of humans that depend on the ocean for food and survival (Fabry et al. 2008).

18.3 TOWARDS A NEW UNDERSTANDING OF “EARTH AND SOCIETY” AS A COUPLED SYSTEM Having examined some of the current risks and future uncertainties for Earth and society that have been discussed in the scientific literature, in this section we explore some of the recent thinking, particularly from the environmental and social sciences on how to re-think Earth–society interactions. We focus on some key concepts such as the concept of “resilience” and use the resilience cube to visualize our concepts. This section addresses the gnosis dimensions of thinking about Earth and society as an adaptive system.

18.3.1 Resilience-Based Approaches One approach that has gained increased attention in the last fifteen years is the resilience approach that emphasizes the coupling of social and ecological processes as part of intertwined social-ecological systems (Holling 1973; Gunderson 1999; Folke 2016). Resilience further emphasizes the capacity of a system to absorb disturbance and reorganize while undergoing change so as to still retain essentially the same function, structure, and feedbacks, and therefore identity (Folke 2016). From a resilience perspective, societies need to develop a different point of reference that instead of predictability and control starts with developing a range of different capacities that are tailored to context, and that emphasize flexibility and adaptability to change; in other words, learning to live with uncertainty, rather than trying to avoid it. Another fundamentally different starting point of resilience, with particular relevance for thinking around Earth–society interactions, is that it integrates a serious consideration of Earth processes in terms of understanding planetary boundaries in societal efforts to achieve human well-being, social and economic development (Steffen et al, 2015). Some of the measures that can increase the adaptiveness of societies include ecosystem restoration, institutional improvement and technical development as part of coping with global change and avoiding transgressing critical planetary boundaries. The concept of ecosystem stewardship is strongly related to resilience because it helps to understand how to foster positive human–nature relations as part of re-thinking the links between Earth and society (Chapin et al. 2010). Stewardship can also be understood as a process that allows one to “shepherd and safeguard the valuables of not just one self but also of others”, and therefore recognizes that human survival depends not only on greater care of humans in isolation (e.g., through

390

Climate Changes in the Holocene

achieving human development goals) but instead, humans in conjunction with taking greater care of the Earth (Folke et al. 2016). Human existence within the Earth requires better stewardship of the various landscapes and seascapes, as well as the ecosystem services that these provide.

18.3.2 Absorptive, Adaptive and Transformative Capacities Resilience comprises a set of three systematic capacities including absorptive, adaptive and transformative (Béné et al. 2014; Folke, 2016). Absorptive capacity is equivalent to the original notion of resilience, which refers to the ability to absorb disturbance while retaining essential structures and functions, or the ability to recover from disturbance. This can be observed in the Earth system as a common phenomenon. In freshwater ecosystems, for example, water bodies have self-purification mechanisms to mitigate the impact from human nutrient enrichment and fish populations can recover from overharvesting when these disturbances are below a certain threshold. For climate change, the rising human-induced airborne CO2 can be slowed down because of the consequent CO2 fertilization effect—higher atmospheric CO2 concentration and temperature accelerate photosynthesis in plants which in return increase the rate of CO2 fixation (Keenan et al. 2016). Adaptive capacity is the capability to respond to perturbation from a changing environment by proactive adjustment and alteration. Gupta et al. (2010) explain adaptive capacity in six dimensions, including: (1) variety, diversity and redundancy; (2) learning capacity; (3) room for autonomous change (which includes the capacity to innovate); (4) leadership; (5) resources; and (6) fair governance. Transformative capacity refers to changing the stability landscape or even to create an entirely new system state. Climate change-induced migration can be an example: with sufficient capacity, the populations at risk may opt to abandon and rebuild their settlements and human–nature couplings in new locations to avoid foreseeable suffering (Methmann and Oels, 2015). Generally, for resilience-based strategies, improving absorptive capacity is for resisting existing hazards, while enhancing adaptive and transformative capacities allows coping with future uncertainties by self-improvement, incrementally or radically (Mao et al. 2017). Attention should be paid to all three capacities, not only the absorptive and adaptive, but also the capability to transform the system to a more favorable trajectory (Chapin et al. 2010). These latter two capacities are going to be increasingly more relevant in dealing with future risks and uncertainties.

18.3.3 Three Perspectives of Human–Nature Coupling and the Resilience Cube We would like to focus on three perspectives that are helpful in understanding human–nature couplings and resilience (Mao et al. 2017). • Perspective 1: The natural subsystem with anthropogenic influence. It highlights the state and change of nature, and its resilience to anthropogenic hazards; however, the human subsystem, the water impact on the human subsystem, or other forms of human impacts on the natural subsystem are not the focus. • Perspective 2: The human subsystem with natural hazards and benefits. It foregrounds the state of human societies and socio-economic resilience to natural disasters, but similarly the water subsystem, human impacts on water subsystem or other forms of water impact on the water subsystem are not emphasized. • Perspective 3: The coupled human–nature system which underscores the feedback and interactions between the two subsystems. We believe that the first two framings emphasize only partially the critical interactions between humans and nature, and it is in fact the third perspective that takes a more holistic view on the indivisible relationship of the two parts and will be increasingly relevant in envisioning Earth and society as an adaptive system (see also Figure 18.1).

What Lies Ahead?

391

FIGURE 18.1  Three perspectives of human–nature coupling. (1) Natural subsystem with human influence, (2) Human subsystem with natural hazards and benefits, (3) Coupled human–nature system. Adapted from Mao et al. (2017).

Scholars such as Mace (2014) find that there is an ongoing shift happening towards the coupled human–nature system. Mace (2014) argues that from “Nature suffering from People” or “People benefiting from Nature”, the current step is to move towards an interactively coupled “People and Nature” paradigm. With the help of the “resilience cube”, Figure 18.2, we can illustrate this transition. The “resilience cube” is a three-dimensional space, constructed by combining all three capacities of resilience along three axes (see also Mao et al. 2017). In the resilience cube

FIGURE 18.2  The resilience cube illustrating the three resilience capacities (adaptive, absorptive, transformative) in relation to the three phases of transition (Nature with Human; Nature for Human; Human and Nature).

392

Climate Changes in the Holocene

(Gillson et al. 2013), the pathways represent a series of structured management interventions, helping to drive systems from a vulnerable state (i.e. bottom left corner) towards a more resilient one (i.e. top right corner). However, the pathways are not always in straight lines—the three capacities usually change at an unequal pace. At the first stage, namely, Nature with Human before intensive modification of environments, human societies mainly relied on natural supply of subsistence (e.g. hunter-gathering), and the Earth was weakly altered by human activities. Because of insufficient social and physical preparedness, the interaction and coupling of human and nature were often vulnerable to disturbances. Therefore, at this stage, absorptive capacity was low, and adaptability was mainly provided by the “naturalness” of ecosystems. Next, with a growing population and increasing societal demands, intentional manipulation of the Earth was made, such as intensive exploitation of natural resources, massive use of fossil fuels and urbanization. This led the system to the second stage, Nature for Human, which explicitly emphasized the benefits people received from nature, and is therefore depicted in a framing of partial interactions between human and nature (i.e., a natural subsystem with human influence). At this stage, expected and known natural disasters are usually defended by hard engineering work (e.g. dams, concrete dikes), which increased absorptive capacity at the expense of decreased natural adaptive capacity. We are now experiencing a transition towards a third stage. In the Human and Nature stage, societal and ecological requirements of living in the Anthropocene are becoming more aligned (Mace, 2014; Mao et al. 2017). This transition is currently ongoing and it will require great efforts and will demand a re-orientation of governance and innovation structures.

18.4 FUTURE DEVELOPMENT PATHWAYS FOR IMPROVED HUMAN–NATURE RELATIONS After describing some of the key concepts, foregrounding the concept of resilience as part of a new kind of shared human knowledge (or gnosis) about Earth and society, in this section, we explore how gnosis may be translated into praxis through placing particular attention on the role of governance and technical innovation.

18.4.1 Governance Implications of Future Development Pathways In order to realize future development pathways that are based on improved human–nature relations, in alignment with the Human and Nature transition, there is a need to re-think our current approaches to governance. In a world that may change both slowly and abruptly in unpredictable directions, traditional governance structures that often assume highly centralized forms of power, often concentrated around the state, make it difficult to govern complex human–nature relationships and changing ecological baseline conditions (Zulkafli et al. 2017). Adaptive governance on the other hand, generates mechanisms for building more flexibility into decisionmaking processes, hence supporting societies to adapt to constantly changing conditions while embedding ecological considerations in governance processes (Karpouzoglou et al. 2016; Folke et al. 2005). In adaptive governance, a more distributed model of power emerges which makes more explicit linkages with local actors, everyday resource management practices, informal institutions and indigenous knowledge systems (Pahl-Wostl 2009). Adaptive governance is further based on polycentric institutional arrangements that comprise “a mosaic of nested sub-units” of decision making rather than a fully integrated, hierarchical whole (Lankford and Hepworth 2010). In practical terms, what this means is that this model of governance recognizes a higher degree of variability in the way in which societies organize collective actions and the roles multiple actors—e.g., citizens, NGOs, the private sector, and the state—have to play as part of finding solutions for dealing better with environmental change and uncertainty.

What Lies Ahead?

393

18.4.2 Recent Innovations for Connecting Social and Natural Systems Recent innovations in low-cost and open information and communication technologies (ICTs) offer new opportunities for fostering connections between social and natural systems (Karpouzoglou et al. 2017). For example, wireless sensor networks can be built with affordable environmental sensors (e.g., for measuring hydrological, metalogical and physio-chemical parameters), hardware platforms (e.g., Arduino and Raspberry Pi), and communication units (e.g., Xbee). These new instruments can be custom-designed for different contexts and scenarios, especially for monitoring conventionally data-scarce regions, and for early warning of environmental disasters (Buytaert et al. 2014; Paul et al. 2017; Mao et al. 2018). Based on the technological innovations, an integrated solution, namely Environmental Virtual Observatories (EVOs), can be developed for sustainability research and management. EVOs combine environmental monitoring with data-processing, communication and visualization (Beven et al. 2012). The aim of EVOs is to broaden the boundaries of environmental monitoring so that different levels of governance and different types of user requirements are more effectively considered (Karpouzoglou et al. 2016). In the recent study application in Huamantanga by Zulkafli et al. (2017), a practical methodology for EVO deployment was considered in the Andean highlands of Peru. Andean agro-pastoralist communities residing in the Andean district of Huamantanga have been experiencing different types of vulnerabilities as a result of changing climatic conditions. In particular, availability of irrigation water and fertile land for animal grazing and crop production are increasingly affected by a changing climate. At the same time, Huamantanga is situated in the Chillon river basin which is one of the main water basins providing drinking water to the capital of Peru, Lima; thus, what happens with the water in Huamantanga also has consequences for supplying water to the city. Therefore local, regional and national actors are all interested in environmental monitoring of land and water in Huamanatanga, albeit they each have different data requirements and expectations. It is in such a context where EVOs can be particularly effective tools since they can help assimilate different types of data to suit the requirements of different users. Zulkafli et al. 2017 developed a methodology for EVO deployment whereby different types of data were considered, including rainfall and water runoff from local monitoring stations, land cover data from remote sensing, and local sources of knowledge derived from interviews, videos and field research using a participatory approach. What the Huamantanga example shows is that the way science will inform decision making needs to transition towards a more adaptive Earth and society system. It is observed in particular that the general public will play an increasingly significant role by performing so-called “citizen science” activities. Based on new innovations, the citizen scientists can contribute in the production of new knowledge through techniques such as crowd-sourced environmental sensing and provide a valuable perspective of observations that may be ignored within traditional scientific structures (Buytaert et al. 2014). Citizen participation is encouraged to take place with close collaboration of environmental professionals throughout the data workflow to co-design of different tools, in order to co-generate knowledge that is contextually relevant and useful to guide local decision making (Grainger et al. 2016; Zulkafli et al. 2017). These low-cost technologies and citizen science practices may shift the data collection, processing and use from a number of centralized institutions to a network of polycentric and diverse actors that is compatible with adaptive governance. This change is believed to have great potential in resilience enhancement across scales and to address the different capacities: absorptive, adaptive and transformative (Buytaert et al. 2016).

18.5 CONCLUSIONS Earth and society can be understood as a complex adaptive system that has successfully responded to adversity over the millennia. However, uncertainties around the future of the climate coupled with ecosystem degradation means that adaptive capacity of the Earth needs to be improved.

394

Climate Changes in the Holocene

In this chapter we have examined both sources of gnosis and praxis that can help navigate an uncertain future while restoring balance across ecological and societal functions. From the point of view of gnosis, we have looked at new ways of conceptualizing human–nature relations as part of coupled systems thinking. At this moment, we are at a breaking point in history where we are experiencing a transition from the linear one-way “Nature suffering from People” or “People benefiting from Nature” relationship to a third stage of Human and Nature. In this stage, the concept of resilience becomes important because it highlights the strong inter-linkages between ecological and societal systems. Given that these interactions are dynamic in both space and time, we require different sets of capacities to navigate Earth and society. We have highlighted three crucial capacities, absorptive, adaptive and transformative and used the resilience canvas to visualize how these capacities interact as Earth and society evolve into the Human and Nature stage. Earth and society as coupled systems ultimately open up considerations about how to build more sustainable and resilient future pathways. This relates to the level of praxis, and in this chapter, we focused on two aspects of praxis that are related to the need to change governance models and the role of institutional/technical innovations. In this context, we have highlighted adaptive governance as one alternative governance model that focuses explicitly on embedding ecosystems in key governance decisions and examines the role of multiple actors, beyond just the state. Furthermore, we have looked at new innovations in environmental data and processing as well as the role of citizen science and illustrated how these new innovations are highly compatible with adaptive governance and the multi-actor approach. This story of praxis and gnosis and their relationship will be partly told by history and depends on humanity’s shared efforts in utilizing absorptive, adaptive and transformative capacities to develop new positive associations between human development and the natural world. A growing community of scholars use the term of a “good Anthropocene” to catalyze new visioning processes, ideas and practices that can steer humanity towards a new type of social contract that restores balance between the human and the non-human world (Bennet et al. 2016). As part of a “good Anthropocene”, humans become better guardians of Earth and society and on the basis of absorptive, adaptive and transformative capacities, steer the transition to the Human and Nature phase. The future does not have to be dystopian, but the future partly depends on what kind of development pathways we will follow to address critical global challenges. A crucial task ahead of us is to take some of these insights and tools forward so that we can begin to re-imagine how to sustain societies that can live and prosper alongside essential ecosystems.

REFERENCES Béné, C., A. Newsham, M. Davies, M. Ulrichs, and R. Godfrey-Wood. 2014. Review article: Resilience, poverty and development. Journal of International Development doi:10.1002/jid.2992. Bennett, E.M., M. Solan, R. Biggs, T. McPhearson, A.V. Norström, P. Olsson, L. Pereira, et al. 2016. Bright spots: Seeds of a good Anthropocene. Frontiers in Ecology and the Environment 14(8): 441–448. Beven, K., W. Buytaert, and L.A. Smith. 2012. On virtual observatories and modelled realities (or why discharge must be treated as a virtual variable). Hydrological Processes 26(12): 1906–1909. Buytaert, W., A. Dewulf, B. De Bièvre, J. Clark, and D.M. Hannah. 2016. Citizen science for water resources management: Toward polycentric monitoring and governance? Journal of Water Resources Planning and Management 142(4): 1816002. Buytaert, W., Z. Zulkafli, S. Grainger, L. Acosta, T.C. Alemie, J. Bastiaensen, B. De Bievre, et al. 2014. Citizen science in hydrology and water resources: Opportunities for knowledge generation, ecosystem service management, and sustainable development. Frontiers in Earth Science 2(October): 1–21. Chapin, F.S, S.R. Carpenter, G.P. Kofinas, C. Folke, N. Abel, W.C. Clark, P. Olsson, et al. 2010. Ecosystem stewardship: Sustainability strategies for a rapidly changing planet. Trends in Ecology and Evolution 25(4): 241–249. Fabry, V.J., B.A. Seibel, R.A. Feely, and J.C. Orr. 2008. Impacts of ocean acidification on marine fauna and ecosystem processes. ICES Journal of Marine Science 65(3): 414–432.

What Lies Ahead?

395

Folke, C., T. Hahn, P. Olsson, and J. Norberg. 2005. Adaptive governance of social-ecological systems. Annual Review of Environment and Resources 30(1): 441–473. Folke, C. 2016. Resilience (Republished). Ecology and Society 21(4). Folke, C., R. Biggs, A.V. Norström, B. Reyers, and J. Rockström. 2016. Social-ecological resilience and biosphere-based sustainability science. Ecology and Society 21(3). Gillson, L., T.P. Dawson, S. Jack, and M.A. McGeoch. 2013. Accommodating climate change contingencies in conservation strategy. Trends in Ecology and Evolution 28(3): 135–142. Grainger, S., F. Mao, and W. Buytaert. 2016. Environmental data visualisation for non-scientific contexts: Literature review and design framework. Environmental Modelling and Software 85: 299–318. Gunderson, L. 1999. Resilience, flexibility and adaptive management—Antidotes for spurious certitude? Conservation Ecology 3: 23–24. Gupta, J., C. Termeer, J. Klostermann, S. Meijerink, M. van den Brink, P. Jong, S. Nooteboom, et al. 2010. The adaptive capacity wheel: A method to assess the inherent characteristics of institutions to enable the adaptive capacity of society. Environmental Science and Policy 13(6): 459–471. Harari, Y. 2015. Homo Deus: A Brief History of Tomorrow. London: Harvill Secker. Holling, C.S. 1973. Resilience and stability of ecological systems. Annual Review of Ecology and Systematics 4: 1–23. IRENA. 2018. Renewable Power Generation Costs in 2017. Abu Dhabi: International Renewable Energy Agency. Karpouzoglou, T., and J. Barron. 2014. A global and regional perspective of rainwater harvesting in subSaharan Africa’s rainfed farming systems. Journal of Physics and Chemistry of the Earth 72–75: 43–53. Karpouzoglou, T., A. Dewulf, and J. Clark. 2016. Advancing adaptive governance of social-ecological systems through theoretical multiplicity. Environmental Science and Policy 57: 1–9. Karpouzoglou, T., L. Pereira, and S. Doshi. 2017. Bridging ICTs with governance capabilities for food-energywater sustainability. In Food, Energy and Water Sustainability: Emergent Governance Strategies, edited by L.M. Pereira C.A. McElroy, A. Littaye, and A.M. Girard. London: Earthscan. Karpouzoglou, T., Z. Zulkafli, S. Grainger, A. Dewulf, W. Buytaert, and D.M. Hannah. 2016. Environmental virtual observatories (EVOs): Prospects for knowledge co-creation and resilience in the information age. Current Opinion in Environmental Sustainability 18: 40–48. Keenan, T.F., I.C. Prentice, J.G. Canadell, C.A. Williams, H. Wang, M. Raupach, and G.J. Collatz. 2016. Recent pause in the growth rate of atmospheric CO2 due to enhanced terrestrial carbon uptake. Nature Communications 7: 13428. Lankford, B., and N. Hepworth. 2010. The cathedral and the bazaar: Monocentric and polycentric river basin management. Water Alternatives 3(1): 82–101. Mace, G.M. 2014. Whose conservation? Changes in the perception and goals of nature conservation require a solid scientific basis. Science 245(6204): 1558–1560. Mao, F., J. Clark, T. Karpouzoglou, A. Dewulf, W. Buytaert, and D.M. Hannah. 2017. HESS Opinions: A conceptual framework for assessing socio-hydrological resilience under change. Hydrology and Earth System Sciences 21(7): 3655–3670. Mao, F., J. Clark, W. Buytaert, S. Krause, and D.M. Hannah. 2018. Water sensor network applications: Time to move beyond the technical? Hydrological Processes 32: 2612–2615. Methmann, C., and A. Oels. 2015. From ‘fearing’ to ‘empowering’ climate refugees: Governing climateinduced migration in the name of resilience. Security Dialogue 46(1): 51–68. Meyer, H., and R. Peters. 2016. A Plea for Putting the Issue of Urbanizing Deltas on the New Urban Agenda. Delta Alliance: TU Delft. Pahl-Wostl, C. 2009. A conceptual framework for analysing adaptive capacity and multi-level learning processes in resource governance regimes. Global Environmental Change 19(3): 354–365. Paul, JD., W. Buytaert, S. Allen, J.A. Ballesteros-Cánovas, J. Bhusal, K. Cieslik, and J. Clark. 2017. Citizen science for hydrological risk reduction and resilience building. Wiley Interdisciplinary Reviews: Water 5: e1262. Renaud, F.G., J.P. Syvitski, Z. Sebesvari, S.E. Werners, H. Kremer, C. Kuenzer, R. Ramesh, et al. 2013. Tipping from the Holocene to the Anthropocene: How threatened are major world deltas? Current Opinion in Environmental Sustainability 5(6): 644–654. Reynolds, J.F., D.M. Smith, E.F. Lambin, B.L. Turner, M. Mortimore, S.P. Batterbury, T.E. Downing, et al. 2007. Global desertification: Building a science for dryland development. Science 316(5826): 847–851. Riehl, S. 2017. Regional environments and human perception: The two most important variables in adaptation to climate change. In Höflmayer. The Late Third Millennium in the Ancient Near East: Chronology, C14, and Climate Change. Chicago: The Oriental Institute.

396

Climate Changes in the Holocene

Rockström, J., O. Gaffney, J. Rogelj, M. Meinshausen, N. Nakicenovic, and H.J. Schellnhuber. 2017. A roadmap for rapid decarbonization. Science 355(6331): 1269–1271. Scheffer, M., S. Barrett, S.R. Carpenter, C. Folke, A.J. Green, M. Holmgren, T.P. Hughes, et al. 2015. Creating a safe operating space for iconic ecosystems. Science 347(6628): 1317–1319. Söderholm, P., R. Hildingsson, B. Johansson, J. Khan, and F. Wilhelmsson. 2011. Governing the transition to low-carbon futures: A critical survey of energy scenarios for 2050. Futures 43(10): 1105–1116. Steffen, W., P.J. Crutzen, and J.R. McNeill. 2007. The anthropocene: Are humans now overwhelming the great forces of nature. AMBIO: A Journal of the Human Environment 36(8): 614–621. Steffen, W., K. Richardson, J. Rockström, S.E. Cornell, I. Fetzer, E.M. Bennett, R. Biggs, et al. 2015. Planetary boundaries: Guiding human development on a changing planet. Science 347(6223): 1259855. United Nations. 2014. World urbanization prospects: The 2014 Revision. Department of Economic and Social Affairs.Population Division, United Nations. Zulkafli, Z., K. Perez, C. Vitolo, W. Buytaert, T. Karpouzoglou, A. Dewulf, B. De Bièvre, et al. 2017. Userdriven design of decision support systems for polycentric environmental resources management. Environmental Modelling and Software 88: 58–73.

19

Epimetron Michel Crucifix

Yesterday my seven-year-old daughter asked me, “What are humans for?” “What do you mean?” I replied. “Mum told me that bees pollinate flowers, mosquito feed frogs; so what are humans for?” Kids are usually very good at asking tough questions, and I must confess I had to pause a minute on this one (wouldn’t you?). Physicists avoid asking what something is for, and prefer answering what something does. However, physiologists, ecologists, anthropologists—in fact, everyone using everyday language—admit that things can have a function. Bees have a function because they fit into a global system, a global metabolism in which materials are recycled and signals are emitted and recognised. Do humans? Struck by the urgency of the question, I answered my daughter that humans have not really found their place yet. Their population is still expanding, using more resources, and have not formed a global, stable network of interactions with their environment. On a second thought, though, history gives us plenty of examples of reasonably perennial societies. French have invented the beautiful, earthy word of “terroir”. It conveys the powerful idea of an alliance between human traditions and the cycles of Nature. In it one can feel—smell!—how the Bordeaux grapes, the chestnut trees of the Cévennes, the Pyrenean dogs exchanged services with the monks, farmers and shepherds who needed them. It is all about food, lives, and cycles. Can we ever imagine a global-scale terroir? This could be the motivating question of this book. Chapter after chapter, we have seen the Pleistocene as the theatre of mankind’s birth. After billions of years of history which saw the fabrication of enzymes, organisms, ecosystems, the production of the human brain was a new installment in the rise of complexity on Earth—another level in the capacity of living beings to process resources and extract energy in an ever-changing environment. Humans have provided the material for yet higher organization levels: the tribe, the city, the civilization, and now the global economy. It seems to be a lesson of history that any developing system is exposed to phases of contraction and reorganization. As we are reminded by Middleton (cf. Chapter 13), citing Tainter, the crisis is “a rapid, significant loss of an established level of sociopolitical complexity”. Or, as summed up by Lucero and Tarmon (cf. Chapter 8), kings lose power—enacting a societal transformation—but farmers remain and reorganise themselves. Climatic events such as long droughts have sometimes triggered crises and even societal collapses. Scholars are careful not to see the climatic event as their deep cause: the crisis is the sudden revelation of an imbalance accumulated over time, such as the increased reliance on a rarefied resource. As time went on the mechanisms and the nature of interactions between humans and the environment evolved (cf. Karpouzoglou and Mao, Chapter 18). Crisis after crisis, technological progress, trade, animal domestication, the discovery of fossil fuel energy resources and information technologies have generated increasingly complex societies. With large-scale transformative projects such as dams and cities, societies improved their resilience to local climatic conditions, but they increased their reliance on a global network. In a sense, the complexification of societies has changed the rules which determine their evolution. Paleolithic and neolithic civilizations exploited ecological niches, which varied over time (cf. Sala, Chapter 15). Climatic events kept having acute consequences on populations even well into the industrial era. Wagner and Zorita (Chapter 5), for example, explain how in 1820 the Tambora 397

398

Climate Changes in the Holocene

eruption caused famine all over the World. Nowadays, a global food network protects the richer countries against a bad summer, but that network is not serving everyone on Earth fairly. It might also turn out to be vulnerable. Indeed, the technosphere in which an ever-increasing fraction of the human population live should not let us forget that humans run on food, and, schematically put, industrial food is the product of oil and phosphorous (Cordell et al. 2009). Oil supplies the energy needed for the production of nitrate fertilizers, farming, fish boats, food transport, plastic packaging, and processing; phosphorous is used as a fertilizer. Not only are these resources exhaustible, but the waste products of the whole human industry built on the top of agriculture has leveraged its pressure on the environment, causing serious threats to the functioning of natural ecosystems: climate change, fish stock depletion, eutrophication of coastal ecosystems, etc. (Rockstrom et al. 2009). These are obviously global issues. As humans became globally interconnected, with food and raw material travelling worldwide, and with values propagated by marketing reaching an ever-increasing fraction of the population, individuals have developed a dependence on a system organized at the global scale. Yet, this global system is in a state of overshoot: it consumes more resources than is sustainable and, worryingly, it seems to have remained largely blind to this notion of overshoot. Global markets keep working as if resources were infinite. But we, individual humans, should not be blind. As long demonstrated by the paintings, sculptures and engravings of the paleolithic, humans have long since grasped the ability to think, imagine, and discourse about the different levels of organization they are embedded in. The realization of the dangerous position in which global civilization has put itself must be for us a call for action. Science, art, and politics are the mechanisms humans can use to favor the emergence of global awareness and anticipation at the global scale. This is a tough challenge. The agreements adopted in the Convention of Parties in Paris in December 2015 to hold “the increase in the global average temperature to well below 2°C above pre-industrial level and to pursue efforts to limit the temperature increase to 1.5°C above pre-industrial levels” are a political fact. As reminded by Guillemot (2017), a survey published by the British newspaper The Guardian in 2009 suggested that 85% of climate scientists did not believe, however, that it would be possible to limit warming to 2°C. Reasons for such skepticism are easily found in current assessments about climate change: about two-thirds of the available budget of emissions for keeping warming to below 2°C have already been emitted (Rogelj et al. 2016; Kriegler et al. 2018). In other words, the 1.5°C target will be pulverized unless we replace about 85% of our current primary energy needs by carbon-neutral alternatives in the next decade or so, or if the next generation finds a way to remove carbon dioxide from the atmosphere (cf. Chapter 17). The latter is a dangerous bet indeed: however bright our children may be, they will have to cope with the laws of thermodynamics (Hansen et al. 2017). The Paris Agreement has therefore caused some turmoil among climate scientists, some applauding the political recognition of the risks of climate change (Schellnhuber et al. 2016), other expressing skepticism or even anger about talks unsupported by a consequent action plan (Le Page, 2015). As scholars swiftly observed (Hulme 2016; Guillemot 2017), this is the relationship between knowledge and policy which is highlighted here. The proliferation of scientific articles, special issues, and the IPCC being commissioned for the preparation of a special report “on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways” show that the scientific production system has absorbed the political shock. Scientists feel they need to remain actively publishing (publish or perish), but they also wonder to what extent climate researchers and research funders should bend their research agendas, programs and projects towards short-term policy-oriented questions (Hulme 2016). Researchers from developing countries, on the other hand, remind us that their populations are at the forefront in the coming environmental crisis, and urge us to envisage all options, including geo-engineering (Cao 2018). In fact, climate scientists’ attempts to predict how the Earth system will evolve over the next century under this or that scenario will be of limited value unless historians, archaeologists and

Epimetron

399

anthropologists put as much wisdom towards understanding how millennial societies can adapt, migrate, morph into each other in the face of rapid changes. Together, we need to develop the tangible elements that might help humanity to solve the difficult puzzle of reducing its global energy and resource needs without crashing. Is a soft landing scenario actually possible? Perhaps the bigger question is whether global politics and economy can ever be molded into a sustainable relationship with the Earth system. Karpouzoglou and Mao (cf. Chapter 18) present society as a “self-adaptive system”. Others have used similar words. Kim and Mackey (2013) describe environmental law as a “complex adaptive system”. The concept of self-adaptive system appeared in the research on emergence and organization of living systems developed around the buzz of the Santa Fe Institute in the 1980s. These concepts may be connected to a deeper epistemological movement focusing on relations, causal pathways, and anticipation—in a word, complexity. Hence, while we witness a knowledge gap between science and politics, bridges appear between the “two cultures” of sciences and humanities. Our challenge will be to cross language barriers. May this book be a contribution in that direction.

REFERENCES Cao, L. 2018. The effects of solar radiation management on the carbon cycle. Current Climate Change Reports 4(1): 41–50. doi: 10.1007/s40641-018-0088-z. Cordell, D., J. Drangert, and S. White. 2009. The story of phosphorus: Global food security and food for thought. Global Environmental Change 19(2): 292–305. doi: 10.1016/j.gloenvcha.2008.10.009. Guillemot, H. 2017. The necessary and inaccessible 1.5°C objective. In Globalising the Climate COP21 and the Climatisation of Global Debates, edited by X. Yi-Chong, J. Foyer, and E. Morena, 39–56. London: Routledge. Hansen, J., M. Sato, P. Kharecha, K. von Schuckmann, D.J. Beerling, J. Cao, S. Marcott, et al. 2017. Young people’s burden: Requirement of negative CO2 emissions. Earth System Dynamics 8(3): 577–616. doi: 10.5194/esd-8-577-2017. Hulme, M. 2016. 1.5°C and climate research after the paris agreement. Nature Climate Change 6(3): 222–224. doi: 10.1038/nclimate2939. Kim, R.E., and B. Mackey. 2013. International environmental law as a complex adaptive system. International Environmental Agreements: Politics, Law and Economics 14(1): 5–24. doi: 10.1007/s10784-013-9225-2. Kriegler, E., G. Luderer, N. Bauer, L. Baumstark, S. Fujimori, A. Popp, J. Rogelj, et al. 2018. Pathways limiting warming to 1.5°C: A tale of turning around in no time? Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences 376(2119): 20160457. doi: 10.1098/rsta.2016.0457. Le Page, M. 2015. Paris climate deal is agreed – but is it really good enough? New Scientist, 12 December. Rockstrom, J., W. Steffen, K. Noone, A. Persson, S.F. Chapin, E.F. Lambin, T.M. Lenton, et al. 2009. A safe operating space for humanity. Nature 461(7263): 472–475. doi: 10.1038/461472a. Rogelj, J., M. den Elzen, N. Höhne, T. Fransen, H. Fekete, H. Winkler, R. Schaeffer, et al. 2016. Paris agreement climate proposals need a boost to keep warming well below 2°C. Nature 534(7609): 631–639. doi: 10.1038/nature18307. Schellnhuber, H.J., S. Rahmstorf, and R. Winkelmann. 2016. Why the right climate target was agreed in Paris. Nature Climate Change 6: 649–653. doi: 10.1038/nclimate3013.

Index A

B

Active regions, 109 Adaptations BCE hydraulic oases sites, 4th millennium, 259 BCE sepulchral landscapes, 5th millennium, 257–259 hydraulic habitation landscapes, 5th millennium, 257–259 integrated and holistic approaches, 265 sepulchral oases sites, 4th millennium, 259 Adaptive system, 387–394 Advanced microwave sounding units (AMSUs), 364 Advanced Very High Resolution Radiometer (AVHRR), 365 Aerosol Optical Depth (AOD), 127 African Humid Period (AHP), 17, 18, 20, 187, 190 African monsoon, 16–19 AHP, see African Humid Period (AHP) Ain e-Raml, 206, 207, 214 Allostratigraphic deposits, 65 American Anthropological Association Statement on Humanity and Climate Change, 175 AMH, see Anatomically modern humans (AMH) AMOC, see Atlantic Meridional Overturning Circulation (AMOC) AMOC. See Atlantic Meridional Overturning Circulation (AMOC) AMO, see Atlantic Multidecadal Oscillation (AMO) AMSUs, see advanced microwave sounding units (AMSUs) Amur pottery, 346 Anatomically modern humans (AMH), 148, 149 Ancient-DNA genetics, 157–161 findings from, 159–160 Anomalies, 46–47 Anthropocene, 96 Anthropology, 67–68, 157 AOD, see Aerosol Optical Depth (AOD) Arab Spring, 285 Archaeohydrology, 265 Archeology, 67–68, 157 Holocene Central Sahara, 183–197 Arctic Circumpolar Cultural Zone, 358 Arctic Circumpolar Cultural Zone (ACCZ), 348 Arid Central Asia, monsoon and Westerlies, interchange, 14–16 Aridity, 254, 349, 350, 351 Artificial landscape alteration, 260 Asian summer monsoon (ASM), 12–14 ASM, see Asian summer monsoon (ASM) Astronomical or orbital forcing, 28, 79–82 Atlantic Meridional Overturning Circulation (AMOC), 129, 334, 339 Atlantic Multidecadal Oscillation (AMO), 129 Atlantic period (7.2–5.2 ka BP), 340 Australopithecines, 145–147 AVHRR, see Advanced Very High Resolution Radiometer (AVHRR)

Babylon, 4 Baraba steppe, 355 Beringia, 347 Bioclimatology, 327–329 Biogenic carriers, 62–63 Bipedalism, 146 Blytt–Sernander scheme, 10 Boreal-Atlantic transition, 355 Boreal forest, 332 Boreal period (9.7–7.2 ka BP), 339 Bronze Age, 222 4.2 ka BP event, Eastern Mediterranean collapse, 276–279 4.2 ka BP event, Greece, 274–276 background, collapses, 272–273 civilizations, 27, 271–287 climate-collapse-migration, 280–282 collapse, defined, 272 EBA collapse, Greece, 273–274 LBA collapses, Eastern Mediterranean, 279–280 LBA collapses, Greece, 279–280 opportunities and issues, 283–286

C Capture technologies, CO2, 375–378 oxy-fuel combustion capture, 377–378 post-combustion technologies, 377 pre-combustion capture, 376–377 Carbon-14, 49 Carbon dioxide, 21, 89 Carbon dioxide capture storage (CCS), 375–380 capture technologies, 375–378 geological storage, 378 implementation scale, 380–382 implications and barriers, 382 Carbon dioxide utilization (CCU), 375–380 coal seams, 379 conversion into chemicals and fuels, 379–380 direct utilization, 379 enhanced oil and coal-bed methane recovery, 379 implementation scale, 381–382 implications and barriers, 382 microalgae, biofuels, 380 mineral carbonation (MC), 380 Causality, 312 CCS, see Carbon dioxide capture storage (CCS) CCU, see Carbon dioxide utilization (CCU) Celali revolution, 135 Cenozoic context, 77–79 Cenozoic temperature trends, 7–8 Chalcolithic period, 222 Chronology, 48–51 Circum-Caspian Cultural Zone, 355 Cis-Baikal case, 349 Classic Maya society, 165–177

401

402 ancient Maya water management and quality, 171–174 climate change impact, 174–175 of Southern lowlands, 168–174 Clean energy, 373–374 2020 Climate and energy package, 374 Climate-collapse-migration, 280–282 critique, 280–282 Climate feedback, 20 Climate fluctuations Cenozoic temperature trends, 7–8 geological background of, 6–10 global climate changes, Holocene, 9–10 Holocene subdivisions, 10 Climate forcings, 122–129 externally forced variations, 124–128 internal variations, 128–129 Climate forcings, 250 Climate, interacting, 249–253 Climate, main characteristics, 131–134 Climate model, 9, 22, 283, 369 Climate-monitoring satellites carbon dioxide concentrations, monitoring changes, 369–370 climate research observations, future, 370–371 contemporary era of, 366–367 equatorial cloud dynamics, terra and aqua, 368 lidar, global atmosphere, 368 observational investigations, 367–370 pre-history of, 363–366 Climate parameters, 63–64 Climate, pleistocene context, 121–122 Climate reconstructions, 41, 44–45 anomalies, 46–47 chronology and dating, 48–51 methods, 130–131 past climate and environmental information, 47–48 weather vs. climate, 44 Climate sensitivity, 6 Climate system, 22, 23, 26 Climate variability, 6 Climatic precession, 79 Coal, 374 Cold steppe, 331, 332, 336, 337, 341, 342, 343, 355 Complex adaptive system, 399 Conifers, 332 Corals, 54 Crete populations, 160–161 Culture, 249–253 Cyprus populations, 160–161

D Dalton minimum, 112, 125 Dansgraad-Oeschger events, 92–93 cause of, 92–93 European vegetation changes, 93 and Greenland stratigraphy, 92 Date palm, 235 Deglaciation, 94–96 early anthropocene hypothesis, 95–96 Degradation, 248, 253–255 Dendroclimatology, 51 Desert, 333 Desertification, 184

Index Desert, onset of, 194–195 Dyuktai culture, 247, 348

E EAM, see Ethiopian African monsoon (EAM) Early Acacus hunter-gatherers, 187–189 Early anthropocene hypothesis, 95–96 Early Helladic (EH) II and III, 274–276 Earth, 387 climate, 107–117 climate system, 5–6 Earth and society absorptive, adaptive and transformative capacities, 390 coupled system, 389–392 human-nature coupling and resilience cube, 390–392 human-nature relations, development pathways, 392–393 resilience-based approaches, 389–390 risks and uncertainties, 388–389 Earth system models (ESM), 11, 130 Earth System Models of intermediate complexity (EMICS), 130 EASM, see East Asian Summer Monsoon (EASM) East Asian Summer Monsoon (EASM), 12, 14, 334 Eastern Sahara early settlements, 203–207 flimsy structures, 205–207 in late pleistocene, 201–203 reoccupation of, 203–207 EBA collapse, Greece, 273–274 EBMs, see Energy Balance Models (EBMs) Eccentricity, 87, 95 Ecosystem stewardship, 389 Ecotone, 333, 334, 341 Egyptian Western Desert, 203 EH, see Early Helladic (EH) II and III El Niño–Southern Oscillation (ENSO), 6, 53, 129, 167 EMICS, see Earth System Models of intermediate complexity (EMICS) Energy Balance Models (EBMs), 130 Energy center trench, 233 Engineering artificial landscape alteration, 260 catchments, preparing, 260 changing watersheds, 260 dams and diversion canals, 261 groundwater management, 260 (pre-) historic arid land hydraulic structures, 260–261 (pre-) historic arid land water management, 259–260 landscape, modifying, 260 Rasif case study, 261–262 troughs, 261 water management and hydraulic features, 259–262 wells and techniques, 260–261 ENSO, see El Niño–Southern Oscillation (ENSO) Environmental studies developments in, 11–23 human adaptation potential, 25–26 monsoon rainfall, paleoclimatic studies, 12–20 reconstructions and modelling, 11 Eocene–Oligocene transition, 78 Epimetron, 397–399 Epipalaeolithic industry, 207

403

Index Epipalaeolithic occupation, 210 ERTS-A, 363 ESM, see Earth system models (ESM) Ethiopian African monsoon (EAM), 18 Ethnic fragmentation, 194–195 European summer temperature, last 2,100 years, 132–134 Eve theory, 148

F Faculae, 109 Falaj irrigation, 236 Farafra Oasis, 206 Feedback loops, 389 Fennoscandian ice sheet, 337 Funerary practices, 195–197

G Galactic cosmic rays (GCRs), 113–114, 116 GCMs, see General Circulation Models (GCMs) GCRs, see Galactic cosmic rays (GCRs) General Circulation Models (GCMs), 11, 130 Genetic history, 11 Geo-engineering, 398 Geoethical Promise, 283, 285 Geological storage, CO2, 378, 380–381 CO2-enhanced oil recovery, 381 deep saline formations, 378 safety of, 381 GHG, see Greenhouse gas (GHG) emissions Glacial cycles, late pleistocene, 88–91 deglaciation mechanisms, 90–91 glacial inception, 89–90 interglacials diversity, 88–89 Glacial geomorphology, 64–65 Global mean surface temperatures (GMST), 22, 23 Global monsoon belt, 19–20 Global warming hiatus, 22–23 GMST, see global mean surface temperatures (GMST) GRACE, see Gravity Recovery and Climate Experiment (GRACE) Gravity Recovery and Climate Experiment (GRACE), 367 Greenhouse gas (GHG) emissions, 5, 373 Green Sahara, building of, 187–190 diversification and sedentism, 189–190 milking, 192–194 specialization and logistic mobility, 187–189 Groundwater management, 260

H Heinrich events, 93–94 High resolution climate reconstruction, 121–136 High-resolution paleoclimate records, 7 Historical climatology, 42 Historical information, proxy indicators, 65–68 anthropology, 67–68 archeology, 67–68 early instrumental and documentary evidence, 67 visual art and discursive data, 66–67 Holocene (11.7–0 ka BP), 338 Holocene Central Sahara, 183–197 concepts and contexts, 184–186

desert, revenge of, 194–197 Green Sahara, building of, 187–190 pleistocene, Arid end, 186–187 Saharan herders, 190–194 Holocene Thermal Optimum, 336, 340 Hominids, 144–147 Homo erectus, 147–148 Homo sapiens, 143–153 early hominids, 144–147 homo erectus, 147–148 human origins and migration, 144–151 later migrations, 151–153 Human adaptation, Arabia, 221–241 Bronze Age, 227–231 early farming communities, 227–231 early water management, 229–231 floodwater harvesting genesis, 227–229 floodwaters, harnessing, 237–239 Green Arabia to desert, 223 holocene, environmental change, 223 hydraulic techniques, blooming, 231–239 Iron Age, 231–239 land of starvation, 222–223 pastoralism to agriculture, 226–227 water management development, phasing, 239–240 water management, prehistory, 223–226 Human cultures, 328 Humanities, ancient DNA studies, 23–24 Human migrations, 159 Human-nature relations, development pathways, 392–393 governance implications, 392 social and natural systems, innovations, 393 Human Niche Construction process, 222 Human phylogenetics, 157 Human phylogeography, approach, 158 Hydraulic cultures, 247–266 Hydraulic engineering, 266 Hydraulic technologies, 221–241 hydrological response, tropics, 94 Hydrology, 247–266 Hyperthermals, 7

I IEA, see International Energy Agency (IEA) Indian summer monsoon (ISM), 12, 13, 19 Indus River Basin, 293–312 IntCal13, 49 interbiotic zone, 334, 341, 346 International Energy Agency (IEA), 383 International Energy Association, 374 Intertropical Convergence Zone (ITCZ), 5, 13, 16, 17, 19, 27, 94, 168, 194, 201, 223 Iranian Plateau, 293–312 abrupt events, 299–300 archaeological data, 300–301 Bronze Age, 308–311 climate change and human responses, linking, 301–302 late Chalcolithic, 306–308 long-term trends, 298–299 modern precipitation, 294–297 nature, scale, limitations and potential, 300–301 neolithic transition, early holocene, 302–306 paleoclimate proxy records, 297–298

404 spatial and temporal climatic context, 294–300 theoretical and practical challenges, 301–302 Iron Age, 222, 231–239 Irtysh-Ob interfluves, 351, 355 ISM, see Indian summer monsoon (ISM) ITCZ, see Intertropical Convergence Zone (ITCZ)

K 4.2 Ka BP event, Eastern Mediterranean collapse, 276–279 Akkadian Empire, 277–278 Egypt, 277 Levant, 276 Kamchatka peninsula, 347 Karstic topography, 173 Kelteminar culture, 353 Khabrats, 260 Kitoi culture, 350–351

L Lake sediment laminations, 53 Laki eruption, 127 LALIA, see Late Antique Little Ice Age (LALIA) Land–sea breeze hypothesis, 13 Land use, 249–253 Last 2,000 years climate forcings in, 122–129 climate, main characteristics, 131–134 climate, pleistocene context, 121–122 climate reconstruction methods, 130–131 high resolution climate reconstruction of, 121–136 late holocene climate implications, societal changes, 134–136 Last glacial cycle, 91–94 Dansgraad-Oeschger events, 92–93 glacial advances, glacial inception, 91–92 Heinrich events, 93–94 Last Glacial Maximum (LGM), 336, 337, 342 Last glacial maximum (LGM), 18, 53, 121, 201, 203, 335–336 Late Acacus foragers, 189–190 Late Antique Little Ice Age (LALIA), 131, 132 Late holocene climate implications, societal changes, 134–136 mobility and exchanges, 211–213 Late Neolithic cultures, 350 Late pastoral herders, 194–195 LBK, see Linearbandkeramik (LBK) culture Levantine neolithic package, 213–214 LGM, see Last glacial maximum (LGM) LIA, see Little Ice Age (LIA) Linearbandkeramik (LBK) culture, 151 Little Ice Age (LIA), 9, 16, 42, 125 Loess, 65 Lowland tropics, 167–168

M Marine interstadials, 90, 92 Masāfī palm grove, 233 MAT, see modern analogue technique (MAT) Maunder minimum, 112, 114, 125

Index MCA, see Medieval Climate Anomaly (MCA) Medieval Climate Anomaly (MCA), 125, 131 Medieval Warm Period (MWP), 9, 16 Mediterranean islands, 160–161 “Megalithic” structures, 197 Mesoamerica, 165–177 climate change and, 166–167 Meteorology, 3 Microfossil assemblages, 60–61 Middle East, 353 Middle-holocene settlements Dakhla, 210–211 Farafra, 209–210 Middle-Late Bronze epoch, 354 Middle Miocene Optimum, 80 Mid-Eocene Climatic Optimum, 7 Mid-pleistocene transition, 85–87 Milancovic theory, 124 Mitochondrial DNA (mtDNA), 158–159, 208 Mobile pastoralism, 253, 255–259 Modern analogue technique (MAT), 63 Modern-DNA genetics, 157–161 mono-biotic” tundra, 341 Monsoon translocation, 19 MtDNA, see mitochondrial DNA (mtDNA) Multidecadal variability, 22 MWP, see Medieval Warm Period (MWP)

N NAO, see North Atlantic Oscillation (NAO) National Oceanic and Atmospheric Administration (NOAA), 364 Neolithic Age civilizations, 27 Neolithic stone-working technique, 348 neolithization, 346 Nile Valley, 201–215 NOAA, see National Oceanic and Atmospheric Administration (NOAA) Nomadism, 194–197 North and Central Asia, 327 case studies, 342 cultural dynamics and early pottery emergence in Lower Amur Basin, 345–347 Early Atlantic Period, 348–351 Northern Far East, 347–349 Siberian Paleolithic refuge during last glacial period, 342–345 trans-urals forest-steppe belt, 351–355 climate-envrionmental changes and human responses climate and environment evolution, 335–340 cultural evolution through adaptation, 340–342 global considerations, 334–335 modern climate and environment of climate, 330–331 geography, 33–334, 329–330 North Arabian mid-holocene, 247–266 North Atlantic Oscillation (NAO), 9, 122, 128 Northern Arabia, mid-holocene climate, 253–255 North Pacific Ocean, 334 Northwest Arabia, 6th–5th millennium B.C., 226–227 Nubian desert, cattle problem, 207–209

405

Index O

Q

Oases, 194–197 Northern Arabia, Iron Age, 232 Oasisation, 252 Optically stimulated luminescence (OSL), 240 Organic biomarkers, 61–62 Orogenesis, 5 Osipovka culture (14–10.3 ka BP), 346 OSL, see Optically stimulated luminescence (OSL)

Qanāt technology, 240 Qas, 260 Qulban Beni Murra, 226, 251, 257

P PAGES Working Group, 88 Palaeo-Bedouins, 263 Palaeocene–Eocene Thermal Maximum, 7 Paleoclimatology, 41 archaeobotany, 24–25 Paleogeography, 5 Paleolakes, 65 Paleolimnology, 53 Paleosoils, 65 Paris Agreement, 388 Pastoral land use, 247–266 Past solar variability, reconstructions, 112–114 Peninsula of Arabia, 19 Penumbra, 109 Perihelion, 79 Phylogeography, 157 Piora Oscillation, 299 Pit-Comb Ware, 352 Plage, 109 Plate tectonics, 5 Pleistocene, Arid end, 186–187 Pleistocene glaciations, 77–96 astronomical forcing, 79–82 cenozoic context, 77–89 deglaciation, 94–96 early pleistocene, 82–85 entry into pleistocene, 82 glacial cycles, late pleistocene, 88–91 last glacial cycle, 91–94 mid-pleistocene transition, late pleistocene, 85–87 pleistocene humid periods, North Africa, 87–88 Pleistocene humid periods, 12, 87–88 Pliocene–Pleistocene transition, 7 Population movements, 157–161 Postglacial climate amelioration, 329 pottery, 344, 346 PPNB, see Pre-Pottery Neolithic B (PPNB) settlement Preboreal period (10.7–9.7 ka BP), 338 Pre-Pottery Neolithic B (PPNB) settlement, 224, 225, 239 Proto-Elamite Phenomenon, 306, 307 Proto-oasis land use, 247–266 Proxies, 4 Proxy indicators, 41–68 climate reconstructions, 44–51 geomorphological and sedimentary features, 64–65 historical information, 65–68 seasonal variations, climate changes, 51–55 sediment cores, 55–64

R Radiocarbon (14C) dating, 48–51, 232 Rainfall, bimodal phase, 209–211 Rajajil, 226, 251, 257 Rasif, 226, 227, 251, 257 RCM, see regional climate models (RCM) Record, 42 regime shift, 333–334 Regional climate models (RCM), 130, 131 Resilience, 388

S Sabkhas, 260 Saharan herders cattle herders, middle pastoral, 192–194 climate variations and cultural trajectories, 190–194 early pastoral, first herders, 190–192 Scanning multichannel microwave sensor (SMMR), 363, 364 Sclerochronology, 53–54 SDGs, see Sustainable Development Goals (SDGs) Seasonality, 304 Seasonal variations, climate changes, 51–55 corals, 54 ice cores, 52–53 lake sediment laminations, 53 paleolimnology, 53 sclerochronology, 53–54 speleothem, 54–55 tree-rings, 51–52 varves, 53 Sea Surface Temperature (SST), 334 Sea surface temperatures (SST), 168 Sedentarization, 304 Sedentary gardening, 255–259 Sediment cores, proxies, 55–64 climate parameters, 63–64 isotopes and chemical composition, biogenic carriers, 62–63 lake sediments and wetlands, microfossil data, 60 microfossil assemblages, 60–61 organic biomarkers, 61–62 pollen grains and spores, 58–60 transfer functions or analogue techniques, 63–64 Self-adaptive system, 399 SEP, see solar energetic particle (SEP) Siberdik, 348 Siberia, 328 climate, 330–331 geography, 329–330 vegetation cover biomes, 331–333 regime shift, 333–334 Slave trade, 153

406

Index

SMMR, see scanning multichannel microwave sensor (SMMR) SMOW, see Standard Mean Ocean Water (SMOW) Social stratification, 194–197 funerary practices and, 195–197 Sociohydraulic trajectory, 248 Solar activity, 114–116 Solar constant, 109 Solar energetic particle (SEP), 116 Solar influence mechanisms, climate, 116–117 Solar ionizing radiation, 110 Solar irradiance, 109–112 variability, 107–117 Solar magnetism, 114 Solar magnetograms, 112 South Arabian irrigation systems, 238, 239 Southeast Arabia Iron Age, 232–237 runoff harvesting, oasis of Masâfî, 236–237 wells and Aflâj, groundwater exploitation, 232–236 Southwest Arabia early water management, 229–231 floodwater harvesting genesis, 227–229 Space weather, 110 Spectral solar irradiance (SSI), 109, 114 Speleothem, 54–55 Spörer minimum, 114 SSI, see Spectral solar irradiance (SSI) SST. See Sea Surface Temperature (SST) SST, see Sea surface temperatures (SST) Standard Mean Ocean Water (SMOW), 53, 62 Steppe, 331–333 Subatlantic (2.5–0 ka BP) climate, 340 Subboreal period (5.2–2.5 ka BP), 340 Summer insolation, 79 Sun, 79, 107–109 Sunlight, 107–109 Sunspots, 109 Sustainable Development Goals (SDGs), 388

Temperate mixed forest, 332 Terrestrial carbon cycle, 20–22 Terrestrial ecosystems, 22 Tipping points, 388 Tobol-Irtysh interfluves, 355 Total solar irradiance (TSI), 109, 114 Tracer, 42 Transitional climate zone (TCZ), 14 TSI, see Total solar irradiance (TSI) Tundra-steppe, 331, 336, 337, 355

T

Yamnaya culture, 352 Y chromosomes, 158–159 YD, see Younger Dryas (YD) Younger Dryas (YD), 203, 335–338

Tambora eruption, 125 TCZ, see Transitional climate zone (TCZ)

U Umbra, 109 Uralian gate, 352, 353 Urban diaspora, 176 Ushki culture, 347

V Varves, 53 Vienna Meteorological Congress of 1873, 4 Volcanic degassing, 78 Volcanic eruptions, 4, 11 Volcanism, 91 Voskarides Konstantinos, 24 Vulnerability, 248, 249–253

W Water mastership, 239 Water vulnerability, 265 Wādī Abu Tulayha, 224–226 Weather, 4, 44 Weighted partial least square (WPLS), 64 Whole-genome sequencing, 161 World Meteorological Organization, 130 WPLS, see Weighted partial least square (WPLS)

Y