185 97 22MB
English Pages 407 Year 2019
ENVIRONMENT AND SOCIETY IN THE LONG LATE ANTIQUITY
ENVIRONMENT AND SOCIETY IN THE LONG LATE ANTIQUITY EDITED BY
ADAM IZDEBSKI AND MICHAEL MULRYAN
LEIDEN | BOSTON
Cover illustration: Interior visualisation of the Alacami (Richard Bayliss)
Typeface for the Latin, Greek, and Cyrillic scripts: “Brill”. See and download: brill.com/brill-typeface. ISBN 978-90-0438379-1 (hardback) ISBN 978-90-0439208-3 (e-book) Copyright 2019 by Koninklijke Brill NV, Leiden, The Netherlands. Koninklijke Brill NV incorporates the imprints Brill, Brill Hes & De Graaf, Brill Nijhoff, Brill Rodopi, Brill Sense, Hotei Publishing, mentis Verlag, Verlag Ferdinand Schöningh and Wilhelm Fink Verlag. All rights reserved. No part of this publication may be reproduced, translated, stored in a retrieval system, or transmitted in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without prior written permission from the publisher. Authorization to photocopy items for internal or personal use is granted by Koninklijke Brill NV provided that the appropriate fees are paid directly to The Copyright Clearance Center, 222 Rosewood Drive, Suite 910, Danvers, MA 01923, USA. Fees are subject to change. This book is printed on acid-free paper and produced in a sustainable manner.
In memory of Mark Whittow (1957–2017), inspiring scholar of Late Antiquity and Byzantium
Contents Acknowledgements ......................................................................................................... xi List of Contributors ........................................................................................................ xii
Introductory Essay Setting the Scene for an Environmental History of Late Antiquity ................... 3 Adam Izdebski
Bibliographic Essays The Environmental History of the Late Antique Eastern Mediterranean: a Bibliographic Essay ....................................................................................................... 17 Lucas McMahon and Abigail Sargent
The Environmental History of the Late Antique West: a Bibliographic Essay ...................................................................................................................................... 31 Merle Eisenberg, David J. Patterson, Jamie Kreiner, Ellen F. Arnold, and Timothy P. Newfield
Regional Vegetation Histories: Overview of the Pollen Evidence Revisiting the Beyşehir Occupation Phase: Land-Cover Change and the Rural Economy in the Eastern Mediterranean during the First Millennium AD ........................................................................................................ 53 Neil Roberts
Regional Vegetation Histories: An Overview of the Pollen Evidence from the Central Mediterranean ................................................................................. 69 Katerina Kouli, Alessia Masi, Anna Maria Mercuri, Assunta Florenzano, and Laura Sadori
A Late Antique Vegetation History of the Western Mediterranean in Context ........................................................................................................................... 83 José Antonio López-Sáez, Sebastián Pérez-Díaz, Didier Galop, Francisca Alba-Sánchez and Daniel Abel-Schaad
Vegetation and Land-Use Change in Northern Europe during Late Antiquity: A Regional-Scale Pollen-Based Reconstruction ..................... 105 Jessie Woodbridge, Neil Roberts and Ralph Fyfe
Case Studies – West Hadrian’s Wall in Context: A Multi-Proxy Palaeoenvironmental Perspective from Lakes .................................................................................................. 121 Petra Dark
viii
Contents
Variation in the Continuity of Land-Use Patterns through the First Millennium AD in Lowland Britain ................................................................. 135 Stephen Rippon and Ralph Fyfe
Late Antique Environment and Economy in the North of the Iberian Peninsula: The Site of La Tabacalera (Asturias, Spain) ....................................... 155 Leonor Peña-Chocarro, Almudena Orejas Saco del Valle, Yolanda Carrión Marco, Sebastián Pérez-Díaz, José Antonio López-Sáez, and Carmen Fernández Ochoa
Olive Cultivation and Olive Products in Southern Apulia (6th–11th c.) ....... 172 Giovanni Stranieri
Case Studies – East Environment, Climate and Society in Roman and Byzantine Butrint ........... 185 Mario Morellón, Gaia Sinopoli, Adam Izdebski, Laura Sadori, Flavio Anselmetti, Richard Hodges, Eleonora Regattieri, Bernd Wagner, Brunhilda Brushulli and Daniel Ariztegui Some Thoughts on Climate Change, Local Environment, and Grain Production in Byzantine Northern Anatolia ......................................................... 200 John Haldon Antioch in the Sixth Century: Resilience or Vulnerability? ............................. 207 Lee Mordechai Human and Deltaic Environments in Northern Egypt in Late Antiquity ... 224 Penelope Wilson
General Surveys Climatic Changes and Their Impacts in the Mediterranean during the First Millennium AD ...................................................................................................... 247 Inga Labuhn, Martin Finné, Adam Izdebski, Neil Roberts and Jessie Woodbridge Mysterious and Mortiferous Clouds: The Climate Cooling and Disease Burden of Late Antiquity .............................................................................................. 271 Timothy P. Newfield Invisible Environmental History: Infectious Disease in Late Antiquity ...... 298 Kyle Harper Settlement, Land Use and Society in the Late Antique Mediterranean, 4th–7th c. An Overview ................................................................................................. 314 Alexandra Chavarría, Tamara Lewit and Adam Izdebski Modelling the Supply of Wood Fuel in Ancient Rome ...................................... 330 Benjamin Graham and Raymond Van Dam
Contents
Rye’s Rise and Rome’s Fall: Agriculture and Climate in Europe during Late Antiquity ................................................................................................................. 342 Paolo Squatriti
Concluding Essays Contours of Environmental Change and Human Response in Late Antiquity ................................................................................................................. 355 Kyle Harper The Environmental Turn: Roll Over Chris Wickham? ......................................... 361 Mark Whittow
Conclusion Catastrophes Aside: Environment and the End of Antiquity .......................... 367 Adam Izdebski Abstracts in French ........................................................................................................ 371 Indices ................................................................................................................................ 377 Late Antique Archaeology ........................................................................................... 387
ix
Acknowledgements The conference which inspired this issue and volume 11 met at the Society of Antiquaries in London, in September 2016, under the title ‘Environment and Society in the First Millennium AD’. It was held in conjunction with the Princeton Institute for International and Regional Studies, Climate Change and History Research Initiative, and the Jagiellonian University in Krakow. The conference was organised by Adam Izdebski and Luke Lavan, assisted by Michael Mulryan. Financial support for these two volumes came from the Ministry of Science and Higher Education of the Republic of Poland, through the National Programme for the Development of the Humanities for the years 2016–19. Our thanks also go out to the many people with whom we have discussed different aspects of this project for their advice and encouragement, and to the referees who read the articles contained here and offered many helpful comments. Adam Izdebski and Michael Mulryan
Jagiellonian University in Krakow, University of Kent
List of Contributors Daniel Abel-Schaad is a researcher specialising in the field of paleobotany, who has developed his career both in public and private institutions. His main interests are plant ecology from a palaeoecological perspective, and its implications on the conservation and management of relict species in Mediterranean mountains. Francisca Alba-Sánchez is Professor at the University of Granada, Spain. Her research expertise is in palaeoecology, conservation biology, plant-environment relationships and ecological niche modeling. She has focused on glacial refugia and postglacial spread for Mediterranean and North African relict plant/vertebrates species. Flavio Anselmetti is a geologist at the Institute of Geological Sciences and the Oeschger Centre for Climate Change Research at the University of Bern. He specialises in sediment-based reconstructions of climate change, natural hazards, as well as human-environment interactions. Daniel Ariztegui is Professor of Limnogeology and Geomicrobiology at the Department of Earth Sciences of the University of Geneva (Switzerland). He uses marine and lacustrine sediments as archives of (paleo) environmental changes, as well as of present and former anthropogenic impact on the environment. He is a sedimentologist and geochemist by training, and is particularly interested in understanding biogeochemical cycles in lakes. He was recently elected president of the International Association of Sedimentologists (IAS). Ellen Arnold is an Associate Professor of History at Ohio Wesleyan University. She is the author of Negotiating the Landscape: Environment and Monastic Identity in the Medieval Ardennes (2013) and is currently working on a cultural history of early medieval rivers. Brunhilda Brushulli is an oil and gas engineer and obtained a Masters in Geosciences in the Faculty of Geology and Mining at the Polytechnic University of Tirana (Albania). She works as a geological engineer for the Directorate of Hydrogeology in the Albanian Geological Survey (Tirana). She has participated in several projects focused on the assessment and monitoring of groundwater, gas and oil resources in Albania. She also analysed the sediments and pore waters of Lake Butrint, under the supervision of Prof. Flavio Anselmetti, at the Swiss Federal Institute of Aquatic Science and Technology in 2012. Her recent publications include co-authorship of “Human-climate interactions in the central Mediterranean region during the last millennia: the laminated record of Lake Butrint (Albania)”, Quaternary Science Reviews (2016). Yolanda Carrión Marco is a researcher at the Department of Prehistory, Archaeology and Ancient History of the University of Valencia, Spain. She specialises in charcoal analysis and her research deals with woodland exploitation, landscape reconstruction, land use and the history of species. Her recent work includes the study of Middle and Upper Palaeolithic charcoal records from Iberian and North African sites.
List of Contributors
Alexandra Chavarría is Associate Professor of Medieval Archaeology in the Department of Cultural Heritage, University of Padua, Italy. She specialises in the study of late antique and early medieval urban and rural landscapes, settlements, burials and churches. Her recent publications include A la sombra de un imperio. Reyes, obispos e iglesias en la Hispania tardoantigua (2018), and the edited volume Ricerche sul centro episcopale di Padova : ricerche 2011–2012 (2017). Petra Dark is an archaeological scientist based in the Department of Archaeology, University of Reading. She specialises in the analysis of pollen and other plant remains for reconstructing human-environment interactions on a variety of spatial and temporal scales. Her publications include The Environment of Britain in the First Millennium AD (2000) and, as co-author, The Landscape of Roman Britain (1997). Merle Eisenberg is a post graduate research associate and lecturer at Princeton University. He studies Late Antiquity and the Early Middle Ages along with environmental history. He has forthcoming publications on the Justinianic Plague in Past & Present and Byzantine and Modern Greek Studies. Carmen Fernández Ochoa is Professor of Archaeology at the Universidad Autónoma de Madrid. She has directed the excavation of the city of Gijón since 1982, and also directs several other excavation and research projects on Roman Spain. Her work has focused on the study of different aspects of the Romanisation process of Hispania, especially in the the north-west of the peninsula and Asturias. She has published more than a hundred works on this subject. Martin Finné is a researcher at the Department of Archaeology and Ancient History, Uppsala University. His main research focus is on Holocene climate in the eastern Mediterranean and the paleoclimatology and socio-environmental dynamics of southern mainland Greece. His recent publications include the peer-reviewed paper “Late Bronze Age climate change and the destruction of the Mycenaean Palace of Nestor at Pylos”, PLOS ONE (2017), and the book chapter “Climate change and ancient societies—facing up to the challenge of chronological control”, in The Resilience of Heritage: Cultivating a Future of the Past (2018). Assunta Florenzano is a palynologist and archaeobotanist. She is a post-doc researcher and Adjunct Professor of Botany at the University of Modena and Reggio Emilia (Italy). Her research interests are Quaternary palynology and paleoecology, pollen analysis from archaeological deposits, anthropogenic impact on vegetation, and the study of cultural landscape transformations in the Mediterranean basin. Her recent publications focus on archaeopalynological evidence of cultural landscapes in the central Mediterranean, with emphasis on nonpollen palynomorphs, like coprophilous fungi, as indicators of pastoral activities. Ralph Fyfe is a Professor at the University of Plymouth, UK. He specialises in the analysis of pollen from sedimentary basins. He works across a range of spatial scales, from the site-level (on archaeological projects) to national and continental-scale synthesis of vegetation patterns across Europe.
xiii
xiv Didier Galop is a researcher at the GEODE – Université Toulouse Jean Jaures (France). He specialises in the study of the temporal dynamics of vegetation, and the anthropisation of high mountain environments in south-western Europe. Benjamin Graham is an Assistant Professor in the Department of History at the University of Memphis. He studies the Mediterranean environment in Late Antiquity and the Early Middle Ages. He is presently writing a book about the production and consumption of olives in early medieval Italy. John Haldon is a history Professor at Princeton University and Director of the Princeton Climate Change and History Research Initiative. His research focuses on the social, economic, institutional and cultural history of the medieval eastern Roman (Byzantine) empire, in particular in the period from the 7th to the 12th c.; on state systems and structures across the European and Islamic worlds from late ancient to early modern times; on the impact of environmental stress on societal resilience in pre-modern social systems; and on the production, distribution and consumption of resources in the late ancient and medieval world. His most recent books include A Tale of Two Saints. The Martyrdoms and Miracles of Sts. Theodore ‘the Recruit’ and ‘The General’ (2016) and The Empire That Would Not Die: the Paradox of Byzantine Survival, 640–740 (2016). Kyle Harper is Senior Vice President and Provost and Professor of Classics and Letters at the University of Oklahoma. His work has focused on the social and environmental history of the Roman empire and Late Antiquity. His books include Slavery in the Late Roman World, AD 275–425 (2011); From Shame to Sin: The Christian Transformation of Sexual Morality in Late Antiquity (2013); and The Fate of Rome: Climate, Disease, and the End of an Empire (2017). Richard Hodges is President of The American University of Rome. He has been on the faculty at the University of Sheffield and the University of East Anglia, as well as Director of the Prince of Wales’ Institute of Architecture, and Williams Director of the University of Pennsylvania Museum. He was Scientific Director of the Butrint Foundation. His most recent books include Dark Age Economics: a New Audit (2012) and the The Archaeology of Mediterranean Placemaking (2016). Adam Izdebski is adiunkt at the Institute of History of the Jagiellonian University in Krakow. From the summer of 2018, he is also the leader of the ‘Byzantine Resilience’ Independent Max Planck Research Group at the Max Planck Institute for the Science of Human History in Jena. He specialises in late antique, Byzantine and environmental history. His recent publications include A Rural Economy in Transition. Asia Minor from Antiquity into the Early Middle Ages (2013), and a series of papers co-authored in the special issue of the Quaternary Science Reviews (2016) devoted to climatic and environmental history of the Mediterranean, which he also co-edited. He is a member of Princeton University’s Climate Change and History Research Initiative. Katerina Kouli is a palynologist, and is Assistant Professor of Palaeontology-Palaeobotany-Geoarchaeology at the National and Kapodistrian University of Athens (Greece). Her research is focused on Quaternary vegetation of the Mediterranean, and a main part of her work considers
List of Contributors
List of Contributors
human impact on vegetation and the study of Holocene cultural and natural landscapes. She has participated in several research projects of both geological and archaeological interest, including numerous archaeological excavations. Her work has been published in international scientific journals, book chapters and conference proceedings. Jamie Kreiner is Associate Professor of History at the University of Georgia. Her current research highlights the relationships between early medieval politics, religion, and agriculture. She is the author of Legions of Pigs in the Early Medieval West (forthcoming with Yale University Press) and The Social Life of Hagiography in the Merovingian Kingdom (2014). Inga Labuhn is a postdoctoral researcher at the Department of Geography, University of Bremen, Germany. The present work was carried out at the Department of Geology, Lund University, Sweden. She specialises in Holocene climate reconstructions based on proxy records from tree rings, lake sediments and speleothems. She is the lead author of “French summer droughts since 1326 CE: a reconstruction based on tree ring cellulose δ18O”, Climate of the Past 12 (2016) 1101–17: doi: 10.5194/cp-12-1101-2016; and “Holocene hydroclimate variability in central Scandinavia inferred from flood layers in contourite drift deposits in Lake Storsjön”, Quaternary (2018) 1.2: doi: 10.3390/quat1010002. Tamara Lewit is an Honorary Fellow in the School of Historical and Philosophical Studies at The University of Melbourne, and a Fellow of the Society of Antiquaries, London. She specialises in the study of Roman and late antique farming, rural settlement, technology, and trade. Her recent publications include “The second sea: exchange between the Mediterranean and the Black Sea in Late Antiquity”. Post-Classical Archaeologies 5 (2015) 149–74, and “Oil and wine press technology in its economic context: screw presses, the rural economy and trade in Late Antiquity”, Antiquité Tardive 20 (2012) 137–49. José Antonio López-Sáez is a Tenured Scientist in the Institute of History at the Spanish National Research Council (CSIC), Madrid, Spain. His research interests include: palaeopalynology and archaeopalynology (pollen, spores, non-pollen palynomorphs) from different periods (Palaeolithic to Middle Ages) and different sedimentary contexts (lakes, peat bogs, archaeological sites, palaeosoils, etc.); and prehistoric agriculture, with particular emphasis on its origins and expansion in the Mediterranean, Mesoamerica, and high-mountain environments during the Holocene. Alessia Masi has a post-doc position at Sapienza University in Rome. She specialises in the archaeobotanical and palynological analysis of Mediterranean sites mainly located in Italy, Turkey and the Balkan peninsula. She is one of the pioneers in the use of stable isotopes in archaeobotany to infer paleoclimate and cultivation practices. She deals with the reconstruction of past vegetation, its evolution, and the uses of plants, with particular attention to the effects of human impact and population resilience. Lucas McMahon is a doctoral candidate in the Department of History at Princeton University. His research focuses on Byzantine political and economic history of the 7th through 12th c., and his dissertation examines the dissemination and movement of information within and around the Byzantine empire.
xv
xvi Anna Maria Mercuri is a biologist, palynologist and archaeobotanist. She is Associate Professor of Systematic Botany and head of the Laboratory of Palynology and Palaeobotany at the University of Modena and Reggio Emilia (Italy). She is the coordinator of the Group of Palynology and Palaeobotany of the Italian Society of Botany. Her main research topics focus on long-term human impact and global change in the Mediterranean and Saharan areas (including two papers in the journal Nature Plants). She has directed European and national projects on past/present vegetation and cultural landscape development, and on the integration between palaeoecology and archaeology in the Mediterranean basin. Lee Mordechai is the inaugural Postdoctoral Fellow in Byzantine Studies at the University of Notre Dame. He studies the interaction between premodern complex societies in the eastern Mediterranean and the environment. His recent publications include “History meets palaeoscience: consilience and collaboration in studying past societal responses to environmental change”, Proceedings of the National Academy of Sciences (2018) and “Short-term cataclysmic events in premodern complex societies”, Human Ecology (2018). Mario Morellón geologist, is a lecturer of Geodynamics at the University of Cantabria (Santander, Spain). He investigates lake and estuarine sediments as archives of past climate changes and human-environment interactions over the last millennia from the late glacial to the Anthropocene. His recent publications include: “Early Holocene humidity patterns in the Iberian Peninsula reconstructed from lake, pollen and speleothem records”, Quaternary Science Reviews (2018) and lead author for “Human-climate interactions in the central Mediterranean region during the last millennia: the laminated record of Lake Butrint (Albania)”, Quaternary Science Reviews (2016). Michael Mulryan is Honorary Research Fellow at the University of Kent (UK) and Associate Researcher at the École Normale Supérieure, Paris. He is the volume editor of Late Antique Archaeology and its supplementary series. He was a DFG Research Fellow at the University of Tübingen in 2018, and was Deputy Director of the University of Kent Ostia Field Project. His interests lie in late antique urban topography. His recent publications include Spatial Christianisation in Context: Strategic Intramural Building in Rome From the 4th–7th c. AD (2014). He is currently compiling a volume on the archaeology of late antique Ostia. Timothy P. Newfield is an Assistant Professor in the Departments of History and Biology at Georgetown University. An environmental historian and historical epidemiologist, his recent journal articles have addressed late antique animal disease ( Journal of Roman Archaeology), 1st millennium volcanism (Geology; Climatic Change), Frankish malaria (Early Medieval Europe), the need for collaborative multidisciplinary scholarship in environmental history ( Journal of Interdisciplinary History), and premodern societal resilience to climatic change (PNAS). Almudena Orejas Saco del Valle is a researcher at the department of Archaeology of the Institute of History, Spanish Council for Scientific Research. She is a specialist in landscape archeology, applied to the Roman provinces. She works on several regions of the north-western Iberian Peninsula, particularly in Roman gold mining areas. She has participated and chaired national and international projects and networks (COST Actions, JPI-Heritage). She is co-editor of the volume La Fábrica de Tabacos de Gijón. Arqueología e historia de un espacio milenario/
List of Contributors
List of Contributors
The Tobacco Factory of Gijón. Archaeology and History of a Millenary Space (2015) and she is currently coordinating further research at the Gijón Tobacco Factory (2016–18). David J. Patterson is a PhD Candidate at the University of Michigan in Ann Arbor. His research focuses on the environmental history of early medieval Europe, and particularly on weather and climate in the Frankish world. He is interested not only in the ground realities of weather and climate, but also in the cultural perception of meteorological phenomena and medieval ways of thinking about weather. Leonor Peña-Chocarro is a researcher at the Instituto de Historia (Spanish National Research Council) in Madrid. She works on plant macro-remains (seeds and fruits) from different periods, with a special interest in the origins and development of agriculture. She has recently edited a three volume book series: EARTH: the Dynamics of Non-Industrial Agriculture: 8000 Years of Resilience and Innovation (2014 and 2015). Sebastián Pérez-Díaz is an environmental archaeologist specialising in the synchronism between cultural and environmental changes. He has focused his research on: the synergisms between sociocultural, climatic, and environmental processes during the Upper Pleistocene; on the first farmers and the spread of agriculture; and on the management of historical landscapes. Eleonora Regattieri is a junior researcher at the Earth Science Department of the University of Pisa. The focus of her research is the reconstruction of past climate and environmental variability in the Mediterranean region, through the geochemical and chronological study of continental carbonates (speleothem and lake sediment). Her recent publications include: “A MIS 9/MIS 8 speleothem record of hydrological variability from Macedonia (FYROM)”, Global and Planetary Change (2018), and “A Last Interglacial record of environmental changes from the Sulmona Basin (central Italy)”, Palaeogeography, Palaeoclimatology, Palaeoecology (2017). Stephen Rippon is Professor of Landscape Archaeology at the University Exeter, UK. His research focuses on the development of the countryside across the Roman and Medieval periods, and the evolution of regional variation in landscape character. His recent publications include Kingdom, Civitas and County (2018), The Fields of Britannia (2015) and Making Sense of an Historic Landscape (2012). Neil Roberts is Professor of Physical Geography, at the University of Plymouth, UK. He is the author of The Holocene: an Environmental History (3rd edn. 2014), and is editor of Quaternary Science Reviews. He researches the environmental history of the Mediterranean via the natural ‘archive’ preserved in lake sediments, including diatom, stable isotope, pollen and micro-charcoal analyses. These enable the reconstruction of past climate changes and landscape transformations brought about by human activities. Laura Sadori is a palynologist and archaeobotanist, and Full Professor of Systemtic Botany at the University ‘La Sapienza’, Rome (Italy). She is vice-president of the International Federation of Palynological Societies. She carries out palaeoecological studies aimed at evaluating the degree of human impact and climate change in the Mediterranean basin,
xvii
xviii collaborating with geologists, archaeologists and historians. She has published more than 150 scientific articles, and is included in the list of Top Italian Scientists. Abigail Sargent is a Ph.D student in the History Department at Princeton University. She focuses on legal status and rural communities in high medieval north-west Europe. She will be publishing the article “The ecology of the Crusader states”, in A Companion to the Environmental History of Byzantium (2018) this year. Gaia Sinopoli is an Italian palynologist and archaeobotanist with a Ph.D in Earth Sciences, curriculum Environment and Cultural Heritage. She has studied the pollen from the sediments of Lake Butrint (Albania) to reconstruct the environmental changes that occurred during the last four millenia in the area. Her recent publications have been focused on the climate reconstruction of the Last Interglacial Complex, on the basis of high resolution pollen data from the sediments of Lake Ohrid (Albania/FYROM), the oldest lake in Europe. Paolo Squatriti teaches pre-modern European history at the University of Michigan. He specialises in the medieval environmental history of Italy. Cambridge University Press released the paperback edition of his Landscape and Change in Early Medieval Italy: Chestnuts, Economy, and Culture in 2017. Giovanni Stranieri is a researcher and Lecturer in Medieval Archaeology at the Centre d’histoire, archéologie, littératures médiévales (CIHAM/UMR 5648), Lyon (France). He currently works on southern Italy, the Adriatic region and Auvergne (France). His research interests are focused on landscape archaeology, agrarian production, land divisions and spatial organisation in the Middle Ages. Raymond Van Dam is Professor Emeritus in the Department of History at the University of Michigan. His research focuses on Roman, Early Medieval, and Early Byzantine history, and his books include Rome and Constantinople: Rewriting Roman History During Late Antiquity (2010) and Remembering Constantine at the Milvian Bridge (2011). Bernd Wagner is Assistant Professor for Quaternary Geology at the Institute of Geology and Mineralogy, University of Cologne, Germany. The focus of his studies is palaeoenvironmental reconstruction based on lake sediment successions from the polar regions to the Mediterranean. He has more than 100 publications in international, peer-reviewed journals. Most recently, he compiled a special issue of the journal Biogeosciences (2016) that examined the environmental and evolutionary history of Lake Ohrid, Europe’s oldest lake. Mark Whittow (1957–2017) was University Lecturer in Byzantine Studies and a Fellow of Corpus Christi College, Oxford. He was a medieval historian and archaeologist, specialising in the Mediterranean and Byzantine worlds, AD 500–1300. His major publications included The Making of Orthodox Byzantium, 600–1025 (1996), and “The Late Roman/Early Byzantine Near East”, in The New Cambridge History of Islam 1 (2010).
List of Contributors
List of Contributors
Penelope Wilson is Associate Professor of Egyptology in the Department of Archaeology, Durham University, UK. She specialises in Egyptian settlement archaeology in the Nile Delta from the earliest Neolithic periods to the modern day, as well as Ptolemaic hieroglyphic texts. She directs excavations at Sais (Sa el Hagar) as the Delta Survey project of the Egypt Exploration Society. Her recent publications include: Sais I: The Ramesside to Third Intermediate Period (2011), Sais II: the Prehistoric Period (2014) and, with D. Grigoropoulos, The West Delta Regional Survey: Beheira Kafr el Sheikh Provinces (2009). Jessie Woodbridge is a Post-Doctoral Research Fellow at the University of Plymouth (UK). She is currently conducting research that aims to reconstruct changes in European land cover throughout the Holocene using pollen data. Her research background is based on palaeoenvironmental reconstruction, specifically in relation to understanding human impacts and past climate trends.
xix
Introductory Essay
Setting the Scene for an Environmental History of Late Antiquity Adam Izdebski Abstract Environmental history is a well-established discipline that until recently focused mainly on the modern era and was dominated by historians. Numerous scholars agree today that this needs to change: a focus on Late Antiquity can help this happen. To make it possible, we should concentrate our efforts on three parallel projects. First, make late antique studies more interdisciplinary, i.e. joining the efforts of historians, archaeologists and natural scientists. Second, look at Late Antiquity and the early Middle Ages as a source of case studies that are relevant to the central themes of environmental history. Third, use environmental history as a new framework that has the potential to modify our vision of the 1st millennium AD, by getting us closer to the actual experience of the people who lived this past.
This book is quite unusual in its attempt to address two scholarly communities that normally have nothing to do with each other: scholars of Late Antiquity and (modern) environmental historians. Exchanges between these two groups are so infrequent that most environmental historians probably only have a vague idea of what Late Antiquity is, while ‘late antiquers’ have little awareness of what environmental historians actually do. This introduction will attempt to alleviate these inequalities. It will briefly explain what the notion of ‘Late Antiquity’ entails, in cultural-historical as well as environmental terms. At the same time, this essay is meant to offer a brief sketch of the themes that are central to the field of environmental history. It will also explain why these themes are relevant for the study of Late Antiquity, and how Late Antiquity can help us to deepen our understanding of the issues that have been keeping environmental historians busy for several decades. Before continuing, however, it is important to explain briefly why both groups will benefit if they start a meaningful dialogue. In fact, it is the same reason that is relevant for both: there is a surprisingly high number of similarities between the socio-environmental developments that occurred in Late Antiquity and those that we see in the Early Modern period, the era in human history that, until recently, was providing the vast majority of case studies that environmental historians
have been working on.1 What is more, research on late antique landscapes and environments, mostly done by archaeologists and palaeoecologists, has produced such a wealth of data that—when we combine them with the huge amounts of diverse textual sources that survived from the period—we are indeed able to ask ‘modern’ questions when studying this ‘remote’ period. In other words, the scholars of Late Antiquity are in a position to start addressing new types of topics, synthesising their evidence in a novel way, and filling important gaps in their knowledge of the period. On top of that, as will be shown in this volume, environmental studies can actually help them answer old questions—addressed repeatedly by generations of scholars—with new data and approaches. Environmental historians, on the other hand, can benefit from this research effort and a conversation with ‘late antiquers’ by deepening, relativising and refining our understanding of the socioenvironmental phenomena that have been preoccupying us ever since the discipline of environmental history came into being. Late Antiquity itself is a problematic notion, in a state of constant flux, and, until recently, has rarely been connected to anything environmental. More than a hundred years ago, the term was first coined by art historians, trying to escape the all-powerful narrative of decline and fall that was used to frame the final centuries of the Roman empire: in politics as well as in arts.2 This insistence on rejecting the notions of decline, or even any idea of significant rupture, remained characteristic for all those who unite under the aegis of Late Antiquity. The period encompasses centuries on both sides of the ‘great divide’ between antiquity and the Middle Ages, conventionally set on the year when the (western) Roman empire formally disappeared (AD 476). In its most generous interpretations, Late Antiquity extends from the late 2nd c. AD until the 8th c. AD. In this way, not only is the continuity of the Roman empire in the
1 For a discussion of the similarities between the Roman (including late antique) world and the modern period, in the economic and cultural context, see Pitts and Versluys (2015). 2 Giardina (1999); reprinted in English translation in Cameron (2013).
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_002 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 3–13
4 East put to the front of the historical narrative, but the emphasis is actually put on structural continuities, that encompass the post-Roman kingdoms in the West, the Persian empire of the Sassanians in the East, and finally even the Arabic empire of the Umayyads and early Abbasids. The period that is defined in this way stands out for achievements in institution building and administration, economic complexity, and, above all, religious and cultural innovation. Adopting the framework of Late Antiquity, in the end, is a way to show that the Abrahamic religions (Judaism, Christianity, and Islam) share a common late antique heritage that is formative to their identities. All three were either created or significantly reshaped in this period, in constant dialogue with each other, and in many ways their late antique configurations persist until today.3 Any discussions of geography in the context of Late Antiquity still tend to focus on culture and, at best, geopolitics. None of the recent companions, guides or handbooks on Late Antiquity contain a chapter that discusses what environmental historians call the ‘environment’.4 In a way reminiscent of the historical geography that was characteristic for the early Annales school, authors writing about the geography of Late Antiquity are busy showing how physical phenomena (seas, rivers, mountains, fertile plains, etc.) shaped political boundaries. Others focus on purely cultural geography, offering overviews of the spatial distribution of the different cultural components of the late antique world. Such introductions serve to orientate the author and the reader in physical space, but they do not take account of non-human actors in history, which is, to put it very simply, the task of environmental history. The fact that the geography of Late Antiquity is defined by politics, religion and culture means that this world is environmentally extremely diverse. It centres on the Mediterranean—in itself a very diverse region: 3 This is visible in both the earliest, foundational works on Late Antiquity, such as Brown (1971), and in the most recent studies, in which the notion of Late Antiquity is picked up and further elaborated on by scholars of Islam (e.g. Schmidt et al. (2016)). 4 Rousseau and Raithel (2009); Johnson (2012). In this last volume, “environmental aspects” are discussed in the context of rural history, but they do not form a topic in its own right (they are mentioned in Cam Grey’s chapter “Concerning rural matters”, on pp. 625–66). The degree to which cultural studies used to dominate the field’s understanding of the environment is well-illustrated by the fact that although the seminal Guide to Late Antiquity (co-edited by the field’s founder, Peter Brown (Bowersock et al. (1999)) does include chapters on “habitat” (by Y. Hirschfeld) and “landscape” (by R. Webb)), this habitat is understood as human dwellings of all kinds (so nothing to do with ecology). Similarly, landscape is discussed in the context of artistic representation, as a scenery in which stories, depicted or described, take place.
Izdebski
consider Morocco on the one hand, and the Balkans on the other—but it extends into northern Europe, as far as Scotland and Ireland, and into Central Asia, at least as far as Iran. There is no doubt that at some point in Late Antiquity these areas formed a cultural commonwealth, particularly if we think of it as underpinned either by Christianity or by Islam, or by both at the same time toward the end of the period. However, one can also rightly say that ecologically these lands are so diverse and disconnected that they can hardly be treated as a single unit of analysis, for the purpose of environmental history. On the other hand, the same could be said about Europe or China, or even North America, and this does not prevent anyone from writing environmental histories of these huge geographical entities.5 In the case of Late Antiquity, defining the spatial boundaries for writing environmental history through culture rather than through physical geography, has important consequences. It means that we focus in particular on how a specific cultural and social system interacted with the diverse environments it was part of. In this book, therefore, we actually set out to explore what such diversity, paired with relative socio-cultural unity, means to our understanding of humanity’s interaction with the environment in the course of history. When discussing ‘late antique’ environments, we should return again to the question of continuity, so central to any attempts at defining Late Antiquity. While not rejecting the notions of gradual developments in many spheres, culture and religion in particular, and accepting that social and economic structures tended to remain relatively stable, in this volume we also insist that Late Antiquity was a period of reconfiguration and major change, in particular when it comes to the environment. The great empires of the end of antiquity disintegrated, changed their extent, yielded themselves to new imperial entities; trade routes and local economies were reshaped, world systems fragmented and re-emerged in new configurations. As environmental historians well know, these processes are always combined with ecological change. What is more, the natural environment itself was not stable, and complex socio-economic and political systems interacted with shifting climates and disease ecologies. If writing an environmental history of Late Antiquity is both possible and necessary today, it is only thanks to the tremendous amount of work already done on the Mediterranean, Middle Eastern and European environments of the 1st millennium AD by numerous scholars, coming from a variety of academic backgrounds. As this
5 Well-known examples include: Elvin (2004); Merchant (2007); Miller (2007); Wills (2013); Hoffmann (2014).
Setting the Scene for an Environmental History of Late Antiquity
research is summarised in detail in the bibliographic essay that follows this introduction, here let us just sketch briefly some major developments. What is crucial, such developments do not necessarily stem from a clearly expressed interest in the environment. Rather, environmental questions are discussed in connection with the research agendas of each of the disciplines that developed within the larger framework of late antique studies over the past half a century. Thus, among historians, those who show the strongest interest in any natural phenomena are social and economic historians. In fact, the debate on land use and the organisation of rural labour (including a number of complex legal issues related to this topic) predates the birth of late antique studies: byzantinists were discussing these problems already at the end of the 19th c. Further, historians as well as archaeologists have also studied the movement of goods—be it through tax or trade—between different parts of the Roman empire. This, in fact, is paving the way for exploring how exchange patterns created connections between different regional environments and how they impacted the local ecologies (a topic still to be properly explored). More recently, a number of historians (many of whom collaborate with scientists) have been increasingly focusing on the Justinianic Plague, and although much remains to be done, the level of understanding of the plague we have today has greatly increased from a few years ago; the same applies to the closely related field of climate history. As regards classic archaeological topics, there is a very strong interest in the environment among those archaeologists who conduct surface survey projects. In their case, the focus on settlement patterns extends naturally into thinking about the landscape. Consequently, most surveys nowadays include notable environmental components, be it archaeobotany or archaeozoology, carried out on nearby excavations; at times, this also involves semi-independent palynological projects (a study of local vegetation history based on pollen deposited in lake or peat sediments). This results in a very good understanding of how people interacted with their natural environments at a local level, allowing us to better comprehend the actual experience of living in a specific landscape. As such, late antique studies has helped answer several environmental questions for some time, and several key aspects of human interaction with the rest of the natural world have already been explored. To some extent, therefore, it remains a paradox that this entire research effort did not lead to the emergence of a fully-fledged environmental history of Late Antiquity. There are a number of reasons that explain this impasse. The current situation can be explained by the fact that most environmental scholarship during
5
the last two decades was produced within archaeology; until recently, it was receiving relatively little attention from historians, and those historians who now collaborate with scientists are still rather few. The disciplinary divides have been quite important in preventing a major debate from taking place. In addition, the majority of ancient and medieval historians often remain sceptical of scientists’ ability to produce useful reconstructions of past environments, with sufficient chronological and spatial precision. On top of that, the discipline of environmental history itself, both in Europe and the USA, focused on the modern era, and on the industrial world in particular, with almost no attention being paid to earlier periods. As a result, even practitioners of history, as long as they maintained their focus on the 1st millennium AD, had little chance of learning about the on-going debates among their colleagues working on environmental matters. Thus, among all these participants in a potential discussion, there was not enough interest to engage in a serious exchange. The result was a growing body of fascinating scholarship on late antique environments, but hardly any debate as to what this work actually tells us about the nature of social-ecological phenomena in Late Antiquity. This volume hopes to change this situation. We want to show the ways in which the themes that environmental historians have been focusing on over the last half a century can become of interest for those studying Late Antiquity and the 1st millennium AD. Scholars of environmental history have developed historiographical models that could be successfully deployed in premodern historical contexts. On the one hand, doing that will make it possible to add a new dimension to our narrative of the transitional era linking antiquity and the Middle Ages. We will gain a more holistic understanding of the social processes and everyday experiences of human life. Equally, we are gradually recognising that taking the environment into account is becoming a necessity. The Roman empire did expand under specific climatic conditions, in many respects, conditions that were quite advantageous. Roman society was transforming Mediterranean environments in unprecedented ways. Justinian was confronted with a climatic downturn and an outbreak of a major pandemic. Several earlier tendencies were reversed, and new anthropogenic environments emerged in the process of this transition. We should not see such events as merely mortality figures, changes to taxation systems or trade patterns, we should understand their environmental dimension in its entire complexity. Environmental history provides ready models, narrative structures and questions that make this task possible. At the same time, the field of environmental history will benefit from embracing Late Antiquity. The
6 1st millennium AD has something important to say in the context of all the major themes of environmental history, and will help its practitioners to liberate themselves from their very limiting, modern examples. Ancient societies achieved unprecedented levels of social-economic complexity and ‘globalisation’, not much different from those achieved in the Early Modern period.6 The late antique era was the culmination of a centuries-long process, and our parallel focus on the Early Middle Ages makes it possible to see what follows after the trends toward complexity and globalisation disappear, something we have not yet experienced on such a scale in the modern era. Therefore, there is much to learn for environmental historians from late antique and early medieval case studies. The 1st millennium AD will help them to relativise the judgements they make on how highly complex societies manage their environments; ancient and medieval historians can show that we are not confined to early modern or almost contemporary examples! In fact, studying the late antique and early medieval evidence brings the promise of helping us understand better the environmental change that the post-industrial world has been experiencing over the past decades. Finally, since we are dealing with a past that is much farther away than the last few centuries of the modern era, there is no other way to study the environmental history of Late Antiquity than to do it interdisciplinarily. We need the evidence and approaches coming from archaeology and environmental science as much as we need those that come with historical research and historiographical models. Each of them has their own role to play, and each fills an important gap. Together, they offer a more complete and holistic narrative about the past.7 Consequently, one of the major aims of this volume is to facilitate and intensify the interdisciplinary dialogue. The structure of this book is determined by these three closely related goals. The editors did everything they could to gather a group of authors among whom history, archaeology and science would be represented on equal terms. Sharing results and approaches across the disciplinary divides was one of the major concerns of this volume. Therefore, after a bibliographic essay that complements this introduction, there follows a special section devoted to scientific research on the vegetation history of the 1st millennium AD. Four overview papers
6 Geraghty (2007); Pitts and Versluys (2015); see also the work of the Oxford Roman Economy Project and their numerous publications: http://www.romaneconomy.ox.ac.uk/, accessed 8 August 2017. 7 Achieving this is not an easy task: see Izdebski et al. (2016) for a recent and comprehensive discussion.
Izdebski
present the results of the work of generations of scientists, on the eastern, central and western Mediterranean, as well as on northern Europe. These papers set the scene for case studies from all parts of the Roman (and post-Roman) world. These studies represent contrasting environments, different historical moments, and diverse research methodologies. The volume contains both historical and archaeological papers that are concerned with environmental questions, as well as papers that engage with evidence coming from a number of disciplines, including a classic multi-archive and multi-proxy study of a small locality in the Balkans (by Morellón et al.). While representative of most recent trends in environmental science, this particular paper shows how state-of-the-art scientific research can easily be connected to historical and archaeological projects, something that this volume hopes to encourage in the field of late antique studies. The final section presents a number of overview papers on key aspects of Mediterranean environmental history during the 1st millennium AD: climate, landscape and disease. It focuses on the way in which the Roman world underwent the transition from its culmination phase in Late Antiquity into its early medieval disintegration. Finally, a series of brief concluding essays tries to tease out a few lessons that could be learnt from this volume, both for the study of Late Antiquity, and for the field of environmental history at large. The academic discipline that is known today as environmental history had different origins on each side of the Atlantic. In the United States, it emerged in the 1960s in response to the birth of ecological consciousness in North America. Initially, it was not a research programme, but rather a way of teaching American history that responded to the issues that were hotly debated on campuses across the United States. It took a few years before the first conferences were organised and pioneering monographs and articles were published, which finally helped the ‘environmental history project’ assume the shape of a fully-fledged academic discipline. What is important is that the origins of this intellectual endeavour can be traced back not only to the North American environmental movements of the 1960s and 1970s, but also to the revolution in historiography that was taking place in the United States at the same time. Environmental history was establishing itself in academia along with histories of gender, slavery, and labour; seeing nature as another subaltern that should be given a ‘voice’ in modern historiography. In Europe, the interest in the environment and history started in a different way. Already at the end of the 19th c., there developed a strong tradition of historical geography. It focused on reconstructing the space of
Setting the Scene for an Environmental History of Late Antiquity
European cities and the countryside in the past. Such interests often encouraged more daring questions into the role of natural factors in shaping the course of history, prefiguring later debates on historical determinism. This early work entered a new stage with the groundbreaking study of the Mediterranean by Fernand Braudel. Published soon after the Second World War (and revised some ten years later), it originated in the debates that were taking place in the French Annales school already in the interwar period. From the very beginning, this new current in French and European historiography aimed to approach history in a holistic way, which assumed an interest in the role of geographical factors in historical processes. Consequently, Braudel devoted a large portion of his study of the 16th c. Mediterranean to investigating the geographical foundations of the region’s history. He thus laid the groundwork for the historical study of the ways in which nature ‘sets the scene’ for human action. A number of historians across Europe continued his pioneering work, looking into different natural factors (such as climate) that mattered for human societies in the past. Gradually, they began to depart from Braudel’s static approach and adopted more dynamic visions of the relationship between society and the environment in the historical past.8 As research on environmental history was progressing on both sides of the Atlantic, there emerged a number of themes that continue to influence the debates among environmental historians and historical geographers. Historically, the most prominent of these was the interest in the relationship between economic activity and environmental change. This, of course, is a complex issue in itself, and has been approached in a number of different ways. One of the fist scholars to consciously write ‘environmental’ history, Donald Worster, made this question one of his central concerns. Influenced by postwar debates on the limits to economic growth, and by the ecological challenges of the 1970s, he focused on the ways in which capitalism brought about environmental change and reshaped ecological mentalities in North America in the 19th and 20th c. In a number of works, he showed that careless or ignorant use of environmental resources (such as soils or plants) and the transformation of nature associated with the unending search for profit, had damaging consequences. In his view, modern capitalism, which knows no limits, is unsustainable. Consequently, environmental change associated with
8 For relevant references, see the detailed introductions to environmental history that have been published in recent years. The most useful ones include: Winiwarter and Knoll (2007); Herrmann (2013); Isenberg (2014); Quennet (2014).
7
the expansion of modern economic systems ultimately leads to a catastrophe.9 Views similar to Worster’s were widespread among the first generation of environmental historians, to the extent that they influenced their views of ancient history. Donald Hugh’s pioneering overview of Greek and Roman environmental history constantly concerns itself with the ‘damage’ that the ancients did to their environments, making their lifestyle unsustainable in the long run. In his opinion, this was one of the major factors that brought about the collapse of Graeco-Roman civilisation.10 Our volume expressly challenges this view of ancient environmental history; seeing the past through the lens of catastrophes, be they political or environmental, is not a way to discover human experience. As the vegetation history overviews in the first section of this book make clear (by Roberts et al., Kouli et al., López-Sáez et al., and Woodbridge et al.), even though there was a common theme to all the landscape transformation that was taking place across the Roman world in Late Antiquity, it is impossible to detect a single trajectory that one could describe as irreversibly damaging the environment. The same conclusions can be reached if we approach this question through archaeological evidence from surface surveys and archaeobotanical studies, as is demonstrated by Chavarria et al. in their paper on the settlement and land use in the Mediterranean in Late Antiquity and the early Middle Ages. They show that throughout these two periods, local and regional environments were in a state of constant change, responding to shifts in economic and political conditions; this process is depicted in further detail in the context of rye’s spread across early medieval Europe in a paper by Squatriti. Finally, Graham and Van Dam show that there is actually evidence that makes it possible to directly contradict the ‘unsustainability and collapse’ hypothesis. Late Romans actually did care for their environmental resources, and the collapse of the imperial political order might actually have resulted in greater environmental transformation (and an unsustainable use of resources) than if the empire had survived. In fact, focusing on the question of the ecological sustainability of historical economic systems is just one way of writing about the relationship between environment and economy in the past. Economic activity, including profit-driven conquests—among the most impressive being the discovery and colonisation of the Americas by early modern Europeans, dubbed
9 Worster (1979), (1988), (1992). 10 Hughes (1993).
8 the ‘Columbian exchange’11—creates ecological connections between regions and places that hitherto would have remained isolated. As a result, trade and economic expansion unifies not only local economic systems, but also local ecologies. Each new connection of that type brings about an environmental change in all ‘participating’ environments. Depending on the context and the species involved, sometimes these changes are so subtle that they escape human perception, while at other times they completely alter the old ways of living, either by reshaping plant and animal landscapes, or by creating conditions that encourage the spread of virulent diseases. In the end, the trend toward increasing economic-environmental connectivity leads to an ecologically globalised and unified world, such as we know from our own experience in the 21st c. Late Antiquity had its own ‘Columbian exchange’, with all of its complications. The papers by Stranieri and Morellón et al. show how well-tuned the local environments of the central Mediterranean were to changing patterns of exchange and trade. As waves of globalisation came and went—connecting southern Italian and Albanian ecologies to different central points of the Mediterranean exchange networks, be it Constantinople, Rome or somewhere else—local populations transformed their environments, adapting to the new conditions and reconfiguring local plant landscapes. For example, rye’s career in the early Middle Ages, described by Squatriti, actually shows the same phenomenon, but on a more general scale. On the microbial level, it is of course no coincidence that Late Antiquity saw the outbreak of the first known pandemic, the Justinianic Plague. As Harper and Newfield explain in their papers, the scale of the disease events in Late Antiquity was proportional to the extent of the ecological connections that entwined the Middle East, the Mediterranean and western Europe at that point of history. Yet another approach to studying the economy and environment in the past, is to focus on how the economy influences human perceptions of nature and human connections to it. Again, environmental historians have mostly focused on the industrial period; the most famous study of this type deals with the development of Chicago’s food supply and food market in the 19th c. In this process, nature became commodified: plant and animal products, such as grain, were no longer real plant elements or whole organisms that were connecting a consumer with a specific farmer and a specific environment. Instead, they were becoming abstract commodities that were no longer part of the natural 11 Crosby (1972), (1986).
Izdebski
world. Local environments started becoming elements of a huge para-industrial machine that was transforming vast tracts of North American landscapes.12 While this type of study is easily conceivable for Constantinople or other Mediterranean mega-cities in Late Antiquity, it is yet to be undertaken. In this book, this topic is present in the papers by Stranieri and Morellón et al., showing Constantinople’s impact on distant environments connected to the city through maritime networks; the paper by Mordechai also includes a discussion of the potential impact of Antioch on Syria’s environment. The ‘machine’ metaphor also appears in environmental history in other contexts. It is common in studies that explore links between technology and the environment. Water is often described as a part of nature that easily becomes integrated into human machines. Initially, this phenomenon was believed to have begun with the first wave of industrialisation, and culminated in turning whole rivers into reservoirs of hydro-energy, or coolants for nuclear plants and other factories.13 Recent work is pointing out that this process actually started much earlier, both in terms of practical solutions and theoretical thinking.14 In the case of this book, such a natural machine can be said to emerge in the case of the Nile Delta, described by Wilson. Here, all the different deltaic environments were connected into a single environmental infrastructure, or living machine, meticulously managed by local human populations.15 The issues of technology and the environment also encompass the crucial question of energy regimes and transitions in history: the study of the ways in which different societies harvested energy for everyday and industrial use. Major works have focused on the transitions from solar regimes (tapping resources from living plants and animals) to fossil fuels, yet there clearly is much scope for the study of how different societies of the pre-industrial era were meeting their energy needs and solving energy bottlenecks.16 In this volume, crucial aspects of the energy regime characteristic of Late Antiquity, and its sustainability, are addressed by Graham and Van Dam. After the economy and technology, another important human phenomenon, whose environmental dimensions have been studied by historians, are social institutions. This particular debate began with studies focusing on
12 As described in Cronon (1991). 13 Steinberg (1991); White (1995). 14 Hoffmann (2014) would see the beginning of this process in the Early Modern period. 15 For the notion of environmental infrastructure, see Kreike (2013). 16 Sieferle (1990); Burke (2009); Bulliet (2012).
Setting the Scene for an Environmental History of Late Antiquity
9
the Anglo-Saxon colonisation of New England in the 17th and 18th c. William Cronon, one of the pioneers of environmental history in the United States, demonstrated that each of the populations that were living in New England in the first decades of colonisation—native Americans as well as English settlers—were constructing their own productive landscapes. These landscapes were sustained by legal and social institutions, in particular by ideas of ownership and organisation of labour. On the one hand, these institutions were shaping the landscape, on the other, once in place, such productive landscapes required these institutions and determined them. Without proper institutions, these productive landscapes could not have existed, and neither the English nor the Americans could have maintained their lifestyles. Crucially, in the same way that the social institutions of the natives and of the settlers were mutually exclusive, also their landscapes could not co-exist within the same territory. In fact, both sides regarded the landscapes created by the ‘other’ as wild (non-cultivated) or absurd (unsustainable). As a result, the technological and economic advantages that the English settlers enjoyed finally led to a profound transformation of New England’s landscape in the course of the 17th c., and the early disappearance of native American ecologies and lifestyles in this part of North America.17 While Cronon mainly focused on legal institutions and local politics, other environmental historians have shown that many other social institutions also have their environmental dimensions. For instance, this same process of environmental change in colonial New England also involved a conflict between native American and English understandings of gender roles, as well as of general worldview and religious customs, all of them having an impact on how people interacted with the environment.18 Late Antiquity, with all its cultural, political and religious diversity, also provides an obvious era for asking these types of questions, as much as earlier periods of Graeco-Roman Antiquity. In this volume, links between social institutions and landscapes loom large in the regional vegetation history overview papers by Roberts, Kouli et al., López-Sáez et al., and Woodbridge et al. These links are also addressed in papers that focus on the late antique-early medieval transition: Rippon and Fyfe’s study of post-Roman Britain, Chavarria et al.’s review of Mediterranean settlement and land use histories, and in Squatriti’s chapter on the spread of rye in post-Roman Europe. Thinking about the ways in which the economy, social roles and institutions influence landscape and resource
use, leads us to another major theme of environmental history: the protection and conservation of nature. To some extent, this was the original focus for the very first generation of environmental historians in the United States; a debate on the role of nature for American identity and the development of nature conservation actually pre-dated the birth of environmental history as an academic discipline.19 The idea of the ‘wilderness’ permeates this volume in many ways. Several papers, for instance the study of a late Roman and postRoman well deposit in the city of Gijón in Spain, by Peña-Chocarro et al., shows the process of rewilding what used to be anthropogenic, managed landscapes. The vegetation history overviews could also be read in this way: to show how the profoundly anthropogenic landscapes of Late Antiquity receded into a state of secondary ‘wilderness’ as the Roman order disappeared across Europe and the Mediterranean. In fact, nature preservation or conservation—at first glance, essentially an environment-focused activity— always has a political dimension. In historical times, there was no pristine nature that was ‘untouched by man’, so some human groups always needed to be deprived of their environment and alienated from the ecological basis of their lifestyle, in order to create environments that could be regarded as worthy of special protection. Even more, nature preservation can become a powerful tool of social domination, when groups in power want to control and ‘rationally’ manage natural resources, such as forests or watersheds: as happened, for instance, in the Maghreb during the times of French colonisation.20 In other words, the protection of nature and natural resources can be seen as a social conflict, which, in the end, has the power to determine access to environmental resources, and define what is in need of protection. What needs protection does not have to be a ‘wild’ environment though, it could well be a cultivated landscape that is of utmost importance for the state’s existence. Such a situation is described for Byzantine Anatolia in Haldon’s chapter, in particular in the context of the eastern Roman taxation system in the Early Middle Ages. The assumption that, in historical times already there was no ‘pristine’ environment, no place on earth untouched by human action, can in fact be considered a central tenet of environmental history. Consequently, a number of historians have focused on researching anthropogenic environments that existed and developed in the past, in order to understand both the ways in which they were created, and the mechanisms of
17 Cronon (1983). 18 Merchant (1989).
19 Worster (1988); McNeill (2003). 20 Davis (2007).
10 (human) coping with new environmental conditions that were (often unexpectedly) emerging as a result of human activity. Two such man-made environments stand out as especially important for the field of environmental history: frontiers and cities. The former, best defined as areas where new ways of exploiting environmental resources and managing landscapes meet old ways of doing things, was first identified as an historical-environmental phenomenon in the context of Anglo-American colonial expansion in the central areas of North America. There existed a clearly identifiable frontier of English (or French) agriculture that was gradually moved westward. As a result, it created areas where indigenous environmental traditions mixed with new strategies, that focused on intensive, market- oriented cultivation.21 Movement and mixing are in fact central to the idea of an environmental frontier, a dynamic place where indigenous human ecologies meet with external, outside influences. Environmental historians quickly noticed that this phenomenon was not limited to colonial and 19th c. North America, but was occurring in many parts of the world, in particular in the Early Modern period. As a result, a number of major works on environmental frontiers focused on countries such as early modern China or Japan, and even Europe itself.22 In Late Antiquity, such environmental frontiers existed in several places, in particular at the interface of the Roman empire and the outer ‘barbarian’ world, as suggested by Woodbridge et al. in their analyses of northern Europe’s vegetation history. It would also be interesting to ask how this frontier changed and moved ‘inwards’ (if at all!) as Germanic groups migrated into the Roman empire at the end of antiquity. An example of how the interplay of Roman presence and local ecological traditions could look like, is also provided by Dark in her study of the Hadrianic frontier in northern Britain. Cities are of course different, as, unlike frontiers, they are not mobile. Rather, the uniqueness of the city as an anthropogenic environment stems from the fact that the intensive use of local and outside natural resources takes place continuously in the same place. This results in a pollution problem, which urban authorities and town-dwellers have been confronted with from antiquity up to the present day. Living in a polluted environment requires adaptation strategies, but even if these are in place, pollution can have still serious consequences for human health. It was this coping
21 White (1991). 22 A global overview focusing on the Early Modern period is Richards (2003); China: Marks (1998); Japan: Walker (2001); Europe: Warde (2006); Appuhn (2009).
Izdebski
with an urban environment that was often the catalyst for something we would call ecological, or even protoecological, awareness in different historical cultures. In this context, while American environmental historians tended to focus mainly on the development of sanitary ideas in the late industrial era,23 in Europe historians also studied other types of pollution (different from waste) and historical recycling strategies.24 The question of urban waste disposal, and the degree to which this process was interwoven with processes of ecological succession, forms the backbone of Peña-Chocarro et al.’s paper in this volume, which presents a study of a well deposit from the Cantabrian city of Gijón. Pollution is of course not the only phenomenon that makes urban environments so special. Cities were also centres of consumption, which means that they were developing complex and intensive ecological connections with their immediate hinterlands, as well as with more distant places. In some cases this resulted in the emergence of truly global cities, and late antique Constantinople surely counted among them. Cities acted as hotspots of ecological connectivity, and at the same time they were major causal factors behind landscape change that was occurring in regions connected to them. In this book, the role of (mega-)cities as catalysts of regional landscape change is addressed in Mordechai’s study of 6th c. Antioch, in the context of this city’s coping with earthquake-related crises. Finally, there also exists a strong tradition of seeing nature as an exogenous agent or factor that intervenes in the course of human history. While focusing on such research questions carries a greater risk of slipping into environmental determinism and mono-causal explanations, proper consideration of both social and natural factors makes it possible to show that the actual impact of any (natural) exogenous shock is to a great degree socially constructed. In this book, the complexities of this problem are demonstrated by Mordechai in the case of Antioch, a major urban centre which benefited from state-funded resilience measures in response to a series of damaging earthquakes that hit the city in the course of the 6th c. To some extent, climate history can also be seen as representing a focus on exogenous forces of nature that ‘intervene’ in the course of human events. The field of climate history concerns itself first of all with reconstructing climate conditions in the past. This is how the field emerged in the 1970s,25 and still today exact and chronologically precise reconstructions of
23 Tarr (1996); Melosi (2005). 24 See a detailed literature review in Quennet (2014). 25 Le Roy Ladurie (1967).
Setting the Scene for an Environmental History of Late Antiquity
past temperatures and precipitation remain a major concern, even if nowadays this is more the task of natural scientists than historians. In the second place, climate history involves the study of social aspects of climate: its perceptions as well as its role in different human activities, not necessarily assuming that a climatic shift could ‘change’ the course of history.26 In fact, historical studies arguing for a major role for climatic change during ‘great’ historical episodes are rare. On the one hand, the causal mechanisms involved were highly complex, on the other, historical societies often proved resilient to climatic changes.27 Historical studies of climate’s social dimension now tend to avoid deterministic reasoning, as it obscures the more contingent and conjunctural nature of the links between climate and society. Climatic, or more broadly, environmental determinism, is more likely to occur in large-scale overviews.28 In fact, even Braudel’s magisterial study of the Mediterranean did not escape deterministic thinking, as he was focusing on the ‘limiting’ and ‘restricting’ role of geographical factors in Mediterranean history, instead of seeing human adaptation to ecological change or the active creation of new environments.29 Climate history permeates this volume from its beginning to the end. This has to do with both the recent flourishing of climate-related research on Late Antiquity, and with the fact that (changing) climate is a factor that has to be taken into account, or at least discussed, in the context of a number of other major environmental phenomena, such as disease or landscape change. In order to provide our readers with the necessary background for understanding these different debates, one of the papers, by Labuhn et al., reviews all the palaeoclimatic evidence that is currently available for the 1st millennium AD. Mediterranean. The authors discuss the regionality of Mediterranean climate histories and the consequences of this fact for our understanding of the climate’s role in the history of Late Antiquity. Apart from that, the potential impact, or lack thereof, that climatic changes could have on the life of late antique and early medieval societies, are addressed in more detail in five papers. Most importantly, climate emerges as a major, if not yet fully understood, factor controlling the occurrence of human and animal disease, as demonstrated 26 Pfister (2007), (2010). 27 A well-received monograph arguing for a climatic role in the 17th c. crisis of the Ottoman empire, is White (2011); for more detailed discussions of the links between climate and history, see Haldon et al. (2014); Izdebski et al. (2016). 28 Such as Diamond (1997). 29 As demonstrated by Horden and Purcell (2000) and in particular by Grove and Rackham (2001).
11
by Harper as well as Newfield. Its role in economic and landscape history is less obvious. Rippon and Fyfe show that there is actually no grounds to assume that climate played a role in the transformation of Britain’s landscapes once this region ceased to be part of the Roman empire in the 5th c. Similar claims of no direct impact of climatic change on the landscape, can be made for Anatolia and much of the Middle East, as summarised in this volume by Roberts and Haldon. Interestingly, at the same time, Haldon points out that the specific climatic conditions of 7th and 8th c. Anatolia might have increased the productivity of cereal cultivation, which at that time became the main focus of Byzantine agriculture. In addition, Morellón et al. show that some Mediterranean environments—humid coastal plains, for instance—are largely resilient to climatic fluctuations by their very nature. In the case of their study area, the impact of increased precipitation was felt mainly by the residents of Butrint, who had to cope with the rising water table in the city’s suburbs. This short overview of the major themes of environmental history—and their treatment in this volume—would not be complete if it did not discuss a major bias shared by much environmental history, at least until quite recently. Inspired by the ecological controversies of the 1960s, the field was for a long time struggling over the ways to morally evaluate human influence on the environment; its practitioners were often inclined to see it in purely negative terms. The pioneers of the field devoted much attention to showing how American capitalism at its different stages led to environmental degradation, the eradication of traditional practices (believed to be more sustainable), and, ultimately, ecological catastrophe. As a result, the field became pervaded with a declentionist narrative: it is a belief that each new human intervention in matters of the ‘environment’ worsens the situation.30 In this vision, the 20th c. emerges as the culmination of this process, a point of no return, and the historian’s task is to explain how we arrived at this deplorable situation.31 Human impact, however, can be seen as something positive, or at least neutral: as another ecological process that started with the spread of the human species across the globe several hundred thousand years ago. It does lead to the creation of a deeply anthropogenic world, but this does not mean that the anthropogenic world
30 For a detailed discussion of environmental history’s coping with the challenge of a declentionist narrative, see the introduction to Isenberg (2014); for the capitalist context, see Uekoetter (1998). 31 For the debates on the 20th c. in environmental history, see for instance Pfister and Brimblecombe (1990); McNeill (2000).
12
Izdebski
should be regarded as something negative; neither is the task just to explain the ‘roots of the current crisis’. The notion of the Anthropocene points in fact to the positive ways of seeing human activity within the broader natural environment. Despite the fact that many scientists tend to consider the Anthropocene to have begun rather recently—with the industrial revolution or with the changes in technology and energy consumption that occurred in the mid 20th c.—environmental archaeologists argue that the construction of man’s own ecological niches had already been transforming the Earth for thousands of years.32 Thus, environmental history is gradually reorienting itself, from tracing human mismanagement of nature, to describing and explaining the different ways in which human societies shaped their relationship with the environment: and how this relationship changed societies themselves. As this introductory essay makes clear, there is huge potential for the field of the environmental history of Late Antiquity to grow, and increase our understanding of how the place of humans in the natural world changed over time. In order for this potential to be realised, more work is needed and this volume hopes to help it grow. From this essay, four recommendations emerge as crucial for this process: 1.
2.
3.
Writing an environmental history of Late Antiquity cannot be achieved without collaboration between historians, archaeologists and natural scientists. All types of evidence are needed to reconstruct these distant social-ecological phenomena, and the skills of all three professions are necessary to make sense out of them. Historians in particular should become more involved in this process. The debate on different environmental aspects of late antique history has been on-going for a few decades now, but it has been largely taking place within the field of archaeology. It should be opened up to other disciplines, as well as to new theoretical frameworks. These new approaches will in turn inspire completely new research questions and agendas. Environmental data should be treated just like every other source of information on the past. Historians, for example, can query this abundant material in order to find answers to questions they have been asking for generations, and which they were unable to address properly due to the deficiencies of the written or material evidence.
32 Boivin et al. (2016); Ellis et al. (2016).
4. As scholars of Late Antiquity, we should strive to take this opportunity to change and refine the explanatory frameworks we use for talking about what we already think we know. This volume, as well as several other recent publications, make clear that we finally ought to factor in the environment—climate, disease, vegetation, ecological processes etc.—into the mainstream narrative of Late Antiquity. Acknowledgements The author acknowledges funding received from the Ministry of Science and Higher Education, Poland (National Programme for the Development of the Humanities, 2016–19). This text benefited greatly from the time spent at the Institute of Advanced Study in Princeton in the academic year 2017–18. Bibliography Appuhn K. R. (2009) A Forest on the Sea: Environmental Expertise In Renaissance Venice (Baltimore 2009). Boivin N. L., Zeder M. A., Fuller D. Q. et al. (2016) “Ecological consequences of human niche construction: examining long-term anthropogenic shaping of global species distributions”, Proceedings of the National Academy of Sciences 113 (2016) 6388–96. doi: 10.1073/pnas.1525200113 (accessed November 2017). Bowersock G. W., Brown P., Grabar O. (1999) edd. Late Antiquity: a Guide To The Postclassical World (Cambridge, Mass. 1999). Brown P. (1971) The World of Late Antiquity: From Marcus Aurelius to Muhammad (London 1971). Bulliet R. W. (2012) “History and animal energy in the arid zone”, in Water on Sand, ed. A. Mikhail (Oxford 2012) 51–67. Burke E. (2009) “The big story. Human history, energy regimes, and the environment”, in The Environment and World History, edd. E. Burke and K. Pomeranz (Berkley-London 2009) 33–53. Cameron A. (2013) ed. Late antiquity on the Eve of Islam (Farnham 2013). Cronon W. (1991) Nature’s Metropolis : Chicago and the Great West (New York 1991). Cronon W. (1983) Changes in the Land: Indians, Colonists, and the Ecology of New England (New York 1983). Crosby A. W. (1986) Ecological Imperialism : the Biological Expansion of Europe, 900–1900 (New York 1986). Cronon W. (1972) The Columbian Exchange: Biological and Cultural Consequences of 1492 (Westport, Connecticut 1972). Davis D. K. (2007) Resurrecting the Granary of Rome: Environmental History and French Colonial Expansion in North Africa (Athens, Ohio 2007). Diamond J. M. (1997) Guns, Germs, and Steel: the Fates Of Human Societies (New York 1997). Ellis E., Maslin M., Boivin N. and Bauer A. (2016) “Involve social scientists in defining the Anthropocene”, Nature News 540 (2016) 192. doi: 10.1038/540192a (accessed November 2017). Elvin M (2004) The Retreat of the Elephants: an Environmental History of China (New Haven, Connecticut 2004).
Setting the Scene for an Environmental History of Late Antiquity Geraghty R. M. (2007) “The impact of globalization in the Roman empire, 200 BC–AD 100” The Journal of Economic History 67 (2007) 1036–61. Giardina A. (1999) “Esplosione di Tardoantico”, Studi Storici 40 (1999) 157–80. Grove A. T. and Rackham O. (2001) The Nature of Mediterranean Europe: an Ecological History (New Haven, Connecticut 2001). Haldon J., Roberts N., Izdebski A. et al. (2014) “The climate and environment of Byzantine Anatolia: integrating science, history and archaeology”, Journal of Interdisciplinary History 45 (2014) 113–61. Herrmann B. (2013) Umweltgeschichte—Eine Einführung in Grundbegriffe (Berlin 2013). Hoffmann R. C. (2014) An Environmental History of Medieval Europe (Cambridge 2014). Horden P. and Purcell N. (2000) The Corrupting Sea: a Study of Mediterranean History (Oxford 2000). Hughes J. D. (1993) Pan’s Travail: Environmental Problems of the Ancient Greeks and Romans (Baltimore 1993). Isenberg A. C. (2014) ed. The Oxford Handbook of Environmental History (Oxford 2014). Izdebski A., Holmgren K., Weiberg E. et al. (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 (2016) 5–22. doi: 10.1016/j.quascirev.2015.10.038 (accessed November 2017). Johnson S. F. (2012) ed. The Oxford Handbook of Late Antiquity (New York 2012). Kreike E. (2013) Environmental Infrastructure in African History: Examining the Myth of Natural Resource Management in Namibia (Cambridge 2013). Le Roy Ladurie E. (1967) Histoire du climat depuis l’an mil (Paris 1967). Marks R. (1998) Tigers, Rice, Silk, and Silt: Environment and Economy in Late Imperial South China (Cambridge 1998). McNeill J. R. (2003) “Observations on the nature and culture of environmental history”, History and Theory 42 (2003) 5–43. doi: 10.1046/j.1468-2303.2003.00255.x (accessed November 2017). McNeill J. R. (2000) Something New Under the Sun : an Environmental History of the Twentieth-Century World (London 2000). Melosi M. V. (2005) Garbage in the Cities : Refuse, Reform, and the Environment (Pittsburgh, revised edn. 2005). Merchant C. (2007) American Environmental History: an Introduction (New York 2007). Merchant C. (1989) Ecological Revolutions: Nature, Gender, And Science in New England (Chapel Hill, North Carolina 1989). Miller S. W. (2007) An Environmental History of Latin America (New York 2007). Pfister C. (2010) “The vulnerability of past societies to climatic variation: a new focus for historical climatology in the twenty-first century”, Climatic Change 100 (2010) 25–31. doi: 10.1007/s10584010-9829-2 (accessed November 2017).
13
Pfister C. (2007) “Climatic extremes, recurrent crises and witch hunts strategies of european societies in coping with exogenous shocks in the late sixteenth and early seventeenth centuries”, The Medieval History Journal 10 (2007) 33–73. doi: 10.1177/097194580701000202 (accessed November 2017). Pfister C. and Brimblecombe P. (1990) edd. The Silent Countdown: Essays in European Environmental History (Berlin 1990). Pitts M. and Versluys M. J. (2015) edd. Globalisation and the Roman world: World History, Connectivity and Material Culture (Cambridge 2015). Quennet G. (2014) Qu’est-ce que l’histoire environnementale (Seyssel 2014). Richards J. F. (2003) The Unending Frontier: an Environmental History of the Early Modern World (Berkeley 2003). Rousseau P. and Raithel J. (2009) edd. A Companion to Late Antiquity (Malden, Mass. 2009). Schmidt N., Schmid N. K. and Neuwirth A. (2016) edd. Denkraum Spätantike. Reflexionen von Antiken im Umfeld des Koran (Wiesbaden 2016). Sieferle R. P. (1990) “The energy system—a basic concept of environmental history”, in The Silent Countdown: Essays in European Environmental History, edd. P. Brimblecombe and C. Pfister (Berlin 1990) 9–20. Steinberg T. (1991) Nature Incorporated: Industrialization and the Waters of New England (New York 1991). Tarr J. A. (1996) The Search For the Ultimate Sink: Urban Pollution in Historical Perspective (Akron, Ohio 1996). Uekoetter F. (1998) “Confronting the pitfalls of current environmental history: an argument for an organisational approach”, Environment and History 4 (1998) 31–52. Walker B. L. (2001) The Conquest of Ainu Lands: Ecology and Culture in Japanese Expansion, 1590–1800 (London 2001). Warde P. (2006) Ecology, Economy and State Formation in Early Modern Germany (Cambridge-New York 2006). White R. (1995) The Organic Machine (New York 1995). White R. (1991) The Middle Ground: Indians, Empires, and Republics in the Great Lakes Region, 1650–1815 (New York 1991). White S. (2011) The Climate of Rebellion in the Early Modern Ottoman Empire (Cambridge 2011). Wills J. (2013) US Environmental History: Inviting Doomsday (Edinburgh 2013). Winiwarter V. and Knoll M. (2007) Umweltgeschichte: eine Einführung (Cologne 2007). Worster D. (1992) Rivers of Empire: Water, Aridity, and the Growth of the American West (New York 1992). Worster D. (1988) “Doing environmental history”, in The Ends of the Earth: Perspectives on Modern Environmental History, ed. D. Worster (Cambridge 1988) 289–307. Worster D. (1979) Dust Bowl: the Southern Plains in the 1930s (New York 1979).
Bibliographic Essays
The Environmental History of the Late Antique Eastern Mediterranean: a Bibliographic Essay Lucas McMahon and Abigail Sargent Introduction Environmental history is a relatively young field, one still in the process of being defined. In North America it has developed to some extent as a discrete (and selfreflective) sub-discipline of history. Elsewhere, however, it remains more nebulous, consisting of studies that use a variety of methods to approach sets of questions related to the environment in the past. The first concern of environmental historians was (and continues to be) to call attention to the central role that the environment has played in human societies throughout history, leading to two broad questions. First, what did the natural world present as the context—the environment—for human decisions and actions? Historical climate is the most obvious topic to be pursued under this heading, along with natural disasters like earthquakes, tsunamis, volcanoes, and even floods. Similarly, disease is often construed as an environmental force, bringing humans into detrimental contact with other organisms. The natural world is usually more beneficent, of course, and answering this question of context also leads to assessments of the resources available to past societies: minerals, soils, landscapes, and even plants. The second question concerns how humans have used the resources available to them, and sometimes, how they have responded to the challenges thrown at them. Environmental historians therefore investigate many different types of resource exploitation, from mining to fishing to agriculture. Beyond these direct uses of the natural world, they also study the ways human settlement patterns and the cultivation of woodlands, pastures, and other semi-domesticated spaces, have been shaped by the resources and constraints in a given region. Studying humans in a landscape quickly brings up a second consideration which environmental history has embraced vigorously: it becomes obvious that humans do not simply use, but actually create, their environments; the relationship is reciprocal. Direct impacts from agriculture and from deliberate landscape changes (terracing, irrigation, land clearance and such) are the most obvious ways in which humans can change their environments. More recently, scholarly focus has shifted
to changes that are unintentional, subtler, or harder to trace; mostly in vegetation (beyond deforestation), but also in soils at the chemical level. Human influences on local climates are now the subject of several studies. It can be helpful to think of late antique environmental history as usually operating within some combination of these three related questions: what, physically, was the environment of a given past society?; how did that society use or respond to that environment?; and how did that society shape its environment? Most pre-modern environmental historians occupy themselves with these kinds of questions. They have not constructed the kind of theoretical frameworks that have become fashionable in modern and especially American environmental history, which are reframing the definitions of nature and culture, as in Cronon (1996), Kreike (2013) and Boomgaard (2001). The mirror image of the kind of environmental history outlined above, which addresses human conceptions of the natural world, for example in Douthwaite (2002), has also been mostly neglected for Late Antiquity (but see Glacken (1967), Redon (2008) and Delort (1995)). Because it touches on almost every physical element of human civilisation, and the field is still young, deciding what studies count as environmental history is perilous. The bibliographies presented here are therefore not exhaustive. Instead, they include those works which unmistakably focus on environmental history, along with some of those in fields which address similar concerns. Likewise, in a field that asks many questions that are not constrained by political boundaries in time or space, it can be difficult to set the necessary boundaries around these bibliographies. In general, ‘Late Antiquity’ has been loosely taken to comprise the period from AD 300 to 800, and the ‘eastern Mediterranean’ includes those areas that were under the control of the eastern Roman empire in AD 395. Sometimes books or articles have been included which fall outside of these geographic and temporal boundaries because they illustrate an important and relevant topic which has not been studied within it. In each section, the most important or useful studies have been listed first; the others follow alphabetically.
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_003 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 17–30
18 General syntheses: Arnold E. F. (2008) “An introduction to medieval environmental history”, History Compass 6 (2008) 898–916; Horden P. and Purcell N. (2000) The Corrupting Sea: a Study of Mediterranean History (Oxford 2000); Bevilacqua P. (2010) “The distinctive character of Italian environmental history”, in Nature and History in Modern Italy, edd. M. Armiero and M. Hall (Athens 2010) 15–32; Cronon W. (1996) ed. Uncommon Ground: Rethinking the Human Place in Nature (New York 1996); McNeill J. R. (1992) The Mountains of the Mediterranean World: an Environmental History (Cambridge 1992); Glacken C. J. (1967) Traces on the Rhodian Shore: Nature and Culture in Western Thought from Ancient Times to the End of the Eighteenth Century (Berkeley 1967); Boomgaard P. (2001) Frontiers of Fear: Tigers and People in the Malay World, 1600–1950 (New Haven, Connecticut 2001); Daim F. (2016) ed. Byzanz. Historisch-kulturwissenschaftliches Handbuch (Stuttgart 2016); della Dora V. (2016) Landscape, Nature, and the Sacred in Byzantium (Cambridge 2016); Delort R. (1995) “Percevoir la nature au moyen âge”, in Campagnes médiévales: l’homme et son espace: études offertes à Robert Fossier (Paris 1995) 31–43; Douthwaite J. V. (2002) The Wild Girl, Natural Man, and the Monster: Dangerous Experiments in the Age of Enlightenment (Chicago 2002); Harris W. V. (2013) ed. The Ancient Mediterranean Environment Between Science and History (Leiden 2013); Kelly G. (2004) “Ammianus and the great tsunami”, JRA 94 (2004) 141–67; Kreike E. (2013) Environmental Infrastructure in African History: Examining the Myth of Natural Resource Management in Namibia (Cambridge 2013); Philippson A. (1939) Das byzantinische Reich als geographische Erscheinung (Leiden 1939); Redon O. (2008) Des forêts et des âmes: espace et société dans la Toscane médiévale. Études rassemblées par Laurence Moulinier-Brogi (Saint-Denis 2008); Sallares R. (2009) “Environmental history”, in A Companion to Ancient History, ed. A. Erskine (Oxford 2009); Squatriti P. (1998) Water and Society in Early Medieval Italy, AD 400–1000 (Cambridge 1998); Telelis I. (2014) “Environmental history and Byzantine studies. A survey of topics and results”, in Aureus: Volume Dedicated to Professor Evangelos K. Chrysos (Athens 2014) 737–60; Thommen L. (2012) An Environmental History of Ancient Greece and Rome (Cambridge 2012); Thonemann P. (2011) The Maeander Valley: a Historical Geography from Antiquity to Byzantium (Cambridge 2011); Vita-Finzi C. (1969) The Mediterranean Valleys: Geological Changes In Historical Times (Paris 1969); White S. (2011) “Middle East environmental history: ideas from an emerging field”, World History Connected 8.2 (2011); Winiwarter V. and Knoll M. (2007) Umweltgeschichte. Eine Einführung (Cologne 2007).
Agriculture Agricultural history is one of the oldest branches of the discipline which is relevant to environmental history. Even those works of agricultural history which cannot be categorised as environmental history are usually closely intertwined with it for at least three reasons. First, people working on agricultural history have spent decades studying the ways in which environmental factors shaped human settlement and society. Second, agricultural historians work on the relationship between humans and their environments in the opposite direction; that is, tracing the impacts of human activities on landscapes and ‘natural’ features. Third, these scholars have already done a great deal of preliminary groundwork— delineating agricultural regimes, estimating carrying capacities, considering fertilisation, and so on—
McMahon and Sargent
without which the work of environmental historians would be both poorer and more difficult. Such studies have obvious relevance to environmental history. The works of Lemerle (especially 1979) and Kaplan (especially 1992) are seminal, and even Ostrogorsky (1941) is still cited, though some of his conclusions have been revised. Decker (2009), is recent and important, discussing both agriculture and rural landscapes. These books incorporate, and often focus on, questions about land tenure, rural social structures, power, and mutuality. While these concerns technically fall outside the realm of environmental history, they are an essential background for understanding the ways in which humans related to the non-human elements in their environments (see Sarris (2006) and (2012), Kaplan (1981), Lefort (1993) and (2002), and Svoronos (1956)). Lepelley (2001) contains a useful section on rural life, but which is an example of a largely legal and economic, rather than environmental, perspective. Studies that encompass the agricultural regime as a whole, either for a large region or a small case study, generally fall into two camps. While those already mentioned consider change over time, they are primarily interested in understanding the structures of rural life and agriculture. Other works focus on the changes to those structures, especially in relation to historical events or climate. For examples of this method see: Haldon (2016), Izdebski (2013), Arthur et al. (2012), Djamali et al. (2009), and Watson (1983), although some of Watson’s more sweeping arguments have been refuted. Those which give climate central importance will be discussed later. Agricultural regimes: Arthur P., Fiorentino G. and Grasso A. M. (2012) “Roads to recovery: an investigation of early medieval agrarian strategies in Byzantine Italy in and around the eighth century”, Antiquity 86 (2012) 444–55; Decker M. (2009) Tilling the Hateful Earth: Agricultural Production and Trade in the Late Antique East (New York 2009); Izdebski A. (2013) A Rural Economy in Transition: Asia Minor from Late Antiquity into the Early Middle Ages (Warsaw 2013); Kaplan M. (1992) Les hommes et la terre à Byzance du VIe au XIe siècle: propriété et exploitation du sol (Paris 1992); Lemerle P. (1979) The Agrarian History of Byzantium from the Origins to the Twelfth Century: the Sources and Problems (Galway 1979); Ostrogorsky G. (1941) “Agrarian conditions in the Byzantine empire in the Middle Ages”, in The Cambridge Economic History of Europe from the Decline of the Roman Empire (Cambridge 1941) 194–223; Dalby A. (2011) ed. Geoponika: Farm Work: a Modern Translation of the Roman and Byzantine Farming Handbook (Totnes 2011); Ashburner W. (1910) “The farmer’s law”, JHS 30 (1910) 85–108; Ashburner W. (1912) “The farmer’s law (continued)”, JHS 32 (1912) 68–95; Banaji J. (2009) “Aristocracies, peasantries and the framing of the early Middle Ages”, Journal of Agrarian Change 9 (2009) 59–91; Cappers R. T. J. (2006) Roman Foodprints at Berenike: Archaeobotanical Evidence of Subsistence and Trade in the Eastern Desert of Egypt (Los Angeles 2006); Djamali M. 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”, JAS 36 (2009) 1364–75;
Environmental History of the Late Antique Eastern Mediterranean
19
Green J. A. (1986) The Carthaginian Countryside: Archaeological Reconnaissance in the Hinterland of Ancient Carthage (Ph.D. diss., Univ. of Chicago 1986); Haldon J. (2016) The Empire that Would Not Die: the Paradox of Eastern Roman Survival, 640–740 (Cambridge, Mass. 2016); Huntington E. (1917) “Climate change and agricultural exhaustion as elements in the fall of Rome”, The Quarterly Journal of Economics 31.2 (1917) 173–208; Kaplan M. (1981) “Les grands propriétaires de Cappadoce (VIe–XIe siècles)” in Le aree omogenee della Civiltà Rupestre nell’ambito dell’Impero Bizantino: la Cappadocia. Atti del quinto convegno internazionale di studio sulla Civiltà Rupestre medioevale nel Mezzogiorno d’Italia, ed. C. D. Fonseca (Galatina 1981) 125–58; Kazhdan A. P. and Kaplan M. (1994) “One more agrarian history of Byzantium”, Byzantinoslavica 55 (1994) 66–95; Koder J. (1994) “Historical aspects of a recession of cultivated land at the end of the late antiquity in the east Mediterranean”, in Evaluation of Land Surfaces Cleared from Forests in the Mediterranean Region during the Time of the Roman Empire, ed. B. Frenzel (New York 1994) 157–67; Lefort J. (1993) “Rural economy and social relations in the countryside”, DOP 47 (1993) 101–13; Lefort J. (2002) “The rural economy, seventh–twelfth centuries”, in The Economic History of Byzantium: From the Seventh Through the Fifteenth Century, ed. A. Laiou (Washington, D.C. 2002) 231–310; Lepelley C. (2001) Aspects de l’Afrique romaine: les cités, la vie rurale, le christianisme (Bari 2001); McClellan M. C. and Rautman M. L. (1995) “Where have all the farmers gone? The Cypriot countryside in the seventh to tenth centuries”, in Visitors, Immigrants, and Invaders in Cyprus, ed. P. W. Wallace (Albany, New York 1995) 85–86; Patlagean É. (1977) Pauvreté économique et pauvreté sociale à Byzance, 4e–7e siècles (Paris 1977); Sarris P. (2006) Economy and Society in the Age of Justinian (Cambridge 2006); Sarris P. (2012) “Large estates and the peasantry in Byzantium c. 600–1100”, RBPhil 90 (2012) 429–50; Svoronos N. G. (1956) “Sur quelques formes de la vie rurale à Byzance: petite et grande exploitation”, Annales. Histoire, Sciences Sociales 11 (1956) 325–35; Tate G. (1989) “Les campagnes de la Syrie du Nord à l’époque proto-byzantine”, in Hommes et richesses dans l’empire Byzantin, vol. 1 (Ve–VIIe siècles) (Paris 1989) 63–77; Teall J. L. (1971) “The Byzantine agricultural tradition”, DOP 25 (1971) 33–59; Watson A. M. (1983) Agricultural Innovation in the Early Islamic World: the Diffusion of Crops and Farming Techniques, 700– 1100 (Cambridge 1983).
around the organisation of fields, and the practical side of their cultivation, including what tools were used, how boundaries were determined and kept, and so on: Bryer (2002), Völling (2002) and Martin (2003). While these questions have sometimes led to a purely social or political history, such an approach has also produced works with important insights into environmental history. This is especially the case when they consider irrigation, or other techniques of ‘improvement’: Christensen (1993), Mayerson (1962), Frumkin (1998) and Rahimi-Laridjani (1988). Studies of arable agriculture began by focusing on what limited information was available from documentary and legal sources, and supplementing this with chance references in historical or literary texts. Yet, the contribution of archaeology has also been indispensable almost from the beginning, for which see Mayerson (1962), Völling (2002), and Vigne and Valladas (1996). More recently, palynological analyses have shed light on the plants grown, and the extent of their cultivation: Leroy (2010) and Izdebski et al. (2015). An area of research that has been largely neglected for Late Antiquity is agricultural failure, although Stathakopoulos (2004) and (2012) are exceptions. Nevertheless, studies which address famines, either shortly before or shortly after the period, are useful for basic information (Garnsey (1998)), as are those works looking at western Europe, such as Newfield (2013). Some work that focuses on economic and demographic shifts at the end of antiquity, as a result of raids or population movements, also addresses crop failure (Haldon (2007)), though usually without exploring the social implications.
Arable Agriculture Largely because of the unquestioned status of grain as the staple of the eastern Mediterranean through the ages, arable agriculture has received more scholarly attention than any other type of agriculture. One set of questions which has occupied scholars has to do with how much of a given crop was necessary, how much was grown, and how those levels of production were sustained: Hillman (1973) and Müller (1993). Information about crops in classical antiquity remains relevant for Late Antiquity, for which see Foxhall and Forbes (1982) and Jasny (1944). Sometimes such studies originate in military history, since the ability to feed the army, especially in the Roman empire, was a crucial concern: Erdkamp (2002). Any study of grain production tends to cite Teall (1959), who comes from a largely economic, rather than agricultural, perspective. Yet, the article remains important for having mined the textual sources for information on the demand for, and trade in, grain. Many studies of arable agriculture centre themselves
Arable agriculture: Briant P. (2001) Irrigation et drainage dans l’Antiquité. Qanats et canalisations souterraines en Iran, en Egypte et en Grèce (Paris 2001); Izdebski A., Koloch G. and Słoczyński T. (2015) “Exploring Byzantine and Ottoman economic history with the use of palynological data: a quantitative approach”, JÖB 65 (2015) 67–110; Teall J. L. (1959) “The grain supply of the Byzantine empire, 330–1025”, DOP 13 (1959) 87–139; Canard M. (1959) “Le riz dans le Proche Orient aux premiers siècles de l’islam”, Arabica 6 (1959) 113–31; Christensen P. (1993) The Decline of Iranshahr: Irrigation and Environments in the History of the Middle East, 500 BC to AD 1500 (Odense 1993); Bryer A. (2002) “The means of agricultural production: muscle and tools”, in The Economic History of Byzantium: From the Seventh Through the Fifteenth Century, ed. A. Laiou (Washington, D.C. 2002) 101–13; Erdkamp P. (2002) “The corn supply of the Roman armies during the Principate (27 BC–235 AD)”, in The Roman Army and the Economy, ed. P. Erdkamp (Amsterdam 2002) 47–69; Foxhall L. and Forbes H. A. (1982) “Σιτομετρεία: the role of grain as a staple food in classical antiquity”, Chiron 12 (1982) 41–90; Frumkin A., Greenbaum N. and Schick A. P. (1998) “Paleohydrology of the northern Negev: comparative evaluation of two catchments”, in Water, Environment and Society in Times of Climatic Change, edd. A. S. Issar, and N. Brown (Dordrecht 1998) 97–111; Hillman G. (1973) “Agricultural productivity and past population potential at Aşvan: an exercise in the calculation of carrying capacities”, AnatSt 23 (1973) 225–40; Hirschfeld Y.
20 (1996) “The importance of bread in the diet of monks in the Judean desert”, Byzantion 66 (1996) 143–55; Jasny N. (1944) The Wheats of Classical Antiquity (Baltimore 1944); Leroy S. A. G. (2010) “Pollen analysis of core DS7–1SC (Dead Sea) showing intertwined effects of climatic change and human activities in the Late Holocene”, JAS 37 (2010) 306–16; Martin J.-M. (2003) “L’espace cultivé”, in Uomo e spazio nell’alto Medioevo, vol. 1, ed. O. Capitani (Spoleto 2003) 239–300; Mayerson P. (1962) “The ancient agricultural regime of Nessana and the central Negev”, in Excavations at Nessana (Auja Hafir, Palestine), ed. D. Colt (London 1962) 211–69; Müller A. E. (1993) “Getreide für Konstantinopel. Überlegungen zu Justinians Edikt XIII als Grundlage für Aussagen zur Einwohnerzahl Konstantinopels im 6. Jahrhundert”, JÖB 43 (1993) 1–20; Rahimi-Laridjani F. (1988) Die Entwicklung der Bewässerungslandwirtschaft im Iran bis in sasanidisch-frühislamische Zeit (Wiesbaden 1988); Vigne J.-D. and Valladas H. (1996) “Small mammal fossil assemblages as indicators of environmental change in northern Corsica during the last 2500 years”, JAS 23 (1996) 199–215; Völling T. (2002) “Early Byzantine agricultural implements from Olympia (5th/6th centuries AD)”, in Πρωτοβυζαντινή Μεσσήνη και Ολυμπία, Αστικός και αγροτικός χώρος στη Δυτική Πελοπόννησο, Πρακτικά του Διεθνούς Συμποσίου, Αθήνα 29–30 Μαΐου 1998 / Early Christian Messene and Olympia. Urban and Agrarian Area in the Western Peloponnese. Acts of the International Symposium, Athens, 29–30 May 1998, edd. P. G. Themelis and V. Konti (Athens 2002) 195–207; Zuckerman C. (2016) “On a bountiful harvest at Antioch of Pisidia (with special regard to the Byzantine modios and to the Mediterranean diet)”, in Le saint, le moine et le paysan: mélanges d’histoire byzantine offerts à Michel Kaplan (Paris 2016) 731–51. Famine: Stathakopoulos D. (2012) “Death in the countryside: some thoughts on the effects of famine and epidemics”, AnTard 20 (2012) 105–14; Garnsey P. (1998) “Famine in history”, in Cites, Peasants and Food in Classical Antiquity: Essays in Social and Economic History, ed. W. Scheidel (Cambridge 1998) 272–92; Haldon J. F. (2007) “‘Cappadocia will be given over to ruin and become a desert’. Environmental evidence for historically-attested events in the 7th–10th centuries”, in Byzantina Mediterranea. Festschrift für Johannes Koder zum 65. Geburtstag (Vienna 2007) 215–30; Holman S. R. (1999) “The hungry body: famine, poverty, and identity in Basil’s Hom. 8”, Journal of Early Christian Studies 7 (1999) 337–63; Newfield T. P. (2013) “The contours, frequency and causation of subsistence crises in Carolingian Europe (750–950)”, in Crisis alimentarias en la Edad Media: modelos, explicaciones y representaciones, ed. P. Benito i Monclús (Milenio 2013) 117–72; Stathakopoulos D. C. (2004) Famine and Pestilence in the Late Roman and Early Byzantine Empire: a Systematic Survey of Subsistence Crises and Epidemics (Aldershot 2004).
Other Plant Cultivation Grain may have been the staple, but other crops were also significant. Studies continue to mention in passing vegetable gardens, beans and pulses, and fruit trees, but rarely focus on their cultivation, largely because they are all but invisible in the sources. More attention—though still less than might be expected—has been paid to two quintessentially Mediterranean crops: the olive tree and the grapevine, for which see, for example, Mitchell (2005) and Vermoere (2003). Mattingly (1988) focuses on the western Mediterranean, but also addresses the centrality of trade in the production of olive oil. Papers from a symposium on wine and oil production were published in Amouretti and Brun (1993), covering antiquity to the modern period, but included several useful and extremely detailed papers on Late Antiquity.
McMahon and Sargent Viticulture and Oleoculture: Amouretti M.-C. and Brun J.-P. (1993) edd. La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991) (Athens 1993); Mitchell S. (2005) “Olive cultivation in the economy of Roman Asia Minor”, in Patterns in the Economy of Asia Minor, edd. C. Katsari and S. Mitchell (Swansea 2005) 83–113; Vermoere M. et al. (2003) “Modern and ancient olive stands near Sagalassos (south-west Turkey) and reconstruction of the ancient agricultural landscape in two valleys”, Global Ecology and Biogeography 12 (2003) 217–36; Hitchner R. B. (1993) “Olive production and the Roman economy: the case for intensive growth in the Roman empire”, in La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991), edd. M.-C. Amouretti, and J.-P. Brun (Athens 1993) 499–508; Matijasic R. (1993) “Oil and wine production in Istria and Dalmatia in classical antiquity and the early Middle Ages”, in La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991), edd. M.-C. Amouretti and J.-P. Brun (Athens 1993) 247–61; Mattingly D. J. (1985) “Olive oil production in Roman Tripolitania”, in Town and Country in Roman Tripolitania, edd. D. J. Mattingly, and D. J. Buck (Oxford 1985) 27–46; Mattingly D. J. (1988) “Oil for export? A comparison of Libyan, Spanish and Tunisian olive oil production in the Roman empire”, JRA 1 (1988) 33–56; Tchernia A. (1993) “Le vignoble italien du Ier siècle avant notre ère au IIIe siècle de notre ère: répartition et évolution”, in La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991), edd. M.-C. Amouretti and J.-P. Brun (Athens 1993) 283–96.
Husbandry In the Mediterranean, animal husbandry is never far from crop cultivation, either historically or historiographically. Animals shared space with crops—sometimes harmoniously, sometimes not—but perhaps just as importantly they also shared nutrients, either by fertilizing the fields themselves or because people grew crops to feed them: Wilkinson (1982). Shiel (2006) also provides a detailed scientific assessment of this relationship. Animals were less important for cultivating the ground in the Mediterranean than in northern Europe, but still provided traction in some areas: Schneider (1985) and Fumagalli (1985). Husbandry is frequently referenced in general works, especially in archaeological reports, and osteoarchaeology also continues to be an important source of information, for which see Arbuckle (2009), Ferro (1999), Baker (2011), Cartledge et al. (1992) and De Cupere (2001). Studies devoted to husbandry are common, largely because of the importance of pastoralism in parts of Anatolia and the Levant, especially linked to nomadic or semi-nomadic lifestyles (see Kaplan (2011), Lewit (2009), Tchernov and Horwitz (1990), Thonemann (2011) and Gabba (1985)). Work on Spain can also be useful for comparative purposes, such as in Wickham (1985) and López-Merino et al. (2009). While archaeological and historical works do investigate the environmental impact of grazing herds and their human custodians, much research has instead approached husbandry from the perspective of its contribution to human diet, such as in Fuller et al. (2012).
Environmental History of the Late Antique Eastern Mediterranean Husbandry: L’uomo di fronte al mondo animale nell’alto Medioevo, 7–13 aprile 1983 (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985); Kaplan M. (2011) “L’activité pastorale dans le village byzantin du VIIe au XIIe siècle”, in Animals and Environment in Byzantium (7th–12th c.), edd. E. Anagnostakis, T. Kolias and E. Papadopoulou (Athens 2011) 407–20; Shiel R. S. (2006) “Nutrient flows in pre-modern agriculture in Europe”, in Soils and Societies: Perspectives from Environmental History, edd. J. R. McNeill and V. Winiwarter (Knapwell 2006) 216–42; Anagnostakis E. T. Kolias and E. Papadopoulou (2011) edd. Animals and Environment in Byzantium (7th–12th c.) (Athens 2011); Arbuckle B. S. (2009) “Chalcolithic caprines, dark age dairy, and Byzantine beef”, Anatolica 35 (2009) 179–224; Baker P. (2011) “Assessment of animal bones excavated in 2004–2005 at Nogara”, in Nogara. Archeologia e storia di un villaggio medievale (scavi 2003–2008), ed. F. Saggioro (Rome 2011) 107–21; Cartledge J., Clark G. and Higgins V. (1992) “The animal bones: a preliminary assessment of the stock economy”, in Excavations at Otranto, vol. 2: The Finds (Galatina 1992) 317–35; De Cupere B. (2001) Animals at Ancient Sagalassos: Evidence of the Faunal Remains (Turnhout 2001); Ferro A. M. (1999) “La fauna”, in San Michele di Trino (VC): dal villaggio romano al castello medievale, ed. M. M. Negro Ponzi Mancini (1999); Fuller B. T. et al. (2012) “Isotopic reconstruction of human diet and animal husbandry practices during the Classical-Hellenistic, imperial, and Byzantine periods at Sagalassos, Turkey”, American Journal of Physical Anthropology 149 (2012) 157–71; Fumagalli V. (1985) “Gli animali e l’agricoltura”, in L’uomo di fronte al mondo animale nell’alto Medioevo, 7–13 Aprile 1983 (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985) 579–609; Gabba E. (1985) “La transumanza nell’Italia Romana”, in L’uomo di fronte al mondo animale nell’alto Medioevo, 7–13 Aprile 1983, (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985) 373–89; Kroll H. (2010) Tiere im Byzantinischen Reich. Archäozoologische Forschungen im Überblick (Mainz 2010); Lewit T. (2009) “Pigs, presses and pastoralism: farming in the fifth to sixth centuries AD”, Early Medieval Europe 17 (2009) 77–91; LópezMerino L. et al. (2009) “2000 years of pastoralism and fire shaping high-altitude vegetation of Sierra de Gredos in central Spain”, Review of Palaeobotany and Palynology 158 (2009) 42–51; Magness J. (2010) “Early Islamic pottery: a revolution in diet and dining habits?”, in Proceedings of the 6th International Congress of the Archaeology of the Ancient Near East, 5–10 May 2009, “Sapienza”, Universita Di Roma, edd. P. Matthiae, F. Pinnock, L. Nigro and N. Marchetti (Wiesbaden 2010) 129–42; Muthesius A. (1989) “From seed to Samite: aspects of Byzantine silk production”, Textile History 20 (1989) 135–49; Navarro T., Alados C. L. and Cabezudo B. (2006) “Changes in plant functional types in response to goat and sheep grazing in two semiarid shrublands of SE Spain”, Journal of Arid Environments 64 (2006) 298–322; Schneider W. C. (1985) “Animal Laborans: das Arbeitstier und sein Einsatz in Transport und Verkehr der Spätantike und des frühen Mittelalters”, in L’uomo di fronte al mondo animale nell’alto Medioevo, 7–13 Aprile 1983, (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985) 457–578; Tazima Y. (1984) “Silkworm moths”, in Evolution of Domesticated Animals, ed. I. Mason (London 1984) 416–24; Tchernov E. and Horwitz L. K. (1990) “Herd management in the past and its impact on the landscape of the southern Levant”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/ BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg, and W. van Zeist (Rotterdam 1990) 207–16; Thonemann P. (2011) The Maeander Valley: a Historical Geography from Antiquity to Byzantium (Cambridge 2011); Wickham C. (1985) “Pastoralism and underdevelopment in the early Middle Ages”, in L’uomo di fronte al mondo animale nell’alto medioevo, 7–13 aprile 1983 (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985) 400–55;
21
Wilkinson T. J. (1982) “The definition of ancient manured zones by means of extensive sherd-sampling techniques”, JFA 9 (1982) 323–33.
Supplying the Cities The domination of the Mediterranean basin by a few large urban areas resulted in the shaping of the environment in order to support their substantial populations. Teall (1959) and Sirks (1991) remain essential reading for understanding the scale of the needs of the major urban centres of Rome and Constantinople. Koder (1993) discusses the amount of land needed to supply Constantinople with vegetables, and argues that the capital could be provided for by its immediate hinterland. Koder (1995) is an English summary of his 1993 monograph. Erdkamp (2013) and Tuck (2013) give a sense of the difficulties in provisioning Rome, and some suggestion as to the burden major cities placed upon their hinterlands. Linn (2012) examines the grain supply of Rome in the mid 5th c. in the context of the disruption created by the Vandal conquest of Africa, and the city’s ability to supply itself from its hinterland. Cities required not only food and fuel, but also water; the final report of Crow et al. (2008) on the aqueduct system of Constantinople reveals the full extent of the network to the north of the city. While not primarily concerned with the environment, Kehoe (1988) provides information on the scale of development in North Africa in order to provision Rome, but should be read alongside the second and third chapters of Dossey (2010) for an understanding of the local context. The impact of major urban centres upon the environment has been studied for other periods, notably London (Galloway et al. (1996)), Istanbul (Murphey (1987)), and Venice (Appuhn 2009)). Also useful, from a comparative perspective, are the papers in Marin and Virlouvet (2003), and the paper of Hoffmann (2007) provides interesting ways of thinking about the effects of urban areas on the environment. Cities: Erdkamp P. (2013) “The food supply of the capital”, in The Cambridge Companion to Ancient Rome, ed. P. Erdkamp (Cambridge 2013) 262–77; Sirks B. (1991) “The size of the grain distribution in imperial Rome and Constantinople”, Athenaeum 79 (1991) 215–37; Teall J. L. (1959) “The grain supply of the Byzantine empire, 330– 1025”, DOP 13 (1959) 87–139; Linn J. (2012) “The Roman grain supply, 442–455”, Journal of Late Antiquity 5.2 (2012) 298–321; Barnish S. J. B. (1987) “Pigs, plebeians and potentes: Rome’s economic hinterland, c. 350–600 A.D.”, PBSR 55 (1987) 157–85; Barthel S. and Isendahl C. (2013) “Urban gardens, agriculture, and water management: sources of resilience for long-term food security in cities”, Ecological Economics 86 (2013) 224–34; Crow J., Bardill J., and Bayliss R. (2008) The Water Supply of Byzantine Constantinople (London 2008); Dossey L. (2010) Peasant and Empire in Christian North Africa (Los Angeles 2010); Durliat J. (1990) De la ville antique à la ville byzantine: le problème des subsistances (Rome 1990); Durliat J. (1995) “L’approvisionnement de Constantinople”, in Constantinople and its Hinterland: Papers from the Twenty-seventh Spring Symposium of Byzantine Studies, Oxford, April 1993, edd. C. Mango and G. Dagron
22 (Aldershot 1995) 19–33; Graham S. (2013) “Counting bricks and stacking wood: providing the physical fabric”, in The Cambridge Companion to Ancient Rome, ed. P. Erdkamp (Cambridge 2013) 278– 96; Green J. A. (1986) The Carthaginian Countryside: Archaeological Renaissance in the Hinterland of Ancient Carthage (Ph.D. diss., Univ. of Chicago 1986); Hoffmann R. (2007) “Footprint metaphor and metabolic realities: environmental impacts of medieval European cities”, in Natures Past: The Environment and Human History, ed. P. Squatriti (Ann Arbor 2007) 288–325; Izdebski A. (2016) “Byzantine Miletus: environmental history of the hinterland”, AA (2016/2) 270–80; Kehoe D. P. (1988) The Economics of Agriculture on Roman Imperial Estates in North Africa (Göttingen 1988); Koder J. (1993) Gemüse in Byzanz: die Versorgung Konstantinopels mit Frischgemüse im Lichte der Geoponika (Vienna 1993); Koder J. (1995) “Fresh vegetables for the capital”, in Constantinople and its Hinterland: Papers from the TwentySeventh Spring Symposium of Byzantine Studies, Oxford, April 1993, edd. C. Mango and G. Dagron (Aldershot 1995) 49–56; Porath J. (2002) “The water-supply of Caesarea: a reassessment”, in The Aqueducts of Israel, edd. D. Amit, Y. Hirschfeld and J. Patrich (Portsmouth, Rhode Island 2002) 104–29; Raban A. (1996) “The inner harbour basin of Caesarea: archaeological evidence for its gradual demise”, in Caesarea Maritima: a Retrospective after Two Millennia, edd. A. Raban and K. Holum (Leiden 1996) 628–66; Tuck S. L. (2013) “The Tiber and river transport”, in The Cambridge Companion to Ancient Rome, ed. P. Erdkamp (Cambridge 2013) 229–45; Galloway J. A., Keene D. and Murphy M. (1996) “Fuelling the city: production and distribution of firewood and fuel in London’s region, 1290–1400”, Economic History Review 49.3 (1996) 447–72; Appuhn K. (2009) A Forest on the Sea: Environmental Expertise in Renaissance Venice (Baltimore 2009); Murphey R. (1987) “Provisioning Istanbul: the state and subsistence in the early modern Middle East”, Food and Foodways 2 (1987) 217–63; Marin B. and Voulouvet C. (2003) edd. Nourrir les cités de Méditerranée: antiquité-temps modernes (Paris 2003).
Landscape History Landscape history is closely related to agricultural history and, like agricultural history, is essential to environmental history but not entirely subsumed into it. Landscape and agriculture are difficult to separate, because people often write about landscapes in terms of agriculture, that is fields, vineyards, terraces, and pastures. Many scholars seek to understand changing landscapes in order to see how agriculture shaped, or was shaped by, those changes. As in agricultural history, social questions, regarding especially land tenure and rights to pasture, but also attempts to extrapolate population trends from land use, have dominated the scholarship: Tsafrir (1995), Lefort et al. (2005) and Lefort (2006). Studies centred around these questions stray quickly into purely legal history, so they remain largely outside the scope of this essay. For example, Dossey (2010) addresses rural settlement patterns, but concentrates on social and political issues, rather than environmental ones. Nevertheless, there is a longstanding tradition of considering the impact of landscapes on human activities and settlement patterns, such as in Geyer (2002) and Harrison (2001). Relevant studies with this
McMahon and Sargent
focus—some leaning in an environmental direction and others remaining primarily social history—may be found in an earlier volume of this series: Bowden et al. (2004). Notably comprehensive works are those which investigate the ways in which humans and non-human elements worked in tandem to construct landscapes of spaces and boundaries, settlements, fields, and woodlands. These include Turner and Crow (2010), Varinlioğlu (2007) and Roberts (1990). Landscape Histories: General: Bowden W., Lavan L. and Machado C. (2004) Recent Research On The Late Antique Countryside (Late Antique Archaeology 2) (Leiden 2004); Lefort J. (2006) Société rurale et histoire du paysage à Byzance (Paris 2006); Lefort J., Morrisson C. and Sodini J.-P. (2005) Les villages dans l’empire byzantin: IVe–XVe siècle (Paris 2005); Behre K.-E. (1990) “Some reflections on anthropogenic indicators and the record of prehistoric occupation phases in pollen diagrams from the Near East”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. EntjesNieborg, and W. van Zeist (Rotterdam 1990) 219–30; Bottema S. and Woldring H. (1990) “Anthropogenic indicators in the pollen record of the eastern Mediterranean”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/ BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg, and W. van Zeist (Rotterdam 1990) 231–64; Commito A. R. (2014) Southwest Asia Minor and Northwest Syria at the End of Antiquity: a View from the Countryside (Ph.D. diss., Univ. of Chicago 2014); Dallman P. R. (1998) Plant Life in the World’s Mediterranean Climates: California, Chile, South Africa, Australia, and the Mediterranean Basin (Berkeley 1998); Geyer B. (2002) “Physical factors in the evolution of the landscape and land use”, in The Economic History of Byzantium: From the Seventh Through the Fifteenth Century, ed. A. Laiou (Washington, D.C. 2002) 115–20; Herrmann S. M. and Hutchinson C. F. (2005) “The changing contexts of the desertification debate”, Journal of Arid Environments 63 (2005) 538–55; Koder J. (1984) Der Lebensraum der Byzantiner. Historisch-geographischer Abriß ihres mittelalterlichen Staates im östlichen Mittelmeerraum (Graz-Vienna-Cologne 1984); Nicholson S. E. (2011) Dryland Climatology (Cambridge 2011). Anatolia: Eastwood W. J., Roberts N. and Lamb H. F. (1998) “Palaeoecological and archaeological evidence for human occupance in southwest Turkey: the Beyşehir Occupation Phase”, AnatSt 48 (1998) 69–86; Argant J. (2003) “Données palynologique”, in La Bithynie au moyen âge, edd. B. Geyer and J. Lefort (Paris 2003); Belke K., Koder J., Hild F. and Soustal P. (2000) Byzanz als Raum. Zu Methoden und Inhalten der historischen Geographie des östlichen Mittelmeerraumes (Vienna 2000); Bintliff J. (2000) “Reconstructing the Byzantine countryside: new approaches from landscape archaeology”, in Byzanz als Raum. Zu Methoden und Inhalten der historischen Geographie des östlichen Mittelmeerraumes, edd. K. Belke, J. Koder, F. Hild, and P. Soustal (Vienna 2000) 37–63; Bottema S. and Woldring H. (2003) “The vegetation history of east-central Anatolia in relation to archaeology: the Eski Acigöl pollen evidence compared with the near eastern environment”, Palaeohistoria 43 (2003) 1–34; Bottema S. et al. (1993) “Late Quaternary vegetation history of northern Turkey”, Palaeohistoria 35 (1993) 13–72; Decker M. and Cooper J. E. (2012) Life and Society in Byzantine Cappadocia (New York 2012); Harrison M. (2001) Mountain and Plain: From the Lycian Coast to the Phrygian Plateau in the Late Roman and Early Byzantine
Environmental History of the Late Antique Eastern Mediterranean Period (Ann Arbor 2001); Roberts N. (1990) “Human-induced landscape change in south and southwest Turkey during the later Holocene”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg and W. van Zeist (Rotterdam 1990) 53–67; Roberts N. (2009) “Contexts of human interaction: geology, geography, geomorphology and environment: paleolimnological investigations in Paphlagonia”, in At Empire’s Edge: Project Paphlagonia: Regional Survey in North-Central Turkey, edd. R. Matthews and C. Glatz (London 2009) 64–73; Varinlioğlu G. (2007) “Living in a marginal environment: rural habitat and landscape in southeastern Isauria”, DOP 61 (2007) 287–317; Vermoere M. (2004) Holocene Vegetation History in the Territory of Sagalassos (Southwest Turkey): a Palynological Approach (Turnhout 2004); Vermoere M. et al. (2001) “Modern pollen studies in the territory of Sagalassos (southwest Turkey) and their use in the interpretation of a Late Holocene pollen diagram,”, Review of Palaeobotany and Palynology 114 (2001) 29–56; Wilkinson T. J. and Algaze G. (1990) Town and Country in Southeastern Anatolia. (Chicago 1990). Other: Turner S. and Crow J. (2010) “Unlocking historic landscapes in the eastern Mediterranean: two pilot studies using historic landscape characterisation”, Antiquity 84 (2010) 216–29; Allevato E. et al. (2012) “Persistence of the cultural landscape in Campania (southern Italy) before the AD 472 Vesuvius eruption: archaeoenvironmental data”, JAS 39 (2012) 399–406; Armstrong P. (1996) “The survey area in the Byzantine and Ottoman periods”, in The Laconia Survey: Continuity and Change in a Greek Rural Landscape, edd. W. G. Cavanagh, J. Crouwel, R. W. V. Catling and G. Shipley (London 1996) 339–402; Athanasiadis N. et al. (2000) “Palynological study of Holocene sediments from Lake Doirani in northern Greece”, Journal of Paleolimnology 24 (2000) 331–42; Bintliff J. (2013) “The contribution of regional surface survey to Byzantine landscape history in Greece”, in Exempli gratia: Sagalassos, Marc Waelkens and Interdisciplinary Archaeology (Leuven 2013) 127–39; Bottema S. (2016) “Pollen analytical investigations in Thessaly (Greece)”, Palaeohistoria 21 (2016) 19–40; Brogiolo G. P. (1999) “Nuove ricerche sulla campagna dell’Italia settentrionale tra tarda antichità e altomedioevo”, in Archéologie des espaces agraires méditerranéens au moyen âge, ed. A. Bazzana (Madrid 1999) 153–66; Cooper J. P. (2014) The Medieval Nile: Route, Navigation, and Landscape in Islamic Egypt (New York 2014); Dossey L. (2010) Peasant and Empire in Christian North Africa (Berkeley 2010) Dunn A. (2004) “Continuity and change in the Macedonian countryside from Gallienus to Justinian”, in Recent Research on the Late Antique Countryside, edd. W. Bowden, L. Lavan and C. Machado (Late Antique Archaeology 2) (Leiden 2004) 535–86; Hoffman E. S. (1981) “Plant remains from Vandal and Byzantine deposits”, in Excavations at Carthage 1977, Conducted by the University of Michigan, vol. 6, ed. J. Humphrey (Ann Arbor 1981) 259–68; MacAdam H. I. (1994) “Settlements and settlement patterns in northern and central Transjordania, ca. 550 - ca. 750”, in The Byzantine and Early Islamic Near East II: Land Use and Settlement Patterns. Papers of the Second Workshop on Late Antiquity and Early Islam (London 25th– 27th April 1991), edd. G. R. D. King and A. Cameron (Princeton 1994) 49–93; Rackham O. (1996) “Observations on the historical ecology of Laconia”, in The Laconia Survey: Continuity and Change in a Greek Rural Landscape, edd. W. G. Cavanagh, J. Crouwel, R. W. V. Catling and G. Shipley (London 1996) 73–119; Tsafrir Y. (1995) “Some notes on the settlement and demography of Palestine in the Byzantine period: the archaeological evidence”, in Retrieving the Past: Essays on Archaeological Research and Methodology in Honor of Gus W. Van Beek (Winona Lake, Indiana 1995) 269–83.
23
More recent research, influenced in part by the environmental history of other regions and periods, has focused instead on the ways human actions have changed elements of the landscape, like soils, water tables, and ‘wild’ plant communities: Kouli (2012), Gerasimidis (2000) and Knipping et al. (2008). Such elements had previously been accepted as products of an autonomous ‘nature’ that was not shaped by human actions. Several scho lars have focused on changes in soils: Van Berghem and Fiselier (1996), Winiwarter (2006), Goldberg and BarYosef (1990), Bruins (1990), and Geyer (1999). Eastwood et al. (1999) uses natural archives to trace changes in plant communities brought about by human and nonhuman activity, and is an example of how broad these studies can be, providing information about agriculture, landscape, and climate. Yet many studies have ignored human activity, limiting themselves to reconstructing the plant communities of a given region, without any consideration of human impacts: Panagiotopoulos (2013). Landscape Change: General: Bakker J. et al. (2013) “Climate, people, fire and vegetation: new insights into vegetation dynamics in the eastern Mediterranean since the 1st century AD”, Climate of the Past 9 (2013) 57–87; Cheyette F. L. (2008) “The disappearance of the ancient landscape and the climatic anomaly of the early Middle Ages: a question to be pursued”, Early Medieval Europe 16 (2008) 127–65; Christie N. (2004) “Landscapes of change in Late Antiquity and the early Middle Ages: themes, directions, and problems”, in Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages, ed. N. Christie (Burlington, Vermont 2004) 1–37; Christie N. (2004) Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages (Burlington, Vermont 2004); Dusar B. et al. (2011) “Holocene environmental change and its impact on sediment dynamics in the eastern Mediterranean”, Earth-Science Reviews 108 (2011) 137–57; Laiou A. E. (2005) “The Byzantine village (5th–14th century)”, in Les villages dans l’empire byzantin: IVe–XVe siècle, edd. J. Lefort, C. Morrisson and J.-P. Sodini (Paris 2005) 31–54; Wagstaff J. M. (1985) The Evolution of Middle Eastern Landscapes: an Outline to AD 1840 (London 1985); Winiwarter V. (2006) “Prolegomena to a history of soil knowledge in Europe”, in Soils and Societies: Perspectives from Environmental History, edd. J. R. McNeill, and V. Winiwarter (Knapwell 2006) 177–215; Van Berghem J.-W. and Fiselier J. (1996) “Soils and land use potential”, in The Laconia Survey: Continuity and Change in a Greek Rural Landscape, edd. W. G. Cavanagh, J. Crouwel, R. W. V. Catling and G. Shipley (London 1996) 57–71. Anatolia: England A. et al. (2008) “Historical landscape change in Cappadocia (central Turkey): a palaeoecological investigation of annually laminated sediments from Nar lake”, The Holocene 18 (2008) 1229–45; Knipping M., Müllenhoff M. and Brückner H. (2008) “Human induced landscape changes around Bafa Gölü (western Turkey)”, Vegetation History and Archaeobotany 17 (2008) 365–80; Bakker J. et al. (2012) “Man, vegetation and climate during the Holocene in the territory of Sagalassos, western Taurus Mountains, SW Turkey”, Vegetation History and Archaeobotany 21 (2012) 249–66; Bottema S., Woldring H. and Aytug B. (1986) “Palynological investigations on the relations between prehistoric man and vegetation in Turkey: the Beyşehir Occupation Phase”, in OPTIMA: Proceedings of the Fifth Meeting at Istanbul, 8–15 September 1986, edd. H. Demiriz and N. Özhatay (Istanbul 1986) 315–28; Brückner H. (2003) “Delta evolution and culture—aspects of geoarchaeological research in Miletos
24 and Priene”, in Troia and the Troad: Scientific Approaches, edd. G. A. Wagner, E. Pernicka, and H.-P. Uerpmann (Berlin 2003) 121–42; Cavanagh W. G. et al. (1996) The Laconia Survey: Continuity and Change in a Greek Rural Landscape (London 1996); Eastwood W. J. et al. (1999) “Holocene environmental change in southwest Turkey: a palaeoecological record of lake and catchment-related changes”, Quaternary Science Reviews 18 (1999) 671–95; Gerritsen F. et al. (2008) “Settlement and landscape transformations in the Amuq Valley, Hatay”, Anatolica 34 (2008) 241–314; Geyer B. (1999) “Erosion et histoire du paysage en Bithynie médiévale”, Castrum 5 (1999) 241–54; Haldon J. et al. (2009) “Integrating paleoecological and archaeo-historical records: land use and landscape change”, in Archaeology of the Countryside in Medieval Anatolia, edd. T. Vorderstrasse and J. Roodenberg (Leiden 2009) 45–70; Izdebski A. (2012) “The changing landscapes of Byzantine northern Anatolia”, Archaeologia Bulgarica 16 (2012) 47–66; Izdebski A. (2013) “The economic expansion of the Anatolian countryside in Late Antiquity: the coast versus inland regions,” in Local Economies? Production and Exchange of Inland Regions in Late Antiquity (Late Antique Archaeology 10) (Leiden-Boston 2013) 343–76; Kaniewski D. et al. (2007) “A high-resolution Late Holocene landscape ecological history inferred from an intramontane basin in the western Taurus Mountains, Turkey”, Quaternary Science Reviews 26 (2007) 2201–18; Roberts N. et al. (2005) “Late Holocene landscape disturbance and recovery: an east Mediterranean case study”, Conference Poster at PAGES: Second Open Science Meeting 10–12 August 2005 Beijing, China (2005): http://www.pages-igbp.org/download/docs/meetingproducts/posters/2005-osm2/Roberts-N.pdf (last accessed April 2017); Rose BC (2011) “Troy and the Granicus River valley in Late Antiquity”, in Archaeology and the Cities of Asia Minor in Late Antiquity, edd. O. Dally and C. J. Ratté (Ann Arbor 2011) 151–72; Vanhaverbeke H. et al. (2009) “What happened after the 7th century AD? A different perspective on Post-Roman rural Anatolia”, in Archaeology of the Countryside in Medieval Anatolia, edd. T. Vorderstrasse and J. Roodenberg (Leiden 2009) 177–90; Vermoere M. et al. (2002) “Late Holocene local vegetation dynamics in the marsh of Gravgaz (southwest Turkey)”, Journal of Paleolimnology 27 (2002) 429–51. Other: Arthur P. (2004) “From vicus to village: Italian landscape 400–1000”, in Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages, ed. N. Christie (Burlington, Vermont 2004) 103–33; Gerasimidis A. (2000) “Palynological evidence for human influence on the vegetation of mountain regions in northern Greece: the case of Lailias, Serres” in Landscape and Land Use in Postglacial Greece, edd. P. Halstead and C. Frederick (Sheffield 2000); Kouli K. (2012) “Vegetation development and human activities in Attiki (SE Greece) during the last 5,000 years”, Vegetation History and Archaeobotany 21 (2012) 267–78; Baruch U. (1990) “Palynological evidence of human impact on the vegetation as recorded in Late Holocene lake sediments in Israel”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. EntjesNieborg, and W. van Zeist (Rotterdam 1990) 283–93; Bottema S. (2015) “Palynological investigations in Greece with special reference to pollen as an indicator of human activity”, Palaeohistoria 24 (2015) 257–88; Bruins H. J. (1990) “The impact of man and climate on the central Negev and northeastern Sinai deserts during the Late Holocene”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg, and W. van Zeist (Rotterdam 1990) 87–99; Casana J. (2007) “Structural transformations in settlement systems of the northern Levant”, AJA 111 (2007) 195–221; Dunn A. (2004) “Continuity and change in the Macedonian countryside from Gallienus to Justinian”, in Recent Research on the Late
McMahon and Sargent Antique Countryside, edd. W. Bowden, L. Lavan, and C. Machado (Late Antique Archaeology 2) (Leiden 2004) 535–86; Goldberg P. and Bar-Yosef O. (1990) “The effect of man on geomorphological processes based upon evidence from the Levant and adjacent areas”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg and W. van Zeist (Rotterdam 1990) 71–86; de Klerk P. et al. (2009) “Vegetation history and environmental development since ca 6000 cal yr BP in and around Ispani 2 (Kolkheti lowlands, Georgia)”, Quaternary Science Reviews 28 (2009) 890–910; Leone A. and Mattingly D. J. (2004) “Vandal, Byzantine, and Arab rural landscapes in North Africa”, in Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages, ed. N. Christie (Burlington, Vermont 2004) 135–62; Lucke B. et al. (2005) “The abandonment of the Decapolis region in northern Jordan—forced by environmental change?”, Quaternary International 135 (2005) 65–81; Panagiotopoulos K. et al. (2013) “Vegetation and climate history of the Lake Prespa region since the Lateglacial”, Quaternary International 293 (2013) 157–69; Stone D. (1997) The Development of an Imperial Territory: Romans, Africans, and the Transformation of The Rural Landscape of Tunisia (Ph.D. diss., Univ. of Michigan 1997); Tinner W. et al. (2003) “Climatic change and contemporaneous land-use phases north and south of the Alps 2300 BC to 800 AD”, Quaternary Science Reviews 22 (2003) 1447–60.
Trees and Woodland Scholars have for a long time recognised the difference between very obviously cultivated trees (usually olive) and woodlands that might have been heavily managed, but were not planted and tended in the same way. Interest in Mediterranean woodlands owed its early popularity to a strand of scholarship that assumed that the natural state of the landscape was dense forest cover, a state only disrupted by human activity, as seen in Thirgood (1981) and Harrison (1992). More recently such an assumption has been seriously questioned, and mostly discarded: Grove and Rackham (2003), Davis (2007), and Rackham (2002). Scholarship now tends to focus on the interaction between humans and woodlands over time, especially through woodland management and the use of its products, as seen in Dunn (1992), Andreolli (2002) and Allevato et al. (2009). Though his study of Renaissance Venice, Appuhn (2009) has been influential in this shift of scholarship towards the Mediterranean. Squatriti’s work on chestnut trees in Italy (Squatriti (2010) and (2013)) demonstrates the possibilities for this type of research. Trees and Woodland: Grove A. T. and Rackham O. (2003) The Nature of Mediterranean Europe: an Ecological History (New Haven, Connecticut 2003); Harris W.V. (2013) “Defining and detecting Mediterranean deforestation, 800 BCE to 700 CE”, in The Ancient Mediterranean Between Science and History, ed. W. V. Harris (Leiden 2013) 173–94; Squatriti P. (2013) Landscape and Change in Early Medieval Italy: Chestnuts, Economy, and Culture (Cambridge 2013); Appuhn K. R. (2009) A Forest on the Sea: Environmental Expertise in Renaissance Venice (Baltimore 2009); Accorsi C. et al. (1996) “Holocene forest pollen vegetation of the Po plain: northern Italy (Emilia Romagna data)”, Alliona 24 (1996) 233–76; Akkemik Ü. and
Environmental History of the Late Antique Eastern Mediterranean Kocabaş U. (2014) “Woods of Byzantine trade ships of Yenikapı (Istanbul) and changes in wood use from the 6th to 11th century”, Mediterranean Archaeology and Archaeometry 14 (2014) 301–11; Allevato E. et al. (2012) “Persistence of the cultural landscape in Campania (southern Italy) before the AD 472 Vesuvius eruption: archaeoenvironmental data”, JAS 39 (2012) 399–406; Allevato E., Ermolli E. R. and di Pasquale G. (2009) “Woodland exploitation and Roman shipbuilding: preliminary data from the shipwreck Napoli C (Naples, Italy)”, Méditerranée [Online] 112 (2009): doi:10.4000/mediterranee.3095 (last accessed April 2017); Andreolli B. (2002) “L’uso del bosco e degli incolti”, in Storia dell’agricoltura Italiana, edd. G. Pinto and M. Ambrosoli (Florence 2002) 123–44; Bargioni E. and Zanzi Sulli A. (1998) “The production of fodder trees in Valdagno, Vicenza, Italy”, in The Ecological History of European Forests, edd. C. Watkins and K. Kirby (Wallingford 1998) 43–52; Bevilacqua P. (1992) Terre del grano, terre degli alberi. L’ambiente nella storia del Mezzogiorno (Rionero in Vulture 1992); Bottema S. (2000) “The Holocene history of walnut, sweetchestnut, manna-ash and plane tree in the eastern Mediterranean”, Pallas (2000) 35–59; Davis D. K. (2007) Resurrecting the Granary of Rome: Environmental History and French Colonial Expansion in North Africa (Athens 2007); Dunn A. (1992) “The exploitation and control of woodland and scrubland in the Byzantine world”, Byzantine and Modern Greek Studies 16 (1992) 235–98; Fall P. L. (1990) “Deforestation in southern Jordan: evidence from fossil hyrax middens”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape: Proceedings of the INQUA/BAI Symposium on the Impact of Ancient Man on the Landscape of the Eastern Mediterranean Region and the Near East: Groningen, Netherlands, 6–9 March 1989, edd. S. Bottema, G. Entjes-Nieborg and W. van Zeist (Rotterdam 1990) 271– 81; Fonti P. et al. (2006) “Tree rings show competition dynamics in abandoned Castanea sativa coppices after land-use changes”, Journal of Vegetation Science 17 (2006) 103–12; Harrison R. P. (1992) Forests: the Shadow of Civilization (Chicago 1992); Hemphill P. (1988) “Deforestation and re-forestation in a central Italian hinterland: land usage during and after the Roman occupation”, in First Millennium Papers. Western Europe in the First Millennium AD, edd. R. F. J. Jones, J. H. Bloemers, S. L. Dyson and M. Biddle (Oxford 1988) 147–58; Hughes J. D. and Thurgood J. V. (1982) “Deforestation in ancient Greece and Rome: a cause of collapse”, Ecologist 12 (1982) 196–208; Redon O. (2008) Des forêts et des âmes: espace et société dans la Toscane médiévale. Études rassemblées par Laurence Moulinier-Brogi. (Saint-Denis 2008); Squatriti P. (2010) “Trees, nuts, and woods at the end of the first millennium: a case from the Amalfi coast”, in Ecologies and Economies in Medieval and Early Modern Europe: Studies in Environmental History for Richard C. Hoffmann (Leiden 2010) 22–43; Thirgood J. V. (1981) Man and the Mediterranean Forest: a History of Resource Depletion (London 1981); Wickham C. (1990) “European forests in the early Middle Ages: landscape and land clearance”, in L’ambiente vegetale nell’alto medioevo, Marzo-5 Aprile 1989 (Settimane di studio del Centro italiano di studi sull’alto medioevo 30) (Spoleto 1990) 479–548.
Uncultivated Resources The study of fishing has not resonated with scholars of Late Antiquity to the extent that it has for scholars of the Medieval West, and the result is a rather limited bibliography. One explanation for this is the preoccupation in late antique studies with cultural and religious history, a point exemplified by Rousselle (1999). Yet, there is evidence that points to intensive exploitation of fish resources in Late Antiquity (Luff and Bailey (2000) and Stieglitz (1998)). The 10th c. Book of the Eparch provides a window into fishing regulations around Constantinople
25
(Dagron (1995) and Maniatis (2010)) but the applicability of this research for the late antique city is unclear. Interest in mining is long-standing, but little work has been done that is primarily concerned with Late Antiquity and the environment. The exception to this is the excavation done at the mining town of Bir Umm Fawakhir in Egypt (see Meyer (1993), (2003), and (2011)), where the effect of mining on the local environment during this period is taken into consideration. Roman mining is also visible in climate indicators, such as ice cores and speleothems, as a result of atmospheric pollution during the smelting of the metals. Fishing and Mining: Dagron G. (1995) “Poissons, pêcheurs et poissonniers de Constantinople”, in Constantinople and its Hinterland: Papers from the Twenty-seventh Spring Symposium of Byzantine Studies, Oxford, April 1993, edd. G. Dagron, and C. Mango (Aldershot 1995) 57–73; Healy J. F. (1978) Mining and Metallurgy in the Greek and Roman World (London 1978); Afanasyeva T. and Ivanov S. A. (2013) “Unexpected evidence concerning gold mining in early Byzantium”, GRBS 53 (2013) 138–44; Bryer A. A. M. (1982) “The question of Byzantine mines in the Pontos: Chalybian iron, Chaldian silver, Koloneian alum and the mummy of Cheriana”, AnatSt 32 (1982) 133–50; Delrieux F. (2008) “L’exploitation de la mer et ses implications économiques, politiques, et militaries dans le golfe de Bargylia en Carie à l’epoque gréco-romaine”, in Ressources et activités maritimes des peuples de l’antiquité, ed. J. Napoli (Boulogne-surmer 2008) 273–93; Edmonson J. C. (1989) “Mining in the late Roman empire and beyond: continuity or disruption?”, JRS 79 (1989) 84–102; Hunt C. O., Gilberston D. D. and El-Rishi H. A. (2007) “An 8000-year history of landscape, climate, and copper exploitation in the Middle East: the Wadi Faynan and the Wadi Dana National Reserve in southern Jordan”, JAS 34.8 (2007) 1306–38; Irabien M. J. et al. (2012) “Historical signature of Roman mining activities in the Bidasoa estuary (Basque Country, northern Spain): an integrated micropalaeontological, geochemical and archaeological approach”, JAS 39 (2012) 2361–70; Luff R. M., and Bailey G. N. (2000) “Analysis of size changes and incremental growth structures in African catfish Synodontis schall (schall) from Tell el-Amarna, Middle Egypt”, JAS 27 (2000) 821– 35; Maniatis G. C. (2000) “The organizational setup and functioning of the fish market in tenth-century Constantinople”, DOP 54 (2000) 13–42; Martínez Cortizas A. et al. (2013) “Atmospheric Pb pollution in N Iberia during the Late Iron Age/Roman times reconstructed using the high-resolution record of La Molina mire (Asturias, Spain)”, Journal of Paleolimnology 50 (2013) 71–86; Matschke K.-P. (2002) “Mining”, in The Economic History of Byzantium: From the Seventh Through the Fifteenth Century, ed. A. Laiou (Washington, D.C. 2002) 115–20; McFarlane D. A. et al. (2014) “A speleothem record of early British and Roman mining at Charterhouse, Mendip, England”, Archaeometry 56 (2014) 431–43; Meyer C. (1998) “Gold miners and mining at Bir Umm Fawakhir”, in Social Approaches to an Industrial Past: the Archaeology and Anthropology of Mining, edd. A. B. Knapp, V. C. Pigott and E. W. Herbert (London 1998) 259–75; Meyer C. et al. (2011) Bir Umm Fawakhir, vol. 2: Report on the 1996–1997 Survey Seasons (Chicago 2011); Meyer C. et al. (2003) “Ancient gold extraction at Bir Umm Fawakhir”, JARCE 40 (2003) 13–53; van Neer W. and Wouters W. (2007) “Salted fish products from the Coptic monastery at Bawit, Egypt: evidence from the bones and texts”, in The Role of Fish in Ancient Time: Proceedings of the 13th Meeting of the ICAZ Fish Remains Working Group in October 4th–9th, Basel/Augst 2005, ed. H. Hüster-Plogmann (Rahden 2007) 147–62; Pyatt F. B. (2000) “An imperial legacy? An exploration of the environmental impact of ancient metal mining and smelting in southern Jordan”, Journal of
26 British Studies 39 (2000) 771–78; Ramin J. (1977) La technique minière et métallurgique des anciens (Brussels 1977); Rousselle A. (1999) “Fish”, in Late Antiquity: a Guide to the Postclassical World, edd. G. Bowersock, P. Brown and O. Grabar (Cambridge, Mass. 1999) 450–51; Russell R. J. (1954) “Alluvial morphology of Anatolian rivers”, Annals of the Association of American Geographers 44 (1954) 363–91; Shaw I., Bunbury J. and Jameson R. (1999) “Emerald mining in Roman and Byzantine Egypt”, JRA 12 (1999) 203–15; Stieglitz R. R. (1998) “A late Byzantine reservoir and piscina at Tel Tanninim”, IEJ 48 (1998) 54–65; Vryonis S. (1962) “The question of the Byzantine mines”, Speculum 37.1 (1962) 1–17; Zug Tucci H. (1985) “Il mondo medievale dei pesci tra realtà e immaginazione”, in L’uomo di fronte al mondo animale nell’alto medioevo, (Settimane di studio del Centro italiano di studi sull’alto medioevo 31) (Spoleto 1985) 291–360.
Plague and Disease A significant bibliography has been produced concerning the so-called ‘Justinianic plague’. Of particular importance are Stathakopoulos (2004), Harper (2017), and the work of the contributors in Little (2007). See especially Stathakopoulos (2000) for a summary of the bibliography and historiography up to that point. Hays (2007) and Sallares (2007) engage with the historiography on plague studies. The debate on the role of climate is ongoing (see Horden (2005) and Little (2011)) but few studies deal with the topic directly. The epidemiology of the plague has been of interest, with Drancourt et al. (2007) and Harbeck et al. (2013) doing foundational work on identifying Yersinia pestis as the causative factor, with McCormick (2007) providing an historical perspective. The transmission of the disease remains to be studied, however. Historians have been quite negative about what archaeological evidence can provide to help understand the plague: see for example Kulikowski (2007) and Kennedy (2007). As with scholarship on the 14th c. plague, some questions have been raised about the role of rats, for example in McCormick (2003), Stathakopoulos (2011) and Harper (2017). For the connection between climate variability and disease, see Harper (2017). The area of greatest interest to scholars is that of plague reception and adaptation. This line of interest tends to downplay what can be understood about the epidemiology of the disease, in favour of understanding how people responded to it, as seen in: Horden (2005), Meier (2016), Morony (2007), Stoclet (2006), Dooley (2007), Stathakopoulos (2007), and Dols (1971). Recent studies using genetic data are particularly important for the plague in the sixth century. Yersinia pestis DNA was found in graves in Aschheim Bavaria: Wiechmann and Grupe (2005). Subsequent work has confirmed the role of Yersinia pestis, notably Harbeck (2013), Wagner (2014), and Feldman (2016). Further
McMahon and Sargent
connections have also been drawn between climate change and plague (Xu 2015). The effects of the plague outbreak in the 6th c. remain a matter of debate, although the view that it caused serious economic and demographic upset is common: Meier (2016), Sarris (2002), Stathakopoulos (2012), Harrison (1993), and Zuckerman (2016). In contrast to earlier scholarship, the plague is no longer thought to have had major repercussions for the strength of the army in the Roman East (Teall (1965), cf. Whitby (1995)), although Woods (2007) makes a case for plague directly causing the Roman defeat by the Arabs at Yarmouk. Regional studies have up to this point been limited, although work has been done on opposite ends of the Mediterranean, namely Syria and Spain: Morony (2007) and Kulikowski (2007) respectively. Plague: Sarris P. (2002) “The Justinianic plague: origins and effects”, Continuity and Change 17.2 (2002) 169–82; Stathakopoulos D. C. (2004) Famine and Pestilence in the Late Roman and Early Byzantine Empire: a Systematic Survey of Subsistence Crises and Epidemics (Aldershot 2004); Little L. K. (2007) ed. Plague and the End of Antiquity: the Pandemic of 541–750 (Cambridge 2007); Horden P. (2005) “Mediterranean plague in the age of Justinian”, in The Cambridge Companion to the Age of Justinian, ed. M. Maas (Cambridge 2005) 134–60; Dols M. W. (1971) “Plague in early Islamic history”, JAOS 94.3 (1971) 371–83; Hays J. (2007) “Historians and epidemics: simple questions, complex answers”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 33–56; Kennedy H. (2007) “Justinianic plague in Syria and the archaeological evidence”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 87–95; Kulikowski M. (2007) “Plague in Spanish Late Antiquity”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 150–70; Little L. (2011) “Plague historians in lab coats”, PastPres 213 (November 2011) 267–90; McCormick M. (2003) “Rats, communications, and plague: toward an ecological history”, The Journal of Interdisciplinary History 34.1 (2003) 1–25; Morony M. (2007) “‘For whom does the writer write?’: the first bubonic plague pandemic according to Syriac sources”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 59–86; Stathakopoulos D. (2000) “The Justinianic plague revisited”, Byzantine and Modern Greek Studies 24 (2000), 256–76; Stathakopoulos D. (2011) “Invisible protagonists: the Justinianic plague from a zoocentric point of view”, in Animals and Environment in Byzantium (7th–12th c.), edd. I. Anagnostakis, T. G. Kolias and E. Papadopoulou (Athens 2011) 87–95; Woods D. (2007) “Jews, rats, and the battle of Yarmūk”, in The Late Roman Army in the Near East from Diocletian to the Arab Conquest: Proceedings of a Colloquium Held at Potenza, Acerenza and Matera, Italy (May 2005), edd. A. Lewin, and P. Pellegrini (Oxford 2007) 367–76. Social and economic effects: Harper K. (2017) The Fate of Rome: Climate, Disease, and the End of an Empire (Princeton 2017); Stoclet A. (2007) “Consilia humana, ops divina, superstitio: seeking succor and solace in times of plague, with particular reference to Gaul in the early Middle Ages”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 135–49; Harrison D. (1993) “Plague, settlement and structural change at the dawn of the Middle Ages”, Scandia 59.1 (1993) 15–48; Meier M. (2016) “The ‘Justinianic plague’: the economic consequences of the pandemic in the eastern Roman empire and its cultural and religious effects”, Early Medieval Europe 24.3 (2016) 267–92; Stathakopoulos D. (2007) “Making
Environmental History of the Late Antique Eastern Mediterranean use of the plague: readings in sixth century history”, in Byzantina Mediterranea: Festschrift für Johannes Koder zum 65. Geburtstag (Vienna 2007) 633–39; Teall J. L. (1965) “The barbarians in Justinian’s armies”, Speculum 40 (1965) 294–322; Whitby M. (1995) “Recruitment in Roman armies from Justinian to Heraclius (ca. 565–615)”, in The Byzantine and Islamic Near East III: States, Resources, and Armies, ed. A. Cameron (Princeton 1995) 61–124. Yersinia pestis: Feldman M. et al. (2016) “A high-coverage yersinia pestis genome from a sixth-century Justinianic plague victim”, Molecular Biology and Evolution 33.11 (2016) 2911–23; Xu L. et al. (2015) “The trophic responses of two different rodent–vector–plague systems to climate change”, Proceedings of the Royal Society of London B: Biological Sciences 282 (2015): doi: 10.1098/rspb.2014.1846 (last accessed April 2017); Wiechmann I. and Grupe G. (2005) “Detection of Yersinia pestis DNA in two early medieval skeletal finds from Aschheim (Upper Bavaria, 6th century AD)”, American Journal of Physical Anthropology 126 (2005) 48–55; Wagner D. M. et al. (2014) “Yersinia pestis and the plague of Justinian 541–543 AD: a genomic analysis”, Lancet Infectious Diseases 14 (2014) 319–26; Drancourt M. et al. (2007) “Yersinia pestis orientalis in remains of ancient plague patients”, Emerging Infectious Diseases 13.2 (February 2007) 332–33; Harbeck M. et al. (2013) “Yersinia pestis DNA from skeletal remains from the 6th century AD reveals insights into Justinianic plague”, PLOS Pathogens 9.5 (May 2013) 1–8; McCormick M. (2007) “Toward a molecular history of the Justinianic pandemic”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 290–312; Sallares R. (2007) “Ecology, evolution, and the epidemiology of plague”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 231–89.
Other Diseases The plague has dominated the discussion of disease in the Mediterranean basin. Current research on other diseases tends to be focused on northern Europe and pushes the boundaries of what constitutes Late Antiquity. Malaria (Newfield (2017), Gowland and Wester (2012), Knottnerus (2002), and Sallares and Gomzi (2001)) and leprosy (Bolsden (2008), Rubini and Zaio (2009)) have attracted most of the attention. The work of Newfield has provided some important insights on both animal and human diseases. Disease: Scheidel W. (1994) “Libitina’s bitter gains: seasonal mortality and endemic disease in the ancient city of Rome”, Ancient Society 25 (1994) 151–75; Newfield T. (2017) “Malaria and malaria-like disease in the early middle ages”, Early Medieval Europe 25.3 (2017) 251–300; Boldsden J. L. (2008) “Leprosy in the early Medieval Lauchheim community”, American Journal of Physical Anthropology 135 (2008) 301–10; Stathakopoulos D. (2012) “Death in the countryside: some thoughts on the effects of famine and epidemics”, AnTard 20 (2012) 105–14; Gowland R. L. and Wester A. G. (2012) “Morbidity in the marshes: using spatial epidemiology to investigate skeletal evidence for malaria in Anglo-Saxon England (AD 410–1050)”, American Journal of Physical Anthropology 147 (2012) 301–11; Knottnerus O. S. (2002) “Malaria around the North Sea: a survey”, in Climate Development and History of the North Atlantic Realm, edd. G. Wefer, W. Berger, K.-E. Behre, and E. Jansen (Berlin 2002) 339–53; Rubini M. and Zaio P. (2009) “Lepromatous leprosy in an early mediaeval cemetery in central Italy (Morrione, Campochiaro, Molise, 6th–8th century AD)”, JAS 36 (2009) 2771–79; Sallares R. and Gomzi S. (2001) “Biomolecular archaeology of malaria”, Ancient Biomolecules 3 (2001) 195–213; Sallares R. (2002) Malaria and Rome: a History of Malaria in Ancient Italy (Oxford 2002).
27
Animal disease: Newfield T. (2012) “A great Carolingian panzootic: the probable extent, diagnosis and impact of an early ninth-century cattle pestilence”, Argos 46 (2012) 200–10; Gillmor C. (2005) “The 791 equine pestilence and its impact on Charlemagne’s army”, Journal of Medieval Military History 3 (2005) 23–45; Newfield T. (2013) “Early Medieval epizootics and landscapes of disease: the origins and triggers of European livestock pestilences, 400–1000 CE”, in Landscapes and Societies in Medieval Europe East of the Elbe: Interactions Between Environmental Settings and Cultural Transformations, edd. S. Kleingärtner, T. Newfield, S. Rossignol, and D. Wehner (Toronto 2013) 73–113; Newfield T. (2015) “Human—bovine plagues in the early Middle Ages”, Journal of Interdisciplinary History 46.1 (2015) 1–38.
Climate Climate history is in many ways the new star of environmental history. The idea that climatic change, with variations in temperature, rainfall, and storm patterns, were important to past societies is not new in itself. But, in the absence of recently-developed scientific techniques, such inquiries have been carried out based on documentary records, as in Le Roy Ladurie (1971). In the past two decades, however, the study of past climates has been invigorated, largely because of the scientific community’s interest in the topic (see Behringer (2009)). It has been triggered, and given urgency, partly by its potential to illuminate climate change in the present (Guiot et al. (2012)), but the techniques it has developed and the climate histories it has traced (even for predictive purposes) are resources too rich for environmental historians to ignore. Incorporating that research can be challenging. A great deal of scientific work on past climates operates on scales that are hardly useful to the late antique historian or archaeologist. They speak in global terms or trace cycles across millennia; for instance, Finné et al. (2011) covers the eastern Mediterranean over the last 6000 years. Because of the nature of the evidence, and the significance of small changes over long periods, even studies that are more oriented to a human timeframe usually consider at least several centuries, such as Jones et al. (2006). Such studies can be extremely useful, but more immediately practical are those which focus closely on individual sites or regions, or which address a few centuries at most, as in Neumann (1991), Akcer (2011) and Pearson et al. (2012). Where proxy data allow for extremely precise dating, the results can be particularly useful as well, seen in England et al. (2008), Haldon (2007) and Kaniewski et al. (2007). Most of these studies seek to establish the rate and direction of change over time in various climatic elements: usually precipitation and temperature. A number of scholars have begun to combine these scientifically-derived data with information from archaeological and historical sources, to determine human responses to changing climates, as in Bulliet (2009), Haldon et al. (2014), Issar and Zohar
28 (2004) and Sadori et al. (2016). Many scholars have been particularly interested in identifying the climate of the Early Byzantine period, and hypothesising its role in the ‘end of antiquity’, such as Koder (1996), and Izdebski (2011). Recently, scholars have identified a period of cooling and climatic instability in the 6th and early 7th c., naming it the “Late Antique Little Ice Age”: Büntgen et al. (2016). Debate about the extent, uniformity, and consequences of this climatic change are set to continue for many years, with McCormick et al. (2012) and Manning (2013) being recent examples of such a discussion. The search for the explanation of the AD 536 dust veil is a particularly active aspect of late antique climate analysis, apparent in Gräslund and Price (2012), Larsen et al. (2008), and Ferris et al. (2011). Most scholars have agreed that it was caused by a volcanic eruption, but Farquharson (1996) contends that it was a cloud of water vapour created by an asteroid falling into the ocean. Climate histories of late antique eastern Mediterranean: General: Büntgen U. et al. (2016) “Cooling and societal change during the Late Antique Little Ice Age from 536 to around 660 AD”, Nature Geoscience 9 (2016) 231–36; Izdebski A. (2011) “Why did agriculture flourish in the late antique east? The role of climate fluctuations in the development and contraction of agriculture in Asia Minor and the Middle East from the 4th till the 7th c. AD”, Millennium 8 (2011) 291–312; Issar A. S. and Zohar M. (2004) Climate Change: Environment and Civilization in the Middle East (Berlin 2004); Manning S. (2013) “The Roman world and climate: context, relevance of climate change, and some issues”, in The Ancient Mediterranean Environment between Science and History, ed. W. V. Harris (Leiden 2013) 103–70; McCormick M. et al. (2012) “Climate change during and after the Roman empire: reconstructing the past from scientific and historical evidence”, Journal of Interdisciplinary History 43 (2012) 169–220; Baillie M. G. L. (2008) “Proposed re-dating of the European ice core chronology by seven years prior to the 7th century AD”, Geophysical Research Letters 35 (2008) doi: 10.1029/2008GL034755 (last accessed April 2017); Decker M. J. (2016) The Byzantine Dark Ages (New York 2016); Ellenblum R. (2012) The Collapse of the Eastern Mediterranean. Climate Change and the Decline of the East, 950–1072 (Cambridge 2012); Izdebski A. et al. (2016) “The environmental, archaeological and historical evidence for regional climatic changes and their societal impacts in the eastern Mediterranean in Late Antiquity”, Quaternary Science Reviews 136 (2016) 189–208; Koder J. (1996) “Climatic change in the fifth and sixth centuries?”, in The Sixth Century: End or Beginning?, edd. P. Allen, and E. Jeffreys (Brisbane 1996) 270–85; Kuniholm P. et al. (2011) “A 2367 year oak tree-ring chronology from 98 sites for the Aegean, East Mediterranean, and Black Seas: the Marmaray contribution”, in Koç Symposium Volume, ed. N. Ergin (Leuven 2011) 47–90; Kuniholm P. I. (1990) “Archaeological evidence and non-evidence for climatic change”, in The Earth’s Climate and Variability of the Sun Over Recent Millennia, edd. J. Runcorn, and J.-C. Pecker (London 1990) 645–55; Roberts C. N., Zanchetta G. and Jones M. D. (2010) “Oxygen isotopes as tracers of Mediterranean climate variability: an introduction”, Global and Planetary Change 71 (2010) 135–40; Schilman B. (2002) “Sealand paleoclimate correlation in the eastern Mediterranean region during the Late Holocene”, Israel Journal of Earth Sciences 51 (2002) 181–90; Stathakopoulos D. (2003) “Reconstructing the climate of the Byzantine world: state of the problem and case studies”, in People
McMahon and Sargent and Nature in Historical Perspective, edd. J. Laszlovszky, and P. Szabó (Budapest 2003) 247–61; Telelis I. (2004) Μετεωρολογικά φαινόμενα και κλίμα στο Βυζάντιο / Meteōrologika phainomena kai klima sto Vyzantio (Athens 2004); Telelis I. (2005) “Historical-climatological information from the time of the Byzantine empire (4th–15th-centuries AD)”, History of Meteorology 2 (2005) 41–50; Telelis I. (2008) “Climatic fluctuations in the eastern Mediterranean and the Middle East AD 300–1500 from Byzantine documentary and proxy physical paleoclimatic evidence—a comparison”, JÖB 58 (2008) 167–207; Xoplaki E. et al. (2016) “The medieval climate anomaly and Byzantium: a review of the evidence on climatic fluctuations, economic performance and societal change”, Quaternary Science Reviews 136 (2016) 229–52. Anatolia: England A. et al. (2008) “Historical landscape change in Cappadocia (central Turkey): a palaeoecological investigation of annually laminated sediments from Nar lake”, The Holocene 18 (2008) 1229–45; Haldon J. F. (2007) “‘Cappadocia will be given over to ruin and become a desert’. Environmental evidence for historicallyattested events in the 7th–10th centuries”, in Byzantina Mediterranea. Festschrift für Johannes Koder zum 65. Geburtstag (Vienna 2007) 215– 30; Haldon J. et al. (2014) “The climate and environment of Byzantine Anatolia: integrating science, history, and archaeology”, Journal of Interdisciplinary History 45 (2014) 113–61; Kaniewski D. et al. (2007) “A high-resolution Late Holocene landscape ecological history inferred from an intramontane basin in the western Taurus Mountains, Turkey”, Quaternary Science Reviews 26 (2007) 2201–18; Pearson C. L. et al. (2012) “Dendroarchaeology of the mid-first millennium AD in Constantinople”, JAS 39 (2012) 3402–14; Akcer O. (2011) Late Holocene Climatic Records of Kucukcekmece Lagoon, Yenicaga, Uludag Glacial and Bafa Lakes (Western Turkey) (Ph.D. diss., Istanbul Technical University/Eurasia Institute of Earth Sciences 2011); Auzépy M.-F. (2008) “State of emergency (700–850)”, in The Cambridge History of the Byzantine Empire, c. 500–1492, ed. J. Shepard (Cambridge 2008) 251–91; Bakker J. et al. (2011) “Numerically derived evidence for LateHolocene climate change and its impact on human presence in the southwest Taurus Mountains, Turkey”, The Holocene 22 (2011) 425–38; Bottema S., Woldring H. and Aytug B. (1993) “Late Quaternary vegetation and climate of southwestern Turkey. Part II”, Palaeohistoria 35/36 (1993) 13–72; Eastwood W. J. et al. (1999) “Holocene environmental change in southwest Turkey: a palaeoecological record of lake and catchment-related changes”, Quaternary Science Reviews 18 (1999) 671–95; Eastwood W. J. et al. (2007) “Holocene climate change in the eastern Mediterranean region: a comparison of stable isotope and pollen data from Lake Gölhisar, southwest Turkey”, Journal of Quaternary Science 22 (2007) 327–41; Haldon J. (2016) The Empire that Would Not Die: the Paradox of Eastern Roman Survival, 640–740 (Cambridge, Mass. 2016); Jones M. D. et al. (2006) “A high-resolution Late Holocene lake isotope record from Turkey and links to North Atlantic and monsoon climate”, Geology 34 (2006) 361–64; Leroy S. et al. (2002) “Abrupt environmental changes within a Late Holocene lacustrine sequence south of the Marmara Sea (Lake Manyas, N-W Turkey): possible links with seismic events”, Marine Geology 190 (2002) 531–52; Ülgen U. B. et al. (2012) “Climatic and environmental evolution of Lake Iznik (NW Turkey) over the last ~4700 years”, Quaternary International 274 (2012) 88–101; van Zeist W., Woldring H. and Stapert D. (1975) “Late Quaternary vegetation and climate of southwestern Turkey”, Palaeohistoria 17 (1975) 53–143; Woodbridge J. and Roberts N. (2011) “Late Holocene climate of the eastern Mediterranean inferred from diatom analysis of annually-laminated lake sediments”, Quaternary Science Reviews 30 (2011) 3381–92. Other: Bulliet R. W. (2009) Cotton, Climate, and Camels in Early Islamic Iran: a Moment in World History (New York 2009); Neumann J. (1991) “Climate of the Black Sea region around 0 CE”, Climatic Change 18 (1991) 453–65; Sadori L. et al. (2016) “Climate, environment and society in southern Italy during the last 2000 years. A review of the
Environmental History of the Late Antique Eastern Mediterranean environmental, historical and archaeological evidence”, Quaternary Science Reviews 136 (2016) 173–88; Bar-Matthews M. et al. (1999) “The eastern Mediterranean paleoclimate as a reflection of regional events: Soreq cave, Israel”, Earth and Planetary Science Letters 166 (1999) 85–95; Leroy S. A. G. (2010) “Pollen analysis of core DS7–1SC (Dead Sea) showing intertwined effects of climatic change and human activities in the Late Holocene”, JAS 37 (2010) 306–16; Neumann F. et al. (2006) “Holocene vegetation and climate history of the northern Golan heights (near east)”, Vegetation History and Archaeobotany 16 (2006) 329–46; Neumann F. H. et al. (2010) “Vegetation history and climate fluctuations on a transect along the Dead Sea west shore and their impact on past societies over the last 3500 years”, Journal of Arid Environments 74 (2010) 756–64. Effects of volcanoes: Arjava A. (2005) “The mystery cloud of 536 CE in the Mediterranean sources”, DOP 59 (2005) 73–94; Baillie M. G. L. (1994) “Dendrochronology raises questions about the nature of the AD 536 dust-veil event”, The Holocene 4 (1994) 212–17; Baillie M. G. L. (2010) “Volcanoes, ice-cores and tree-rings: one story or two?”, Antiquity 84 (2010) 202–15; Churakova (Sidorova) O. V. et al. (2014) “A cluster of stratospheric volcanic eruptions in the AD 530s recorded in Siberian tree rings”, Global and Planetary Change 122 (2014) 140–50; Farquharson P. (1996) “Byzantium, planet Earth and the solar system”, in The Sixth Century: End or Beginning?, edd. P. Allen and E. Jeffreys (Melbourne 1996) 263–69; Fei J., Zhou J. and Hou Y. (2007) “Circa A.D. 626 volcanic eruption, climatic cooling, and the collapse of the eastern Turkic empire”, Climatic Change 81 (2007) 469–75; Ferris D. G. et al. (2011) “South pole ice core record of explosive volcanic eruptions in the first and second millennia AD and evidence of a large eruption in the tropics around 535 A.D.”, Journal of Geophysical Research 116 (2011) D17308 doi: 10.1029/2011JD015916 (last accessed April 2017); Gräslund B. and Price N. (2012) “Twilight of the gods? The ‘dust veil event’ of A.D. 536 in critical perspective”, Antiquity 86 (2012) 428–43; Larsen L. B. et al. (2008) “New ice core evidence for a volcanic cause of the AD 536 dust veil”, Geophysical Research Letters 35 (2008) doi: 10.1029/2007GL032450 (last accessed April 2017); McCormick M., Dutton P. E. and Mayewski P. A. (2007) “Volcanoes and the climate forcing of Carolingian Europe, A.D. 750–950”, Speculum 82 (2007) 865–95; Stothers R. B. and Rampino M. R. (1983) “Volcanic eruptions in the Mediterranean before A.D. 630 from written and archaeological sources”, Journal of Geophysical Research 88 (1983) 6357–71.
All of the new works on climate history, whether from a primarily historical perspective, or from a largely scientific one, are indebted to the proliferation of new scientific techniques, for extracting information about past climates from natural archives. Archaeological and historical sources continue to be used (as in Hirschfeld (2004), Domínguez-Castro et al. (2012), Koenig (1991) and Widell 2007), but are also supplemented by evidence from ice cores, speleothems, tree rings, and (with care) ancient pollen from lake beds. Pollen analysis is used extensively to understand the responses of plant communities (and sometimes human cultivators) to climate fluctuations, and can even reflect political change and warfare: see Haldon (2007). Databases containing information from these palaeoenvironmental proxy sources are now available online. Along with new techniques (or the increased availability of old methods and data) come, of course, debates over methods. These include questions about the best laboratory methods to extract such information, and the best ways to interpret
29
the resulting data to determine climatic conditions, for which see: Briffa (2000), and Bradley (2014), which has gone through three editions since 1999. Questions also arise as to the appropriate use of such data for historical research, see in Manning (2013), Rotberg and Rabb (1981) and Brázdil et al. (2005). Currently, the pendulum of historical thought has swung away from the euphoric hope that climate could explain every previously unexplainable event, or movement in history, to a more moderate view. This view holds that climate can influence, but rarely determine, the decisions of stubborn human agents: Rosen (2007), McIntosh et al. (2000) and Hulme (2011). These methodological developments are evident from the foregoing bibliography; the studies listed below are more specifically concerned with explaining or debating techniques, or use particularly novel methods. Additionally, since climate histories demand broad perspectives, the final section lists useful studies which are temporally or geographically just beyond the scope of this volume. Databases and working groups for palaeoenvironmental data sets: European Pollen Database (EPD): http://www.europeanpollendata base.net. National centres for environmental information: https:// www.ncdc.noaa.gov/data-access/paleoclimatology-data/datasets; McCormick M et al. (2013) “Geodatabase of historical evidence on Roman and Post-Roman climate”, hdl:1902.1/22615, Harvard Dataverse, V4: https://dataverse.harvard.edu/dataset.xhtml?persiste ntId=hdl:1902.1/22615 (last accessed April 2017). Methodologically useful climate histories: Bradley R. S. (2014) Paleoclimatology: Reconstructing Climates of the Quaternary (San Diego 2014); Brázdil R. et al. (2005) “Historical climatology in Europe—the state of the art”, Climatic Change 70 (2005) 363–430; Briffa K. R. (2000) “Annual climate variability in the Holocene: interpreting the message of ancient trees”, Quaternary Science Reviews 19 (2000) 87–105; Hulme M. (2011) “Reducing the future to climate: a story of climate determinism and reductionism”, Osiris 26 (2011) 245– 66; McIntosh R. J., Tainter J. A. and McIntosh S. K. (2000) The Way the Wind Blows: Climate Change, History, and Human Action (New York 2000); Alexandre P. (1987) Le climat en Europe au moyen âge: contribution à l’histoire des variations climatiques de 1000 à 1425, d’après les sources narratives de l’Europe occidentale (Paris 1987); Behringer W. (2009) A Cultural History of Climate (Cambridge 2009); Brázdil R. (2003) “Historical climatology and its progress after 1990”, in People and Nature in Historical Perspective, edd. J. Laszlovszky and P. Szabó (Budapest 2003) 197–227; Brooke J. L. (2014) Climate Change and the Course of Global History: a Rough Journey (Cambridge 2014); Carey M. (2012) “Climate and history: a critical review of historical climatology and climate change historiography”, WIREs Climate Change 3 (2012) 233–49; Domínguez-Castro F. et al. (2012) “How useful could Arabic documentary sources be for reconstructing past climate?”, Weather 67 (2012) 76–82; Dean J. R. et al. (2013) “Palaeo-seasonality of the last two millennia reconstructed from the oxygen isotope composition of carbonates and diatom silica from Nar Gölü, central Turkey”, Quaternary Science Reviews 66 (2013) 35–44; Glaser R. and Riemann D. (2009) “A thousand-year record of temperature variations for Germany and central Europe based on documentary data”, Journal of Quaternary Science 24 (2009) 437–49; Guiot J. (2012) “A robust spatial reconstruction of April to September temperature in Europe: comparisons between the medieval period and the recent warming with a focus on extreme values”, Global and Planetary
30 Change 84–85 (2012) 14–22; Griggs C. et al. (2007) “A regional highfrequency reconstruction of May–June precipitation in the North Aegean from oak tree rings, A.D. 1089–1989”, International Journal of Climatology 27 (2007) 1075–89; Haldon J. F. (2007) “‘Cappadocia will be given over to ruin and become a desert’. Environmental evidence for historically-attested events in the 7th–10th centuries”, in Byzantina Mediterranea. Festschrift für Johannes Koder zum 65. Geburtstag (Vienna 2007) 215–30; Hirschfeld Y. (2004) “A climatic change in the Early Byzantine period? Some archaeological evidence”, PEQ 136 (2004) 133–49; Koenig D. (1991) Medieval Winters in Baghdad: a Study in the History of Climate Between 296 A.H. and 493 A.H. Based on Ibn al-Jawzi’s Kitab al-Muntazam (M.A. diss., Univ. of Columbia 1991); Kress A. et al. (2014) “Swiss tree rings reveal warm and wet summers during medieval times”, Geophysical Research Letters 41 (2014) doi: 10.1002/2013GL059081 (last accessed April 2017); Le Houérou H. N. (1982) “The arid bioclimates in the Mediterranean isoclimatic zone”, Ecologica Mediterranea 8 (1982) 103–14; Le Roy Ladurie E. (1971) Times of Feast, Times of Famine: the History of Climate Change since the Year 1000 (Garden City, New York 1971); Luterbacher J. et al. (2012) “A review of 2000 years of paleoclimatic evidence in the Mediterranean”, in The Climate of the Mediterranean Region, ed. P. Lionello (Oxford 2012) 87–185; Manning S. (2013) “The Roman world and climate: context, relevance of climate change, and some issues”, in The Ancient Mediterranean Environment between Science and History, ed. W. V. Harris (Leiden 2013) 103–70; McIntosh R., Tainter J. A. and McIntosh S. K. (2000) “Climate, history, and human action”, in The Way the Wind Blows: Climate Change, History, and Human Action, edd. R. J. McIntosh, J. A. Tainter and S. K. McIntosh (New York 2000) 1–42; Moreno A. et al. (2012) “The medieval climate anomaly in the Iberian peninsula reconstructed from marine and lake records”, Quaternary Science Reviews 43 (2012) 16–32; Pfister C. (1985) Klimageschichte der Schweiz 1525–1860. Das Klima der Schweiz von 1525–1860 und seine Bedeutung in der Geschichte von Bevölkerung und Landwirtschaft (Bern 1985); Roberts N. (2011) “‘Living with a moving target’: longterm climatic variability and environmental risk in dryland regions”, in Sustainable Lifeways: Cultural Persistence in an Ever-Changing Environment, edd. N. Miller, K. Moore and K. Ryan (Philadelphia 2011) 13–38; Rotberg R. I. and Rabb T. K. (1981) Climate and History: Studies in Interdisciplinary History (Princeton 1981); Ruddiman W. F. (2010) Plows, Plagues, and Petroleum: How Humans Took Control of Climate (Princeton 2010); Steig E. (1999) “Mid-Holocene climate change”, Science 286 (1999) 1485–87; Vogel B. (2009) “Bibliography of recent literature in the history of meteorology: twenty six years, 1983–2008”, History of Meteorology 5 (2009) 23–125; Widell M. (2007) “Historical evidence for climate instability and environmental catastrophes in northern Syria and the Jazira: the Chronicle of Michael the Syrian”, Environment and History 13 (2007) 47–70.
McMahon and Sargent Peripherally-relevant climate histories: Rosen A. M. (2007) Civilizing Climate: Social Responses to Climate Change in the Ancient Near East (Lanham, Maryland 2007); Boomer I. et al. (2009) “Advances in understanding the Late Holocene history of the Aral Sea region”, Quaternary International 194 (2009) 79–90; Büntgen U. et al. (2011) “2500 years of European climate variability and human susceptibility”, Science 331 (2011) 578–82; Cook E. R. et al. “Old world megadroughts and pluvials during the common era”, Science Advances (Nov. 2015) doi: 10.1126/sciadv.1500561 (last accessed April 2017); D’Arrigo R. et al. (2001) “1738 years of Mongolian temperature variability inferred from a tree-ring width chronology of Siberian pine”, Geophysical Research Letters 28 (2001) 543–46; Finné M. et al. (2011) “Climate in the eastern Mediterranean, and adjacent regions, during the past 6000 years—a review”, JAS 38 (2011) 3153–73; Francke A. et al. (2013) “A Late Glacial to Holocene record of environmental change from Lake Dojran (Macedonia, Greece)”, Climate of the Past 9 (2013) 481–98; Göktürk O. M. et al. (2011) “Climate on the southern Black Sea coast during the Holocene: implications from the Sofular Cave record”, Quaternary Science Reviews 30 (2011) 2433–45; Graham N. E. et al. (2011) “Support for global climate reorganization during the ‘Medieval Climate Anomaly’ ”, Climate Dynamics 37 (2011) 1217–45; Gunn J. D. (2000) The Years without Summer: Tracing AD 536 and its Aftermath (Oxford 2000); Kuzucuoğlu C. et al. (2011) “Mid- to LateHolocene climate change in central Turkey: the Tecer Lake record”, Holocene 21 (2011) 173–88; Netser M. (1998) “Population growth and decline in the northern part of Eretz-Israel during the historical period as related to climatic changes”, in Water, Environment and Society in Times of Climatic Change, edd. A. S. Issar and N. Brown (Dordrecht 1998) 129–45; Nichol K. and Küçükuysal C. (2013) “Emerging multiproxy records of Late Quaternary palaeoclimate dynamics in Turkey and the surrounding region”, Turkish Journal of Earth Sciences 22 (2013) 126–42; Noorollahi D. et al. (2011) “Climatic and environmental reconstruction based on stable isotopes of Parishan Lake (Iran)”, Journal of Rangeland Science 1 (2011) 203–16; Rita F. D. and Magri D. (2009) “Holocene drought, deforestation and evergreen vegetation development in the central Mediterranean: a 5500 year record from Lago Alimini Piccolo, Apulia, southeast Italy”, The Holocene 19 (2009) 295–306; Roberts N. et al. (2012) “Palaeolimnological evidence for an east—west climate see-saw in the Mediterranean since AD 900”, Global and Planetary Change 84–85 (2012) 23–34; Squatriti P. (2010) “The floods of 589 and climate change at the beginning of the Middle Ages: an Italian microhistory”, Speculum 85 (2010) 799–826; Touchan R. et al. (2007) “May–June precipitation reconstruction of southwestern Anatolia, Turkey during the last 900 years from tree rings”, Quaternary Research 68 (2007) 196–202; Tvauri A. (2014) “The impact of the climate catastrophe of 536–537 AD in Estonia and neighbouring states”, Estonian Journal of Archaeology 18 (2014) 30–56.
The Environmental History of the Late Antique West: a Bibliographic Essay Merle Eisenberg, David J. Patterson, Jamie Kreiner, Ellen F. Arnold, and Timothy P. Newfield Introduction Premodern environmental history is a young subfield with deep roots. While less scholarship deals directly with the environmental history of Late Antiquity than with the Middle Ages or early modernity, a plethora of late antique articles and books have appeared over the last 50 years that are relevant to the field. Of course, few of the authors might identify themselves as environmental historians. Fewer yet, were trained as environmental historians attuned to the dynamic relationship people have always had with the natural world, or were accustomed to interdisciplinarity, and taught from books like Worster (1977), Merchant (1980), and White (1983). Nevertheless, studies on late antique agrarian technology, military rationing, population trends, Mediterranean trade, and many other topics, ultimately research humanecological interaction. Environmental history might have only came into being in the 1970s, but late antique environmental history stretches back long before that. Environmental history grew exponentially in the 1990s (McNeill (2003); Winiwarter (2004)) and that growth has not slowed. The best introductions with an emphasis on premodernity and the West, are Winiwarter and Knoll (2007), Arnold (2008), and Hoffmann (2014). For research on the late antique East, see the accompanying essay in this volume by McMahon and Sargent (2018). The field continues to expand and evolve at a rapid rate, of which this volume, we hope, is an important part. A number of key themes are addressed below. While our coverage is not exhaustive, the works introduced here will be of use to students and scholars of Late Antiquity. A final prefatory observation seems warranted. When written sources are thin, dramatic environmental change is often positioned as a major force of cultural, economic and social evolution. The natural sciences certainly offer an insight on moments when documentary evidence is scarce, but environmental variables, whether faintly described by historians or only visible in natural archives, should not acquire a prominent role because of the slightness of the written sources. Great caution is required if documentary silence amplifies the way one treats the scientific evidence. In this regard, the climatic variability and plagues experienced in Late Antiquity, and in the mid 6th c. especially, have garnered much attention. Popular and scholarly work has argued that widespread plague mortalities and dramatic climate deterioration effectively ended antiquity
(in the East and West) and ushered in the Middle Ages. Keys (1999), Rosen (2007) and Harper (2017) are a few examples of this approach. The scope and severity of the events in question—the 536–50 temperature downturn and the Justinianic Plague in particular—will be debated for years to come. In the meantime, it is important to emphasise, that ‘catastrophism’ and ‘collapsology’ neither drive nor define most research on Late Antiquity’s environmental history, nor should they. Environmental historians pursue many strands of human-natural world interaction; climate and disease, whether conceived of as dramatic drivers of change or not, are but two. General works: Worster D. (1977) Nature’s Economy: a History of Ecological Ideas (San Francisco 1977); Merchant C. (1980) The Death of Nature: Women, Ecology, and the Scientific Revolution (San Francisco 1980); White R. (1983) The Roots of Dependency: Subsistence, Environment, and Social Change Among the Choctaws, Pawnees and Navajos (Lincoln 1983); Winiwarter V. and Knoll M. (2007) Umweltgeschichte: Eine Einführung (Vienna 2007); Arnold E. F. (2008) “An introduction to Medieval environmental history”, History Compass 6 (2008) 898–916; Hoffmann R. (2014) An Environmental History of Medieval Europe (Cambridge 2014); McMahon L. and Sargent A. (2018) “Environmental history of the late antique eastern Mediterranean: a bibliographic essay”, in this volume; McNeill J. (2003), “Observations on the nature and culture of environmental history”, History and Theory 42.4 (2003) 4–43; Winiwarter V. et al. (2004) “Environmental history in Europe from 1994 to 2004: enthusiasm and consolidation”, Environment and History 10 (2004) 501– 30; Keys D. (1999) Catastrophe: an Investigation into the Origins of Modern Civilization (London 1999); Rosen W. (2007) Justinian’s Flea: the First Great Plague and the End of the Roman Empire (New York 2007); Harper K. (2017) The Fate of Rome: Climate, Disease, and the End of an Empire (Princeton 2017).
Demography Premodern demography remains an imprecise field, and its numbers are often best viewed as ‘numerical metaphors’; good for thinking with, rather than hard realities: Hopkins (1998). Settlement archaeology and documentary evidence can be supplemented with modern ‘life model tables’ that supply a realistic ranges of estimates: Frier (2000); Parkin (1992). With some idea of a region’s carrying capacity or a city’s ghost acreage, one can also get a rough idea of population potential. Evidence of late antique population decline in the cities and countryside of the West has prompted lively scholarly discussion. Estimates of this decline can be staggering, particularly for Rome. Having achieved a population of perhaps 1 million during the Pax Romana, Rome’s population was roughly halved by the beginning of the 5th c., and by the end of the 6th c. it was as low as
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_004 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 31–50
32 60,000: Durliat (1990). By the 9th c., Rome may have been yet half that size: Verhulst (2002). Imperial ideology and redistribution likely accounted for Rome’s artificially massive size at its height, but as imperial power shifted east, so too did people; Rome and Constantinople were on opposite demographic trajectories in Late Antiquity: Van Dam (2010). In the ancient West, however, Rome marks a very exceptional case. The overwhelming majority of the population was rural, or based in much smaller settlements. Much has been written on the emptying of the Roman countryside and the decline of the Roman city in Late Antiquity: Potter (1976); Christie and Loseby (1996); Loseby (2000); Francovich and Hodges (2003); Ward-Perkins (2005); Christie (2006), (2010). Where the archaeological record suggests pronounced settlement abandonment, changes in social and material culture may offer explanations, rather than demographic collapse, see for example Lewit (2003); Favory et al. (2003). Scholars often draw connections between population decline and environmental ‘causes’ such as plague and climate change: Cheyette (2008). The demographic impact of the Antonine (AD 165–82/156–89), Cyprianic (ca. 249–70) and Justinianic plagues (ca. 541–749) are highly uncertain, and 6th c. declines in Italy and Africa could just as easily be ascribed to conflict as to emerging disease and subsistence crises: Wickham (2005). The most secure archaeological cases for large regional declines in rural population (ca. 50%) come from northern Gaul and eastern England in the 5th c.—so before the Justinianic plague—leading Wickham to conclude that local (rather than universal) explanations must be sought for late antique population erosion. Just as environmental changes have been positioned as causes of demographic contraction, so too have they been posited as the result of shifting population patterns; for some examples, see ‘Trees and Woodland’, below. For Italy, Lo Cascio and Malanima (2005) caution against catastrophist interpretations, arguing instead for a gradual (and spatiotemporally unsteady) population decline from the 3rd to 7th c. Others have focused on the mechanisms of decline, for example increased mortality rather than reduced fertility: Barbiera and Dalla-Zuanna (2009). On reduced fertility through late marriage and birth control in the Post Roman period, see Wickham (2005) 538, and more generally Riddle (1997). Population history: Hopkins K. (1998) “Christian number and its implications”, Journal of Early Christian Studies 6 (1998) 185–226; Frier B. (2000) CAH 11: the High Empire AD 70–192 (Cambridge 2000) 787–811; Parkin T. (1992) Demography and Roman Society (Baltimore 1992); Durliat J. (1990) De la ville antique à la ville byzantine: le problème des subsistances (Paris 1990); Verhulst A. (2002) The Carolingian Economy (Cambridge 2002); Cheyette F. L. (2008) “The disappearance of the ancient landscape and the climatic anomaly of the Early Middle Ages: a question to be pursued”, Early Medieval Europe
Eisenberg et al. 16.2 (2008) 127–65; Van Dam R. (2010) Rome and Constantinople: Rewriting Roman History During Late Antiquity (Waco, Texas 2010); Liebeschuetz J. H. W. G. (2001) Decline and Fall of the Roman City (Oxford 2001); Wickham C. (2005) Framing the Early Middle Ages: Europe and the Mediterranean 400–800 (Oxford 2005) 548ff.; Lo Cascio E. and Malanima P. (2005) “Cycles and stability: Italian population before the demographic transition (225 BC–AD 1900)”, Rivista di Storia Economica 21 (2005) 197–332; Barbiera I. and Dalla-Zuanna G. (2009) “Population dynamics in Italy in the Middle Ages: new insights from archaeological findings”, Population and Development Review 35 (2009) 367–89; Potter T. (1979) The Changing Landscape of South Etruria (London 1979); Christie N. and Loseby S. (1996) edd. Towns in Transition: Urban Evolution in Late Antiquity to the Middle Ages (Aldershot 1996); Loseby S. (2000) “Urban failures in late-antique Gaul”, in Towns in Decline, AD 100–1600, ed. T. Slater (Aldershot 2000) 72–95; Francovich R. and Hodges R. (2003) Villa to Village: the Transformation of the Roman Countryside (London 2003); Christie N. (2004) ed. Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages (Aldershot 2004); Ward-Perkins B. (2005) The Fall of Rome and the End of Civilization (Oxford 2005); Christie N. (2006) From Constantine to Charlemagne: an Archaeology of Italy, AD 300–850 (Aldershot 2006); Christie N. (2010) The Fall of the Western Roman Empire: an Archaeological and Historical Perspective (London 2010); Christie N. and Augenti A. (2012) edd. Vrbes Extinctae: Archaeologies of Abandoned Classical Towns (Aldershot 2012); Livi Bacci M. (2000) The Population of Europe: a History (Hoboken, New Jersey 2000); Bachrach B. (1999) “Early Medieval military demography: some observations on the methods of Hans Delbrück”, in The Circle of War in the Middle Ages: Essays on Medieval Military and Naval History, edd. D. J. Kagay and L. J. A. Villalon (Rochester, New York 1999) 3–20; Russell J. (1985) The Control of Late Ancient and Medieval Population (Philadelphia 1985; orig. publ. 1958); WittwerBackofen et al. (2008) “Basics in paleodemography: a comparison of age indicators applied to the early medieval skeletal sample of Lauchheim”, American Journal of Physical Anthropology 137 (2008) 384–96; Brothwell D. (1972) “Palaeodemography and earlier British populations”, WorldArch 4 (1972) 75–87; Lewit T. (2003) “‘Vanishing villas’: what happened to élite rural habitation in the West in the 5th–6th c?”, JRA 16 (2003) 260–74; Riddle J. (1997) Eve’s Herbs: a History of Contraception and Abortion in the West (Cambridge, Mass. 1997); Favory F. et al. (2003) “La dynamique spatio-temporelle de l’habitat rural gallo-romain”, in Archéologie et systèmes socioenvironnementaux: Études multiscalaires sur la vallée du Rhône dans le programme ARCHAEOMEDES, edd. S. van der Leeuw, F. Favory, J.-L. Fiche (Paris 2003).
Religion and Environment The religious transformations of Late Antiquity (traced in Brown (2013)) comprise yet another central issue warranting more attention. The extent to which they played a role in the changed uses of, and interactions with, the environment remains debated. White Jr. (1967) advanced the idea that Christianity fundamentally transformed the relationship between humans and nature, to the detriment of both. As Livingstone (1994) has summarised (updated in Hoffmann (2014)), White Jr. has had many critics. Le Goff (1980) has examined the changes Christianity brought to human relations with the environment around them, especially as some regions in the West were transformed, metaphorically, into eastern deserts. Monastic ideals of desert living could fundamentally alter debates and episcopal ways of living, as
The Environmental History of the Late Antique West
Leyser (2006) has shown. For a case study on how religious identity was influenced by the environment in the early medieval West see Arnold (2012). Several works examine, more generally, how people conceptualised the environment itself. The classic survey of premodernity remains Glacken (1967), and more recently Coates (1998) and Hughes (2014). Several works also frame plants (Marder (2014)) and animals (Salisbury (2011), Steel (2011)) as central to understandings of the past. Christianity and conceptualising nature: White L. Jr. (1967) “The historical roots of our ecological crisis”, Science 155. 3767 (1967) 1203–1207; Livingstone D. (1994) “The historical roots of our ecological crisis: a reassessment”, Fides et Historia 26.1 (1994) 38–55; Whitney E. (2013) “The Lynn White Thesis: Reception and Legacy”, Environmental Ethics 35, 313–331; Brown P. (2013) The Rise of Western Christendom: Triumph and Diversity, AD 200–1000. (Chichester, 10th anniversary rev. edn. 2013); Le Goff J. (1980) Time, Work and Culture in the Middle Ages (transl. A. Goldhammer) (Chicago 1980); Arnold E. (2012) Negotiating the Landscape: Environment and Monastic Identity in the Medieval Ardennes (Philadelphia 2012); Leyser C. (2006) “The uses of the desert in the sixth-century West”, Church History and Religious Culture 86.1 (2006) 113–34; Glacken C. J. (1967) Traces on the Rhodian Shore; Nature and Culture in Western Thought from Ancient Times to the End of the Eighteenth Century (Berkeley 1967); Coates P. (1998) Nature: Western Attitudes Since Ancient Times. (Cambridge 1998); Hughes J. D. (2014) Environmental Problems of the Greeks and Romans: Ecology in the Ancient Mediterranean (Baltimore, 2nd edn. 2014); Marder M. (2014) The Philosopher’s Plant: an Intellectual Herbarium. (New York 2014); Salisbury J. (2011) The Beast Within: Animals in the Middle Ages (New York, 2nd edn. 2011); Steel K. (2011) How to Make a Human: Animals and Violence in the Middle Ages (Columbus, Ohio 2011); Herlihy D. (1980) “Attitudes toward the environment in medieval society”, in Historical Ecology: Essays on Environment and Social Change (Port Washington, New York 1980) 100–16; Oelschlaeger M. (1991) The Idea of Wilderness: From Prehistory to the Age of Ecology (New Haven, Connecticut 1991); Epstein S. (2012) The Medieval Discovery of Nature (Cambridge 2012).
Agriculture Arable Agriculture Britain For Anglo-Saxon England arable agriculture has been studied in some detail, although often in local contexts. The best overview of existing paradigms and their problems can be found in Oosthuizen (2016). A useful overview of all types of grain that was grown here, and current literature on them, is in Moffett (2011), while McKerracher (2016) reanalyses existing theories on the consumption of various types of grain. The article by Jones, Straker, and Davis (1991) examines plant use, and the essays edited by Rackham (1994) provide a useful archaeological synthesis on rural and urban sites. On landscape archaeology, broadly construed, including discussions of cereals and field systems, see the essays in Higham and Ryan (2010). For an examination of the long 8th c., see now McKerracher (2018). An in-depth
33 discussion of farms and farming techniques for the entire Anglo-Saxon period can be found in Banham and Faith (2014). The consumption of arable agriculture has been well-studied in Banham (2004) and Hagen (2006). For a broader discussion of settlements, society, and the consumption of goods see Hamerow (2002) and (2012). An annually updated bibliography on agricultural history is in McShane (2018), who compiles new works that includes a wide range of late antique, early medieval, and Anglo-Saxon references. Britain: Oosthuizen S. (2016) “Recognizing and moving on from a failed paradigm: the case of agricultural landscapes in AngloSaxon England c. AD 400–800”, Journal of Archaeological Research 24.2 (2016) 179–227; Moffett L. (2011) “Food plants on archaeological sites”, in The Oxford Handbook of Anglo-Saxon Archaeology, edd. H. Hamerow, D. A. Hinton, and S. Crawford (Oxford 2011) 346–60; McKerracher M. (2016) “Bread and surpluses: the Anglo-Saxon ‘bread wheat thesis’ reconsidered”, Environmental Archaeology 21.1 (2016) 88–102; Jones G., Straker V., and Davis A. (1991) “Early medieval plant use and ecology’, in Anglo-Saxon and Norman London 2: Finds and Environmental Evidence, ed. A. Vince (London 1991) 347–88; Rackham J. (1994) ed. Environment and Economy in AngloSaxon England (York 1994); Higham N. J. and Ryan M. J (2010) edd. The Landscape Archaeology of Anglo-Saxon England (Rochester, New York, 2010); McKerracher M. (2018) Farming Transformed in AngloSaxon England: Agriculture in the Long Eighth Century (Oxford 2018); Banham D. and Faith R. (2014) Anglo-Saxon Farms and Farming (Oxford 2014); Banham D. (2004) Food and Drink in Anglo-Saxon England (Stroud 2004); Hagen A. (2006) Anglo-Saxon Food and Drink: Production, Distribution and Consumption (Hockwold cum Wilton, Norfolk 2006); Hamerow H. (2002) Early Medieval Settlements: the Archaeology of Rural Communities in North-West Europe, 400–900 (Oxford 2002); Hamerow H. (2012) Rural Settlements and Society in Anglo-Saxon England (Oxford 2012); McShane P. (2018) “Annual list of publications on agrarian history, 2016”, Agricultural History Review 66.1 (2018) 135–51.
Italy and Spain There is abundant scholarship for this region on Roman farming and the imperial grain supply. Erdkamp (2005) attempts to estimate Italian yields and the size of the peninsula’s agricultural labour force. The classic work on Roman farming technology is White (1970), who is concerned with antiquity but it is also useful for later periods. Work on Rome’s grain supply naturally focuses on North Africa and Sicily (for which see Soraci (2011)) and on political and economic, rather than environmental, questions. However, Shaw (2013) is indispensable for the logistics of Mediterranean cereal harvests in the late empire, and includes Italian harvesting contracts in his appendix. There are several useful overviews of late antique and early medieval cereal cultivation with Italian or Iberian material examined, such as Jasny (1944) and more recently Comet (2004). Arthur (2004) surveys broad developments in the Italian countryside from AD 400–1000. A useful survey of agricultural production and land use in late antique Iberia can now be found
34 in Fernández (2017). The heterogeneity of Italy’s and Spain’s agricultural landscapes, however, makes local studies of specific regions necessary. Vaccaro (2013) treats the grain heartland of interior Sicily and its wider economic links, while Buonincontri et al. (2014) shed light on shifting cultivation patterns in central Italy. Due to a shortage of documentary sources, archaeology has always been vital to the study of Italian and Iberian agriculture. See, for example: Jarman (1976); Ampolo (1980); Bosqued and Pellicer (2000); Castanyer and Tremoleda (2006). Diet, famine, and agricultural failure have received some attention for the Roman period (Garnsey (1988)), but less so for Late Antiquity and the Early Middle Ages (Montanari 1979). In some cases, new techniques have offered fresh insight on old questions. Palynological and archaeobotanical studies can illuminate shifting patterns of plant use (Alonso (2008)), although much remains to be done for Late Antiquity (Peña-Chocarro et al. (2017)). In general, scholars have pointed to the diminishing status of barley in the Post-Classical period, the gradual replacement of emmer wheat with durum, and the rise of rye (Sallares (2008)). Local studies are of great importance here as well, since no general pattern emerges for the entire western Mediterranean or Europe. Castiglioni and Rottoli (2010), for example, argue that sorghum made inroads in northern Italy during the Middle Ages. Watson (1977) famously argued for the westward diffusion of sorghum and a number of other important crops within the early caliphate, although the “Arab agricultural revolution” thesis has not been without critics (Squatriti (2014)). Italy and Spain: Comet G. (2004) “Les céréales du bas-empire au moyen âge”, in The Making of Feudal Agricultures?, edd. M. Barcelo and F. Siguat (Leiden 2004) 120–76; Arthur P. (2004) “From vicus to village: Italian landscapes, A.D. 400–1000”, in Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages, ed. N. Christie (London 2004) 103–33; Fernández D. (2017) Aristocrats and Statehood in Western Iberia, c. 300–600 C.E. (Philadelphia 2017); Christie N. (1996) “Barren fields? Landscapes and settlements in Late Roman and Post-Roman Italy”, in Human Landscapes in Classical Antiquity: Environment and Culture, edd. G. Shipley and J. Salmon (London 1996) 254–84; Erdkamp P. (2005) The Grain Market in the Roman Empire: a Social, Political, and Economic Study (New York 2005); Sallares R. (2008) “Ecology”, in The Cambridge Economic History of the Greco-Roman World, edd. W. Scheidel, I. Morris, and R. Saller (New York 2008) 15–37; Soraci C. (2011) Sicilia frumentaria. Il grano siciliano e l’annona di Roma V a.C.-V d.C. (Rome 2011); Vaccaro E. (2013) “Patterning the late antique economies of inland Sicily in a Mediterranean context”, in Local Economies?: Production and Exchange of Inland Regions in Late Antiquity, ed. L. Lavan (Late Antique Archaeology 10) (Leiden-Boston 2013) 259–313; Bianchi G. and Grassi F. (2012) “Sistemi di stoccaggio nelle campagne italiane (secc. VII–XIII): l’evidenza archeologica dal caso di Rocca degli Alberti in Toscana”, in Horrea, Barns and Silos: Storage and Incomes in Early Medieval Europe, edd. G. Bianchi et al. (Bilbao 2012) 58–77; Buonincontri et al. (2014) “Farming in a rural settlement in central
Eisenberg et al. Italy: cultural and environmental implications of crop production through the transition from Lombard to Frankish Influence”, Vegetation History and Archaeobotany 23 (2014) 775–88; Castiglioni E. and Rottoli M. (2010) “Il sorgo (Sorghum bicolor) nel medioevo in Italia settentrionale”, Archeologia medievale 37 (2010) 485–96; Cracco Ruggini L. (1995) Economia e società nell’Italia annonaria: rapporti fra agricoltura e commercio dal IV al VI secolo d.C. (Bari 1995); De Robertis F. (1972) La produzione agricola in Italia dalla crisi del III secolo all’età dei Carolingi (Rome 1972); Ampolo C. (1980) “Le condizioni materiali della produzione. agricoltura e paesaggio agrario”, Dialoghi d’Archeologia NS 1 (1980) 15–46; Jarman H. (1976) “The plant remains”, in A Faliscan Town in South Etruria, Excavations at Narce 1966–71 (London 1976) 308–10; Montanari I. (1975) “Cereali e legumi nell’alto medioevo in Italia nord, secoli IX–X”, Rivista Storica Italiana 87 (1975) 439–88; Garnsey P. (1988) Famine and Food Supply in the Graeco-Roman World: Responses to Risk and Crisis (New York 1988); Montanari M. (1979) L’ alimentazione contadina nell’alto Medioevo (Naples 1979); Jones R. (1988) “The end of the Roman countryside in the Iberian peninsula”, in First Millennium Papers: Western Europe in the First Millennium, ed. R. Jones (BAR-IS 401) (Oxford 1988) 159–73; Bosqued C. and Pellicer R. (2000) El yacimiento romano de La Torrecilla: de villa a Tugurium (Madrid 2000); Castanyer P. and Tremoleda J. (2006) “La villa de Vilauba. De la antigüedad tardía hasta el abandono final”, Anejos de AEspA 39 (2006) 133–51; Alonso N. (2008) “Crops and agriculture during the Iron Age and Late Antiquity in Cerdanyola del Vallès (Catalonia, Spain)”, Vegetation History and Archaeobotany 17.1 (2008) 75–84; Peña-Chocarro et al. (2017) “Roman and medieval crops in the Iberian Peninsula: a first overview of seeds and fruits from archaeological sites”, Quaternary International (2017) doi: https://doi.org/10.1016/j.quaint.2017.09.037 (in press); Watson A. (1974) “The Arab agricultural revolution and its diffusion, 700–1100”, Journal of Economic History 34.1 (1974) 8–35; Squatriti P. (2014) “Of seeds, seasons, and seas: Andrew Watson’s medieval agrarian revolution forty years later”, The Journal of Economic History 74.4 (2014) 1205–20; Hernández J. et al. (2002) “De Roma al medievo. Estructuras de hábitat y evolución del paisaje vegetal en el Territorio de Salamanca”, Zephyrus 55 (2002) 283–309.
Gaul/Francia Gaul and Francia can be roughly subdivided between research on the Roman, Merovingian, and Early Carolingian periods. The foundational studies of Bloch (1931) and Duby (1962) begin with the Early Middle Ages, and in many respects continue to be useful overviews. A detailed history of agriculture in the Loess Belt west of the river Rhine can be found in Bakels (2009). On premodern grain storage, see Sigaut (1980). Landscape transformations are central to questions of arable agriculture, which is broadly surveyed in Loveluck (2013). This is more broadly connected to climate, as Cheyette (2008) summarises. The connection between land, aristocracies, and peasants has been central to questions of arable land in Francia, for which see Banaji (2009). Various publications have examined the broader late antique agricultural countryside, such as Alfonso (2007), Grey (2012), and Chavarría Arnau and Lewit (2004) on archaeology. An overview of the late antique period in Gaul specifically can be found in van Ossel and Raynaud (2012); for Mediterranean France there is Durand and Leveau
The Environmental History of the Late Antique West
(2004), and Roller (2005) for northern Gaul. The transition from Late Roman to early medieval Gallic agriculture has been the focus of various studies, especially Crabtree (2010) on archaeology, as well as Henning (2004) on debates over continuities and discontinuities between Germanic and Roman systems. Regional studies proliferate and can be found, for example, in Nice (1994), Louis (2004), Petit (2006) and Wilkin (2011), but similar studies can be found in other places. Specific studies on Merovingian Gaul have examined both material culture evidence and literary sources. Henning (1996) provides an overview alongside Reigniez (2004) for material culture, while diets and their impact on agriculture is noted in Pearson (2001). The Church was a significant factor in systems of arable agriculture, which Lebecq (2000) has discussed. Carolingian Francia has had extensive work, especially the connections between arable land and the transformation of economic systems. An overview of agriculture in this period can be found in Devroey and Wilkin (2014) along with broader economic discussions in Verhulst (2002), Innes (2009), and Sonnlechner (2004). The expansion of agriculture east of the Rhine increasingly becomes more important as well, as Meier (2011) has noted. Gaul/Francia: Bloch M. (1931) Les caractères originaux de l’histoire rurale française (Oslo 1931); Duby G. (1962) L’économie rurale et la vie des campagnes dans l’Occident médiéval (Paris 1962); Bakels C. (1980) The Western European Loess Belt: Agrarian History, 5300 BC–AD 1000. (New York 2009); Sigaut F. (1980) “Significance of underground storage in traditional systems of grain production”, Developments in Agricultural Engineering 1 (1980) 3–14; Loveluck C. (2013) Northwest Europe in the Early Middle Ages, c. AD 600–1150: a Comparative Archaeology (Cambridge 2013); Cheyette F. L. (2008) “The disappearance of the ancient landscape and the climatic anomaly of the Early Middle Ages: a question to be pursued”, Early Medieval Europe 16.2 (2008) 127–65; Banaji J. (2009) “Aristocracies, peasantries and the framing of the Early Middle Ages”, Journal of Agrarian Change 9.1 (2009) 59–91; Alfonso I. (2007) “Exploring national historiographies of the medieval countryside: an introduction”, in The Rural History of Medieval European Societies: Trends and Perspectives, ed. I. Alfonso (Turnhout 2007) 1–19; Grey C. (2012) “Concerning rural matters”, in The Oxford Handbook of Late Antiquity, ed. S. F. Johnson (Oxford 2012) 625–66; Chavarría Arnau A. and Lewit T. (2004) “Archaeological research on the late antique countryside: a bibliographical essay”, in Recent Research on the Late Antique Countryside, edd. W. Bowden et al. (Leiden 2004) 3–51; Ossel, von P. and Raynaud C. (2012) “L’économie rurale et les productions en gaule durant l’antiquité tardive: moyens et techniques de production”, Antiquité Tardive 20 (2012) 151–59; Durand A. and Leveau P. (2004) “Farming in Mediterranean France and rural settlement in the Late Roman and Early Medieval periods: the contribution from archaeology and environmental sciences in the last twenty years (1980–2000)”, in The Making of Feudal Agricultures?, edd. M. Barceló and F. Sigaut (Leiden 2004) 177–253; Roller O. (2005) “Die Landwirtschaft der Spätantike in Obergermanien und dem Ostteil der Gallia Belgica”, in Imperium Romanum: Römer, Christen, Alamannen—Die Spätantike am Oberrhein, edd. M. Geiberger et al. (Stuttgart 2005) 146–49; Crabtree P. J. (2010) “Agricultural innovation and socio-economic change in early medieval Europe: evidence from
35 Britain and France”, World Archaeology 42.1 (2010) 122–36; Henning J. (2004) “Germanisch-romanische Agrarkontinuität und diskontinuität im nordalpinen Kontinentaleuropa—Teile eines Systemwandels? Beobachtungen aus archäologischer Sicht”, in Akkulturation: Probleme einer germanisch-romanischen Kultursynthese in Spätantike und frühem Mittelalter, edd. D. Hägermann et al. (Berlin 2004) 396–435; Nice A. (1994) “L’habitat Mérovingien de GoudelancourtLes-Pierrepont (Aisne). Aperçu provisoire d’une unité agricole et domestique des VIe et VIIe siècles”, Revue Archéologique de Picardie 1–2 (1994) 21–63; Louis E. (2004) “A de-Romanised landscape in northern Gaul: the Scarpe Valley from the 4th to the 9th century A.D.”, in Recent Research on the Late Antique Countryside, edd. W. Bowden et al. (Leiden 2004) 479–504; Petit C. (2006) “Géoarchéologie du site antique de Molesme en Vallée de Laigne (Côte-d’Or): mise en évidence de l’impact anthropique sur la sédimentation alluviale”, Gallia 63 (2006) 263–81; Wilkin A. (2011) “Some aspects of productivity in early medieval Europe: the case of eastern Belgium”, in Growth and Stagnation in European Historical Agriculture, edd. M. Olsson and P. Svensson (Turnhout 2011) 35–56; Henning J. (1996) “Landwirtschaft der Franken”, in Die Franken, Wegbereiter Europas: vor 1500 Jahren: König Chlodwig und Seine Erben, ed. A. Wieczorek (Mainz 1996) 774–85; Reigniez P. (2004) “Histoire et techniques: l’outil agricole dans la periode du Haut Moyen-Age (Ve–Xe siècle)”, in The Making of Feudal Agricultures?, edd. M. Barceló and F. Sigaut (Leiden 2004) 33–120; Pearson K. (2001) “Salic law and barbarian diet”, in Law, Society, and Authority in Late Antiquity, ed. R. W. Mathisen (Oxford 2001) 272–85; Lebec, S. (2000) “The role of the monasteries in the systems of production and exchange of the Frankish world between the seventh and the beginning of the ninth centuries”, in The Long Eighth Century, edd. I. L. Hansen and C. Wickham (Leiden 2000) 121–48. Carolingian Francia: Devroey J-P. and Wilkin A. (2014) “Die Landwirtschaft in der Karolingerzeit”, in Karl der Grosse, ed. F. Pohle (Dresden 2014) 86–93; Verhulst A. (2002) The Carolingian Economy (Cambridge 2002); Innes M. (2009) “Framing the Carolingian economy”, Journal of Agrarian Change 9.1 (2009) 42–58; Sonnlechner C. (2004) “The establishment of new units of production in Carolingian times: making early medieval sources relevant for environmental history”, Viator: Medieval and Renaissance Studies 35 (2004) 21–48; Meier T. (2011) “A farewell to the market! Constructing a Carolingian subsistence economy east of the Rhine”, in Processing, Storage, Distribution of Food: Food in the Medieval Rural Environment., edd. J. Klápště and P. Sommer (Turnhout 2011) 285–300.
Other Plant Cultivation Grain has received the greatest scholarly attention of the famed Mediterranean triad: grains, wine, and olive oil. Nevertheless, significant historiography has developed around wine and olive oil, even if late antique material is often peripheral to a classical focus: Rossiter (2008). Studies frequently rely on preserved amphorae, but more recently have looked to the machinery of production: Marzano (2013). The late antique florescence of African oleiculture has attracted particular attention; African oil amphorae increase in late antique Rome’s archaeological record: Graham (2014). Rome’s late reliance on African oil has yet to be fully explained, but some scholarship points to the subsidised distribution of African wheat, which eased the northward passage of other products across the Mediterranean: Wickham (1998). Archaeology hints at the effects of Vandal conquests on the oil trade of North Africa, with African amphorae becoming less common in Italy, and
36 correspondingly more common at Marseille: Loseby (1992). Graham (2014) argues that the breakdown of Mediterranean connectivity made the conspicuous consumption of oil by late antique popes (for illumination rather than food) a strong marker of status. In Iberia, the dating of wine and oil presses suggests a decline in large capital investments after the 1st c. AD, but with a slight resurgence in Late Antiquity from the end of the 3rd c.: Brun (2003); Marzano (2013). The use of vegetables, herbs, and spices in Late Antiquity and the Early Middle Ages has received less scholarly attention. The importance of woodland products, such as chestnuts, has recently been emphasised: Squatriti (2013). The diffusion of plant species from East to West, alongside Arab expansion, has been a topic of historiographic interest since Watson (1977). On archaeobotanical evidence for the cultivation and diffusion of various species, both local and exotic, see Livarda (2011) and Van der Veen (2008), (2014), (2015). Though broad in chronological and geographic scope, the essays in Chevalier et al. (2014) provide an ambitious interdisciplinary model, surveying the diverse relationships between plants and people, using archaeobotanical, ethnographic, and documentary evidence. Oleiculture and viticulture: Graham B. (2014) Profile of a Plant: the Olive in Early Medieval Italy, 400–900 CE (Ph.D. diss., Univ. of Michigan 2014); Rossiter J. (2008) “Wine-making after Pliny: viticulture and farming technology in late antique Italy”, in Technology in Transition: AD 300–650 (Leiden 2008) 93–118; Amouretti M.-C. and Brun J.-P. (1993) edd. La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991) (Athens 1993); Brun, J.-P. (2003) Le vin et l’huile dans la Méditerranée antique: viticulture, oléiculture et procédés de transformation (Paris 2003); Hitchner R. B. (1993) “Olive production and the Roman economy: the case for intensive growth in the Roman empire”, in La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991), edd. M.-C. Amouretti, and J.-P. Brun (Athens 1993) 499–508; Mattingly D. (1988) “Oil for export? A Comparison of Libyan, Spanish and Tunisian olive oil production in the Roman empire”, JRA 1 (1988) 33–56; Mattingly D. (1991) “Fruits of empire— the production of olive oil in Roman Africa”, Research & Exploration 7. 1 (1991) 36–55; Loseby S. (1992) “Marseille: a late antique success story?”, JRS 82 (1992) 165–85; Tchernia A. (1993) “Le vignoble italien du Ier siècle avant notre ère au IIIe siècle de notre ère: répartition et évolution”, in La production du vin et de l’huile en Méditerranée: actes du Symposium International (Aix-en-Provence et Toulon, 20–22 novembre 1991), edd. M.-C. Amouretti and J.-P. Brun (Athens 1993) 283–96; Lewit T. (2009) “Pigs, presses and pastoralism: farming in the fifth to sixth centuries AD”, Early Medieval Europe 17.1 (2009) 77–91; Lewit T. (2012) “Oil and wine press technology in its economic context: screw presses, the rural economy and trade in Late Antiquity”, Antiquité Tardive 20 (2012) 137–149; Keay S. (1984) Late Roman Amphorae in the Western Mediterranean, a Typology and Economic Study: The Catalan Evidence (BAR-IS 196.2) (Oxford 1984); Peña J. (1998) “The mobilization of state olive oil in Roman Africa: the evidence of late 4th c. ostraca from Carthage”, in Carthage Papers: The Early Colony’s Economy, Water Supply, a Public Bath, and the Mobilization of State Olive Oil, edd. J. Peña et al. (JRA Supplement 28) (Portsmouth, Rhode
Eisenberg et al. Island 1998) 116–238; Canetti L. (2007) “Olea sanctorum”, in Olio e vino nell’alto medioevo (Centro italiano di studi sull’alto medioevo. Atti delle settimane 54) (Spoleto 2007) 1335–1415; Martin A. (2008) “Imports at Ostia in the imperial period and Late Antiquity: the amphora evidence from the DAI-AAR excavations”, MAAR supplementary vol. 6 (2008) 105–18; Marzano A. (2013) “Agricultural production in the hinterland of Rome: wine and olive oil”, and “Capital investment and agriculture: multi-press facilities from Gaul, the Iberian Peninsula, and the Black Sea Region”, in The Roman Agricultural Economy: Organization, Investment, and Production, edd. A. Bowman and A. Wilson (Oxford 2013) 85–142; Bourgeon O. (2017) “Baetican olive-oil trade under the late empire: new data on the production of Late Roman amphorae (Dressel 23) in the Lower Genil Valley”, JRA 30 (2017) 517–29; Ruas M.-P. (2005) “Aspects of early medieval farming from sites in Mediterranean France”, Vegetation History and Archaeobotany 14 (2005) 400–415. Other: Squatriti P. (2013) Landscape and Change in Early Medieval Italy: Chestnuts, Economy, and Culture (New York 2013); Watson A. (1974) “The Arab agricultural revolution and its diffusion, 700–1100”, Journal of Economic History 34.1 (1974) 8–35; Van der Veen M. et al. (2008) “New food plants in Roman Britain: dispersal and social access”, Environmental Archaeology 13 (2008) 11–36; Van der Veen M. and Morales J. (2015) “The Roman and Islamic spice trade: new archaeological evidence”, Journal of Ethnopharmacology 167 (2015) 54–63; Van der Veen M. and Livarda A. (2008) “Social access and dispersal of condiments in northwest Europe from the Roman to the Medieval Period”, Vegetation History and Archaeobotany 17 (2008) 201–209; Livarda A. (2011) “Spicing up life in northwestern Europe: exotic food plant imports in the Roman and Medieval World”, Vegetation History and Archaeobotany 20.2 (2011) 143–64; Chevalier A. et al. (2014) edd. Plants and People: Choices and Diversity through Time (Oxford 2014).
Animals in Agriculture and Culture Animals were not only proxies for, but also participants in, environmental history, and to late antique persons they were especially charismatic partners. Writers and preachers liked to think about animals’ features and behaviours as attestations to the work of God, and as metaphors and lessons for humans’ own conduct. This intellectual harvesting of animals was of course concurrent with a very material use of them, domesticated animals in particular (for meat, dairy, wool, parchment, eggs, ploughing, carting, riding, guarding, herding, hunting, and mousing) but also the wild creatures of the grasslands, desert, woods, waters, and skies. Humans were aware of these forms of dependency, as research is just beginning to show, and even in the exegetical traditions of rabbinic Judaism and Christianity there was a distinct sense that animals were constantly shaping human lives and ideas, and not just the other way around. But for now, what Taylor (2018) said of the Middle Ages is also true of Late Antiquity: scholarship has until only recently tended to frame animals as the objects or instruments of human action, rather than asking how animals themselves made a difference in their environments. There are exceptions, such as Shaw (1979) on camels in North Africa, Pluskowski (2006) on wolves in northern and eastern Europe, Pastoureau (2007) on bears
The Environmental History of the Late Antique West
on the continent, and Kreiner (2017) on pigs in Gaul. Salvadori (2017) and O’Connor (2017) have written about the involvement of animals in urban ecologies, though the subject has received considerably more attention from scholars of the High and Late Middle Ages—such as Geltner (2019), Camphuijsen and Coomans (2015), Jørgensen (2013), Ruhland (1999)—largely because animals became more of a problem later on. This was as cities became more densely populated and less ‘green’, in contrast to the relatively open and agrarian towns of Late Antiquity and the Early Middle Ages, on which see, for example, Arthur (2002), and the forthcoming work by Caroline Goodson. Substantial advances in zooarchaeology and settlement archaeology since the 1980s are providing an increasingly sophisticated body of material to help assess how animal husbandry was embedded in material and cultural landscapes. Even on the evidence of bones and teeth alone, zooarchaeologists have pointed to changes in breeding, herd demography, animal uses (whether for labour, meat, or ‘secondary’ products), quality or extent of pasture, animal diet, living conditions, and even animal travel. Much has also been said about late antique changes in animal height and size, particularly of cattle: Audoin-Rouzeau (1998); Kron (2002); MacKinnon (2010); Duval et al. (2013); Holmes (2014); Rizzetto et al. (2017). And when excavators use mesh screens for flotation and sieving, they will almost invariably find fragments of birds, amphibians, reptiles, rodents, fish, and snails. These wild animals can be a valuable indication of local environments, given what we know about the habitats, behaviours, and preferences of their modern counterparts, although these features sometimes change over time. Beyond the bones, when faunal evidence is considered as part of a package of data from a given site—such as seeds, charcoal, pollen, soils and sediment, ceramics and other manufactured products, artisanal spaces, ditches and fencing, storage facilities, housing—the potential for environmental history is much greater. Exemplary studies of this sort include, for example, the projects of Flixborough (Dobeny (2007)), DevelierCourtételle (Guélat et al. (2008)), Miranduolo (Valenti (2008)), Villa Magna (Fentress et al. (2016)), San Giovanni di Ruoti (MacKinnon (2002)), Faragola (Volpe (2009)), Althiburos (Kallala and Sanmartí (2011–16)), Jerba (Fentress et al. (2009)), and Fazzān (Mattingly (2013)). They focus on a site or even region over the long durée, over years of fieldwork, and their directors make a concerted effort to analyse (and publish) the many pieces of their projects as an integrated whole. This obviously requires considerable resources and time, and for some sites such an investigation is impossible. But approaches like these are creating a valuable foundation for asking
37 how animals were part of the dynamic, interdependent, and diverse histories of late antique environments. Animals: Studies cited above listed first; the rest are organised geographically, north to south. Ecocritical views of animals: Spittler J. (2008) Animals in the Apocryphal Acts of the Apostles: the Wild Kingdom of Early Christian Literature (Tübingen 2008); Crane S. (2013) Animal Encounters: Contacts and Concepts in Medieval Britain (Philadelphia 2013) 11–40; Neiss R. R. (2018) “When species meet in the Mishnah”, Ancient Jew Review, 8 May 2018: http://www.ancientjewreview.com/ articles/2018/5/8/when-species-meet-in-the-mishnah; Wasserman M. B. (2017) Jews, Gentiles, and Other Animals: the Talmud after the Humanities (Philadelphia 2017); Berkowitz B. (2015) “Animal”, in Late Ancient Knowing: Explorations in Intellectual History, edd. C. M. Chin and M. Vidas (Berkeley 2015) 36–57; Buell D. K. (2014) “The microbes and pneuma that therefore I am”, in Divinanimality: Animal Theory, Creaturely Theology, ed. S. D. Moore (New York 2014) 63–87. Animal influence and agency: Taylor A. L. (2018) “Where are the wild things? Animals in western medieval European history”, History Compass 16 (2018): https://doi.org/10.1111/hic3.12443; Shaw B. D. (1979) “The camel in Roman North Africa and the Sahara: history, biology, and human economy”, Bulletin de l’Institut Fondamental d’Afrique Noire 41, series B, no. 4 (1979) 663–721; Pluskowski A. (2006) Wolves and the Wilderness in the Middle Ages (Woodbridge 2006); Pastoureau M. (2007) L’ours: histoire d’un roi déchu (Paris 2007); Kreiner J. (2017) “Pigs in the flesh and fisc: an early medieval ecology”, PastPres 236 (2017) 3–42. Cities and animals: Salvadori F. (2017) “Animals in Italian medieval towns: from Late Antiquity to the Middle Ages”, in Animaltown: Beasts in Medieval Urban Space, edd. A. M. Choyke and G. Jaritz (Oxford 2017) 129–45; O’Connor T. (2017), “Animals in urban life in Medieval to Early Modern England”, in The Oxford Handbook of Zooarchaeology, ed. U. Albarella (Oxford 2017) 215–29; Geltner G. (2019) Roads to Health: Infrastructure and Urban Wellbeing in Later Medieval Italy (Philadelphia 2019); Camphuijsen F. and Coomans J. (2015) “De middeleeuwse stad en zijn varkens”, Madoc: Tijdschrift over de Middeleeuwen 28 (2015) 140–48; Jørgensen D. (2013) “Running amuck? Urban swine management in Late Medieval England”, Agricultural History 87 (2013) 429–51; Ruhland F. (1999), “Schweinehaltung in und vor der Stadt”, in Nürnberg: Archäologie und Kulturgeschichte, edd. B. Friedel and C. Frieser (Büchenbach 1999); Arthur P. (2002) Naples: From Roman Town to City-State (London 2002). Osteoarchaeology and environment: European Journal of Archaeology 20.3 (special issue 2017) Animal Husbandry in the Western Roman Empire: a Zooarchaeological Perspective, edd. S. ValenzuelaLama and U. Albarella; Quaternary International 346 (special issue 2014) = Agrarian Archaeology in Early Medieval Europe, ed. J. A. Quirós Castillo; Debates de Arqueología Medieval 3 (special issue 2013) = La ganadería como tema central; Ascough P. et al. (2014) “Stable isotopic (δ13C and δ15N) characterization of key faunal resources from Norse Period settlements in North Iceland”, Journal of the North Atlantic 7 (special issue 2014) 25–42; Wigh B. (2001) Animal Husbandry in the Viking Age Town of Birka and its Hinterland (Birka Studies 7) (Stockholm 2001); Crabtree P. J. (1987) West Stow, Suffolk: Early AngloSaxon Animal Husbandry (East Anglian Archaeology Report 47) (Ipswich 1987); McCormick F. and Murray E. (2007) Excavations at Knowth 3: Knowth and the Zooarchaeology of Early Christian Ireland (Dublin 2007); van der Jagt I. M. M. et al. (2012) “An insight into animal exchange in early medieval Oegstgeest: a combined archaeozoological and isotopic approach”, in A Bouquet of Archaeozoological Studies: Essays in Honour of Wietske Prummel (Groningen 2012) 139–50; Kootker L. M. et al. (2016) “Strontium isoscapes in the Netherlands: spatial variations in 87sr/86sr as a proxy for palaeomobility”, JAS: Reports 6 (2016) 1–16; Grau-Sologestoa I. (2015) The Zooarchaeology of Medieval Álava in its Iberian Context (Oxford 2015); Yvinec J.-H. and Barme M. (forthcoming) “Livestock and the
38 early medieval diet in northern Gaul”, in The Oxford Handbook of the Merovingian World, edd. B. Effros and I. Moreira (Oxford forthcoming 2019); Jiménez-Camino R. et al. (2010) “¿Continuidad o cambio en la dieta entre la población bizantina y paleoandalusí? Aproximación a partir de registro faunístico de dos intervenciones arqueológicas en Algeciras”, in Espacios urbanas en el occidente Mediterráneo (s. VI–VIII), ed. A. García (Toledo 2010) 153–64; García-Blanco V. and Vila S. (2006) “Restos animales y vegetales del yacimiento visigodo de Prado de los Galápagos, interpretación ambiental”, in La investigación arqueológica de la época visigoda en comunidad de Madrid, ed. J. Morín de Pablos (Zona Arqueológica 8) (Alcalá de Henares 2006) vol. 3: 962–72; Buglione A. (2007) “People and animals in northern Apulia from Late Antiquity to the Early Middle Ages: some considerations”, in Breaking and Shaping Beastly Bodies: Animals as Material Culture in the Middle Ages, ed. A. Pluskowski (Oxford 2007) 189–215; MacKinnon M. (2004) Production and Consumption of Animals in Roman Italy: Integrating the Zooarchaeological and Textual Evidence (JRA Supplementary Series 54) (Portsmouth, Rhode Island 2004); Ikeguchi M. (2017) “Beef in Roman Italy”, JRA 30 (2017) 7–37; Buglione A., De Venuto G., and Volpe G. (2016) “Agricoltura e allevamento nella Puglia settentrionale tra età romana e Medioevo: il contributo delle bioarcheologie”, MÉFRA 128/2 (2016): https://mefra.revues. org/3475; MacKinnon M. (2010), “‘Romanizing’ ancient Carthage: evidence from zooarchaeological remains”, in Anthropological Approaches to Zooarchaeology: Colonialism, Complexity, and Animal Transformations, edd. D. Campana et al. (Oxford 2010) 168–77; Humphrey J. H. (1981) ed. Excavations at Carthage 1977 Conducted by the University of Michigan, vol. 6 (Ann Arbor 1981); Barker G. (1977) “Economic life at Berenice: the animal and fish bones, marine molluscs and plant remains”, in Excavations at Sidi Khrebish, Benghazi (Berenice), vol. 2, ed. J. Lloyd (Tripoli 1977). Animals changing size: Audoin-Rouzeau F. “Cheptel antique, cheptel médiéval: mutations ou innovations?”, in L’innovation technique au Moyen Âge. Actes du VIe congrès international d’archéologie médiévale (Caen 1998) 30–34; Kron G. “Archaeozoological evidence for the productivity of Roman livestock farming”, Münstersche Beiträge zur Antike Handelsgeschichte 21.2 (2002) 53–73; Kron G. “Animal husbandry”, in The Oxford Handbook of Animals in Classical Thought and Life, ed. G. Campbell (Oxford 2014) 109–35; MacKinnon M. “Cattle ‘breed’ variation and improvement in Roman Italy: connecting the zooarchaeological and ancient textual evidence”, WorldArch 42.1 (2010) 55–73; Forest V. and Rodet-Belarbi I. (2002) “À propos de la corpulence des bovines en France durant les périodes historiques”, Gallia 59 (2002) 273–306; Colominas L. et al. (2014) “The impact of the Roman empire on animal husbandry practices: study of the changes in cattle morphology in the northeast of the Iberian peninsula through osteometric and ancient DNA analyses”, Archaeological and Anthropological Sciences 6.1 (2014) 1–16; Crabtree P. (2014) “Animal husbandry and farming in East Anglia from the 5th to the 10th centuries CE”, Quaternary International 346 (2014) 102–108; Duval C. et al. “Morphological changes in domestic cattle in Gaul, from the second century BCE to the fifth century AD: diversity of herds in the Seine Valley (France) and northern Gaul”, JAS 40 (2013) 3977–90; Holmes M. (2014) “Does size matter? Changes in the size of animals throughout the English Saxon period (AD 450–1066)”, JAS 43 (2014) 77–90; Rizzetto M. et al. (2017) “Livestock changes at the beginning and end of the Roman period: acculturation, adaptation, and ‘improvement’”, EJA 20 (2017) 535–56. Settlement studies integrating zooarchaeological evidence: Dobney K. et al. (2007) Farmers, Monks, and Aristocrats: the Environmental Archaeology of Early Medieval Flixborough (Oxford 2007) (vol. 3); Guélat M. et al. (2008) Develier-Courtételle: un habitat rural mérovingien, vol. 4: Environnement et exploitation du terroir (Cahier d’archéologie jurassienne 16) (Porrentruy 2008): http://w3.jura.ch/ services/oph/sar/Publications/Publications.htm; Valenti M. (2008) Miranduolo in Alta Val di Merse (Chiusdino, SI): archeologia su un sito
Eisenberg et al. di potere del medioevo toscano (Florence 2008); Fentress E., Goodson C., and Maiuro M. (2016) Villa Magna: an Imperial Estate and its Legacies. Excavations 2006–10 (London 2016); MacKinnon M. R. (2002) The Excavations of San Giovanni di Ruoti, vol. 3: The Faunal and Plant Remains (Toronto 2002); Volpe G. and Turchiano M. (2009) edd. Faragola 1: un insediamento rurale nella Valle del Carapelle (Bari 2009); Kallala N. and Sanmartí J. (2011–16) Althiburos, 2 vols. (Tarragona 2011–16); Fentress E., Drine A., and Holod R. (2009) An Island through Time: Jerba Studies, vol. 1: The Punic and Roman Periods (Portsmouth, Rhode Island 2009) (vol. 2 is in progress); Mattingly D. J. (2013) The Archaeology of Fazzān, vol. 4: Survey and Excavations at Old Jarma (Ancient Garama) (Tripoli-London 2013); Lucas G. (2009) Hofstaðir: Excavations of a Viking Age Feasting Hall in North-Eastern Iceland (Reykjavík 2009); Seaver M. (2016) Meithal: the Archaeology of Lives, Labours and Beliefs at Raystown, Co. Meath (Dublin 2016); Hardy A., Charles B. M., and Williams R. J. (2007) Death and Taxes: the Archaeology of a Middle Saxon Estate Centre at Higham Ferrers, Northamptonshire (Oxford 2007); Hey G. (2004) Yarnton: Saxon and Medieval Settlement and Landscape (Oxford 2004); Ettel P. (2001) Karlburg, Rosstal, Oberammerthal: studien zum frühmittelalterlichen Burgenbau in Nordbayern, 3 vols. (Rahden 2001); Porte P. (2011) Larina de l’antiquité au moyen âge, vol. 1: Études archéologiques et historiques en Nord-Isère (Biarritz 2011); Grassi F. (2013), L’insediamento medievale nelle Colline Metallifere (Toscana, Italia): il sito minerario di Rochette Pannocchieschi dall’VIII al XIV secolo (Oxford 2013); Francovich R. and Valenti M. (2007) Poggio Imperiale a Poggibonsi: il territoria, lo scavo, il parco (Milan 2007); Arthur P. and Bruno B. (2009) Apigliano: un villagio bizantino e medioevale in Terra d’Otranto. L’ambiente, il villaggio, la popolazione (Galatina 2009); Arthur P. Imperiale M. L., and Tinelli M. (2015) Apigliano: un villaggio bizantino e medievale in Terra d’Otranto. I reperti (Lecce 2015); Malalana Ureña A., Barroso Cabrera R., and Morín de Pablos J. (2012), La Quebrada II: un habitát de la tardoantigüedad al siglo XI. La problemática de los “silos” en la Alta Edad Media hispana (Alcázar de San Juan 2012); Sonte D. L. Mattingly D. J., and Ben Lazreg N. (2011) Leptiminus (Lamta) Report No. 3: the Field Survey (Portsmouth, Rhode Island 2011).
Supplying Cities The Roman Mediterranean was a landscape of cities, whose demographic size required the reshaping of the surrounding environment to supply the necessary food, water, and other supplies. Throughout the late antique west, urban centres transformed, although this change differed depending on various factors, including political systems, socio-economic changes, religious and cultural needs, and climate changes. Wickham (2005) and McCormick (2001) are two key existing syntheses regarding the transformation of urbanism and its connectivity to nearby environments. Likewise, Cleary (2013) highlights the differences between various regions in Gaul and Iberia already from the 3rd c., while Liebeschuetz (2001), in contrast, discusses the breakdown of urban centres and their surrounding networks. Useful surveys of various parts of the empire and the surrounding urban landscape are found in the volumes edited by Christie and Loseby (1996) and Christie (2004). The place of the Church in the changing supply networks should not be overlooked, which has been noted in Wood (2013) and (2018). There is also a useful summary of overviews in McMahon and Sargent (2018).
The Environmental History of the Late Antique West Supplying the cities: Wickham C. (2005) Framing the Early Middle Ages: Europe and the Mediterranean 400–800. (Oxford 2005); McCormick M. (2001) Origins of the European Economy: Communications and Commerce, AD 300–900. (Cambridge 2001); Cleary A. S. (2013) The Roman West, AD 200–500: an Archaeological Study. (Cambridge 2013); Liebeschuetz J. H. W. G. (2001) Decline and Fall of the Roman City (Oxford 2001); Christie N. and Loseby S. T. (1996) edd. Towns in Transition: Urban Evolution in Late Antiquity and the Early Middle Ages. (Aldershot 1996); Christie N. (2004) ed. Landscapes of Change: Rural Evolutions in Late Antiquity and the Early Middle Ages (Aldershot-Burlington, Vermont 2004); Wood I. (2013) “Entrusting western Europe to the Church, 400–750”, Transactions of the Royal Historical Society 23 (2013) 37–73; Wood I. (2018) The Transformation of the Roman West (Leeds 2018); McMahon L. and Sargent A. (2018) “Environmental history of the late antique eastern Mediterranean: a bibliographic essay”, in this volume.
North Africa North Africa was a booming economy in Late Antiquity with Carthage at its economic heart. Before the Vandal conquest it supplied Rome with annual grain supply, but even afterward the economy continued to produce significant quantities of foodstuffs, which then shifted to supply local and regional urban networks. Conant (2012) discusses the general economic situation, Bockmann (2013) delves into the archaeological background of the region, and Tedesco (2011) connects numismatic and economic evidence together. Carthage has had significant excavations, which are edited by Humphrey (1976) in seven volumes, with detailed discussions of various features of the urban networks; excavations funded by the British Academy also took place: Hurst et al. (1984–94). Dossey (2010) has synthesised smaller urban networks across North Africa, while Leone (2013) discusses the cultural and religious changes to the urban fabric. North Africa: Conant J. (2012) Staying Roman: Conquest and Identity in Africa and the Mediterranean, 439–700. (Cambridge 2012); Bockmann R. (2013) Capital Continuous: a Study of Vandal Carthage and Central North Africa from an Archaeological Perspective (Wiesbaden 2013); Tedesco P. (2011) “Economia e moneta nell’Africa vandalica”, Annali (Istituto Italiano di Numismatica) 57 (2011) 115–38; Humphrey J. (1976) ed. Excavations at Carthage: Conducted by the University of Michigan. (Ann Arbor 1976); Hurst H. R. et al. (1984–94) Excavations at Carthage: The British Mission, 4 vols. (Sheffield 1984–94); Dossey L. (2010) Peasant and Empire in Christian North Africa (Berkeley 2010); Leone A. (2013) The End of the Pagan City: Religion, Economy, and Urbanism in Late Antique North Africa (Oxford 2013).
Italy and Rome The supplying of Rome with food, water, and supplies has been well-researched in the Roman and late antique period. Rome was certainly an outlier given its large size, and depended upon foodstuffs from a broader region than other cities. For upper and lower estimates of its population respectively see, Scheidel (2004) and Storey (1997). Fundamental to any study of food supply for Rome are Durliat (1990), Sirks (1991) and Garnsey (1988).
39 Johnson (2012) surveys the changes to the urban fabric, economics, and the food supply more broadly. In the late antique period, research has examined the changes to food supplies due to demographic and political instability in Barnish (1987), the essays edited by Harris and Arce (1999), and recently Linn (2012) after the Vandal conquest of North Africa. The knowledge of classical Rome’s water supply and sanitation has been studied in detail in Koloski-Ostrow (2015) and its water supply and aqueducts in Hodge (1992) and Coates-Stephens (1998). On the management of Rome’s water supply see Purcell (1996). For comparisons between developments in Rome and other cities see Bauer (1996). Water supply in the late antique period more broadly has been extensively studied in Squatriti (1998). Other Italian cities, such as Ravenna, and their development based on connections to nearby water sources, have been researched in Squatriti (1992). Italy and Rome: Scheidel W. (2004) “Human mobility in Roman Italy I: the free population”, JRS 94 (2004) 1–26; Storey G. (1997) “The population of ancient Rome”, Antiquity (1997) 966–87; Durliat J. (1990) De la ville antique à la ville byzantine: le problème des subsistances (Rome 1990); Sirks A. J. B. (1991) Food for Rome: the Legal Structure of the Transportation and Processing of Supplies for the Imperial Distributions in Rome and Constantinople (Amsterdam 1991); Garnsey P. (1988) Famine and Food Supply in the Graeco-Roman World: Responses to Risk and Crisis. (Cambridge 1988); Johnson P. (2012) Economic Evidence and the Changing Nature of Urban Space in Late Antique Rome. (Barcelona 2012); Barnish S. J. B. (1987) “Pigs, plebeians and potentes: Rome’s economic hinterland, c. 350–600 AD”, PBSR 55 (1987) 157–85; Harris W. and Arce J. (1999) edd. The Transformations of Vrbs Roma in Late Antiquity (Portsmouth, Rhode Island 1999); Linn J. (2012) “The Roman grain supply, 442–455”, Journal of Late Antiquity 5.2 (2012) 298–321; Koloski-Ostrow A. (2015) The Archaeology of Sanitation in Roman Italy: Toilets, Sewers, and Water Systems (Chapel Hill, North Carolina 2015); Hodge A. T. (1992) Roman Aqueducts and Water Supply. (London 1992); Coates-Stephens R. (1998) “The walls and aqueducts of Rome in the Early Middle Ages, AD 500–1000”, JRS 88 (1998) 166–78; Purcell N. (1996) “Rome and the management of water: environment, culture and power”, in Human Landscapes in Classical Antiquity. Environment and Culture, edd. G. Shipley and J. Salmon (London 1996) 180–212; Bauer F. (1996) Stadt, Platz und Denkmal in der Spätantike: Untersuchungen zur Ausstattung des Öffentlichen Raums in den Spätantiken Städten Rom, Konstantinopel und Ephesos (Mainz am Rhein 1996); Squatriti P. (1998) Water and Society in Early Medieval Italy: AD 400–1000 (Cambridge 1998); Squatriti P. (1992) “Marshes and mentalities in early medieval Ravenna”, Viator: Medieval and Renaissance Studies 23 (1992) 1–16.
Gaul Many of late antique Gaul’s urban centres have received extensive examination, excavation, and analysis, with the results showing that northern regions were more susceptible to change, especially fortification, while the southern regions remained more secure. A good overarching comparative study is Cleary (2013), and for changes between Gaul and Rome, see Goodman (2007). The role of Christianity in this process is central, but differed in various contexts: see Guyon (2013) for a summary.
40 Southern Gaul continued largely unchanged through the political end of the Roman empire but, by the early 5th c., Arles, as the capital of the Gallic provinces, underwent significant changes to its hinterland and supply networks, as noted in Guyon and Heijmans (2013). Marseille serves as a useful case study to understand the transformation in supply networks around southern Gaul, which flourished in the 6th c., but shrank in the 7th and 8th c.: see respectively Loseby (1998) and (2000). Bachrach (2010) has used urban information from Bordeaux to extrapolate transformations in the urban fabric throughout all the cities of Gaul. The developments of emporia in northern Gaul have been the focus of various studies; for an overview see Verhulst (2000). Gaul: Goodman P. (2007) The Roman City and its Periphery: From Rome to Gaul. (New York 2007); Guyon J. (2013) “Les chefs lieux de cités de Gaule méridionale aux Ve et VIe siècles: un espace urbain en mutation”, in Gallien in Spätantike und Frühmittelalter: Kulturgeschichte einer Region, edd. S Diefenbach and G. M. Müller (Berlin 2013) 201–22; Guyon J. and M. Heijmans (2013) edd. L’antiquité tardive en Provence (IVe–VIe siècle): naissance d’une chrétienté (Arles 2013); Loseby S. T. (1998) “Marseille and the Pirenne thesis I: Gregory of Tours, the Merovingian kings and ‘un grand port’”, in The Sixth Century: Production, Distribution and Demand, edd. R. Hodges and W. Bowden (Leiden 1998) 203–29; Loseby S. T. (2000) “Marseille and the Pirenne thesis II: ‘ville morte’”, in The Long Eighth Century, edd. I. Hansen and C. Wickham (Leiden 2000) 167–93; Bachrach B. S. (2010) “The fortification of Gaul and the economy of the third and fourth centuries”, Journal of Late Antiquity 3.1 (2010) 38–64. Verhulst A. (2000) “Roman cities, emporia and new towns (sixth-ninth centuries)”, in The Long Eighth Century, edd. I. Hansen and C. Wickham (Leiden 2000) 105–20.
Iberia Iberia likewise underwent significant transformations over the course of Late Antiquity. Reynolds (2010) discusses the general trends, through ceramics, from the Late Roman to the Post Roman period. Local agents transformed many of these urban networks through their reshaping of the Roman past to meet new needs, as noted in Fernández (2017). The study of Iberia’s cities is examined regionally, with work such as Martínez Jiménez (2017) on the water supply of two cities. Iberia: Reynolds P. (2010) Hispania and the Roman Mediterranean, AD 100–700: Ceramics and Trade (London 2010); Fernández D. (2017) Aristocrats and Statehood in Western Iberia, c. 300–600 C.E. (Philadelphia 2017); Martínez Jiménez J. (2017) “Water supply in the Visigothic urban foundations of Eio (El Tolmo de Minateda) and Reccopolis”, in New Cities in Late Antiquity: Documents and Archaeology, ed. E. Rizos (Turnhout 2017) 233–46.
Britain The Romano-British urban supply networks have been explored in detail. The Roman water supply is discussed in Burgers (2001) and Stephens (1985). For a recent discussion of the grain supply to various urban centres, see the essays in Fulford and Holbrook (2015). On the end of
Eisenberg et al.
Late Roman urban networks, see the overview in Henig (2011). Britain’s urban infrastructure largely collapsed between the Roman and Early Medieval periods. Fleming (2012) has noted the end of metallurgical work and the reuse of existing metal supplies, until urban networks revived centuries later. Crabtree (1996) and Maddicott (2005) have noted the expansion of urban networks and the new roles they played in state formation during the Anglo-Saxon period. An overview of this process as a whole, is outlined in Pestell (2011). There was also an increase in agricultural production and emporia toward the end of the late antique period, for which see Hamerow (2007). There have also been some attempts to model grain supply for the long dureé, as seen in Smith and Kenwood (2011). Britain: Burgers A. (2001) The Water Supplies and Related Structures of Roman Britain (Oxford 2001); Stephens G. R. (1985) “Civic aqueducts in Britain”, Britannia 16 (1985) 197–208; Fulford M. and Holbrook N. (2015) edd. The Towns of Roman Britain: the Contribution of Commercial Archaeology since 1990 (London 2015); Henig M. (2011) “The fate of Late Roman towns”, in The Oxford Handbook of AngloSaxon Archaeology, edd. D. Hinton, S. Crawford, and H. Hamerow (Oxford 2011) 515–33; Fleming R. (2012) “Recycling in Britain after the fall of Rome’s metal economy”, PastPres 217 (2012) 3–45; Crabtree P. (1996) “Production and consumption in an early complex society: animal use in Middle Saxon East Anglia”, WorldArch 28.1 (1996) 58–75; Maddicott J. R. (2005) “London and Droitwich, c. 650–750: trade, industry and the rise of Mercia”, Anglo-Saxon England 34 (2005) 7–58; Pestell T. “Markets, emporia, WICS, and ‘productive’ sites: pre-Viking trade centres in Anglo-Saxon England”, in The Oxford Handbook of Anglo-Saxon Archaeology, edd. D. Hinton, S. Crawford, and H. Hamerow (Oxford, 2011) pp?; Hamerow H. “Agrarian production and the emporia of Mid Saxon England, ca. AD 650–850”, in The Heirs of the Roman West, ed. J. Henning (Berlin 2007) 219–32; Smith D. and Kenward H. (2011) “Roman grain pests in Britain: implications for grain supply and agricultural production”, Britannia 42 (2011) 243–62.
Landscapes Trees and Woodland Environmental historians and ecologists have often categorised eco-regions according to their native species, and this tendency has been particularly pronounced for the Mediterranean. Horden and Purcell (2000) offer maps illustrating the Mediterranean’s characteristic vegetation, especially the olive. Braudel defined the Mediterranean as the region where olives (but not palm trees) thrived in the wild, though Squatriti (2014) has pointed out that the region’s vegetative complexities make generalisation difficult. The extensive history of domesticates in the Mediterranean sometimes obscures the difference between native and introduced species: Zohary, Hopf, and Weiss (2012). A traditional assumption is that the region once supported extensive continuous woodland, and that its contemporary
The Environmental History of the Late Antique West
mosaic of ecosystems (often dominated by shrubby maquis and garrigue) is the result of anthropogenic degradation, through deforestation and subsequent erosion: Thirgood (1981). This assumption has been widely criticised, for example by Grove and Rackham (2003) and Davis (2007), but retains adherents: Hughes (2017). In Iberia, there is some evidence for late antique deforestation linked to grazing patterns: Esparaguerra and Martinez (2001). In western Europe more generally, Late Antiquity and the ‘Dark Ages’ have often been associated with forest renewal, as the reduced agricultural footprint of a declining population allowed woodlands to regenerate, around Naples; see Ermolli et al. (2014). In the English context, Hoskins (1955) envisioned a Post Roman landscape of regenerated forests, only partially cleared in the lowlands by Anglo-Saxons, starting in the mid 5th c. Williamson (2013) embodies the recent movement away from strict categorisations, arguing that prime agricultural land in early medieval England was usually intermixed with woodland, and that undisturbed forest was relatively rare. Hooke (2011) treats not only the uses and distribution of forests in the Anglo-Saxon period, but also their literary and cultural significance. While there is evidence for both forest regeneration and landscape continuity (or even deforestation), recent scholarship has noted how the persistence or growth of woodland does not indicate a lack of human involvement. For Italy, Squatriti (2013) and Mercuri et al. (2017) have emphasised the importance of woodlands in mixed subsistence regimes, with chestnuts particularly well-suited to Postclassical, silvo-pastoral regimes. Decreasing pressure on land in the 5th and 6th c. promoted the development of peasant economies well-suited to local ecologies, frequently exploiting the products of woodlands. The composition of forests could also shift depending on use, with the grazing of pigs and cattle favouring some species over others: Kreiner (2017). Trees and woodland: Wickham C. (1990) “European forests in the Early Middle Ages: landscape and land clearance”, in L’ambiente vegetale nell’alto medioevo (Centro italiano di studi sull’alto medioevo. Atti delle settimane 37) (Spoleto 1990) 479–545; Hooke D. (2011) Trees in Anglo-Saxon England (Suffolk 2011); C. Higounet (1966) ‘Les forêts de l’Europe occidentale du Ve au XIe siècle’, in Agricoltura e mondo rurale in Occidente nell’alto medioevo (Centro italiano di studi sull’alto medioevo. Atti delle settimane 13) (Spoleto 1966) 343–98; Montanari M. and Andreolli B. (1988) edd. Il bosco nel medioevo (Bologna 1988); Squatriti P. (2013) Landscape and Change in Early Medieval Italy: Chestnuts, Economy, and Culture (Cambridge 2013); Squatriti P. (2014) “The vegetative Mediterranean”, in A Companion to Mediterranean History, ed. P. Horden (Hoboken, New Jersey 2014) 26–41; Squatriti P. (2010) “Trees, nuts, and woods at the end of the first millennium: a case from the Amalfi coast”, in Ecologies and Economies in Medieval and Early Modern Europe: Studies in Environmental History for Richard C. Hoffmann (Leiden 2010) 22–43; Grove A. and Rackham O. (2003) The Nature of Mediterranean Europe: an Ecological History
41 (New Haven, Connecticut 2003); Thirgood J. (1981) Man and the Mediterranean Forest: a History of Resource Depletion (London 1981); Thirgood J. and Hughes J. D. (1982) “Deforestation in ancient Greece and Rome: a cause of collapse”, Ecologist 12 (1982) 196–208; Hughes J. D. (2017) “Ancient deforestation revisited”, Journal of the History of Biology 44 (2017) 43–57; Harris W. (2011) “Bois et déboisement dans la Méditerranée antique”, Annales 66.1 (2011) 105–40; Hoskins W. (1955) The Making of the English Landscape (London 1955); Williamson T. (2013) Environment, Society and Landscape in Early Medieval England (Rochester 2013); Zohary D., Hopf M., and Weiss E. (2012) Domestication of Plants in the Old World: the Origin and Spread of Cultivated Plants in West Asia, Europe, and the Nile Valley (New York, 4th edn. 2012); Davis D. (2007) Resurrecting the Granary of Rome: Environmental History and French Colonial Expansion in North Africa (Athens, Ohio 2007); Mercuri A. et al. (2017) “The late antique plant landscape in Sicily: pollen from the agro-pastoral Villa Del Casale”, Quaternary International (October 2017) https://doi.org/10.1016/j .quaint.2017.09.036 (in press); Bernard V. (1998) L’homme, le bois et la forêt dans la France du nord entre le mésolithique et le haut moyenâge (BAR-IS 733) (Oxford 1998). Ermolli R. et al. (2014) “The natural and cultural landscape of Naples (southern Italy) during the GraecoRoman and late antique periods”, Journal of Archaeological Science 42 (2014): 399–411; Allevato E. et al. (2012) “Persistence of the cultural landscape in Campania (southern Italy) before the AD 472 Vesuvius eruption: archaeoenvironmental data”, JAS 39 (2012) 399–406; Lewit T. (2009) “Pigs, presses and pastoralism: farming in the fifth to sixth centuries AD”, Early Medieval Europe 17.1 (2009) 77–91; Kreiner J. (2017) “Pigs in the flesh and fisc: an early medieval ecology”, PastPres 236 (2017) 3–42; Esparraguera G. and Martinez J. (2001) “Structuration du territoire dans le nord-est de l’Hispanie pendant l’antiquité tardive: transformation du paysage et dynamique du peuplement”, in Les campagnes de la Gaule à la fin de l’antiquité (Antibes 2001) 303– 29; Accorsi C. et al. (1996) “Holocene forest pollen vegetation of the Po plain: northern Italy (Emilia Romagna data)”, Alliona 24 (1996) 233–76; Andreolli B. (2002) “L’uso del bosco e degli incolti”, in Storia dell’agricoltura Italiana, edd. G. Pinto and M. Ambrosoli (Florence 2002) 123–44; Hemphill P. (1988) “De-forestation and re-forestation in a central Italian hinterland: land usage during and after the Roman occupation”, in First Millennium Papers. Western Europe in the First Millennium AD, edd. Jones R. et al. (Oxford 1988) 147–58; Janssen C. (1994) “Palynological indications for the extent of the impact of man during Roman times in the western part of the Iberian Peninsula”, in Evaluation of Land Surfaces Cleared from Forests in the Mediterranean Region during the Time of the Roman Empire, edd. B. Frenzel, L. Reisch, and M. M. Weiss (New York 1994) 13–22; Noyé G. (2001) “Economia e società nella Calabria bizantina”, in Storia della Calabria medievale, vol. 2: Culture, arti, tecniche, ed. A. Placanica (Reggio 2001) 579–655.
Uncultivated Resources Fish The historiography of fisheries in the late antique western Mediterranean is slim, although work on Roman fishing often incorporates late antique material (for example Marzano (2013)). The Postclassical period seems to have witnessed dramatic changes in fishing, aquaculture, and salting: Zerbini (2015). Roman law tended to view watercourses as public, but in Late Antiquity and the Early Middle Ages, historians have traced a shift toward privatised ownership and tightly controlled fishing rights: Squatriti (2002). As in other areas of environmental history, the study of rivers and
42 fisheries has moved toward interdisciplinary research, combining the expertise of historians with that of ecologists, hydrologists, and other scientists. Such approaches aim to develop fuller histories of riverine ecosystems as integrated “socio-ecological systems”: Haidvogl et al. (2015). Fishing: Haidvogl G. et al. (2015) “Historical ecology of riverine fish in Europe”, Aquatic Sciences 77 (2015) 315–24; Hoffmann R. (1993) “The protohistory of pike in western culture”, in The Medieval World of Nature: a Book of Essays (New York 1993) 61–76; Hoffmann R. (1996) “Economic development and aquatic ecosystems in medieval Europe”, AHR 101 (1996) 631–69; Hoffmann R. (2000) “Medieval fishing”, in Working with Water in Medieval Europe: Technology and Resource Use (Leiden 2000) 331–93; Squatriti P. (2002) Water and Society in Early Medieval Italy (New York 2002) 97–125; Enghoff I. (1999) “Fishing in the Baltic region from the fifth century B.C. to the sixth century A.D.: evidence from fish bones”, Archaeofauna 8 (1999) 41–85; Enghoff I. (2000) “Fishing in the southern North Sea region from the first to the sixteenth century A.D.: evidence from fish bones”, Archaeofauna 9 (2000) 59–132; Higginbotham J. (1997) Piscinae: Artificial Fishponds in Roman Italy (Chapel Hill, North Carolina 1997); Balon E. (1995) “Origin and domestication of the wild carp, cyprinus carpio: from Roman gourmets to the swimming flowers”, Aquaculture 129 (1995) 3–48; Marzano A. (2013) Harvesting the Sea: the Exploitation of Marine Resources in the Roman Mediterranean (Oxford 2013); Rousselle A. (1999) “Fish”, in Late Antiquity: a Guide to the Postclassical World (Cambridge, Mass. 1999) 450–51; Zug Tucci H. (1985) “Il mondo medievale dei pesci tra realtà e immaginazione”, in L’uomo di fronte al mondo animale nell’alto medioevo (Centro italiano di studi sull’alto medioevo. Atti delle settimane 31) (Spoleto 1985) 291–360; Bernal D. and Sáez A. (2008) “Fish-salting plants and amphorae production in the Bay of Cadiz (Baetica, Hispania). Patterns of settlement from the Punic era to Late Antiquity”, Thinking about Space: Studies in Eastern Mediterranean Archaeology 8, edd. H. Vanhaverbeke et al. (Turnhout 2008) 45–113; Botte E. (2009) Salaisons et sauces de poisons en Italie du sud et en Sicilie durant l’antiquité (Naples 2009); Lagóstena Barrios L. (2001) La producción de salsas y conservas de pescado en la Hispania Romana, II a. C.–VI d. C (Barcelona 2001); Villaverde N. and López F. (1995) “Una nueva factoría de salazones en Septem Fratres (Ceuta). El origen de la iocalidad y la problemática de la industria de salazones en El Estrecho durante el Bajo Imperio”, in Il Congreso internacional El Estrecho de Gibraltar (Ceuta 1990) (Madrid 1995) 455–72; Zerbini (2015) “The late antique economy: primary and secondary production”, in Local Economies?: Production and Exchange of Inland Regions in Late Antiquity, ed. L. Lavan (Late Antique Archaeology 10) (Boston 2015) 61–81 (66–67); Ben-Lazreg N. et al. (1995) “Production et commercialisation des salsamenta de l’Afrique ancienne”, in L’Afrique du nord antique et médiévale I. Production et exportations africaines: actualités archéologiques. Actes di VIe colloque d’histoire et d’archéologie de l’Afrique, Pau, 25–29 oct. 1993 (Paris 1995) 103–42; Charles M. (2010) “Unseemly professions and recruitment in Late Antiquity: Piscatores and Vegetius Epitoma 1.7.1–2”, AJP 521 (2010) 101–20.
Hunting There is a vast historiography on hunting in western Europe for the Late Medieval and Early Modern periods, but it was also an important activity for emperors and aristocrats in Late Antiquity. For the Early Middle Ages, scholars have paid attention to the cultural and political dimensions of the Frankish royal hunt: Goldberg
Eisenberg et al.
(2013). In earlier centuries, the public venationes of the Roman arena brought wild beasts to slaughter for the entertainment of civic crowds. Such spectacles continued into the Vandal period in North Africa, and are frequently depicted on late antique decorative pavements: Sparreboom (2016). Venationes outlasted gladiatorial combat as spectacles in general waned: Bomgardner (2013). Hunting: Sparreboom A. (2016) Venationes Africanae: Hunting Spectacles in Roman North Africa: Cultural Significance and Social Function (Ph.D. diss., Univ. of Amsterdam 2016); Harries J. (2002) “Favor populi: pagans, Christians and public entertainment in late antique Italy”, in Bread and Circuses: Euergetism and Municipal Patronage in Roman Italy, edd. K. Lomas and T. Cornell (New York 2002) 125–41; Bomgardner D. (2013) The Story of the Roman Amphitheater (New York 2013) 197–227; Goldberg E. (2013) “Louis the Pious and the hunt”, Speculum 88 (2013) 613–43; Goldberg E. (forthcoming) In the Manner of the Franks: Hunting, Kingship, and Manhood in the Early Middle Ages, 300–1000 (forthcoming); Godman P. (1990) “The poetic hunt: from Saint Martin to Charlemagne’s heir”, in Charlemagne’s Heir, edd. P. Godman and R. Collins (London 1990) 565–89; Dutton P. (2004) Charlemagne’s Mustache and Other Cultural Clusters of a Dark Age (New York 2004) 43–68; Aymard J. (1951) Essai sur les chasses romaines des origines à la fin du siècle des Antonins (Paris 1951); Anderson J. (1984) Hunting in the Ancient World (Berkeley 1984); Allsen T. (2006) The Royal Hunt in Eurasian History (Philadelphia 2006).
Minerals It has long been assumed that exploitation of mineral resources decreased sharply in Late Antiquity, particularly in the West: Kellens (2008). Archaeological surveys suggest that extraction peaked in the 2nd c. AD in important western mining regions like Iberia and Britain (McCormick (2001)), perhaps due to 3rd c. instability rather than exhaustion of the mines. One question of interest is the degree to which activity increased in the 4th c. Edmondson (1989) remains a useful overview, but systematic surveys of late antique mining are not numerous. Much of our new information comes from proxy indicators. Greenland ice cores, for example, offer insights into changing levels of economic activity over the course of antiquity and the Middle Ages, as the mining and smelting of lead-silver ores leave characteristic signatures in the ice: McConnell et al. (2018). Ice cores have also offered evidence for decreasing copper production in the Post Roman period, with levels rising again in the 9th c. through the exploitation of German ores: Hong and Patterson (1996). Mining: General: McCormick M. (2001) Origins of the European Economy: Communication and Commerce, AD 300–900 (Cambridge 2001) esp. 42–53; Téreygeol F. (2013) ed. Comprendre les savoirfaire métallurgiques antiques et médiévaux (Paris 2013); Kellens N. (2008) “Metal technology in Late Antiquity: a bibliographic note”, in Technology in Transition: AD 300–650, edd. L. Lavan, E. Zanini and A. Sarantis (Late Antique Archaeology 4) (Leiden-Boston 2008) 41– 52; McConnell J. et al. (2018) “Lead pollution recorded in Greenland
The Environmental History of the Late Antique West ice indicates European emissions tracked plagues, wars, and imperial expansion during antiquity”, PNAS (May 2018) doi: 10.1073/ pnas.1721818115; Edmondson J. (1989) “Mining in the later Roman empire and beyond: continuity or disruption?”, JRS 79 (1989) 84–102; Hong S. and Patterson C. (1996) “History of ancient copper smelting pollution during Roman and Medieval times recorded in Greenland ice”, Science 272.5259 (1996) 246–49; Healy J. F. (1978) Mining and Metallurgy in the Greek and Roman World (London 1978); Kent J. and Painter K. (1977) edd. The Wealth of the Roman World: Gold and Silver AD 300–700 (London 1977). Spain: Mil-Homens M. et al. (2017) “Insights of Pb isotopic signature into the historical evolution and sources of Pb contamination in a sediment core of the southwestern Iberian Atlantic shelf”, Science of the Total Environment 586 (2017) 473–84; Irabien M. J. et al. (2012) “Historical signature of Roman mining activities in the Bidasoa Estuary (Basque Country, northern Spain): an integrated micropalaeontological, geochemical and archaeological approach”, JAS 39 (2012) 2361–70; Martínez Cortizas A. et al. (2013) “Atmospheric Pb pollution in N Iberia during the Late Iron Age/Roman times reconstructed using the high-resolution record of La Molina mire (Asturias, Spain)”, Journal of Paleolimnology 50 (2013) 71–86; Larrazabal J. (1996) “Iron mining and metallurgy during the Roman and Medieval periods in Zamora (Spain)”, in The Importance of Ironmaking: Technical Innovation and Social Change, vol. 2, ed. G. Magnusson (Stockholm 1996) 167–75; Domerque C. (1990) Les mines de la péninsule ibérique dans l’antiquité romaine (Rome 1990). Gaul: Domerque C. and Leroy M. (2000) “L’etat de la recherche sur les mines et les metallurgies en Gaule, de l’époque gauloise au haut moyen âge”, Gallia 57 (2000) 310; Benoit P. and Brunstein P. (1983) edd. Mines, carrières, et métallurgie dans la France médiévale (Paris 1983). Italy: Tizzoni M. (1999) La miniera perduta. Cinque anni di ricerche archeometallurgiche nel territorio di Bienno (Breno 1999); La Salvia V. (1995) “Gap or continuity? Mining in Early Middle-Ages Italy”, in The Importance of Ironmaking: Technical Innovation and Social Change, ed. G. Magnusson (Stockholm 1995) 263–71.
Water Water drove the economies and ecologies of the western Mediterranean in Late Antiquity. This has been long recognised, and many famous works on world civilisation have addressed the import of water systems and water control. Wittfogel (1957) argued that the organisation of labour around water resources was at the core of early states and empires such as Rome. This idea has remained a key part of many general works on the history of water. For the case of Rome, the role of the military and the slave economy and the personal agendas of emperors, have been important for understanding the spread of large-scale water engineering projects, and how water control is connected to imperial power: Campbell (2012); Rogers (2013). Recently, the centrality of the state for water control in the West has been challenged by Lucas (2011), who argues that the role of local rather than regional authorities (cities, nobles, villages, etc.) was more important for the building and maintenance of water networks. Scholars continue to emphasise the importance of water in shaping and reflecting cultural values: Squatriti (1998); Jones (2005). For Rome and much of the antique West, bath houses and bathing culture have been seen
43 as crucial: Fagan (1999). Religious historians of Late Antiquity often discuss the importance of baptism and water in religion, while cultural and art historians explore the role of fountains in domestic and urban spaces. The history of technology has been a mainstay of understanding the importance of water in antique societies; in the West this discussion tends to focus on aqueducts, bridges, and ports, and water transportation: Garbrecht (1998); Blair (2007); Mays (2010). Land drainage is a relevant issue for large swaths of the North: Guillerme (1998). The history of irrigation is less prominent for the western Mediterranean than for the East, as irrigation in the North was less ubiquitous, necessary, and more local, given differing rainfall regimes: Leone (2012). The rise of ‘envirotech’ scholarship, that works to bridge socio-cultural research and the history of technology, has been slower to address ancient and late antique history, but will push the study of water technologies in intriguing directions. Environmental history encourages us to see that the relationship between water and human societies works both ways; people shaped water, but water systems asserted their own agency, putting human lives, economies, and agriculture at risk. For example, flooding, and other forms of water disaster, have been explored: Aldrete (2007). Water in the late antique West: Wittfogel K. (1957) Oriental Despotism: a Comparative Study of Total Power (New Haven, Connecticut 1957); Campbell J. B. (2012) Rivers and the Power of Ancient Rome (Chapel Hill, North Carolina 2012); Rogers A. (2013) Water and Roman Urbanism: Towns, Waterscapes, Land Transformation and Experience in Roman Britain (Leiden 2013); Lucas A. (2011) Wind, Water, Work: Ancient and Medieval Milling Technology (Leiden 2011); Squatriti P. (1998) Water and Society in Early Medieval Italy: AD 400–1000 (Cambridge 1998); Jones P. (2005) Reading Rivers in Roman Literature and Culture (Lanham, Maryland 2005); Fagan G. (1999) Bathing in Public in the Roman World (Ann Arbor 1999); Garbrecht G. (1998) ed. Die Wasserversorgung Antiker Städte: Mensch und Wasser, Mitteleuropa, Thermen, Bau/materialien, Hygiene (Mainz am Rhein 1988); Blair J. (2007) ed. Waterways and Canal-Building in Medieval England (Oxford 2007); Mays L. W. (2010) ed. Ancient Water Technologies (DordrechtLondon 2010); Guillerme A. (1998) The Age of Water: the Urban Environment in the North of France, AD 300–1800 (College Station, Texas 1988); Leone A. (2012) “Water management in late antique North Africa: agricultural irrigation”, Water History 4.1 (April 2012) 119–33; Aldrete G. (2007) Floods of the Tiber in Ancient Rome (Baltimore 2010). Other aqueous works: Arnold E. (2014) “Fluid identities: poetry and the navigation of mixed ethnicities in late antique Gaul”, Ecozon@ 5.2 (2014) 88–106; Arnold E. (2017) “Rivers of risk and redemption in Gregory of Tours’ writings”, Speculum 92 (2017) 117–43; Brooks G. (2006) “The ‘Vitruvian mill’ in Roman and Medieval Europe”, in Wind and Water in the Middle Ages: Fluid Technologies from Antiquity to the Renaissance, ed. S. Walton (Tempe, Arizona 2006) 1–38; Campbell J. B. (2012) Rivers and the Power of Ancient Rome (Chapel Hill, North Carolina 2012); Edmonds F. (2007) “Barrier or unifying feature? Defining the nature of early medieval water transport in the North-West”, in Waterways and Canal-Building in Medieval England, ed. J. Blair (Oxford 2007) 21–36; Hoffmann R. (2014) An Environmental History of Medieval Europe (Cambridge
44 2014); Leybourne M. and Gaynor A. (2006) edd. Water: Histories, Cultures, Ecologies. Contemporary Issues (Crawley, Western Australia 2006); Lucas A. (2011) Wind, Water, Work: Ancient and Medieval Milling Technology (Leiden 2011); Mauch C. and Zeller T. (2008) edd. Rivers in History: Perspectives on Waterways in Europe and North America (Pittsburgh 2008); McCormick M. (2001) Origins of the European Economy: Communications and Commerce, AD 300–900 (Cambridge 2001); Molkenthin R. (2006) Strassen aus Wasser: technische, wirtschaftliche und militärische Aspekte der Binnenschifffahrt im Westeuropa des frühen und hohen Mittelalters (Münster 2006); Nielsen I. (1990) Thermae et Balnea: the Architecture and Cultural History of Roman Public Baths (Aarhus 1990); Patterson D. (2013) “Adversos paganos: disaster, dragons, and episcopal authority in Gregory of Tours”, Medieval and Renaissance Studies 44 (2013) 1–28; Squatriti P. (2010) “The floods of 589 and climate change at the beginning of the Middle Ages: an Italian microhistory”, Speculum 85 (2010) 799–826; Pelteret D. (2010) “The role of rivers and coastlines in shaping early English history”, Haskins Society Journal: Studies in Medieval History 21 (2010) 21–46; Rogers A. (2013) Water and Roman Urbanism: Towns, Waterscapes, Land Transformation and Experience in Roman Britain (Leiden 2013); Squatriti P. (2000) ed. Working with Water in Medieval Europe: Technology and Resource-Use (Leiden 2000); Suttor M. (2011) “La navigation sur l’Escaut, des origines au XVIIe siècle”, Revue Du Nord 93, no. 391–92 (2011) 851–67; Suttor M. (2006) Vie et dynamique d’un fleuve. La Meuse de Sedan à Maastricht (des origines à 1600) (Paris 2006); Tvedt T. and Coopey R. (2010) edd. Rivers and Society: From the Birth of Agriculture to Modern Times (London 2010); Walton S. A. (2006) ed. Wind and Water in the Middle Ages: Fluid Technologies from Antiquity to the Renaissance (Tempe, Arizona 2006); Wikander Ö. (2000) ed. Handbook of Ancient Water Technology (Leiden 2000); Yegül F. (1992). Baths and Bathing in Classical Antiquity (New York 1992).
Plague and Disease The Justinianic Plague A massive bibliography exists on the ‘Justinianic Plague’. Almost none of this work has differentiated between its potential effects in the eastern and western Mediterranean, but, problematically, assumes a single outcome across all regions, sub-regions, and microregions. The bibliography for these wider Mediterranean studies can be found in McMahon and Sargent (2018). Western focused examinations remain limited to only a few locations and sources. Kulikowski (2007) examined the evidence for Iberia, while Horden (1992) and Bachrach (2007) examined Gregory of Tours as a literary source for Justinianic plague. There has also been some work on early Ireland, Anglo-Saxon England and plague in MacArthur (1949), Russell (1976), and Maddicott (1997/2007) and Dooley (2007). Scholars have not examined either literary or non-literary sources in detail from North Africa, Italy, the Balkans, or Gaul (except for Gregory of Tours). Scholars tend to assume a uniform demographic collapse across the Mediterranean that applies equally to the West: see Biraben and Le Goff (1969) and (1975), Mitchell (2015), and especially Harper (2017) for a few examples of this. Some have not
Eisenberg et al.
hesitated to put the Justinianic Plague in the far north completely in lieu of written evidence: Seger (1982). Yet, others have argued that the Justinianic Plague did not have grave effects in the West (particularly north of the Mediterranean) after about 600, like Harrison (1993) and Devroey (2009). A forthcoming reappraisal rejects plague as a significant driver of change in Late Antiquity (Mordechai and Eisenberg (2019)). The western discussion on the plague has recently concentrated on archaeological and ancient DNA (aDNA) evidence in great detail, since projects in this region have recorded such data, unlike at eastern sites. In the West, aDNA evidence has found Yersinia pestis at four sites: Sens and Vienne in France and Aschheim and Altenerding in Germany. Drancourt et al. (2004) found the first evidence of plague at Sens, which they subsequently located at Vienne, seen in Drancourt et al. (2007). Prentice et al. (2004), Vergnaud (2005) and Harbeck et al. (2013) cast doubt on these findings, and plague aDNA from Sens and Vienne is now considered problematic. The cemetery at Aschheim has been found to contain a few remains with the plague, first laid out in Wiechmann and Grupe (2005) and then confirmed in Harbeck et al. (2013), and Wagner (2014). A subsequent study at Altenerding in Feldman et al. (2016) likewise found evidence for Y. pestis in Bavaria. Their analysis included a reassessment and correction of the earlier Wagner study. Despite first appearances, the Aschheim and Alterneding strains, found just 20 km apart, are likely the same. There has also been work on mass graves in the West from Late Antiquity, which in some cases align spatiotemporally with known outbreaks of plague, for which see McCormick (2015) and (2016). Some of these sites, such as Vienne, have been studied in detail (see Signoli et al. (2009)), while others await detailed analysis. Burials, like those at Camerton in England are only suggestive of plague or large mortality events: Meaney (1964). Plague: McMahon L. and Sargent A. (2018) “Environmental history of the late antique eastern Mediterranean: a bibliographic essay”, in this volume; Kulikowski M. (2007) “Plague in Spanish Late Antiquity”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 150–70; Horden P. (1992) “Disease, dragons and saints: the management of epidemics in the Dark Ages”, in Epidemics and Ideas: Essays on the Historical Perception of Pestilence, edd. T. Ranger and P. Slack (Oxford 1992) 45–76; Bachrach B. S. (2007) “Plague, population, and economy in Merovingian Gaul”, Journal of the Australian Early Medieval Association 3 (2007) 29–57; MacArthur W. (1949) “The identifications of some pestilences recorded in the Irish Annals”, Irish Historical Studies 6 no. 23 (1949) 169–88; Russell J. (1976), “The earlier Medieval plague in the British Isles”, Viator: Medieval and Renaissance Studies 7 (1976) 65–78; Maddicott J. “Plague in seventh-century England”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 171–214 (a reprint of Maddicott J. (1997) “Plague in seventh-century England”, PastPres 156 (1997) 7–54); Dooley A. (2007) “The plague and its consequences
The Environmental History of the Late Antique West in Ireland”, in Plague and the End of Antiquity: the Pandemic of 541– 750, ed. L. K. Little (Cambridge 2007) 215–30; Biraben J.-N., and Le Goff J. (1969) “La peste dans de haut moyen âge”, AnnÉconSocCiv 24 (1969) 1484–1510 (transl. as Biraben J.-N., and Le Goff J. (1975) “The plague in the Early Middle Ages”, in Biology of Man in History, edd. R. Forster and O. Ranum (Baltimore 1975) 48–80); Mitchell S. (2017) A History of the Later Roman Empire, AD 284–641 (Malden, Mass. 2nd edn. 2015); Harper K. (2017) The Fate of Rome: Climate, Disease, and the End of an Empire (Princeton 2017); Seger T. (1982) “The Plague of Justinian and other scourges: an analysis of the anomalies in the development of the Iron Age population in Finland”, Fornvännen 77 (1982) 184–97; Harrison D. (1993), “Plague, settlement and structural change at the dawn of the Middle Ages”, Scandia 59.1 (1993) 15–48; Devroey J.-P. (2009) “Catastrophe, crise et changement social: à propos des paradigmes d’interprétation du développement médiéval (500–1100)”, in Actes des 9e journées anthropologiques de Valbonne, edd. L. Buchet et al. (Valbonne 2009) 139–61; Mordechai L. and Eisenberg M. (2019) “Rejecting catastrophe: the case of the Justinianic plague” PastPres (2019); Newfield T. (2015) “Human-Bovine Plagues in the Early Middle Ages.” Journal of Interdisciplinary History 46.1 (2015) 1–38; Drancourt M. et al. (2004) “Genotyping, orientalis-like Yersinia Pestis, and plague pandemics”, Emerging Infectious Diseases 10, no. 9 (September 2004) 1585–92; Drancourt M. et al. (2007) “Yersinia pestis orientalis in remains of ancient plague patients”, Emerging Infectious Diseases 13.2 (February 2007) 332–33; Vergnaud G. et al. (2005) “Yersinia pestis genotyping”, Emerging Infectious Diseases 11, no. 8 (August 2005) 1317–19; Prentice M. B. (2004) “Was the Black Death caused by Yersinia pestis?”, The Lancet. Infectious Diseases 4, no. 2 (February 2004) 72; Wiechmann I. and Grupe G. (2005) “Detection of Yersinia pestis DNA in two early medieval skeletal finds from Aschheim (Upper Bavaria, 6th century AD)”, American Journal of Physical Anthropology 126 (2005) 48–55; Harbeck M. et al. (2013) “Yersinia pestis DNA from skeletal remains from the 6th century AD reveals insights into Justinianic plague”, PLOS Pathogens 9.5 (May 2013) 1–8; Wagner D. M. et al. (2014) “Yersinia pestis and the plague of Justinian 541–543 AD: a genomic analysis”, Lancet Infectious Diseases 14 (2014) 319–26; Feldman M. et al. (2016) “A high-coverage yersinia pestis genome from a sixth-century Justinianic plague victim”, Molecular Biology and Evolution 33.11 (2016) 2911–23; McCormick M. (2015) “Tracking mass death during the fall of Rome’s empire (I)”, JRA 28 (2015) 325–57; McCormick M. (2015) “Tracking mass death during the fall of Rome’s empire (II): a first inventory of mass graves”, JRA 29 (2016) 1004–1007. Signoli M. et al. (2009). “Une sépulture de pestiférés du haut moyen âge à Vienne (Isère)”, Archéologie du Midi Médiéval 27, no. 1 (2009) 19–29; Meaney A. (1964) Gazetteer of Early AngloSaxon Burial Sites (London 1964); Foot S. (2010) “Plenty, portents and plague: ecclesiastical readings of the natural world in early medieval Europe”, in God’s Bounty? The Churches and the Natural World, edd. P. Clarke and T. Claydon (Woodbridge 2010) 15–41.
Other Diseases Epidemics A number of pathogens pestered, seriously sickened, and killed late antique peoples. Yersina pestis was but one of those pathogens. Reports of non-plague epidemics, or plagues not thought to be yersinial, abound in late antique sources. However, relatively few of these events have been studied in depth, which has only lent support to the idea that Late Antiquity was an age of plague alone. There are a few surveys of early medieval epidemic diseases, both Y. pestis and others, in the West that touch on Late Antiquity. The exceptions are Bonser (1944) and (1963) for late antique Anglo-Saxons, Welsh
45 and Irish mortalities, and Newfield (2011) for Carolingian mortalities. There are individual case studies of seemingly non-plague epidemics that disseminated in the West in and around, for example, 569–70 (Newfield (2015)), 574 (Newfield (2018)), the 580s (Janssens (1973)), and 785 (MacArthur (1951)). Other case studies tackle epidemics often, but not always, considered to be Y. pestis, like Ireland’s so-called ‘Yellow Plague’ of 664: cf. Shrewsbury (1949); MacArthur (1947), (1949), and (1950); Maddicott (2007). As the above suggests, not all historians agree that every epidemic listed as plague in Biraben and Le Goff’s (1969) seminal study on the Justinianic Plague was true plague. At the same time, epidemics not included in that list, like the Irish plague at the turn of the 8th c., have been considered yersinial: Woods (2011). For a pre-Justinianic plague, which seems not to have occurred, see Todd (1977). Epidemics: Bonser W. (1944), “Epidemics during the Anglo-Saxon period”, Journal of the British Archaeological Association 9.1 (1944) 48–71; Bonser W. (1963) The Medical Background of Anglo-Saxon England: a Study in History, Psychology, and Folklore (London 1963); Newfield T. (2011) The Contours of Disease and Hunger in Carolingian and Early Ottonian Europe (c.750–c.950 CE) (Ph.D. diss., McGill University 2011); Newfield T. (2015) “Human-bovine plagues in the Early Middle Ages”, Journal of Interdisciplinary History 46 (2015) 1–38; Newfield T. (2018) “Mysterious and mortiferous clouds: towards an historical epidemiology of the Late Antique Little Ice Age”, in this volume; Janssens P. (1973) “De morbus desentericus in de Historia Francorum van Gregorius van Tours (VIe eeuw)”, Scientiarum Historia 15 (1973) 217–23; MacArthur W. (1951), “The pestilence called ‘scamach’”, Irish Historical Studies 7 no. 27 (1951) 199–200; Shrewsbury J. (1949) “The yellow plague”, Journal of the History of Medicine and Allied Sciences 4 no. 1 (1949) 5–47; MacArthur W. (1947) “Famine fevers in England and Ireland”, Post-Graduate Medical Journal 23 no. 260 (1947) 283–86; MacArthur W. (1949) “The identifications of some pestilences recorded in the Irish Annals”, Irish Historical Studies 6 no. 23 (1949) 169–88; MacArthur W. (1950) “Comments on Shrewbury’s ‘The yellow plague’”, Journal of the History of Medicine 5 no. 1 (1950) 214–15; Maddicott J. (2007) “Plague in seventh-century England”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. K. Little (Cambridge 2007) 171–214 (a reprint of Maddicott J. (1997) “Plague in seventh-century England”, PastPres 156 (1997) 7–54; Biraben J.-N., and Le Goff J. (1969) “La peste dans de haut moyen âge”, AnnÉconSocCiv 24 (1969) 1484–1510 (transl. as Biraben J.-N., and Le Goff J. (1975) “The plague in the Early Middle Ages”, in Biology of Man in History, edd. R. Forster and O. Ranum (Baltimore 1975) 48–80); Woods D. (2011) “Adomnán, plague and the eastern controversy”, Anglo-Saxon England 40 (2012) 1–13; Todd M. (1977) “Famosa pestis and Britain in the fifth century”, Britannia 8 (1977) 319–25.
Malaria Beyond epidemics, some attention has been given to endemic diseases. In recent years, studies on malaria’s cyclical fevers, visible in Frankish sources, have appeared: Faure and Jacquemard (2014); Newfield (2017). The most recent argues Plasmodium vivax and Plasmodium malariae were comparatively common in the Merovingian kingdom. Similarly, Hassl (2008) has argued that the most benign malaria, P. malariae, was prevalent in late
46 antique Italy. Older work focused on epigraphic evidence, and has looked to illuminate malaria-like seasons of mortality in late antique Italy (Scheidel (1994)) and in Spain (Handley (2003)). An interdisciplinary study of late antique malaria by palaeopathologists Gowland and Western (2012) examined potential skeletal indicators of malaria on 5,802 Anglo-Saxons, arguing that P. vivax was widespread in eastern England from the 5th c. onward. Other work, that also puts bones before texts to discuss malaria’s late antique past, includes Hadjouis’ (1996) analysis of the early medieval cemetery of IvryParmentier, France, and Baggieri and Mallegni’s (2001) study of a child’s skeleton from Pisa, Italy. There is also Facchini’s (2004) assessment of the human remains from Late Roman Ravenna. Two recent interdisciplinary syntheses on malaria deserve special emphasis; both focus on the most acute of the human malarias. Zeigler (2016) surveyed a diverse range of scholarship to illuminate a landscape of Plasmodium falciparum in late antique central Italy, and Marciniak et al. (2018) integrated aDNA, epidemiological, epigraphical, landscape and textual data to implicate P. falciparum in imperial Roman economic and social history. These cross-disciplinary studies consider plasmodial parasites as part of Italian landscapes. Both expand upon the earlier interdisciplinary work of McCormick (1998), Sallares et al. (2004), and Sallares (2006). The latter two studies, in turn, built partially on Sallares (2001), which reported the first aDNA evidence for P. falciparum. Molecular remnants of that parasite, encountered in the bones of a child who was unearthed in an unusually large mid 5th c. infant cemetery in Lugnano, Italy, is thought to be indicative of a malaria epidemic: Birkby (1999); Soren (2003). Additional aDNA evidence for malaria in late antique Europe was only reported in Marciniak et al. (2016). That study identified P. falciparum in two imperial Romans (dated 1st–4th c.). Other work has attempted to work out whether famous figures in Late Antiquity died a malarial death, like Alaric I (Galassi et al. (2016); Faure (2017)), or if people avoided central Italy because of its plasmodial fevers, like Attila: Soren (2003). Speculation on malaria’s late antique and early medieval North Sea history is found in Knotternus (2002). Malaria: Faure E. and Jacquemard N. (2014) “L’émergence du paludisme en Gaule”, Caesadunum 44–45 (2014) 55–69; Newfield T. (2017) “Malaria and malaria-like disease in the Early Middle Ages”, Early Medieval Europe 25 (2017) 251–300; Hassl A. (2008) “Die Malaria im Römischen Kaiserreich: eine Bemerkenwerte Textstelle in den Digesten”, Wiener klinische Wochenschrift 120.4 (2008) 11–14; Scheidel W. (1994) “Libitina’s bitter gains: seasonal mortality and endemic disease in the ancient city of Rome”, Ancient Society 25 (1994) 151–75; Handley M. (2003) Death, Society and Culture: Inscriptions and Epitaphs in Gaul and Spain, 300–750 (Oxford 2003); Gowland R. and Western A. (2012) “Morbidity in the marshes: using
Eisenberg et al. spatial epidemiology to investigate skeletal evidence for malaria in Anglo-Saxon England (AD 410–1050)”, American Journal of Physical Anthropology 147 (2012) 301–11; Hadjouis D. (1996) “Mortalité infantile et enfantine et causes de mortalité: l’exemple de la nécropole médiévale d’Ivry-Parmentier 2 (Val-de-Marne, France)”, Cahiers du Centre de Recherches Anthropologiques 8.1 (1996) 15–26; Baggieri G. and Mallegni F. (2001) “Morphopathology of some osseous alterations of thalassic nature”, Palaeopathology Newsletter 116 (2001) 10–16; Facchini F. et al. (2004) “Cribra orbitalia and cribra cranii in Roman skeletal remains from the Ravenna area and Rimini (I–IV century AD)”, International Journal of Osteoarchaeology 14 (2004) 126–36; Ziegler M. (2016) “Malarial landscapes in late antique Rome and the Tiber Valley’, Landscapes 17 (2016) 139–55; Marciniak S. et al. (2018) “A multi-faceted anthropological and genomic approach to framing Plasmodium falciparum malaria in Imperial period centralsouthern Italy”, Journal of Anthropological Archaeology 49 (2018) 210–24; Birkby W. (1999) “The infant cemetery at Poggio Gramignano: description and analysis”, in A Roman Villa and a Late Roman Infant Cemetery, edd. D. Soren and N. Soren (Rome 1999) 447–530; Soren D. (2003) “Can archaeologists excavate evidence of malaria”, WorldArch 35.2 (2003) 193–209; McCormick M. (1998) “The imperial Edge: Italo-Byzantine identity: movement and integration, A.D. 650–950’, in Studies on the Internal Diaspora of the Byzantine Empire, edd. H. Ahrweiler and A. Laiou (Washington 1998) 17–52; Sallares R. et al. (2004) “Spread of malaria to southern Europe in antiquity: new approaches to old problems”, Medical History 48 (2004) 311–28; Sallares R. et al. (2006) “Role of environmental changes in the spread of malaria in Europe during the Holocene”, Quaternary International 150 (2006) 21–27; Sallares R. and Gomzi S. (2001) “Biomolecular archaeology of malaria”, Ancient Biomolecules 3 (2001) 195–213; Marciniak S. et al. (2016) “Plasmodium falciparum malaria in 1st–2nd century CE southern Italy”, Current Biology 26 (2016) R1220–22; Knotternus O. (2002) “Malaria around the North Sea: a survey”, in Climate Development and History of the North Atlantic Realm, edd. G. Wefer et al. (Bremen 2002) 339–553; Galassi F. (2016) “The sudden death of Alaric I (c.370–410 AD), the vanquisher of Rome: a tale of malaria and lacking immunity”, European Journal of Internal Medicine 31 (2016) 84–87; Faure E. (2017) “The death of Alaric I (c.370–410 AD), the vanquisher of Rome: additional arguments strengthening the possible involvement of malaria”, European Journal of Internal Medicine 37 (2017) e14–15.
Tuberculosis Much less has been researched about tuberculosis, undoubtedly another important contributor to late antique pathogenic loads. For TB, palaeopathologists have taken the lead. This is not unexpected, considering the relative indistinguishability of tuberculosis’ symptoms, the ability of the disease to present in a number of ways, and the ambiguity typical of late antique and early medieval accounts of disease in the West. It must be kept in mind, however, that both Mycobacterium tuberculosis and Mycobacterium bovis mark the bones of few of their victims, and that TB is underrepresented in skeletal assemblages. Tuberculosis may have been common in the Imperial Roman period: Roberts and Buikstra (2008); Eddy (2015). For the Late Imperial period, several site-wide studies, like Lewis’ (2011) analysis of the children of Poundbury Camp, England, have appeared. Regional studies, such as Blondiaux et al.’s (1999) assessment of multiple cemeteries from
The Environmental History of the Late Antique West
late antique and early medieval northern France, and Jakob’s (2009) work on several cemeteries from late antique south-western Germany, are far less common. Reports detailing one or a few people, adults or children, identified as having (potentially) long suffered from TB, are common: consider, for example, Stirland and Waldron (1990), Palfi and Marcsik (1999), WiltschkeSchrotta and Berner (1999), Haas et al. (2000), Rajić and Ujčić (2003), and Bauduer et al. (2014). First-rate tuberculosis studies now combine potential skeletal indicators of tuberculosis with aDNA evidence for the bacterium: Cooper et al. (2016). With varying success, many attempts have been made to isolate tuberculosis aDNA from late antique skeletal assemblages: Müller et al. (2014). Tuberculosis: Roberts C. and Buikstra J. (2008) The Bioarchaeology of Tuberculosis: a Global View on a Re-Emerging Disease (Gainesville, Florida, 2nd edn. 2008); Eddy J. (2015) “The ancient city of Rome, its empire, and the spread of tuberculosis in Europe”, Tuberculosis 95 (2015) S23–28; Lewis M. (2011) “Tuberculosis in the non-adults from Romano-British Poundbury Camp, Dorset, England”, International Journal of Paleopathology 1.1 (2011) 12–23; Blondiaux J. et al. (1999) “Epidemiology of tuberculosis: a 4th to 12th c. AD picture in a 2498-skeleton series from northern France”, in Tuberculosis: Past and Present, edd. G. Pakfi et al. (Budapest 1999) 519–30; Jakob T. (2009) Prevalence and Patters of Disease in Early Medieval Populations: a Comparison of Skeletal Samples of the 5th–8th Centuries AD from Britain to Southwestern Germany (BAR-IS 1959) (Oxford 2009); Stirland A. and Waldron T. (1990) “The earliest cases of tuberculosis in Britain”, JAS 17 (1990) 221–30; Palfi G. and Marcsik A. (1999) “Palaeoepidemiological data of tuberculosis in Hungary”, in Tuberculosis: Past and Present, edd. G. Pakfi et al. (Budapest 1999) 531–40; Wiltschke-Schrotta K. and Berner M. (1999) “Distribution of skeletal tuberculosis in the skeletal material of eastern Austrian sites”, in Tuberculosis: Past and Present, edd. G. Pakfi et al. (Budapest1999) 541–48; Haas C. et al. (2000b) “Molecular evidence for different stages of tuberculosis in ancient bone samples from Hungary”, American Journal of Physical Anthropology 113 (2000) 293–304; Rajić P. and Ujčić Ž. (2003) “Anthropological analysis of the Late Roman/Early Medieval cemetery of Novigrad (Istria)”, Collegium Anthropologicum 27.2 (2003) 803–808; Bauduer F. et al. (2014) “Multiple calvarial lytic lesions: a differential diagnosis from early medieval France (5th to 7th c. AD)”, International Journal of Osteoarchaeology 24 (2014) 665–74; Cooper C. et al. (2016) “Tuberculosis in early medieval Switzerland—osteological and molecular evidence”, Swiss Medical Weekly 146 (2016) 1–11; Müller R. et al. (2014) “Biomolecular identification of ancient mycobacterium tuberculosis complex DNA in human remains from Britain and continental Europe”, American Journal of Physical Anthropology 153 (2014) 178–89.
Leprosy Like tuberculosis, leprosy is encountered in skeletal assemblages from many regions of the Mediterranean and Europe in the late antique period. If more culturally significant than TB (for leprosy in late antique Christian thought, see now Miller and Nesbitt (2014)), leprosy had far less demographic influence than TB or malaria. Palaeopathologists have investigated this faintly communicable disease with weak pathogenicity from
47 a number of angles. Although few regional or site-wide studies have appeared—like Boldsen’s (2008) assessment of the cemetery in Lauchheim, Germany, which was in use between 450–680—detailed analyses of one-off skeletons marked with leprous-like lesions are relatively common: Reader (1974); Manchester (1981), Blondiaux et al. (2002); Holbrook and Thomas (2005); Belcastro (2005); Rubini and Zaio (2009); Watson et al. (2009); Rubini et al. (2014). The best of these analyses now are thoroughly interdisciplinary, and look for leprosy on and in bone: Rubini et al. (2012); Inskip et al. (2015). A collaborative study recently suggested that multiple strains of leprosy were introduced into Europe in Late Antiquity with the “westward migration of peoples from Central Asia” (Donoghue et al. (2015)), but cutting-edge genomic work now identifies the European peninsula as a key region “for the early spread and worldwide dissemination of leprosy”: Schuenemann et al. (2018). Drawing on a genomic study of a 7th c. leper from Hungary, Mendum et al. (2018) disagree. They posit that at least one strain of the disease arrived from Asia, and identify the Avars as a potential conduit. Either way, leprosy may have interacted with tuberculosis in important ways in the Late Roman era. It has long been argued that tuberculosis exposure made one immune to leprosy, but also that lepers were particularly prone to dying from tuberculosis: Manchester (1984b); Hohmann and Vöss-Böhme (2013); Rawson et al. (2014). Some aDNA evidence for co-infections is known: Donoghue et al. (2005). Leprosy: Miller T. and Nesbitt J. (2014) “Leprosy in the Latin West”, in Walking Corpses: Leprosy in Byzantium and the Medieval West, edd. T. Miller and J. Nesbitt (Ithaca, New York 2014) 96–117; Boldsen J. (2008) “Leprosy in the early medieval Lauchheim community”, American Journal of Physical Anthropology 135 (2008) 301–10; Reader R. (1984) “New evidence for the antiquity of leprosy in early Britain”, JAS 1 (1974) 205–207; Manchester K (1981) “A leprous skeleton of the 7th century from Eccles, Kent, and the present evidence of leprosy in early Britain”, JAS 8 (1981) 205–209; Blondiaux J. et al. (2002) “Microscopic study and x-ray analysis of two 5th century cases of leprosy: palaeoepidemiological inferences”, in The Past and Present of Leprosy: Archaeological, Historical, Palaeopathological and Clinical Approaches, edd. C. Roberts et al. (BAR-IS 1054 (Oxford 2002) 105–10; Holbrook N. and Thomas A. (2005) “An early-medieval monastic cemetery at Llandough, Glamorgan: excavations in 1994”, Medieval Archaeology 49 (2005) 1–92; Belcastro M. et al. (2005) “Leprosy in a skeleton from the 7th century necropolis of Vicenne-Campochiaro”, International Journal of Osteoarchaeology 15 (2005) 431–48; G. Pálfi and E. Molnár (2009) “The palaeopathology of specific infectious diseases from southeastern Hungary: a brief overview”, Acta Biologica Szegediensis 53.2 (2009) 111–16; Rubini M. and Zaio P. (2009) “Lepromatous leprosy in an early medieval cemetery in central Italy (Morrione, Campochiaro, Molise, 6th–8th century AD)”, JAS 36 (2009) 2771–79; Watson C. L. et al. (2009) “Single nucleotide polymorphism analysis of European archaeological M. leprae DNA”, PLoS One 5.1 (2009) e7547: https://doi.org/10.1371/journal.pone.0007547; Rubini M. et al. (2014) “Tuberculosis and leprosy in Italy: new skeletal evidence”, HOMO: Journal of Comparative Biology 65 (2014) 13–32; Rubini M. et al. (2012) “Pathological and molecular study of two
48 cases of ancient childhood leprosy from the Roman and Byzantine empires”, International Journal of Osteoarchaeology 24 (2014) 570– 82; Inskip S. et al. “Osteological, biomolecular and geochemical examination of an Early Anglo Saxon case of lepromatous leprosy”, PLoS One 10.5 (2015) e0124282: doi.org/10.1371/journal.pone.0124282; Donoghue et al. (2015) “A migration-driven model for the historical spread of leprosy in medieval eastern and central Europe”, Infection, Genetics and Evolution 31 (2015) 250–56; Schuenemann V. et al. “Ancient genomes reveal a high diversity of mycobacterium leprae in medieval Europe”, PLoS One 14.5 (2018) e1006997: doi.org/10.1371/ journal.ppat.1006997; Mendum T. et al. (forthcoming) “The genome sequence of a SNP type 3k strain of mycobacterium leprae isolated from a seventh-century Hungarian case of lepromatous leprosy”, International Journal of Osteoarchaeology (forthcoming 2018); Manchester K. (1984b) “Tuberculosis and leprosy in antiquity: an interpretation”, Medical History 28 (1984) 162–73; Hohmann H. and Vöss-Böhme A. (2013) “The epidemiological consequences of leprosytuberculosis co-infection”, Mathematical Biosciences 241 (2013) 225– 37; Rawson T. et al. (2014) “Leprosy and tuberculosis concomitant infection: a poorly understood, age-old relationship”, Lepra Review 85 (2014) 288–95; Donoghue H. et al. “Co-infection of mycobacterium tuberculosis and mycobacterium leprae in human archaeological samples: a possible explanation for the historical decline of leprosy”, Proceedings of the Royal Society B 272 (2005) 389–94.
Crop and Livestock Disease Countless other pathogens caused disease on endemic and epidemic scales in the late antique West. Not all of those microorganisms plagued people. Although crop disease has garnered little attention (Bondesson and Bondesson (2014)), much has been said recently about livestock disease, plagues in bovines especially. Surveys of textual evidence for late antique and early medieval livestock disease are found in Newfield (2013) and (2015b), and zooarchaeological studies concerning unusual animal die-offs, possibly epizootics, include Putalet (2013) and Renou (2013). Apparently large outbreaks of bovine disease in AD 569–70, which also seem to have afflicted people, are analysed in Newfield (2015), and an early 9th c. pan-European cattle event is discussed in Newfield (2012). Although most of the evidence available for livestock disease concerns cattle, Gillmor (2005) shines light on a horse epizootic in 791. Another study addresses the 376–86 bovine panzootic, which, though relatively wellknown in the animal health sciences, seems not to have actually occurred: Newfield (2017b). Crop and livestock disease: Bondesson L., and Bondesson T. (2014) “On the mystery cloud of AD 536, a crisis in dispute and epidemic ergotism: a linking hypothesis”, Danish Journal of Archaeology 3 (2014) 61–67; Newfield T. (2013) “Early medieval epizootics and landscapes of disease: the origins and triggers of European livestock pestilences, 400–1000 CE”, in Landscapes and Societies in Medieval Europe East of the Elbe, edd. S. Kleingärtner, T. Newfield, S. Rossignol, and D. Wehner (Toronto 2013) 73–113; Newfield T. (2015b) “Domesticates, disease and climate in early Post-classical europe: the cattle plague of c.940 and its environmental context”, Postclassical Archaeologies 5 (2015) 95–126; Putelat O. (2013) “Archéologie des depots animaux et mortalité extradinaire du cheptel bovin au premier moyen âge: plaidoyer pour une redynamisation et une mise en perspective de
Eisenberg et al. la recherché”, in Les dépots d’ossements d’animaux en France, de la fouille à l’interprétation, edd. G. Auxiette and P. Meniel (Montagnac 2013) 249–74; Renou S. et al. (2013) “Un bilan des connaissances sur les epizooties au moyen âge et un episode de mort extraordinaire de bovines à Luxé (Charente) durant l’époque mérovingienne”, in Les dépots d’ossements d’animaux en France, de la fouille à l’interprétation, edd. G. Auxiette and P. Meniel (Montagnac 2013) 133–43; Newfield T. (2012) “A great Carolingian panzootic: the probable extent, diagnosis and impact of an early ninth-century cattle pestilence”, Argos 46 (2012) 200–12; Gillmor C. (2005) “The 791 equine pestilence and its impact on Charlemagne’s army”, Journal of Medieval Military History 3 (2005) 23–45; Newfield T. (2017b) “Livestock plagues in Late Antiquity, with a disassembling of the bovine panzootic of AD 376– 386”, JRA 30 (2017) 490–508.
Climate and Food Shortages Late Antiquity has been cold for decades. A 2016 dendroclimatological study (Büntgen et al. (2016)) identified a series of unusually cold summers from about 536 to 660 in tree-ring-width (TRW) chronologies from both the Alpine and the Altai mountain ranges. This chilly century was coined the Late Antique Little Ice Age (LALIA). Importantly, the LALIA is not visible as a distinct phenomenon in all natural climate proxies across Late Antiquity. Some have argued that it was a non-event: Helama et al. (2017); Helama et al. (2017b). According to these scholars the LALIA, if it existed, should conclude about 580, since only from 536–80 do most high-resolution temperature proxies register exceptionally cool summers. Nevertheless, there is unusual synchronicity between central European and central Asian TRW series, within which a 130-year LALIA can be seen. There is no reason to doubt the LALIA, but it was not experienced everywhere. Rather, it was an especially cool period that occurred within a longer, more nebulously defined less cool period, that spanned, depending on the proxies, from the 5th to the 8th c. (about 400–765). This longer cool period has many names; the ‘anti-LALIA’ scholars prefer Dark Age Cold Period (DACP), a title first advanced in McDermott et al. (2001). More widely used, if misleading (Squatriti (2010)), is the Vandal Minimum. Others include Migration Pessimum and Early Medieval Cold Period. This cool era is assessed in a longer context in McCormick et al. (2012), Manning (2013) and Luterbacher et al. (2016). The DACP, like the LALIA, is decipherable only in natural archives of climate. These archives, from tree rings to ice cores, and from lake sediments to speleothems (mineral deposits in caves), are influenced by climatic parameters. They can provide information about pre-instrumental variations in temperature and precipitation (not to mention other climate variables), and they do so on various spatiotemporal scales. A proxy, in turn, is an indirect indicator of climate that is recorded in an archive, like tree-ring width. Studies
The Environmental History of the Late Antique West
of trees and speleothems are capable of providing the highest resolution, though lakes that freeze can shine a light on annual climate trends as well. Naturally, proxies cannot tell us everything. Few regions of Europe have good coverage for Late Antiquity, and there is no high-resolution palaeoclimatology for winter temperature or precipitation. What we understand best is summer temperature, thanks to the relative wealth of tree-ring studies. But even for summer conditions, spatial coverage is limited, with long-running dendroclimatological studies available only for parts of central Europe, continental Scandinavia, and Ireland. For the early modern period, data also exists, for instance, for the Balkans, Czech lands, England, France, Morocco, Poland, Scotland and Spain. Written sources provide some assistance. They help us flesh out extreme events, they comment on winter conditions, and they provide a resolution (monthly, weekly, daily, and hourly) that natural proxies cannot. Thin as they are, late antique texts in the West do not permit us to construct anything resembling a temperature or precipitation trend. Historical climatology is only useful for assessing climate in the West from the 15th c. onwards and, even then, the best work integrates natural proxies: Brazdil et al. (2005); Camenisch et al. (2016). Textual evidence for atmospheric clouding was instrumental in early descriptions of the so-called ‘mystery cloud’ of 536 (Stothers (1984); Arjava (2005)), which is now better understood as a 536–50 temperature downturn: for an overview see Newfield (2018b). Five Mediterranean sources observe what is almost certainly dust veiling for 12 to 18 months. Very different events have been suggested to explain it, from a local, climatically inconsequential eruption to an extraterrestrial impact, from far-off major explosive volcanism to a ‘damp fog’. Multiple proxies, however, point to two very large volcanic events in 535/36 and 539/40: Larsen et al. (2008); Baillie (2008). The first may account for the Mediterranean clouding. Both triggered dramatic summer temperature cooling, visible in TRW and Maximum Latewood Density dendro-studies across the northern hemisphere. These events initiated the LALIA. Large volcanic events occurred regularly in Late Antiquity and were more common than in the preceding Roman period and succeeding Early Medieval period: Sigl et al. (2015); Gao et al. (2016); Toohey and Sigl (2017). Some writers have assigned great agency to these events, especially the eruptions of 536 and 540 (Keys (1999); Harper (2017)), which are conceived of as precipitating the Justinianic Plague (Stathakopoulos (2003); McCormick (2007); Newfield (2018)) and causing widespread food shortages. More work is required, though, to tease out the causal mechanisms by which late antique volcanism may have influenced plague occurrence or
49 affected crops. In this regard, it is worth stressing that though the DACP has been characterised as more wet than dry (Helama et al. (2017b) the effects of these large volcanic events on precipitation is little studied. Climate: Büntgen U. et al. (2016) “Cooling and societal change during the Late Antique Little Ice Age from 536 to around 660 AD”, Nature Geoscience 9 (2016) 231–36; Helama S. et al. (2017) “Limited late antique cooling”, Nature Geoscience 10 (2017) 242–43; Helama S. et al. (2017b) “Dark Ages Cold Period: a literature review and directions for future research”, The Holocene 27 (2017) 1600–1606; McDermott et al. (2001) “Centennial-scale Holocene climate variability revealed by a high-resolution speleothem δ18o record from SW Ireland”, Science 295 (2001) 1328–31; Squatriti P. (2010) “The floods of 589 and climate change at the beginning of the Middle Ages: an Italian microhistory”, Speculum 85 (2010) 799–826; McCormick M. et al. (2012) “Climate change during and after the Roman empire: reconstructing the past from scientific and historical evidence”, Journal of Interdisciplinary History 43 (2012) 169–220; Manning S. (2013) “The Roman world and climate: context, relevance of climate change, and some issues”, in The Ancient Mediterranean Environment Between Science and History, ed. W. V. Harris (Leiden 2013) 103–70; Luterbacher J. et al. (2016) “European summer temperatures since Roman times”, Environmental Research Letters 11 (2016) 024001: https://doi .org/10.1088/1748-9326/11/2/024001; Brazdil R. et al. (2005) “Historical climatology in Europe—the state of the art”, Climatic Change 70 (2005) 363–430; Camenisch C. et al. (2016) “The 1430s: a cold period of extraordinary internal climate variability during the Early Spörer Minimum with social and economic impacts in north-western and central Europe”, Climate of the Past 12 (2016) 1–20; Stothers R. (1984) “Mystery cloud of AD 536”, Nature 307 (1984) 344–45; Arjava A. (2005) “The mystery cloud of 536 CE in the Mediterranean sources”, DOP 59 (2005) 73–94; Newfield T. (2018b) “The global climatic downturn of c.535–550: the state of the art”, in The Palgrave Handbook on Climate History, edd. S. White et al. (London 2018) 447–493; Larsen L. et al. (2008) “New ice core evidence for a volcanic cause of the AD 536 dust veil”, Geophysical Research Letters 35 (2008) L0708: doi .org/10.1029/2007GL032450; Baillie M. (2008) “Proposed re-dating of the European ice core chronology by seven years prior to the 7th century AD”, Geophysical Research Letters 35 (2008) L15813: doi .org/10.1029/2008GL034755; Sigl M. et al. (2015) “Timing and climate forcing of volcanic eruptions for the past 2,500 years”, Nature 523 (2015) 543–49; Gao C et al. (2016) “Reconciling multiple ice-core volcanic histories: the potential of tree-ring and documentary evidence, 670–730 CE”, Quaternary International 394 (2016) 180–93; Toohey M. and Sigl M. (2017) “Volcanic stratospheric sulfur injections and aerosol optical depth from 500 BCE to 1900 CE”, Earth System Science Data 9 (2017) 809–31; Keys D. (1999) Catastrophe: an Investigation into the Origins of Modern Civilization (London 1999); Harper K. (2017) The Fate of Rome: Climate, Disease, and the End of an Empire (Princeton 2017); Stathakopoulos D. (2003) “Reconstructing the climate of the Byzantine world: state of the problem and case studies”, in People and Nature in Historical Perspective, edd. J. Laszlovszky and P. Szabó (Budapest 2003) 247–61; McCormick M. (2007) “Towards a molecular history of the Justinianic pandemic”, in Plague and the End of Antiquity: the Pandemic of 541–750, ed. L. Little (Cambridge 2007) 290–312; Newfield T. (2018b) “Mysterious and mortiferous clouds: towards an historical epidemiology of the Late Antique Little Ice Age”, in this volume; Helama S. et al. (2017b) “Dark Ages Cold Period: a literature review and directions for future research”, The Holocene 27 (2017) 1600–1606.
50 Food Shortage The causal mechanisms by which climate change in Late Antiquity accelerated economic or social evolution or impacted crop development, demand more attention. Sen’s (1981) food entitlement theory has not yet been employed to explain subsistence crises in the late antique West. Malthus too has not played a major role in accounts of dearth in the Late Roman or Carolingian West. Rather, exogenous environmental shocks, namely short-term climate variability and extreme weather events, are thought to have generated Late Antiquity’s great episodes of hunger. With few exceptions though (Mathison (1993); Cândido da Silva (2016)), these sorts of events have been little studied. Food crises before the Carolingian period (Verhulst (1965); Bonnassie (1989); Newfield (2013); Cândido da Silva (2014); Ebert (2017)), require more attention. Their scale and severity, and importantly the ability of affected peoples to buffer
Eisenberg et al.
against dearth, and cope with such events, remain poorly understood. Food shortages: Sen A. (1981) Poverty and Famines: an Essay on Entitlement and Deprivation (Oxford 1981); Mathison R. (1993) “Nature or nurture—some perspectives on the Gallic famine of circa A.D. 470”, AncW 24 (1993) 91–105; Cândido da Silva M. (2016) “Os agentes públicos e a fome nos primeiros séculos da Idade Média”, Varia Historia 32 no. 60 (2016) 779–805; Verhulst A. (1965) “Karolingische Agrarpolitik: das Capitulare de Villis und die Hungersnöte von 792/93 und 805/06”, Zeitschrift für Agrargeschichte und Agrarsoziologie 13 (1965) 175–89; Bonnassie P. (1989) “Consommation d’aliments immondes et cannibalisme de survie dans l’Occident du haut moyen âge”, AnnÉconSocCiv 44–45 (1989) 1035–56; Newfield T. (2013) “The contours, frequency and causation of subsistence crises in Carolingian Europe (750–950)”, in Crisis alimentarias en la Edad Media: modelos, explicaciones y representaciones, ed. P. Benito i Monclús (Lleida 2013) 117–72; Cândido da Silva M. (2014) “L’économie morale’ carolingienne (fin VIIIe–début IX siècle)”, Médiévales 66 (2014) 159–78; Ebert S. (2017) “Starvation under Carolingian rule: the famine of 779 and the Annales Regni Francorum”, in Famines During the ‘Little Ice Age’ (1300–1800): Socionatural Entanglements in Premodern Societies, edd. D. Collet and M. Schuh (Cham, Switzerland 2017) 211–30.
Regional Vegetation Histories: Overview of the Pollen Evidence
Revisiting the Beyşehir Occupation Phase: Land-Cover Change and the Rural Economy in the Eastern Mediterranean during the First Millennium AD Neil Roberts Abstract The latter part of the Beyşehir Occupation Phase (BOP) corresponds in space and time to the Late Roman empire in the eastern Mediterranean. The emphasis on tree crops in pollen records, particularly olive trees, implies long-term investment, stable trade networks and regional economic integration. The onset of the BOP was time-transgressive, starting between the Bronze Age and Hellenistic times in different localities. During the mid 1st millennium AD, the BOP came to an end, often abruptly, with a marked decline in agricultural indicators and an increase in forests, implying partial landscape re-wilding. This termination is most commonly dated to the 7th c. AD, coinciding with Arab attacks on Byzantine territory, and this, rather than climate change, seems the most likely explanation for the regional collapse of the rural agrarian system. The end of the BOP marks the transition from Late Antiquity to the Early Medieval era, a transition which appears to have been notably later in date and more dramatic than elsewhere in the Mediterranean.
Introduction In 1970 the first palynological field research programme in south-west Anatolia was carried out by a team from Groningen University, led by Wim van Zeist, which led to sediment cores being taken from a number of marshes and lakes (figs 1 and 2). Subsequent pollen analysis and radiocarbon dating identified a phase of humaninduced land-cover change, particularly in cores taken from the Söğüt marshes and from the south-western corner of Lake Beyşehir.1 This pollen phase had a number of distinctive characteristics. In addition to a decline in forest taxa and an increase in the pollen of grasses (including cereal-type) and ruderals, it was marked by the presence of cultivated trees, such as olive (Olea) and walnut (Juglans). Another distinguishing feature was that it often came to a clear and abrupt end, rather than changing gradually through to more recent times (fig. 3). In the Söğüt and Beyşehir pollen records, anthropogenic
* This chapter is dedicated to the memory of Wim van Zeist (1924– 2016), a true pioneer of eastern Mediterranean palynology. 1 Van Zeist et al. (1975).
indicators disappeared almost completely and were replaced mainly by pine, implying secondary afforestation and a partial re-wilding of the landscape.2 The distinctive phase of arboriculture and other human land use was labelled the ‘Beyşehir Occupation Phase’ (or BOP) after one of the two type-sites.3 The end of the BOP was not dated directly in these initial investigations, but it had an inferred age of the middle of the 1st millennium AD, which corresponds to the time period of Late Antiquity (ca. AD 200–800). The BOP included the period of Classical Antiquity, and its ending may represent Post-Classical economic and demographic decline, at least in the countryside. Subsequent pollen investigations have found many more sites recording the BOP in the Pisidian Lake District of south-west Anatolia.4 Some of these pollen records lie close to, and can be linked directly to, Classical archaeological sites such as Sagalassos and Gölhisar (possibly ancient Sinda, in the territory of Cibyra5). Most of the pollen sites from this region are located in intermontane basins at elevations between 900 and 1200 masl, in the ancient provinces of Lycia and Pisidia. Some of them have well-resolved chronologies with numerous radiocarbon dates; these dates are confirmed at several sites by a volcanic ash layer from the Bronze Age eruption of Santorini (Thera).6 A modified form of the BOP has since been recognised in pollen records from elsewhere in south-west Asia, with pollen taxa indicative of arboriculture, as well as cereals, grazing indicators, weeds, etc. These records derive, in particular, from central and northern Anatolia, Cyprus, and the southern Levant.7 This distinctive land-use phase therefore occurred at pollen sites across much of the eastern Mediterranean region, and its significance was regional, not local. In its latter part 2 Roberts (1990). 3 Bottema et al. (1986). 4 E.g. Bottema and Woldring (1984), (1990); Sullivan (1988); Eastwood et al. (1998). 5 Sagalassos: Vanhaverbeke and Waelkens (2003); Vermoere (2004). Sinda: Hall (1994); Eastwood et al. (1999). 6 Eastwood et al. (1998), (1999). 7 Central and northern Anatolia: Bottema et al. (1993–94); Izdebski (2013a), (2013b). Cyprus: Kaniewski et al. (2013). Southern Levant: e.g. Heim et al. (1997); Neumann et al. (2010).
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_005 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 53–68
54
Roberts
figure 1
Field coring at Lake Beyşehir, 1977 (left, H. Woldring; centre, S. Bottema; right, W. van Zeist).
figure 2 Map of selected pollen sites in Anatolia and the Levant, along with the approximate eastern frontier of the eastern Roman empire in the 6th and 8th c. AD. Inset map shows sites in south-western Anatolia.
55
Revisiting the Beyşehir Occupation Phase
figure 3 Summary pollen diagram from Söğüt, south-west Anatolia, highlighting the Beyşehir Occupation Phase (redrawn from van Zeist et al. (1975); Holocene part only).
(i.e. 4th–7th c. AD) the BOP appears to be spatially congruent with the provinces of the eastern Roman empire in Anatolia, Cyprus and the Levant (fig. 2). In this paper, I will assess what insights pollen records can offer to explain changes in the rural economy and environment in the eastern Mediterranean during the 1st millennium AD. Proxy climate data other than pollen will be used to evaluate the role that climatic changes played in the transformation of societies at the end of antiquity.8 The Character of the Beyşehir Occupation Phase Pollen analysis provides a record of past vegetation and land cover, but it has both strengths and limitations. Some taxa (e.g. pine) produce more pollen grains and/or are dispersed more widely than others (e.g. “cereal-type” pollen). Some, such as the daisy family (Asteraceae/ Compositae, including Liguliflorae and Tubuliflorae) have pollen that is relatively resistant to oxidation, and hence can be over-represented when there is poor pollen preservation. A number of economically useful plants, such as cultivated fig trees (Ficus), are completely absent in the pollen record because they are pollinated by insects rather than by the wind. This means that past pollen percentages cannot be equated to former plant abundances, although methods are now being developed to transform pollen data statistically to overcome some of these distortions.9 Most pollen grains can only be identified to genus not species level, and for some herbs and grasses even genus-level identification is not possible. Thus, Cannabis pollen could represent either rope (i.e. hemp) or dope!
8 McCormick et al. (2012); Haldon et al. (2014). 9 Gaillard et al. (2010).
The pollen diagrams shown in this paper, and others in this volume, show untransformed percentage data, and should be interpreted in the light of the potential biases described above. Accurate interpretation of pollen diagrams therefore requires a thorough knowledge of the local modern vegetation, and an understanding of how this is represented and reflected in the modern pollen rain. Despite these uncertainties, pollen data offer important insights into long-term land-cover changes. In particular, most pollen sequences represent timecontinuous vegetation histories that can span time periods for which historical (e.g. documentary) records may be sparse or lacking.10 There are no ‘Dark Ages’ for palaeoecological evidence, except for hiatuses in sedimentation or samples with poor pollen preservation. For this reason alone, pollen data have a special value for our understanding of the transition from antiquity to the Medieval world, a period whose textual record is far from complete.11 The range of anthropic pollen types found during the BOP at sites in south-west Anatolia is shown in figure 4. As this shows, some taxa, e.g. olive, walnut and cerealtype, are present at almost all sites, although this need not necessarily imply that the crop was grown locally. Olive pollen, for example, can be spread up-valley from coastal locations by wind transport, whereas vine (Vitis) pollen is only dispersed locally. Some pollen records show signs of local specialisation in specific crops, such as manna ash (Fraxinus ornus) at Beyşehir, sweet chestnut (Castanea sativa) at Gölçük, south of the Classical city of Sardis,12 and walnut at Elmalı (ironically now famed for, and named after, its apple orchards). Some
10 Haldon et al. (2014). 11 Izdebski (2013a); Roberts et al. (2018). 12 Sullivan (1988), (1989).
56
Roberts
figure 4 The maximum pollen percentages of different BOP crops or planted trees at selected sites in south-west Anatolia (modified from Bottema and Woldring (1990); Eastwood et al. (1998); additional data from Kaniewski et al. (2007) and Vermoere (2004)).
planted species, such as the shade-giving plane tree (Platanus) have a clear association with human activity, without providing obvious economic benefit. The combination of crop plants represented in the BOP includes the typical Mediterranean triad of olive, wheat and vine that could be processed and consumed locally as oil, bread and wine. However, the range of tree crops strongly suggests that much of the production was for extra-local consumption, as part of wider networks of trade and exchange. This is in good accord with the archaeological and textual evidence, which shows there was large-scale maritime transportation of wine and olive oil in amphorae.13 Manna ash, for example, provides a sugary extract from the sap that is extracted by
making a cut in the bark; today, its production is best known in Sicily, but its economic and medicinal value has been diminished by access to other forms of sugar and medicine. Its value in antiquity must have been as a specialist crop, rather than as a subsistence one. In Anatolia and the Levant, Olea was the most commonly cultivated tree pollen type during the BOP. This includes areas of interior Anatolia where olive trees are not cultivated today, although Olea pollen is sometimes present in surface sediment and pollen trap samples from these areas due to long-distance wind transport (fig. 5). In sites from the Jordan Valley prior to the 7th c. AD, the proportion of olive pollen was commonly over 20%,14 as can be seen in the pollen diagram from
13 Mattingly (1996); Horden and Purcell (2000) 209 ff.
14 Van Zeist et al. (2009); Neumann et al. (2010).
57
Revisiting the Beyşehir Occupation Phase
a
b
c figure 5 Olive pollen percentage maps for the eastern Mediterranean: a) modern samples b) AD 700–900 c) AD 400–600. Primary data source: European Pollen and Modern Pollen Databases. Note that data coverage is not the same for modern and for palaeo-pollen samples; hence, the lack of modern samples for western Turkey or Cyprus does not imply that there are no olive trees growing in these regions.
58
Roberts
figure 6 Birket Ram pollen and magnetic susceptibility diagram (redrawn after Schwab et al. (2004)).
Birket Ram in the Golan Heights (fig. 6). These values are higher than those found today in the same region of the Levant (fig. 5), and are comparable to surface pollen samples in regions such as southern Spain, where olive cultivation is a major agro-industry today.15 This suggests that Palestine was used for large-scale commercial olive oil production for export within the eastern Roman empire, which accords with the archaeological and textual evidence.16 The only other sites where high (>20%) olive pollen values occurred during Late Roman times were in western Anatolia, in the Menderes (Maeander) river valley (today an important olive producing area), and at Lake Iznik, on whose shores was located the city of Nicaea.17 During the post-BOP period (8th–9th c. AD), olive pollen percentages drop dramatically at almost all sites, and none has values as high as those during the BOP (fig. 5), implying an end to large-scale commercial olive production and olive oil trade in the region. The importance of arboriculture during the BOP implies long-term investment in tree crops, whose economic yield could only be estimated decades ahead of the time of planting. This, in turn, implies a confidence in the future and stability of the rural economy. Tree crops are also a feature of pollen records from parts of the central and western Mediterranean during antiquity, for example in the Italian peninsula, where O-J-C (olive-walnut-chestnut) pollen has been used to identify periods of human landscape transformation.18 It
15 Díaz de la Guardia et al. (2003). 16 Waliszewski (2014). 17 Izdebski (2016); Miebach et al. (2016). 18 Mercuri et al. (2013).
is therefore unsurprising that the BOP, at least in its latter part, can be linked to the stable political and economic system that existed under Roman imperial rule. Equally, it is unsurprising that the post-BOP phase saw a shift away from cultivated trees towards crops, such as wheat and barley, that did not rely on long-term investment, complex processing or dependable trading networks. This means not only a decline in the rural economy during and after the period of Late Antiquity, but also a re-localisation and simplification of agrarian production. Does the BOP Termination Mark the End of Antiquity in the Eastern Roman Empire? While it is tempting to make correlations between changes recorded in different natural and historical “archives”, this question can only be answered with confidence if proposed correspondences can be shown to be robust and reliable. This, in turn, requires a critical evaluation of the evidence for synchronism, that is, via chronologies. Whereas most historical events, such the dedication of Byzantium and its renaming as Constantinopolis on 11 May AD 330, are placed in time quite precisely and reliably, this is not true of archaeological or palaeo-environmental evidence. In these cases, chronologies have to be constructed from evidence such as radiocarbon dates that require calibration and that have statistical uncertainties associated with them. In the case of pollen diagrams, radiocarbon dated levels may not correspond to changes in pollen zones, such as the end of the BOP. Instead their ages have to be interpolated between radiocarbon dates above and below the change in the pollen
59
Revisiting the Beyşehir Occupation Phase
assemblage zone, which inevitably reduces dating precision. The time interval between pollen samples also varies greatly, between 200 years for eastern Mediterranean sites during the 1st millennium AD Furthermore, radiocarbon dates can be contaminated by old or young carbon. There is a danger of what Baillie termed “suck-in and smear”,19 in which the dating of events is so imprecise that they can lead to spurious correlations and wrongly inferred causal relationships. As Figure 7 shows, the beginning of the BOP was time-transgressive, with onset dates clustering into three main time intervals; namely the 2nd millennium BC (Middle-Late Bronze Age), the early 1st millennium BC (8th–10th c. BC, corresponding to the early Iron Age), and the mid-1st millennium BC (3rd–6th c. BC, corresponding to the Achaemenid Persian and Hellenistic eras). Sites with a Bronze Age onset for the BOP are concentrated in south-west Anatolia, and correspond to the Lukka lands described in Hittite texts,20 and which were associated in the 1st millennium BC with the Lycians (or Termilae). By contrast, end-dates for the BOP cluster over a much narrower time range, in the middle of the 1st millennium AD. Different age estimates for its onset and termination can be explained in some cases by poor dating control, for example at Söğüt, which is dated by only a single radiocarbon determination during the last 3000 years (see fig. 3). Nonetheless, there is a striking contrast between the patterns shown in figure 7 for the BOP start and end dates, with the former spread out over more than 1300 years, while the latter corresponds to a narrow ‘bell curve’ distribution, consistent with a synchronous termination in the mid 1st millennium AD. In Anatolia, the best-dated pollen sequences for the end of the BOP come from Gravgaz, Bereket, Gölhisar (all in south-west Turkey), Iznik, in northwest Anatolia, and Nar lake in Cappadocia.21 At four of these sites, calculated periods for the end of the BOP fall between the mid 7th and the mid 8th c., whereas the fifth site (Bereket) has a significantly earlier estimated BOP end, of AD 300–400. Bereket lies in the territory of Sagalassos, and the apparently early, and very abrupt, disappearance of pollen from crop plants in the record from this site might plausibly be explained by historical events specific to this Classical city; the city was fortified for defence around this time, for instance. On the other hand, the Gravgaz pollen site lies in the territory of Sagalassos
19 Baillie (1991). 20 Bryce (2005). 21 Gravgaz: Vermoere (2004); Bakker et al. (2012b). Bereket: Kaniewski et al. (2007); Bakker et al. (2013). Gölhisar: Eastwood et al. (1999). Iznik: Miebach et al. (2016); Ülgen et al. (2012). Nar lake: England et al. (2008).
too, and the pollen records from several separate cores here indicate that agriculture continued to flourish for a further two-three centuries, despite an earthquake that damaged Sagalassos in AD 518. Archaeological evidence indicates that Sagalassos was not abandoned until the mid 7th c. following another earthquake, in line with the end-date for the BOP at Gravgaz. The apparently early BOP end-date at Bereket may be partly explained by an abrupt change in the style of sedimentation at this marsh in Late Classical times. There was rapid accumulation of silts and peat here between ca. 400 BC and ca. AD 300, which allowed an exceptionally high-resolution pollen sampling interval, up to 1–2 years, in the BKT1/2 core.22 After this time, however, sedimentation slowed dramatically, so that the mean time interval between AD 300 and 1900 declines to one pollen sample per century in this diagram. This change in sedimentation may have included a depositional lacuna in the BKT1/2 core. In a second core from Bereket (SA09JBDrill02) a better stratigraphic continuity was found for the last two millennia.23 The pollen diagram for this core shows a decline in olive pollen at ca. AD 330, but with a second short-lived peak occurring in the early 7th c. In between the two, the vegetation in this valley was dominated by grasses and other herbs, from which it has been inferred that in this rather peripheral valley, crop cultivation diminished and was largely replaced by pastoralism in the 5th and 6th c.24 A second and more important change in land cover occurred in the early 8th c., with increasing pine pollen and a decrease in anthropic indicators. The Bereket valley thus seems to represent a local variant in the BOP story, but one in which the main landscape abandonment occurred at the same time as elsewhere in south-west Anatolia. Another way to evaluate pollen-inferred land-cover change is by using evidence from systematic regional archaeological site surveys for changes in rural settlement. Izdebski and others summarised the results of six such archaeological surveys from southern Anatolia and the Levant during the 1st millennium AD.25 In four of the surveys, settlement numbers were maintained through until the 7th c., when there was a marked decline, presumably linked to rural depopulation. In the fifth survey (Balboura), a similar decline took place in the 8th c. The Balboura data can be compared with the nearby pollen record from Gölhisar, where tree crop pollen fell to zero values between AD 600 and 735 (fig. 8).26
22 Kaniewski et al. (2007). 23 Bakker et al. (2013). 24 Kaptijn et al (2013). 25 Izdebski et al. (2016). 26 Balboura: Coulton (2012). Gölhisar: Eastwood et al. (1999).
60
Roberts
figure 7 (A) Onset and (B) end dates for the BOP in individual pollen records (see table 1 for details); histograms show the number of sites per century, along with a two-point running mean.
table 1
List of pollen sites with start and end dates for the BOP. STL=Santorini Tephra Layer, ATL=Avellino Tephra Layer. Partly based on Haldon et al. (2014).
Site number
Site name
Estimated start date
Estimated end date
Basis of chronology (number of 14C dates)
Original publication
Anatolia 1 2 3 4 5 6
Bafa (n/Miletus) Köyceğiz Söğüt Gölhisar Pınarbaşı Ağlasun
not clear 13th–16th c. BC 10–14th c. BC 10th-13th c. BC 6th c. BC ~10th c. BC
3rd c. AD 9th c BC 4th–5th c. AD mid-7th–mid-8th c. AD 9th–10th c. AD 7th–8th AD
14C-based (3) 14C-based (2) +STL 14C-based (2) +STL 14C-based (11) +STL 14C-based (2) 14C-based (5)
7
Bereket
4th c. BC
4th c. AD
14C-based (11)
8 9 10 11 12 13 14 15
Gravgaz Beyşehir Gölü I Hoyran Adliye (n/Iznik) Göksü (Iznik) Manyas Küçük Akgöl Melen
4th c. BC ~1800 BC ~4th c. BC not known not known not clear 8th c. BC 20th c. BC
Mid-7th c. AD 4th–6th c. AD 4th–6th c. AD 6th–7th c. AD 6th–7th c. AD 8th c. AD 5th–6th c. AD 10th–11th c. AD
14C-based (7) 14C-based (2) 14C-based (1) 14C-based (2) 14C-based (2) 14C-based (2) 14C-based (2) 14C-based (1)
Knipping et al. (2007) van Zeist et al. (1975) van Zeist et al. (1975) Eastwood et al. (1999) Bottema et al. (1984) Vermoere (2004); Bakker et al. (2012b) Kaniewski et al. (2007); Bakker et al. (2013) Bakker et al. (2012b) van Zeist et al. (1975) van Zeist et al. (1975) Argant (2003) Argant (2003) Leroy et al. (2002) Bottema et al. (1993-94) Bottema et al. (1993-94)
61
Revisiting the Beyşehir Occupation Phase table 1
List of pollen sites with start and end dates for the BOP. STL=Santorini Tephra Layer, ATL=Avellino Tephra Layer (cont.).
Site number
Site name
Estimated start date
Estimated end date
Basis of chronology (number of 14C dates)
Original publication
16 17
Abant Yenicağa
9th c. BC 11th c. BC
11th c. AD 4th c. AD
14C-based (5) 14C-based (7)
18 19 20 21
Ladik Kaz Demiryurt Nar
9th c. BC 4th c. BC not recorded 5th c. BC
8th–9th c. AD 5th–6th c. AD 6th–7th c. AD AD 670s
14C-based (4) 14C-based (2) 14C-based (1) Varve years
22 23
Iznik Sapanca
3rd c. BC not clear
6th–8th c. AD 6th AD
14C-based (14) + ATL 14C-based (11)
Bottema et al. (1993-94) Beug and Bottema (2015) Bottema et al. (1993-94) Bottema et al. (1993-94) Bottema et al. (1993-94) England et al. (2008); Roberts et al. (2016) Miebach et al. (2016) Leroy et al. (2010)
Cyprus 24
Larnaca salt lake
8th c. BC
4th–5th AD
14C-based (3)
Kaniewski et al. (2013)
Birket Ram Ze’elim A-2, Dead Sea Ein Feshkha, Dead Sea DS7, NE Ein Gedi, Dead Sea Kinneret
5–6th c. BC 9–7th c. BC
2nd–4th AD 7th c. AD
14C-based (12) 14C-based (14)
Neumann et al. (2007b) Neumann et al. (2007a)
12–9th c. BC
7th–8th c. AD
14C-based (8)
Neumann et al. (2007a)
4th–2nd c. BC
7th–8th c. AD
14C-based (5)
4th c. BC
2nd–3rd c. AD
14C-based (6)
Neumann et al. (2010), Leroy (2010) Baruch (1986); Schiebel (2013)
Levant 25 26 27 28 29
Although the apparent ca. 100-year time lag between the archaeological and pollen records could be an artefact of dating imprecision, it is equally possible that rural settlement in this rather remote highland location was able to persist for some time after the decline in organised arboriculture in the adjacent valleys. Overall, there is good temporal correspondence between archaeological and palaeoecological evidence from Anatolia and the Levant, in showing a decline in rural settlement at the same time that the BOP came to an end. In a few pollen records (e.g. Bereket) this termination apparently occurred in the 4th–5th c., while Izdebski has proposed that pollen diagrams from western Anatolia (e.g. Bithynia, Miletus) show that the BOP may have ended somewhat later, perhaps in the 11–12th c.27 In the majority of pollen records, however, the BOP is dated as ending during the 7th c., that is, at the same time that the eastern Roman empire came under sustained attack from Arab armies. These military 27 Izdebski (2013a), (2016).
incursions led to the Islamic conquest of Byzantine provinces in Egypt and the Levant. It also led to a period of more than two centuries when central and southwest Anatolia became a frontier zone that was regularly harassed by Arab forces. These lands were marked by insecurity and rural depopulation, in part linked to the system of Byzantine military defence.28 The pollen evidence thus indicates that the agrarian system that had operated during antiquity ended in most of Anatolia and the Levant as the eastern Roman empire came under attack by the new and expansionist caliphates. This would appear to support the view of Henri Pirenne that it was the expansion of Islam, rather than invasion by northern (e.g. Germanic) tribes, that marked the true end of the Classical world system, at least in the eastern Mediterranean.29
28 Haldon and Kennedy (1980); England et al. (2008). 29 Pirenne (1956).
62
Roberts
figure 8 (a) Changes in pollen-inferred land use and climate (inferred from oxygen isotopes) at Gölhisar (core GHA-92) compared with (b) Byzantine rural potsherd (bars) and general site numbers (solid line) in the countryside per century, from the Balboura archaeological survey. The abandonment phase from the 8th to the 11th c. is highlighted. Early Islamic Avlan-ware potsherd counts (mainly from the 11th through the 13th c.) are not shown. (after Haldon et al. (2014); data from Coulton (2012), Eastwood et al. (1999) and (2007)).
The Role of Climatic Change Correlation is not causation, however, and it is important to evaluate other potential explanations for the end of the BOP. In particular, there has been strong recent interest in the role that climate change may have played in the decline of the Roman empire, as well as other major socio-political events during the mid 1st millennium AD.30 For example, a ‘Late Antique Little Ice Age’ has been identified between AD 536 and ca. 660 from tree-ring records in Eurasia, and a link has been made between this period of cool summers and human migrations, political turmoil and wider societal reorganisations at this time.31 Also, there is evidence from both proxy data and documentary records of severe, snowy winter weather during the mid 1st millennium AD.32 The overall decline in temperatures between the 3rd and 6th c. would have been felt especially keenly because unusually warm conditions affected Europe during earlier Roman times, especially during the 1st and 3rd c. AD.33 This climatic cooling may have adversely affected frostsensitive crops in upland regions, such as the Taurus
30 E.g. McCormick et al. (2012). 31 Büntgen et al. (2016). 32 Proxy data: e.g. Dean et al. (2013). Documentary records: e.g. Telelis (2008). 33 Luterbacher et al. (2016).
mountains of south-west Anatolia. The upper agroclimatic limit for olive and vine cultivation here was almost certainly at higher elevations in ‘Roman Imperial’ times than it was after AD 536.34 Thus, climatic cooling is likely to have been partly responsible for the reduction of olive cultivation in upland regions of Anatolia during the 1st millennium AD. However, its effects may have been lagged, with only a few new trees being replanted here during the mid-Byzantine economic recovery of the 10th–11th c. The onset of the Late Antique Little Ice Age also coincided with the first wave of the Justinian plague pandemic (AD 540 onwards),35 which would have had a substantial negative demographic impact, and may have led to partial rural depopulation. Any decline in the Roman imperial economy linked to increasing droughts, floods, snow and frost should, in turn, be registered alongside a corresponding decline in tree crops and other indicators of agrarian prosperity in regional pollen records. However, as noted above, in most eastern Mediterranean pollen records, the BOP did not end in the mid 6th c., but rather one or two centuries later, implying that there was no simple cause-and-effect relationship between climatic deterioration and agrarian decline. Some researchers have attempted to use
34 See Roberts (1990) and England et al. (2008) for discussion of the evidence. 35 Stathakopoulos (2004); Hirschfeld (2006).
Revisiting the Beyşehir Occupation Phase
63
pollen to reconstruct both climate and human impact, by partitioning plant taxa between different controlling agencies via multi-variate statistics.36 However, this approach is potentially hazardous and runs the risk of circular reasoning.37 For instance, many common taxa (e.g. grass pollen) cannot be attributed clearly to just climate or human disturbance, but were likely affected by both agencies, and the same pollen indicator cannot be used to estimate climate as well as land use. Not surprisingly, when this has been done, there is often a close similarity between reconstructed parameters, such as reconstructed rainfall and agrarian indicators. An alternative and more rigorous test of societalclimate relations can be derived by using a multi-proxy approach, notably for lake sedimentary records, which possess both pollen (e.g. crop plants) and independent climate proxies (e.g. stable isotopes). This approach was used at Gölhisar,38 which showed an overall trend from wetter to drier climatic conditions after ca. AD 700, at the same time as rural settlement declined and tree crops diminished (fig. 8). However, this record lacks temporal detail for the 1st millennium AD, partly because the sediments at the bottom of this shallow lake were subject to mixing by wind turbation, leading to smoothing of the stratigraphic record. A more detailed and better-dated sedimentary sequence comes from Nar lake in Cappadocia, for which high-resolution pollen, stable isotopes and other proxies are available.39 Importantly, the sediments of this lake are annuallylaminated, or varved, which not only provides a very well-resolved core chronology, but which were also unaffected by mixing at the mud-water interface. Selected pollen taxa from the Nar record during Late Antiquity are shown in figure 9, which includes the latter part of the BOP, with a strong presence of olive, vine, walnut and cereals. The subsequent phase saw a reduction or disappearance in the pollen of all these crops, that was matched by an increase in the pollen for forest trees (mainly oak), implying the abandonment of agricultural land and woodland regeneration. According to varve counting, the end of the BOP in the NAR01 core took no more than 20 years, between AD 662 and 682, coinciding in time with some of the first major Arab attacks on Byzantine Anatolia.40 Cappadocia was directly exposed to hostile forces from the 650s, that led up to a great invasion in AD 674, which culminated in a four-year siege of
Constantinople. Thus, there is a very close chronological and spatial congruence between historically-attested Arab-Byzantine warfare and pollen-inferred rural abandonment in central Anatolia. What about possible changes in climate at this time? Oxygen-isotope data for Nar, based on sedimentary carbonates, show significant variations during the centuries after AD 300 (fig. 9), which reflect shifts between wetter and drier hydro-climatic conditions. If climate changes had acted as a direct trigger for abrupt rural abandonment, then the most likely mechanism would be an extended period of drought in central Anatolia, where most agriculture has traditionally been rain-fed rather than based on irrigation. A causal linkage between megadrought and social ‘collapse’ has been postulated for a number of complex pre-modern societies.41 However, the Nar Lake stable isotope record shows that dry climatic conditions prevailed between AD 400–500,42 that is, two centuries before the end of the BOP in this pollen sequence. Examination of climatic and cultural history from the same sedimentary archive provides a rigorous test of the hypothesis that the late antique societal crisis was prompted or inflamed by climatic stress, and in this case it can be falsified. That is not to say that climate change played no role in rural abandonment across Anatolia during the mid 1st millennium AD, but any contribution it did make must have been indirect and/ or lagged in time. In the southern Levant, any possible link between shifts in climate and the agrarian economy is currently more difficult to evaluate than in Anatolia. There are well-resolved climate records in this area of the Levant, such as from Soreq cave, and there are several pollen records from the Jordan valley.43 However, these are separate sites dated independently of each other, and correlation between them introduces a chronological uncertainty of at least ±100 years, so making it hard to establish a rigorous temporal sequencing of events. One of the key climate proxies in the Jordan rift comes from fluctuations in the water level of the Dead Sea, a non-outlet hyper-saline lake. Past water levels can be reconstructed directly from radiocarbon dating of former lake shorelines, or indirectly from the sedimentology of lake cores.44 These show a period of high lake levels in the 5th c., which may correspond, at least partly, to a phase of wetter climate in central and western
36 E.g. Bakker et al. (2012a). 37 See Izdebski et al. (2016) for further discussion. 38 By Eastwood et al. (2007). 39 England et al. (2008); Jones et al. (2006); Woodbridge and Roberts (2011). 40 Haldon and Kennedy (1980).
41 E.g. Tainter (1990); DeMenocal (2001); Diamond (2005). 42 Jones et al. (2006). 43 Soreq cave: Orland et al. (2009). Jordan valley: e.g. Neumann et al. (2010). 44 Bookman et al. (2004); Enzel et al. (2003); Migowski et al. (2006).
64
Roberts
figure 9 Multi-proxy records of changes in climate and pollen-inferred land cover during the 1st millennium AD at Lake Nar, central Anatolia (data from Jones et al. (2006); England et al. (2008)).
Anatolia.45 In the southern Levant, it was followed by a period of falling lake levels and inferred climatic desiccation starting in the 7th c. AD, which persisted until later Medieval times. This has fed arguments that drying of the climate prompted the movement of nomads out of the Arabian Desert and into the contiguous, better-watered ‘sown’ lands of the Fertile Crescent. In a sediment sequence taken off Ein-Gedi, on the western shore of the Dead Sea, Leroy was able to study both pollen and infer past lake levels in the same core, without the need for inter-site correlation.46 In core DS7-1SC sediment facies changed from deeper-water laminites to shallower-water halite not prior to, but after the pollen change marking the end of the BOP (fig. 10). On that basis, the shift from olive-cereal cultivation to pastoralism must have taken place a few decades before the transition to a drier climate, which then could not have been the direct cause of land abandonment. In this scenario, the Arab conquest of Palestine at ca. AD 640 would have been responsible for the major change in rural land use recorded as the end of the BOP, but with subsequent climatic deterioration reinforcing this shift and making it more difficult to reverse.
45 Haldon et al. (2014); Izdebski et al. (2016). 46 Leroy (2010).
Post-BOP Landscape Change If the latter part of the BOP pollen phase can be seen as representing the agrarian landscape of the eastern Mediterranean during antiquity, what came afterwards? In most pollen records from this region there was a postBOP decline in tree crops, such as olive and walnut, and an increase in non-cultivated trees, implying that some agricultural land reverted to secondary woodland. In south-west Anatolia, the ‘re-wilding’ of the landscape was dominated by pine, whereas in central Anatolia and the southern Levant, oak was more important in the process of reforestation; deciduous oak in the former, evergreen oak in the latter. One significant feature is that the woodland composition was usually different after the BOP than it had been prior to the BOP. For instance, at Söğüt, pre-BOP forests had included juniper and oak alongside pine, whereas post-BOP vegetation was pinedominated (fig. 3). At Birket Ram in the Golan Heights, the post-BOP vegetation was predominantly evergreen oak, compared to the deciduous oak woodland that had existed at the start of the 1st millennium BC (fig. 6). The magnetic susceptibility profile for this Birket Ram core gives one indication why this may have been the case, namely soil erosion during and after the BOP, which would have changed the growing conditions for secondary woodland. Magnetic susceptibility is a measure of detrital sediment inwash from eroded
65
Revisiting the Beyşehir Occupation Phase
figure 10 Dead Sea pollen record from core DS7–1SC during and after the BOP (modified from Leroy (2010)). Note how the change in sediment lithology, indicative of lower lake levels, occurs after the end of the BOP.
soils and bedrock, and at Birket Ram erosion rates increased sharply as woods were replaced by olive groves and pastureland, especially during the early part of the BOP.47 The same is true in some other pollen records, such as Gravgaz, where sedimentation rates due to soil inwash peaked very markedly at the start of the BOP.48 Doubtless other factors would also have influenced the species composition of the post-BOP woodland, including burning regimes and grazing by livestock herds.49 In the uplands of south-west Anatolia and in the southern Levant, the post-BOP landscape transformation was effectively a permanent one, up until modern times. By contrast, in Cappadocia and north-west Anatolia, land abandonment lasted two-three centuries but eventually gave way to a new form of land use, in the former case based on agro-pastoralism rather than on arboriculture. The Byzantine expulsion of Arab armies from central Anatolia during the later 9th and early 10th c. led to a progressive resettlement of the Cappadocian countryside, reflected in pollen evidence indicating renewed forest clearance, a reoccupation of the land that was probably aided by progressively wetter climatic conditions from AD 860 to 1030 (fig. 9). The mid Byzantine ‘golden age’ after AD 950 is reflected in an increasing proportion of pollen from cereal grain crops at Lake Nar, but relatively few pollen from tree crops. Thus, in central Anatolia, there was a clear rupture between the agrarian landscape of Late Antiquity and that which re-emerged in medieval times.
47 Schwab et al. (2004). 48 Dusar et al. (2012). 49 Roberts (1990).
There is only a single relevant pollen diagram from Cyprus, from Larnaca salt lake, and it is poorly dated for the period of Late Antiquity.50 Nonetheless, it is significant in showing a decline in agrarian pollen (cereals, olive) during the mid 1st millennium AD, but without a complete rupture, as in Cappadocia, nor a major reforesting of the landscape, as seen in south-west Anatolia. This greater continuity in landscape ecology and rural economy is in accord with the distinctive history of the island at this time. Although the Arabs invaded Cyprus in the 650s, in AD 688 the Emperor Justinian II and the Caliph Abd al-Malik reached an unprecedented agreement. For the next three centuries, the island was ruled jointly by both Arabs and Byzantines as a condominium, despite the nearly constant warfare between the two parties on the mainland. The resulting maintenance of relative economic prosperity led to a gradual metamorphosis of the classical agrarian landscape into the medieval one, when the Byzantines once again took sole control of Cyprus. Moving east of the former Roman frontier into those lands under Persian domination, it is not easy to recognise a clear equivalent to the BOP, nor was there any clear rupture in land cover, judging from pollen records, in eastern Anatolia and north-west Iran following the 7th c. Islamic conquest.51 Thus, pollen evidence suggests that the relatively uniform and economically integrated agrarian system of Late Roman times gave way during the 7th c. AD to
50 Kaniewski et al. (2013). 51 Eastern Anatolia: Lake Van: Wick et al. (2003). North-west Iran: Lake Urmia: Bottema (1986); Lake Almalou: Djamali et al. (2009).
66 a much more heterogeneous pattern, with strong interregional differences in landscape and culture across the eastern Mediterranean. Even where the countryside continued to be inhabited after the mid 7th c., there appears to have been a major demographic contraction, and the rural economy no longer focused on large-scale cultivation of tree crops, such as olive. This heterogeneity in rural economic geography is highlighted in the contrasting regional trends for cereal-type pollen found in western and northern Anatolia, on the one hand, and Greece/Macedonia on the other, during the early medieval period.52 Rural economies thus appear to have become increasingly fragmented and independent of each other in the post-BOP world. Conclusion The Beyşehir Occupation Phase (or BOP) was spatiotemporally heterogeneous, and better chronologies for its onset and end are needed to further investigate links with archaeological/historical evidence and/or climatic changes. Nonetheless, it is clear that the latter part of the BOP corresponds in space and time to the Late Roman empire in the eastern Mediterranean. The emphasis on tree crops in pollen records, particularly olive groves, implies long-term investment, stable trade networks and regional economic integration. During the mid 1st millennium AD, the BOP came to an end, often abruptly, with a marked decline in tree crops and other agricultural indicators, and an increase in woodland, implying rural decline or abandonment and partial landscape re-wilding. This termination is most commonly dated to the 7th c. AD, coinciding with the Arab conquest of Byzantine lands in the Levant, and their raiding and invasions of Byzantine Anatolia. In the Levant, this time was also marked by a desiccation of the climate, although it is not clear whether this coincided exactly in time with, and therefore might have triggered, the collapse of the rural agrarian system. In central Anatolia there was no desiccation trend at the time when the BOP ended, so climate change cannot have been a direct cause. Finally, we can ask why the BOP, with its rather abrupt ending, is so much more evident in Anatolia and the Levant than in the equivalent phases of pollen diagrams from Persia, the Balkans or Italy? Although it is not possible to give a definitive answer to this question, it may be precisely because economic prosperity was so marked in the eastern Mediterranean during Late Antiquity,53 notwithstanding the impact of the Plague
52 Xoplaki et al. (2016); Roberts et al. (forthcoming). 53 Izdebski (2011).
Roberts
of Justinian. During the 5th–6th c., a wetter climate allowed agriculture to flourish in agro-climatically marginal lands as far south as the Negev and Libya, beyond limits that were sustainable in the long term.54 Because the countryside remained so prosperous until so late, the ending of the late antique rural economy here— when it finally occurred, following Arab invasions and conquest during the 7th c.—is especially clearly marked in the pollen records. The system rupture consequently appears to have been much greater here than it was either in Persia or in the central/western Mediterranean. In Anatolia and the Levant, it seems that antiquity ended with a bang rather than a whimper. Acknowledgements I am grateful to Warren Eastwood, David Kaniewski, Henk Woldring, Jessie Woodbridge, and the anonymous reviewers of this paper. I am also grateful to the European Pollen Database contributors in providing the pollen data, and to the Plymouth University GeoMapping unit for cartographic assistance, particularly Shaun Lewin and Tim Absalom. Bibliography Argant J. (2003) “Données palynologiques”, in La Bithynie au moyen âge, edd. B. Geyer and J. Lefort (Paris 2003) 175–200. Baillie M. G. L. (1991) “Suck-in and smear: two related chronological problems for the 90s”, Journal of Theoretical Archaeology 2 (1991) 12–16. Bakker J., Paulissen E., Kaniewski D., Poblome J., De Laet V., Verstraeten G. and Waelkens M. (2013) “Climate, people, fire and vegetation: new insights into vegetation dynamics in the eastern Mediterranean since the 1st century AD”, Climate of the Past 9 (2013) 57–87. Bakker J., Kaniewski D., Verstraeten G., De Laet V. and Waelkens M. (2012a) “Numerically derived evidence for Late-Holocene climate change and its impact on human presence in the southwest Taurus Mountains, Turkey”, The Holocene 22 (2012) 425–38. Bakker J., Paulissen E., Kaniewski D., De Laet V., Verstraeten G. and Waelkens M. (2012b) “Man, vegetation and climate during the Holocene in the territory of Sagalassos, western Taurus mountains, SW Turkey”, Vegetation History and Archaeobotany 21 (2012) 249–66. Baruch U. (1986) “The Late Holocene vegetational history of Lake Kinneret (Sea of Galilee), Israel”, Paléorient 12 (1986) 37–48. Beug H. J. and Bottema S. (2015) “Late Glacial and Holocene vegetation history at Lake Yeniçağa, northern Turkey”, Vegetation History and Archaeobotany 24 (2015) 293–301. Bookman R., Enzel Y., Agnon A. and Stein M. (2004) “Late Holocene lake levels of the Dead Sea”, Geological Society of America Bulletin 116 (2004) 555–71.
54 Negev: Rosen (2000). Libya: Gilbertson et al. (2000).
Revisiting the Beyşehir Occupation Phase Bottema S. (1986) “A Late Quaternary pollen diagram from Lake Urmia (northwestern Iran)”, Review of Paleobotany and Palynology 47 (1986) 241–61. Bottema S. and Woldring H. (1990) “Anthropogenic indicators in the pollen record of the eastern Mediterranean”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape, edd. S. Bottema, G. Entjes-Nieborg, and W. van Zeist (Balkema, Rotterdam 1990) 231–64. Bottema S. and Woldring H. (1984) “Late Quaternary vegetation and climate of southwestern Turkey, part II”, Palaeohistoria 26 (1984) 123–49. Bottema S., Woldring H. and Aytuğ B. (1993–94) “Late Quaternary vegetation history of northern Turkey”, Palaeohistoria 35/36 (1993–94) 13–72. Bottema S., Woldring H. and Aytuğ B. (1986) “Palynological investigations on the relations between prehistoric man and vegetation in Turkey: the Beyşehir Occupation Phase”, in Proceedings of the 5th Optima Congress, (Istanbul), edd. H. Demiriz and N. Özhatay (Istanbul 1986) 315–28. Bryce T. R. (2005) The Kingdom of the Hittites (Oxford 2005). Büntgen U., Myglan V. S., Ljungqvist F. C. et al. (2016) “Cooling and societal change during the Late Antique Little Ice Age from 536 to around 660 AD”, Nature Geoscience 9 (2016) 231–36. Coulton J. J. (2012) “Late Roman and Byzantine Balboura”, in The Balboura Survey and Settlement in Highland Southwest Anatolia. 1. Balboura and the History of Highland Settlement, ed. J. J. Coulton (London 2012) 163–84. Dean J. R., Jones M. D., Leng M. J. et al. (2013) “Palaeo-seasonality of the last two millennia reconstructed from the oxygen isotope composition of diatom silica and carbonates from Nar Gölü, central Turkey”, Quaternary Science Reviews 66 (2013) 35–44. DeMenocal P. B. (2001) “Cultural responses to climate change during the Late Holocene”, Science 292 (2001) 667–73. Diamond J. (2005) Collapse. How Societies Choose to Fail or Succeed (London 2005). Díaz de la Guardia C., Alba F., Del Mar Trigo M., Galán C., Ruíz L. and Sabariego S. (2003) “Aerobiological analysis of Olea europaea L. pollen in different localities of southern Spain: forecasting models”, Grana 42 (2003) 234–43. Djamali M., de Beaulieu J.-L., Andrieu-Ponel V., Berberian M., Miller N. F., Gandouin E., Lahijani H., Shah-Hosseini M., Ponel P., Salimian M. and Guiter F. (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”, JAS 36 (2009) 1364–75. Dusar B., Verstraeten G., D’haen K., Bakker J., Kaptijn E. and Waelkens M. (2012) “Sensitivity of the eastern Mediterranean geomorphic system towards environmental change during the Late Holocene: a chronological perspective”, Journal of Quaternary Science 27 (2012) 371–82. Eastwood W. J., Leng M. J., Roberts N. and Davis B. (2007) “Holocene climate change in the eastern Mediterranean region: a comparison of stable isotope and pollen data from Lake Gölhisar, southwest Turkey”, Journal of Quaternary Science 22 (2007) 327–41. Eastwood W. J., Roberts N., Lamb H. F. and Tibby J. C. (1999) “Holocene environmental change in southwest Turkey: a palaeoecological record of lake and catchment-related changes”, Quaternary Science Reviews 18 (1999) 671–96. Eastwood W. J., Roberts N. and Lamb H. F. (1998) “Palaeoecological and archaeological evidence for human occupance in southwest Turkey: the Beyşehir Occupation Phase”, AnatSt 48 (1998) 69–86. England A., Eastwood W. J., Roberts N., Turner R. and Haldon J. F. (2008) “Historical landscape change in Cappadocia (central Turkey): a palaeoecological investigation of annually-laminated sediments from Nar lake”, The Holocene 18 (2008) 1229–45.
67 Enzel Y., Bookman R., Sharon D., Gvirtzman H., Dayan U., Ziv B. and Stein M. (2003) “Late Holocene climates of the Near East deduced from Dead Sea level variations and modern regional winter rainfall”, Quaternary Research 60 (2003) 263–73. Gaillard M.-J., Sugita S., Mazier F. et al. (2010) “Holocene land cover reconstructions for studies on land cover-climate feedbacks”, Climate of the Past 6 (2010) 483–99. Gilbertson D., Hunt C. and Gillmore G. (2000) “Success, longevity, and failure of arid-land agriculture: Romano-Libyan floodwater farming in the Tripolitanian pre-desert”, in The Archaeology of Drylands. Living at the Margin, edd. G. Barker and D. Gilbertson (London–New York 2000) 137–59. Haldon J. F. and Kennedy H. (1980) “The Byzantine—Arab frontier in the eighth and ninth centuries: military organisation and society in the borderlands”, Zbornik Radova Vizantološkog Instituta 19 (1980) 79–116. Haldon J., Izdebski A., Roberts N. et al. (2014) “The climate and environment of Byzantine Anatolia: integrating science, history and archaeology”, Journal of Interdisciplinary History 45 (2014) 113–61. Hall A. S. (1994) “Sinda”, in Studies in the History and Topography of Lycia and Pisidia. In Memoriam A. S. Hall, ed. D. H. French (British Institute of Archaeology at Ankara, Monograph 20) (London 1994) 48–52. Heim C., Nowaczyk N. R., Negendank J. F. W., Leroy S. A. G. and BenAvraham Z. (1997) “Near East desertification: evidence from the Dead Sea”, Naturwissenschaften 1984 (1997) 398–401. Hirschfeld Y. (2006) “The crisis of the sixth century: climatic change, natural disasters and the plague”, Mediterranean Archaeology and Archaeometry 6 (2006) 19–32. Horden P. and Purcell, N. (2000) The Corrupting Sea: a Study of Mediterranean History (Oxford 2000). Izdebski A. (2016) “Byzantine Miletus. Environmental history of the hinterland”, AA (2016.2) 270–80. Izdebski A. (2013a) A Rural Economy in Transition. Asia Minor from Late Antiquity into the Early Middle Ages, (Journal of Juristic Papyrology Supplement Series) (Warsaw 2013). Izdebski A. (2013b) “The economic expansion of the Anatolian countryside in Late Antiquity: the coast versus inland regions”, in Local Economies? Production and Exchange of Inland Regions in Late Antiquity, ed. L. Lavan (Late Antique Archaeology 10) (Leiden 2013) 343–76. Izdebski A. (2011) “Why did agriculture flourish in the late antique East? The role of climate fluctuations in the development and contraction of agriculture in Asia Minor and the Middle East from the 4th till the 7th c. AD”, Jahrbuch zu Kultur und Geschichte des ersten Jahrtausends n. Chr. 8 (2011) 291–312. Izdebski A., Pickett J., Roberts N. and Waliszewski T. (2016) “The environmental, archaeological and historical evidence for regional climatic changes and their societal impacts in the eastern Mediterranean in Late Antiquity”, Quaternary Science Reviews 136 (2016) 189–208. Jones M. D., Roberts N., Leng M. J. and Türkeş M. (2006) “A highresolution Late Holocene lake isotope record from Turkey and links to North Atlantic and monsoon climate”, Geology 34 (2006) 361–64. Kaniewski D., Paulissen E., De Laet V. and Waelkens M. (2007) “3000 years BP high-resolution ecological landscape history inferred from an intramontane basin in the western Taurus mountains, Turkey”, Quaternary Science Reviews 26 (2007) 2201–18. Kaniewski D., Van Campo E., Guiot J., Le Burel S., Otto T. et al. (2013) “Environmental roots of the Late Bronze Age crisis”, PLoS One 8 (8), e71004: http://dx.doi.org/10.1371/journal.pone.0071004 (accessed February 2017). Kaptijn E., Poblome J., Vanhaverbeke H., Bakker J. and Waelkens M. (2013) “Societal changes in the Hellenistic, Roman and Early
68 Byzantine periods. Results from the Sagalassos Territorial Archaeological Survey 2008 (southwest Turkey)”, AnatSt 63 (2013) 75–95. Knipping M., Müllenhof M. and Brückner H. (2007) “Human induced landscape changes around Bafa G.lü (western Turkey)”, Vegetation History and Archaeobotany 17 (2007) 365–80. Leroy S. A. G. (2010) “Pollen analysis of core DS7–1SC (Dead Sea) showing intertwined effects of climatic change and human activities in the Late Holocene”, JAS 37 (2010) 306–13. Leroy S. A. G., Kazancı N., Ileri Ö., Kibar M., Emre O., McGee E. and Griffiths H. I. (2002) “Abrupt environmental changes within a Late Holocene lacustrine sequence south of the Marmara Sea (Lake Manyas, N-W Turkey): possible links with seismic events”, Marine Geology 190 (2002) 531–52. Leroy S. A. G., Schwab M. J. and Costa P. J. M. (2010) “Seismic influence on the last 1500-year infill history of Lake Sapanca (North Anatolian Fault, NW Turkey)”, Tectonophysics 486 (2010) 15–27. Luterbacher J. et al. (2016) “European summer temperatures since Roman times”, Environmental Research Letters 11, 024001: doi:10.1088/1748-9326/11/2/024001 (accessed February 2017). McCormick M., Büntgen U., Cane M. A. et al. (2012) “Climate change during and after the Roman empire: reconstructing the past from scientific and historical evidence”, Journal of Interdisciplinary History 43 (2012) 169–220. Mattingly D. J. (1996) “First fruit? The olive in the Roman world”, Human Landscapes in Classical Antiquity: Environment and Culture 6 (1996) 213–53. Mercuri A. M., Mazzanti M. B., Florenzano A., Montecchi M. C. and Rattighieri E. (2013) “Olea, Juglans and Castanea: the OJC group as pollen evidence of the development of human-induced environments in the Italian peninsula”, Quaternary International 303 (2013) 24–42. Miebach A., Niestrath P., Roeser P. and Litt T. (2016) “Impacts of climate and humans on the vegetation in northwestern Turkey: palynological insights from Lake Iznik since the last Glacial”, Climate of the Past 12 (2016) 575–93. Migowski C., Stein M., Prasad S., Negendank J. F. W. and Agnon A. (2006) “Holocene climate variability and cultural evolution in the Near East from the Dead Sea sedimentary record”, Quaternary Research 66 (2006) 421–31. Neumann F. H., Kagan E. J., Schwab M. J. and Stein M. (2007a) “Palynology, sedimentology and palaeoecology of the Late Holocene Dead Sea”, Quaternary Science Reviews 26 (2007) 1476–98. Neumann F., Schölzel C., Litt T., Hense A. and Stein M. (2007b) “Holocene vegetation and climate history of the northern Golan heights (Near East)”, Vegetation History and Archaeobotany 16 (2007) 329–46. Neumann F. H., Kagan E. J., Leroy S. A. G. and Baruch U. (2010) “Vegetation history and climate fluctuations on a transect along the Dead Sea west shore and their impact on past societies over the last 3500 years”, Journal of Arid Environments 74 (2010) 756–64. Orland I. J., Bar-Matthews M., Kita N. T., Ayalon A., Matthews A. and Valley J. W. (2009) “Climate deterioration in the eastern Mediterranean as revealed by ion microprobe analysis of a speleothem that grew from 2.2 to 0.9 ka in Soreq Cave, Israel”, Quaternary Research 71 (2009) 27–35. Pirenne P. (1956) Mohammed and Charlemagne (transl. Bernard Miall, New York 1956). Roberts N. (1990) “Human-induced landscape change in south and south-west Turkey during the later Holocene”, in Man’s Role in the Shaping of the Eastern Mediterranean Landscape, edd. S. Bottema, G. Entjes-Nieborg, and W. van Zeist (Rotterdam 1990) 53–67. Roberts N., Allcock S. L., Arnaud F., Dean J. R., Eastwood W. J., Jones M. D., Leng M. J., Metcalfe S. E., Malet E., Woodbridge J. and Yiğitbaşıoğlu H. (2016) “A tale of two lakes: a multi-proxy comparison of Late Glacial and Holocene environmental change in
Roberts Cappadocia, Turkey”, Journal of Quaternary Science 31 (2016) 348–62. Roberts N., Cassis M., Doonan O., Eastwood W., Elton H., Haldon J., Izdebski A. and Newhard J. (2018) “Not the end of the world? PostClassical decline and recovery in rural Anatolia”, Human Ecology 46 (2018) 305–22. Rosen S. A. (2000) “The decline of desert agriculture: a view from the classical period Negev”, in The Archaeology of Drylands. Living at the Margin, edd. G. Barker and D. Gilbertson (London–New York 2000) 45–62. Schiebel V. (2013) Vegetation and Climate History of the Southern Levant During the Last 30,000 Years Based on Palynological Investigation (Ph.D. diss., Univ. of Bonn 2013). Schwab M. J., Neumann F., Litt T., Negendank J., and Stein M. (2004) “Holocene palaeoecology of the Golan Heights (Near East): investigation of lacustrine sediments from Birket Ram crater lake”, Quaternary Science Reviews 16–17 (2004) 1723–32. Stathakopoulos D. C. (2004) Famine and Pestilence in the Late Roman and Early Byzantine Empire: a Systematic Survey of Subsistence Crises and Epidemics (Birmingham Byzantine and Ottoman Monographs (Aldershot 2004). Sullivan D. G. (1989) Human-Induced Vegetation Change in Western Turkey: Pollen Evidence from Central Lydia (Ph.D. diss., Univ. of California, Berkeley 1989). Sullivan D. G. (1988) “The discovery of Santorini Minoan tephra in western Turkey”, Nature 333 (1988) 552–54. Tainter J. (1990) The Collapse of Complex Societies (Cambridge 1990). Telelis I. G. (2008) “Climatic fluctuations in the eastern Mediterranean and the Middle East AD 300–1500 from Byzantine documentary and proxy physical paleoclimatic evidence—a comparison”, JÖB 58 (2008) 167–207. Ülgen U. B., Franz S. O., Biltekin D., Çağatay M. N., Roeser P. A., Doner L. and Thein J. (2012) “Climatic and environmental evolution of Lake Iznik (NW Turkey) over the last ~4700 years”, Quaternary International 274 (2012) 88–101. Van Zeist W., Baruch U. and Bottema S. (2009) “Holocene palaeoecology of the Hula area, northeastern Israel”, in A Timeless Vale. Archaeological and Related Essays on the Jordan Valley, in Honour of Gerrit Van Der Kooij on the Occasion of his Sixty-Fifth Birthday (Leiden University Archaeological Studies 19) (Leiden 2009) 29–64. Van Zeist W., Woldring H. and Stapert D. (1975) “Late Quaternary vegetation and climate of southwestern Turkey”, Palaeohistoria 17 (1975) 55–143. Vanhaverbeke H. and Waelkens M. (2003) The Chora of Sagalassos: the Evolution of the Settlement Pattern from Prehistoric until Recent Times (Studies in Eastern Mediterranean Archaeology) (Turnhout 2003). Vermoere M. (2004) Holocene Vegetation History in the Territory of Sagalassos (Studies in Eastern Mediterranean Archaeology 6) (Turnhout 2004). Waliszewski T. (2014) Elaion. Olive Oil Production in Roman and Byzantine Syria-Palestine (Warsaw 2014). Wick L., Lemcke G. and Sturm M. (2003) “Evidence of Lateglacial and Holocene climatic change and human impact in eastern Anatolia: high-resolution pollen, charcoal, isotopic and geochemical records from the laminated sediments of Lake Van, Turkey”, The Holocene 13 (2003) 665–75. Woodbridge J. and Roberts N. (2011) “Late Holocene climate of the eastern Mediterranean inferred from diatom analysis of annuallylaminated lake sediment”, Quaternary Science Reviews 30 (2011) 3381–92. Xoplaki E., Fleitmann D., Luterbacher J., Wagner S., Haldon J., Zorita E., Telelis I., Toreti A. and Izdebski A. (2016) “The Medieval Climate Anomaly and Byzantium. A review of the evidence on climatic fluctuations, economic performance and societal change”, Quaternary Science Reviews 136 (2016) 229–52.
Regional Vegetation Histories: An Overview of the Pollen Evidence from the Central Mediterranean Katerina Kouli, Alessia Masi, Anna Maria Mercuri, Assunta Florenzano, and Laura Sadori Abstract Vegetation patterns during the 1st millennium AD in the central Mediterranean, exhibit a great variability, due to the richness of these habitats and the continuous shaping of the environment by human societies. Variations in land use, witnessed in the pollen record, reflect the role that local vegetation and environmental conditions played in the choices made by local societies. The interdisciplinary study of offsite cores remains the key evidence for palaeoenvironmental transformations mirroring the ‘semi-natural’ vegetation, and revealing temporal fluctuations and the amount of human impact on a regional scale.
Introduction The Mediterranean region is a hotspot of biodiversity, lying in a climatic transitional zone under the influence of both tropical and mid-latitude regimes,1 and therefore highly sensitive to regional and global climate changes.2 During the last 10,000 years, the Mediterranean mosaic of habitats has been continuously transformed by both climatic and cultural changes at the regional, local and micro-habitat scale.3 The central Mediterranean region shows significant latitudinal and altitudinal pattern variety. The Italian peninsula in its western part and the southern Balkan peninsula in its eastern part, with their long coastlines and mountainous chains (the result of recent geological processes), form a diverse Mediterranean landscape with several micro-climates and localised vegetation patterns. In addition, the two large islands of Sicily and Crete, in the south, add even more variables to an already varied set of environments. Several Holocene palaeovegetational studies have been carried out in the central Mediterranean region,4 1 Lionello et al. (2006) 1–26. 2 Rohling et al. (2002) 61–75; Sadori et al. (2016) 1423–37; Triantaphyllou et al. (2014) 1697–1712; Weiberg et al. (2016) 40–65. 3 Mercuri (2014) 1799–1810; Mercuri and Sadori (2014) 507–27; Mercuri et al. (2011) 189–206; Mercuri et al. (2013a) 24–42; Sadori et al. (2010b) 940–951; Sadori et al. (2013a) 141–56. 4 Mercuri et al. (2012) 353–72; Sadori et al. (2011) 117–29; CombourieuNebout et al. (2013) 2023–42.
nevertheless the succession of plant landscapes is far from being clearly understood. Pollen studies record a remarkable temporal and spatial variability. Even though the nature of Early Holocene vegetation is generally related to climate variability,5 distinguishing between climatic and human impact in the plant landscape is difficult for the Mid Holocene and impossible for the Late Holocene period.6 The cumulative activities of humans over thousands of years, coupled with the general trend in aridity of the Late Holocene, as well as short-term climatic fluctuations, have left their imprint on vegetation. The interaction between human societies and vegetation is continuous and diachronic, and has resulted in a palimpsest plant landscape.7 Already from the Bronze Age, human societies steadily increased their knowledge and ability to exploit natural resources,8 increasing the impact of the ‘human ecological factor’ on the landscape. Therefore, any discussion about vegetation history during the 1st millennium AD is a very intricate task. Pollen records covering this period in the central Mediterranean are rare, due mainly to oxidation and erosion, promoted by agricultural practices that have destroyed the upper sediments in drained lakes. Yet, some data from archaeological sites, offering a high number of pollen samples from very local contexts, are available from central Sicily (Villa Romana del Casale, Philosophiana), Sardinia (Sa Osa), and from rural (the Cinigiano area in southern Tuscany) and urban contexts (Mutina and Rome).9 At the same time, some pollen data from harbours (Rome, Naples, Pisa) and delta areas provide us with evidence of ‘semi-disturbed’
5 Kotthoff et al. (2008) 1019–32; Lawson et al. (2005) 873–87. 6 Roberts et al. (2011) 3–12; Sadori et al. (2011) 117–29; Sadori et al. (2013a) 141–56. 7 Mercuri (2008); Mercuri et al. (2010) 888–901; Kouli (2015) 195– 204; Gogou et al. (2016) 209–28. 8 Mercuri et al. (2006) 43–60; Mercuri and Sadori (2014) 507–27; Cremaschi et al. (2015) 264–76. 9 Villa Romana del Casale, Philosophiana: Montecchi and Mercuri (2018) 743–57; Mercuri et al. (2017). Sa Osa: Sabato et al. (2015) 205–15. Cinigiano area: Bowes et al. (2015) 174–84; Rattighieri et al. (2013) 207–15; Vaccaro et al. (2013) 129–79. Mutina and Rome: Bosi et al. (2015) 19–31; Giardini (2007) 301–16.
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_006 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 69–82
70 contexts.10 In this deeply anthropogenic framework, pollen from long-core lakes remains the classical record of key relevance to investigate long-term trends in regional vegetation history. Robust chronology correlation is, though, the ‘Achilles heel’ when trying to compare different pollen records, in an effort to draw a regional image of the vegetation. Herein we will present vegetation development based on ten pollen sites, selected from the available lake cores from central-southern Italy and the southern Balkans-Greece. We will focus on the 300 BC to the AD 1000 time interval. This represents the full late antique period (AD 250–750 for central-southern Italy11), set between the Early Roman and Early Medieval eras.
Kouli et al.
Pollen analysis is an invaluable tool for the study of past plant landscapes and their diachronic succession; it is also of use for gathering local and regional information about vegetation changes and for getting an idea as to the intensity of human land use. Pollen is produced by microsporangia of seed plants and is dispersed mainly by wind and insects. The anoxia of lake sediments is the best environment for the preservation of pollen grains. Pollen walls can survive for hundreds of thousands of years, maintaining a distinctive morphology of the plant that produced it, allowing in some cases an identification at the species level. Fossil pollen preserved in sediments can be extracted and analysed, providing a diachronic picture of the environment. The type and amount of pollen preserved can be extremely informative. Woody, grass, cereal and weed plants can tell us the history of past landscapes, land management, forest clearance and cultivation. In order to evaluate plant cover, selected pollen taxa can be grouped into sums, useful for environmental and land use reconstructions, based on local flora and vegetation (table 1).12 These sums provide rich data on the variable anthropogenic landscape of the Mediterranean area, where humans had a significant impact, and generated impressive patterns of landscape complexity.13 Therefore, a great deal of attention needs to be paid to the information on ‘natural’ vegetation, something which has in fact been influenced/transformed by the
spread of human activity since the Mid-Holocene, at least.14 The main sums that are useful for reconstructing natural plant cover composition, are: the subsets of the woody or Arboreal Pollen (AP); upland forest and mixed deciduous forest pollen (including upland and deciduous woodland taxa, respectively); along with specimens from Mediterranean trees and shrubs (mainly formed by evergreen oaks of the Mediterranean macchia). Pollen from riverine vegetation trees, and hydro-hygrophilous herb taxa represent wetlands; these sums describe the local environmental characteristics of wet habitats, especially reservoir/meadow areas and their near surroundings. The sum of steppe and grasslands indicators suggests an opening of the landscape, dominated by grasses and steppe elements (Non-Arboreal Pollen— NAP). In pollen diagrams, the ‘human impact indicators’ help to elucidate past human activity, and its impact on the environment. Clear signs of anthropogenic land transformation are marked by the curves of Olea, Juglans and Castanea pollen (OJC group); tree crops act as indicators of increasing anthropisation in the Mediterranean area.15 Olea pollen of the domesticated subspecies, is quite indistinguishable from that of the wild olive subspecies,16 and therefore a variable amount of uncultivated olive trees may contribute to the sum. Other cultivated trees and woody plants collected in the wild may be further important anthropogenic markers in pollen records. The clearest signal for human influence on the environment is given by the spread of cereals (Cerealia-type; primary anthropogenic indicators, sensu Behre17) and in general by the presence of cultivated/cultivable herbs. These plants are important evidence for land use and agricultural systems, even if the percentage of their pollen is under-represented in the spectra. Cereals, and most cultivated plants, are in fact low pollen producers,18 so even a small amount of their pollen may be an indication of cultivated fields. Wild synanthropic plants (weeds and ruderals, growing within farming contexts and on open disturbed ground) are commonly included among the human impact indicators (secondary indicators, sensu Behre19). Being involuntarily favoured by the spread of human activities, these herbs are particularly useful for reconstructing the dynamics of past and present
10 Harbours: Mariotti Lippi et al. (2007) 453–65, Pepe et al. (2013) 73–81; Russo Ermolli et al. (2014) 399–411; Sadori et al. (2010a) 3294–3305; Sadori et al. (2015) 435–44; Sadori et al. (2016) 173– 88. Delta areas: Bellotti et al. (2016) 1457–71; Sadori et al. (2016) 173–88; Triantaphyllou et al. (2016) 478–87. 11 Vaccaro (2013) 259–313. 12 Florenzano (2013). 13 Mercuri and Sadori (2014) 183–211.
14 Zohary et al. (2012); Roberts et al. (2011) 3–13. 15 Mercuri et al. (2013a) 24–42. 16 E.g. Messora et al. (2017) 204–14. 17 Behre (1990) 219–30. 18 Hall (1989) 63–69; Dreßler et al. (2006) 25–37. 19 Behre (1990) 219–30.
Reading and Interpreting the Pollen Records
71
Regional Vegetation Histories table 1
Pollen sums useful for palaeoenvironmental and vegetation reconstructions. Asterisks (*) mark the sums used in the present study.
‘Natural’ vegetation
Pollen sum
Taxa included
Arboreal pollen (AP)*
All trees and shrubs
Upland forest vegetation
Abies, Betula, Fagus, Picea, Pinus
Mixed deciduous forest*
Mediterranean trees and shrubs*
Acer campestre-type, Carpinus, Corylus, Fraxinus excelsior-type, Ostrya carpinifolia/Carpinus orientalis, broadleaved Quercus, Tilia, Ulmus Juniperus-type, Myrtus, Phillyrea, Pistacia, Quercus ilex-type
Riverine vegetation (trees)
Alnus, Populus, Salix
Wetland herbs
Alisma-type, Butomus, Callitriche, Ceratophyllum, Cyperaceae, Hydrocharis, Lemna, Lythrum, Myriophyllum, Nuphar, Nymphaea alba-type, Phragmites, Potamogeton, Ranunculus aquatilis group, Typha/Sparganium, Typha latifolia-type Artemisia, Chenopodiaceae, Poaceae wild grass group
Open vegetation—Steppic indicators OJC (Mercuri et al. (2013a))*
Olea, Juglans, Castanea
Other cultivated trees and woody plants collected in the wild
Corylus, Cornus, Morus, Prunus, Vitis
Cultivated/Cultivable herbs
cereals (Avena/Triticum group, Hordeum group, Secale cereale, Zea mays), Cannabis, Linum usitatissimum, Vicia Artemisia, Campanula, Centaurea, Cerastium-type, Convolvulus, Cuscuta europaea-type, Galium-type, Mercurialis, Papaver, Plantago, Polygonum, Rumex, Silene dioica-type, Urticaceae Alchemilla/Aphanes-type, Anagallis, Artemisia, Centaurea, cereals, Cichorieae, Cirsium, Convolvulus, Galium-type, Mercurialis, Papaver rhoeas-type, Plantago, Polygonum, Rumex, Urtica
Weeds and ruderals*
Human impact indicators
Anthropopgenic Pollen Indicators (API) (Behre (1986); Mercuri et al. (2013b)) Local Pastoral Pollen Indicators (LPPI) (Mazier (2007); Mazier et al. (2009))
Asteroideae, Cichorieae, Cirsium, Galium-type, Potentilla-type, Ranunculaceae, Stellaria-type
Regional Human Activity Pollen Indicators (RHAPI) (Mazier (2007); Mazier et al. (2009))
Artemisia, cereals, Chenopodiaceae, Plantago, Rumex, Urtica
Pollen Disturbance Index (PDI) (Kouli (2015))
Centaurea, Cichorieae, Plantago, Polygonum aviculare-type, Pteridium, Ranunculus acris-type, Sarcopoterium, Urtica dioica-type
72 anthropic ecosystems.20 The sum of ‘Anthropogenic Pollen Indicators’ (API) assembles plant taxa directly associated with human activities.21 Seven of them have been found to be quasi-ubiquitous in archaeological sites from the Bronze Age to the Renaissance period in Italy, and to define human landscapes.22 They include: Cerealia-type, Artemisia, Centaurea, Plantago, Trifolium type, Urtica, and Cichorieae, the fenestrate morphotype of Cichorioideae.23 Weeds, ruderals and cultivated plants included in the API sum are both recurring taxa in pollen records from archaeological sites,24 and mark the development of human environments in off-site marine and terrestrial cores.25 Other sums useful for reconstructing anthropogenic landscapes are ‘Local Pastoral Pollen Indicators’ (LPPI) and ‘Regional Human Activities Pollen Indicators’ (RHAPI), indices that can be used to infer past local pastoral activities and background human activities, respectively.26 The ‘Pollen Disturbance Index’ (PDI) sum includes selected anthropogenic indicators (Centaurea, Cichorieae, Plantago, Polygonum aviculare type, Pteridium, Ranunculus acris type, Sarcopoterium, Urtica dioica type), which all correlate with pastoral activities.27 Cichorieae pollen, included in both LPPI and PDI indices, is a good marker for the presence of pastures and open habitats.28 These pollen grains are abundant in such areas due to the action of herbivores on vegetation.29 Key Sites This vegetation overview is based on ten sites (fig. 1, table 2). This includes three extant lakes from the central Italian Peninsula (Lake Albano, Lake Nemi and Lake Vico), three from the Balkan Peninsula (Lake Shkodra in the western part, Lake Orestias and Lake Dojran in the southern part), two sea level sites from the south-eastern Hellenic peninsula (Vravron and Lake Lerna) and two from the islands of Sicily (Lake Pergusa) and Crete (Asi Gonia). The selected sites vary from coastal to mountainous environments, and experience variable climate
20 Brun (2011) 135–42. 21 Behre (1986); Mercuri et al. (2013b) 143–53. 22 Mercuri et al. (2013b) 143–53. 23 Florenzano et al. (2015) 154–65. 24 Florenzano and Mercuri (2012) 750–52; Bosi et al. (2015) 19–31; Bowes et al. (2015) 174–84. 25 Mercuri et al. (2012) 353–72. 26 Mazier (2007); Mazier et al. (2009) 171–88. 27 Kouli (2015) 195–204. 28 Hjelle (1999) 55–81; Mazier et al. (2006) 91–103; Ejarque et al. (2011) 123–39. 29 Florenzano et al. (2015) 154–65.
Kouli et al.
conditions within the Mediterranean climatic gradient. In the selected sites the period of Late Antiquity is fairly well-recorded. Chronologies are based on Accelerator Mass Spectrometry (AMS) radiocarbon dates, and, in some, on tephra layers, while the mean temporal resolution ranges from 20 years per sample, in Asi Gonia, to approximately 100 in Lakes Nemi, Orestias, Vico and Vravron. Lake Lerna represents the lowest resolution site, with a mean of a 150 year interval between pollen samples. For the period 300 BC to AD 1000, selected pollen percentages and pollen groups have been used in an effort to show vegetation development, so as to properly characterise the agrarian landscape in each area (table 1). In our study, the total numbers of Arboreal Pollen (AP), deciduous mesophilous trees and evergreen Mediterranean trees and shrubs (fig. 2) are considered ‘natural/semi-natural’ woodland vegetation. Past agricultural activities are assessed from OJC sum, and cereals curves, while Cichorieae are used as grazing indicators.30 Also, weeds or ruderals inform us as to the overall human impact on vegetation (figs 3–4). In the diagrams, only percentages above 5% are presented. Lake Vico (Lago di Vico, Central Italy) Lake Vico is a horseshoe shaped crater lake (maximum depth ca. 50 m, maximum diameter ca. 5 km, surface area ca. 12 km2) located in central Italy at an altitude of 561 masl. Climatic conditions are submediterranean, with an annual precipitation ca. 1,400–1,600 mm; there is no summer aridity, and the mean annual temperature is between 12 and 13.6 Celsius. Vegetation consists mainly of deciduous oaks and beech forests, while there is cultivation of chestnut and hazel around the caldera today. Deposits from Lake Vico, recovered by two parallel cores (V3 and V4),31 record the vegetation history of the area since the Holocene. The chronology is based on four AMS radiocarbon dates.32 The most characteristic feature of the Late Holocene vegetation, is the overwhelming abundance of deciduous oaks at the expense of other trees.33 The mean temporal resolution for the 1st millennium AD is 100 years. Vegetation patterns appear stable during this interval, with two minor drops of AP, the reductions observed around AD 250 and 550 (fig. 2).
30 OJC: Mercuri et al. (2013a) 24–42. Cichorieae: Florenzano et al. (2015) 154–65. 31 Frank (1969) 67–85; Leroy et al. (1996) 189–201; Magri and Sadori (1999) 247–60. 32 Sulpizio et al. (2008) 263–76; Sevink et al. (2011). 33 Sadori et al. (2011) 117–29.
73
Regional Vegetation Histories
figure 1
Location of the selected central Mediterranean pollen records.
table 2
Location, main characteristics and references for the pollen records considered in the present study. Region
Elevation (masl)
Latitude (N)
Longitude (E)
Temporal reso lution (yrs) 2400800BP interval
Selected key references
Lake Albano
Latium Italy
293
41°45’
12°40’
92
Mercuri et al. (2002); Mercuri et al. (2012, 2013a, 2013b)
Lake Nemi
Latium Italy
320
41°43’
12°42’
112
Mercuri et al. (2002); Mercuri et al. (2011, 2012, 2013a, 2013b)
Lake Vico
Latium Italy
510
42°19’
12°I0’
105
Unpublished
Lake Shkodra
S Balkans, Albania / Montenegro
6
42°12’
19°15’
83
Sadori et al. (2015); Mazzini et al. (2016)
Lake Dojran
S Balkans, FYROM / Greece
144
41° 12’
22° 44’
64
Masi et al. (2018); Thienemann (2017)
Orestias
S Balkans, Greece
620
40°30’
21°15’
91
Kouli (2007); Kouli and Dermitzakis (2010)
Vravron
Attica, Greece
0
37°55’
23°60’
108
Triantaphyllou et al. (2010); Kouli (2012)
Lerna
Peloponnese, Greece
0
37°34’
22°43’
148
Jahns (1993)
Lake Pergusa
Sicily, Italy
667
37°31’
14°18
60
Sadori et al. (2013, 2016)
Asi Gonia
Crete, Greece
780
35°15’
24°16’
18
Atherden and Hall (1999)
74
Kouli et al.
figure 2
Pollen percentage diagram from Lake Albano, Lake Nemi, Lake Vico, Lake Shkodra, Lake Dojran, Lake Orestias, Vravron, Lake Lerna, Lake Pergusa and Asi Gonia. Selected curves: Arboreal pollen (AP), deciduous trees, Mediterranean trees and shrubs (taxa included in each sum are according to table 1).
figure 3
Pollen percentage diagram from Lake Albano, Lake Nemi, Lake Vico, Lake Shkodra, Lake Dojran, Lake Orestias, Vravron, Lake Lerna, Lake Pergusa and Asi Gonia. Selected curves: OJC sum and cereals.
figure 4
Pollen percentage diagram from Lake Albano, Lake Nemi, Lake Vico, Lake Shkodra, Lake Dojran, Lake Orestias, Vravron, Lake Lerna, Lake Pergusa and Asi Gonia. Selected curves: Cichorieae, weeds/ruderals and hemp pollen. (taxa included in weeds/ruderals are according to table 1).
75
Regional Vegetation Histories
Lake Albano (Lago Albano, Central Italy) The Albano crater lake is located in the Albano Hills volcano at 293 masl. It has an ellipsoidal shape covering an area of 6 km2 (dimensions ca. 3.5 × 2.3 km), and 168 m deep. The climate and vegetation are the same as described for the previous lake. The shorelines are quite different though, as the south-western side is more forested and less exploited than the north-western side. Several houses and diffused cultivations are part of the landscape. Borehole PALB94.1E was located in the north-central part of the lake, at a water depth of about 70 m.34 The chronology was assessed based on the basis of (bio-) stratigraphical correlations, radiocarbon dates and tephra layers, and covers the period from the Late Glacial to the Late Holocene (with a Mid-Holocene hiatus). The mean temporal resolution for the 1st millennium AD is 90 years. From 3,600 cal. years BP, corresponding to the Middle Bronze Age, signs of human activities, such as the cereal cultivation, became unquestionable. In fact, a pile dwelling site (called the ‘Villaggio delle Macine’ (Grindstone Village)) developed on the lakeshores. Meanwhile, Olea, grazing indicators and other anthropogenic indicators increased, and forest decreased. At the same time, cultivated trees (Olea, Juglans, and Castanea) spread, together with cereals and weeds, between 2,900 and 2,700 cal. years BP, matching the significant impact that Pre-Roman Iron Age populations had on the central Italian landscape.35 The Roman period is marked by the ‘Cannabis-phase’, the hemp pollen peak (at 1,790 BP), which followed the peak of cultivated trees. Later, the ‘chestnut landscape’ definitively spread at around 700 years BP,36 but with the forest cover displaying an irreversibly decreasing trend. Lake Nemi (Lago di Nemi, Central Italy) The crater Lake Nemi (mean depth ca. 33 m, maximum diameter ca. 1.3 km) is also located in the Alban Hills in central Italy at an altitude of 320 masl, and is part of the same volcanic complex as Lago Albano, sharing with it part of the watershed, although covering a smaller area of 1.8 km2. Located near the Lago Albano, climatic conditions are Mediterranean, with surroundings covered by evergreen woodland and deciduous oak woods. PNEM.1B, a 915 cm core sequence, was obtained from Lake Nemi at a water depth of 30 m. The pollen record covers the last
34 Lowe et al. (1996) 71–98. 35 Follieri et al. (1988) 329–56; Mercuri et al. (2002) 263–76; Mercuri et al. (2013a) 24–42. 36 Mercuri et al. (2012) 353–72.
11,000 years.37 The mean temporal resolution for the 1st millennium AD is 110 years. A general trend towards aridity is especially marked by the tree curve, showing a decreasing trend with oscillations from around 5,700 BP. The AP vegetation, and in particular deciduous oaks, that are the dominant vegetation, exhibit two significant fluctuations, the first at ca. 2,250 years BP and the second, more pronounced, at ca. 1,800 BP. Another minor retreat of deciduous oaks is evident at ca. 1,250 BP (fig. 2). The cultivated trees and cereals, including the Avena/Triticum group become more frequent just after 4,500 cal. yr BP, and forest decrease is evident at ca. 3,600 cal. yr BP. The OJC trees, also favoured by such a climate, expand, together with Roman culture, from about 2,700 BP. The ‘Cannabis-phase’ is here dated at 1,810 BP.38 The presence of the same peaks in the two fairly coeval samples from the Nemi and Albano lakes, proves that the Romans used these large watery places for retting, a practice not reported for Latium by the written sources, and therefore only documented by these two pollen records.39 Judging from the two magnificent Roman ships that were found within it (but destroyed during World War 2 on May 31 1944), we know that Nemi was used to perform naumachie, or re-enactments of naval combat, that were ordered by the emperor Caligula during the 1st c. AD. Since that time, the area has been rich and where agriculture also developed, including the appearance of cereal fields, and probably hemp and tree crop cultivation. Lake Pergusa (Lago di Pergusa, Sicily) Lake Pergusa is located in the interior of Sicily at 667 masl, in the area of the Erei mountains. It is the main endorheic lake of the island, sub elliptical in shape, although quite small. Today the body of water is artificially controlled, and is ca. 1.6 km long, ca. 1 km wide, and is ca. 1.8 km2 in surface area, with a 3.5 m mean water depth. The current vegetation landscape is strongly influenced by humans, with the surroundings mainly cultivated with cereal crops and olive. The natural vegetation is characterised by mixed oak forest, with a Phragmites belt at the lacustrine shore. The entire pollen sequence spans from the last Pleniglacial to the Holocene, and many cores have been analysed. Here we used the high-resolution data from PG2, with samples every ca. 60 years; eight AMS radiocarbon dates and a Sicilian tephra were used to provide
37 Lowe et al. (1996) 71–98. 38 Mercuri et al. (2012) 353–72. 39 Mercuri et al. (2002) 263–76.
76 a reliable chronology.40 A recent detailed study has correlated natural changes with the history of the island, and has shown that the period ca. AD 450–750 was a humid period one.41 This wet period corresponds with an intense agricultural use of the Sicilian landscape, and has been correlated with demographic and economic expansion. Yet, around AD 750, we see a decrease in synanthropic taxa and a recovery of arboreal vegetation. A detailed comparison between vegetation dynamics and the history of Sicily links climatic change to socioeconomic decline. Moreover, as this change occurred prior to the Arab invasion of Sicily in AD 827, such environmental processes may help to explain the collapse of Byzantine society in Sicily. After this event, a longer period of agricultural decline occurred, lasting until around AD 1000, following which we see the first signs of a slow economic recovery. Ongoing archaeobotanical research at nearby Roman and Medieval archaeological sites, suggests that land use changed over centuries.42 A comparison with local realities will help us decide if the pollen trend seen in Pergusa was a regional, potentially climate-induced phenomenon. Although textual evidence records considerable population losses during the later Middle Ages as a result of the Black Death, the effects of the plague are not obviously apparent in the pollen data, except for some short-term fluctuations. Lake Shkodra (Balkan Peninsula, Albania/ Montenegro) Lake Shkodra, the largest natural freshwater lake in the Balkans, is shared between the FYROM (Former Yugoslav Republic of Macedonia) and Albania. The lake surface is only 5–10 masl. The surface area is quite variable, depending on rivers and precipitation. It has a karstic and tectonic origin and elliptical shape (45 km long and ca. 15 km wide). The entire record spans the last 4500 years, and the chronology is quite robust; it is based on 3 Pb-210 dates, 4 radiocarbon AMS dates and 4 well-known tephra layers.43 The temporal resolution is ca. 85 years. The surroundings of the lake consist of some of the richest reservoirs of biodiversity in Europe. The vegetation mainly consists of Mediterranean and Balkan species, although some central European types have their southernmost distribution limit here. Pollen data record the dominance of arboreal vegetation here for all the sequence, as well as a humid phase at
40 Sadori and Narcisi (2001) 655–71; Sadori and Giardini (2007) 173–80; Sadori and Giardini (2008) 229–36; Sadori et al. (2008) 4–14; Sadori et al. (2013b) 1969–84. 41 Sadori et al. (2016) 173–88. 42 Montecchi and Mercuri (2018) 743–57; Mercuri et al. (2017). 43 Sadori et al. (2015) 435–44; Zanchetta et al. (2012) 780–89.
Kouli et al.
ca. AD 650 which followed an aridity crisis at ca. AD 500. In the last seven to eight centuries, the natural forest has in part been replaced by the intense cultivation of chestnut and walnut.44 Lake Dojran (Balkan Peninsula, Greece/FYROM) Lake Dojran is a transboundary lake located in a karstic depression (144 masl) between two mountain chains, at the border between the FYROM (Former Yugoslav Republic of Macedonia) and Greece. It has an elliptic shape, a mean depth of 6.5 m and measures 9 × 7 km (ca. 40 km2 in surface area). Small rivers and groundwater feed the lake, that has only one outlet channel. Cultivated lands surround the lake. Natural vegetation is composed of Mediterranean plants and mixed oak forest. Mountain vegetation is also present, due to the surrounding mountain chains. The recently published45 palynological data cover the entire Holocene in 7 m of sediment; the temporal resolution of the data are ca. 65 years. Organic geochemical and palynological evidence has been used to reconstruct the impact of humans on the landscape. Pollen data point to continuous forestry and agricultural practises and land use from ca. 500 BC to AD 780.46 Lake Orestias (Balkan Peninsula, Greece) Lake Orestias, surrounded by mountains peaking at 2,100 m and located at an altitude of 630 masl, is the remnant of a larger lake system formed in Macedonia during the Neogene. Today it is a small and shallow kidney-shaped lake covering an area of about 32.4 km2 (maximum diameter ca. 32.5 km), with a maximum depth of 9.1 m and a mean depth of 4.5 m. The climate here is described as semi-humid to humid with moderate dry summers, with a mean annual precipitation of ca. 875 mm, showing a double fluctuation within the year (maxima in December and May), while the mean annual temperature is 12.46 Celsius. A significant part of the area is covered by dense woodland, beech forests, with black pine woodland on the uplands and mixed deciduous oak forests at intermediate altitudes. In the lower parts of the basin, cultivated areas and pastures cover most of the area.47 The Late Glacial to Holocene vegetation history of the area is based on the pollen analysis of a single sediment core (G25), recovered from the western part of the lake.48 The chronology is based on three AMS radiocarbon
44 Sadori et al. (2015) 435–44. 45 Masi et al. (2018) 351–67. 46 Francke et al. (2013) 481–98; Thienemann et al. (2017) 1–12. 47 Kouli (2007). 48 Kouli and Dermitzakis (2010) 154–56.
Regional Vegetation Histories
dates performed on pollen grains. Palynological analysis recorded a change in vegetation from open woodland, at the onset of the Holocene, to dense forests, that were exploited by the Neolithic communities of the area.49 The upper part of the record has a temporal resolution of ca. 100 years per sample, which appears to show that the area was already deforested by around 200 BC, with AP values around 20%. The period examined is characterised by a gradual increase in AP values, especially after ca. AD 400, mainly due to pine expansion (fig. 2). A brief temporal expansion of deciduous forest is also recorded between AD 100–400, while a major two-step expansion (maxima ca. AD 650 and 900) of Cichorieae, followed by weeds/ruderals, is evidenced afterwards (fig. 4). Vravron (South Greece) The Vravron wetland is a marshy area located on the eastern coast of the Attica peninsula (eastern Greece). The area, inhabited since the Neolithic, is of great historical significance, largely due to the Archaic and Classical sanctuary of Vravronia Artemis here.50 The climate is typically Mediterranean, with warm, dry summers and mild, humid winters, with the area being exposed to NNE winds, mainly prevailing during the summer. The present vegetation consists of predominantly vine and olive, and shows evidence for other cultivation as well as grazing activities. Scatter natural areas of maquis and more extensive phrygana vegetation are recorded in the landscape, that turned into a rapidly developing urban area during the last few decades. Core VG3 was retrieved from the wetland of Vravron as part of joint environmental study work.51 The pollen assemblage provides a vegetation record for the last 5,000 years, and suggests a close correlation between vegetation development and the human presence in Attica. The chronology is based on four AMS radiocarbon dates,52 providing a temporal resolution of ca. 100 years per sample for the 1st millennium AD. A rather variable landscape, where open Mediterranean pine woods alternated with macchia shrublands and grasslands, together with significant imprints of human activity, is evidenced. Agricultural practices displayed temporal variations, with intensive cereal cultivation during the Bronze Age, the spread of olive exploitation from the Geometric to Classical eras, followed by the gradual abandonment of olive groves during the
49 Kouli (2015) 145–204. 50 For more detail on this area, see the relevant chapters in Aikaterinidis (2001). 51 Triantaphyllou et al. (2010) 267–78. 52 Triantaphyllou et al. (2010) 267–78.
77 Hellenistic period.53 During the 1st millennium AD pine was the main tree species in the area, while both evergreen Mediterranean and thermophilous deciduous trees appear in constant, but low, abundances (fig. 2). The increase in olive and cereal abundances, recorded after ca. AD 600 (fig. 3), is indicative of agricultural activities during the Byzantine period. Lake Lerna (Peloponnese, Southern Greece) The ancient Lake Lerna is an elongated (dimensions ca. 190 × 60 m) drained karstic lake, lying at sea level in the eastern Peloponnese. The climate of the area is thermomediterranean, with a mean temperature of 18 Celsius, and a mean precipitation of ca. 500 mm, while a pronounced arid period from April to October exists. The surrounding area is intensively cultivated, with olive and citrus being the main cultivars; it is also grazed. Phrygana and some Mediterranean evergreen sclerophyllous shrub covers the uncultivated parts of the basin. The pollen profile of Lake Lerna covers the last ca. 7,500 years, based on nine radiocarbon dates. The base of the sequence illustrates a closure of the deciduous woodland before the Bronze Age, while after this the increased human impact on vegetation (mainly extensive olive cultivation and grazing) is recorded.54 The temporal resolution of the studied interval is the lowest in the selected sites, ca. 150 years per sample.55 During the 1st millennium AD the surroundings of Lerna were characterised by an open landscape with evergreen sclerophyllous shrubs and trees, and limited deciduous cover. Pine trees and high altitude firs show a temporal increase ca. 50 BC, though after ca. AD 200 a further decrease in woody vegetation is recorded (fig. 2). Olive cultivation activity appears to be intensive during the Hellenistic and Early Roman periods, but gradually falls away after ca. AD 100 (fig. 3). Asi Gonia (Crete, Greece) Asi Gonia is a small peat bog (dimensions 190 × 60 m), located at an altitude of 780 masl in the White Mountains of central Crete. Due to the nature of the site it accurately records local vegetation patterns. A mix of phrygana and steppe types is the characteristic vegetation around the peat bog, while some tall shrubs and woodland, with evergreen and some deciduous oaks, occur in the wider area, that today is above the altitude limit for olive
53 Kouli (2012) 267–78. 54 Jahns (1993) 187–203. 55 Data extracted from the European Pollen Database: http://www .europeanpollendatabase.net/ (last accessed 5 January 2017).
78 cultivation.56 With seven AMS radiocarbon dates in 4.7 m of deposits, spanning the last 1,500 years, Asi Gonia is the most detailed of the sites examined here, with a temporal resolution of ca. 20 years per sample.57 The pollen record is dominated by oaks (mainly evergreen), with AP exceeding 80% of the total vegetation. The evergreen and deciduous oaks appear to relate to one another, as, during the brief periods where decreased evergreen oaks are observed (e.g. ca. AD 650, 800 and 1000), an expansion of deciduous oaks can be seen (fig. 2). Other tree species seem to have been relatively unimportant in the woodland of the area during the past 1500 years. Vegetation Patterns The vegetation patterns of the sites discussed comprehensively cover both the altitudinal and latitudinal variations evidenced in the central Mediterranean region (fig. 2). Furthermore, most of them are large-scale sites, and thus record the vegetational pollen signal for the region as a whole.58 The areas that appear to have high forest cover (fig. 3) are almost exclusively located in the northern part of our region, namely in the central Italian peninsula and in the southern Balkans (Shkodra and Dojran). In those records the main component of the arboreal vegetation is deciduous forest. Conifers, especially pine woodlands, are also recorded as being more extensive in the coastal site of Shkodra. However, evergreen Mediterranean trees, like holm oaks or pistachio, are limited. This may also partly be an effect of the entomophilous nature of most of this vegetation (e.g. Pistacia, Myrtus, Cistus), resulting in a lower palynological picture compared to oaks and pines. High forest cover is also evidenced from the mountainous Cretan record of Asi Gonia, though in this southernmost site evergreen Mediterranean forest vegetation is dominant. Few temporal vegetation fluctuations are evidenced in the three Latium sites, which are marked by short periods of retreat of deciduous forest (fig. 2). Despite the fact that the Latium sites lay in a relatively small region, only the retreat centred ca. AD 250 appears to be more or less synchronous across the profiles. This forest retreat might have been connected with both a (general) climate oscillation and (local) human activities, and shows certain differences from site to site.59 The records from Lakes Shkodra and Dojran do not show significant fluctuations in the extent of forest cover during the 1st
56 Atherden and Hall (1999) 183–93. 57 Data extracted from the European Pollen Database: http://www .europeanpollendatabase.net/ (last accessed 5 January 2017). 58 Mazier et al. (2012) 38–49. 59 See below ‘The Agrarian Landscape’.
Kouli et al.
millennium AD, though an expansion of deciduous trees, at the expense of pine, is evidenced in Lake Dojran after the 1st c. AD (fig. 2). The general vegetation pattern of well-developed forest vegetation in the northern sites is in contrast with the record from Lake Orestias (western Macedonia, Greece), where the landscape appears to be deforested. Pines are in general the most common trees, with the exception of a 300 year period around AD 300, during which deciduous trees had a small expansion. After ca. AD 400, pines are the dominant trees in an open landscape. This can be also attributed to systematic and continuous human activity in the area, evidenced in the curves of the anthropogenic indicators, like OJC, cereals and Cichorieae (fig. 4). Several important centres, like the Neolithic Dispilio lake settlement,60 Hellenistic Keletron, Roman Diocletianopolis or Byzantine Justinianopolis,61 were located in a continuously inhabited basin, where natural resources have been exploited since the Neolithic.62 An open plant landscape with strong signs of human activity is evidenced in the remaining southern records from the Hellenic peninsula (Vravron and Lerna) as well as from Sicily (Lake Pergusa). Both thermophilous deciduous or evergreen trees are limited. Lake Lerna is diachronically the most deforested landscape studied, providing a picture of phrygana vegetation with some pine. A small advance of deciduous trees between ca. AD 100–350 is followed by a slow continuous decline of the few tree populations. In nearby Vravron, the vegetation patterns seem more balanced, with an alternation of open pine woodlands along with areas of open vegetation. Evergreen Mediterranean and thermophilous deciduous trees are systematically evidenced in small abundances. Around AD 400 and until AD 650, a shortlived increase in Mediterranean vegetation compared to the deciduous tree population is observed, a trend that is reversed thereafter. The differences in vegetation between Lerna and Vravron reflect differences in the human impact on the landscape, as both sites are located in climatically similar conditions on the eastern coast of Greece. Therefore, they need to be examined closely alongside the history of each area. Within our study set, the evidence from Lake Pergusa in Sicily and lakes Albano and Nemi in central Italy suggest that human presence highly influenced the vegetation history in those places. The aridification trend, that started from the middle of the Mid Holocene, exacerbated by water level oscillations and human impact that occurred during the Late Holocene, reduced the lake surface area and vegetation cover around Lake Pergusa. In the last three
60 Karkanas et al. (2011) 83–117; Facorellis et al. (2014) 511–28. 61 Gregory and Wharton (1991). 62 Kouli (2015) 145–204.
79
Regional Vegetation Histories
millennia, the arboreal pollen equates to the herbaceous pollen, indicating opening vegetation; Juglans and Olea, together with high amounts of cereals and Cichorieae, indicate the diversified nature of the land use of the territory. The great variability in the vegetation patterns across the sites discussed, are mainly connected to high environmental variability, such as altitude, basin characteristics, local microclimates, and the soil properties of these central Mediterranean habitats. If we examine each site in isolation, the few dramatic vegetation changes observed should be associated with human activity rather than natural causes. In all Mediterranean regions, the plant landscape was slowly and continuously shaped by human activity throughout most of the Holocene.63 The Agrarian Landscape The pollen records of the central Mediterranean region feature a temporal and spatial variability in plant landscape exploitation as a result of differences in human population, cultures and economies. Cultivated trees, summarised in the OJC curve (fig. 3), play a very significant role in the landscapes of the central Mediterranean region. In the three (mainly deciduous) forest sites from the central Italian peninsula, high abundances of walnut and chestnut are recorded, while wild olive trees were probably only a minor part of the wider domesticated and processed crop. Differences between sites clearly exist. In Lake Albano, mainly chestnut, but also walnut and olive exploitation became evident already from ca. 150 BC. The observed retreat of deciduous trees ca. AD 250 seems to also affect chestnuts, and could indicate a change in human activity, if it correlates with the systematic increase in cereal abundances. In the nearby site of Lake Nemi, chestnut appears in greater numbers only in the bottom part of the record and up to 50 BC, while after AD 100 walnut gradually becomes the dominant tree crop. At both sites, the remarkable synchronous increase in hemp pollen ca. AD 250 (fig. 4) indicates retting activity. Chestnut is also systematically present at the third Latium site of Lake Vico. Lake Vico, the site with the highest elevation, has, in this period, always been largely surrounded by deciduous mesophilous forest. Under those circumstances, the increase of olive after ca. AD 400 is noteworthy. At the Balkan sites of Lake Orestias and Dojran, tree crops are mainly walnut. In Lake Orestias, walnut expands between AD 100–800, while a significant increase in cereal cultivation is also evidenced after AD 100. 63 Mercuri (2014) 1799–1810.
Nevertheless, the most striking feature of the vegetation of Lake Orestias is the major two-step expansion of Cichoriae and ruderals (fig. 4), indicating a move towards animal herding during the Byzantine period. Lake Orestias’ landscape appears deforested and intensively exploited in contrast to the other Balkan sites. At Lake Dojran both walnut and cereals are not as abundant, while at the coastal Lake Shkodra, both olives and walnuts are evidenced (fig. 3). Close to Lake Prespa, human activity, included cereal and walnut cultivation, is apparent, but the impact on natural vegetation is not as profound.64 The southern sites of Vravron and Lerna in the Hellenic peninsula, Pergusa in Sicily, and Asi Gonia in Crete, are characterised by the dominance of olive in the OJC curve, although olive exploitation does not seem to be systematically intensive and continuous over the period. In Vravron, agricultural activities during the Byzantine era are evidenced by the expansion of olive and cereals, recorded after AD 600. In contrast to this, in the area of Lake Lerna, the 1st millennium AD is a period of decline for human activity. The abandonment of olive cultivation is apparent already from the Roman period, but gradually increases. A similar decrease in the grazing indicator Cichorieae also indicates lower land use during this period. Lake Pergusa presents a more complicated record, with repeated fluctuations both in olives and cereals. This should be viewed alongside the historical record for the area, which also testifies to changes in general agriculture practices, as a result of the different cultures who settled in Sicily over the period.65 Finally, in Asi Gonia, the only indicator of human activity is the appearance of olive. Concluding Remarks The 1st millennium AD is a rather short time period in the vegetation history of the Mediterranean, nevertheless a great variety of vegetation patterns is evidenced. Apart from the spatial environmental differences in the region, the main factors shaping plant landscapes were human societies. The variations in land use during the 1st millennium AD, apparent from the pollen records of the central Mediterranean region, are neither peculiar nor unexpected. This landscape variety cannot be understood and interpreted in a uniform way, so a ‘biodiversity of replies’ seems to be the best approach.66 The exact role of local vegetation patterns,
64 Panagiotopoulos et al. (2013) 157–69. 65 See section ‘Lake Pergusa’; Montecchi and Mercuri (2018) 743–57. 66 Mercuri (2014) 1799–810.
80 and the environmental conditions that played a part in the choices made by local societies, should be examined in detail by joint groups of historians, archaeologists and palaeoecologists. The interdisciplinary study of off-site cores remains the key way to examine palaeoenvironmental transformations mirroring the ‘semi-natural’ vegetation, and to reveal temporal variations and the degree of human impact on a regional scale. In contrast to earlier periods within the Holocene, the numerous extant written sources of Late Antiquity, and later periods, can provide us with a new ‘piece in the puzzle’, adding information about the human ecological factor, its activities and choices. However, as the Cannabis-phase has revealed, the direct ‘plant fingerprint’ is irreplaceable and may provide new pieces of evidence. In this way, the need for an interdisciplinary evaluation of the records is as important as the need for more, well-dated, high-resolution palaeovegetation records. Acknowledgments The authors wish to thank John Haldon, Adam Izdebski and Luke Lavan, convenors of the LAA 2016 Conference, for the stimulating interdisciplinary discussion that resulted from the meeting, and Michael Mulryan for his constructive comments and editing of the manuscript. The pollen data from Lerna and Asi Gonia were taken from the European Pollen Database (http://www.europeanpollendatabase.net). Bibliography Aikaterinidis G. (2001) ed. Mesogaia, History and Civilization of Mesogaia Attika (Athens 2001). Atherden M. A. and Hall J. A. (1999) “Human impact on vegetation in the White Mountains of Crete since AD 500”, Holocene 9 (1999) 183–93. Behre K. E. (1990) “Some reflections on anthropogenic indicators and the record of prehistoric occupation phases in pollen diagrams from the Near East”, in Man’s Role in the shaping of the Eastern Mediterranean Landscape, eds. S. Bottema, G. Entjes-Nieborg and W. van Zeist (Rotterdam 1990) 219–30. Behre K. E. (1986) Anthropogenic Indicators in Pollen Diagrams (Rotterdam 1986). Bellotti P., Calderoni G., Dall’Aglio P. L. et al. (2016) “Middle to Late Holocene environmental changes in the Garigliano delta plain (central Italy): which landscape witnessed the development of the Minturnae Roman colony?”, Holocene 26 (2016) 1457–71. Bosi G., Mercuri A. M., Bandini Mazzanti M., Florenzano A., Montecchi M. C., Torri P., Labate D. and Rinaldi R. (2015) “The evolution of Roman urban environments through the archaeobotanical remains in Modena, northern Italy”, JAS 53 (2015) 19–31. Bowes K., Mercuri A. M., Rattighieri E., Rinaldi R., ArnoldusHuyzendveld A., Ghisleni M., Grey C., Mackinnon M. and Vaccaro E. (2015) “Palaeoenvironment and land use of Roman
Kouli et al. peasant farmhouses in southern Tuscany”, Plant Biosystems 149 (2015) 174–84. Brun C. (2011) “Anthropogenic indicators in pollen diagrams in eastern France: a critical review”, Vegetation History and Archaeobotany 20 (2011) 135–42. Combourieu-Nebout N., Peyron O., Bout-Roumazeilles V., Goring S., Dormoy I., Joannin S., Sadori L., Siani G. and Magny M. (2013) “Holocene vegetation and climate changes in the central Mediterranean inferred from a high-resolution marine pollen record (Adriatic Sea)”, Climate of the Past 9 (2013) 2023–42. Cremaschi M., Zerboni A., Charpentier V., Crassard R., Isola I., Regattieri E. and Zanchetta G. (2015) “Early-Middle Holocene environmental changes and Pre-Neolithic human occupations as recorded in the cavities of Jebel Qara (Dhofar, southern Sultanate of Oman)”, Quaternary International 382 (2015) 264–76. Dreβler M., Selig U., Dörfler W., Adler S., Schubert H. and Hübener T. (2006) “Environmental changes and the Migration Period in northern Germany as reflected in the sediments of Lake Dudinghausen”, Quaternary Research 66 (2006) 25–37. Ejarque A., Miras Y., and Riera S. (2011) “Pollen and non-pollen palynomorph indicators of vegetation and highland grazing activities obtained from modern surface and dung datasets in the eastern Pyrenees”, Review Palaeobotany Palynology 167 (2011) 123–39. Facorellis Y., Sofronidou M. and Hourmouziadis G. (2014) “Radiocarbon dating of the Neolithic lakeside settlement of Dispilio, Kastoria, northern Greece”, Radiocarbon 56 (2014) 511–28. Florenzano A. (2013) Evolution of a Mediterranean landscape as shown by the archaeo-environmental reconstruction of Lucanian sites (Ph.D. diss., Univ. of Modena and Reggio Emilia 2013). Florenzano A. and Mercuri A. M. (2012) “Palynology of archaeological sites: the example of economy and human impact of the Metaponto area (6th–1st century BC)”, Rendiconti Online Societa Geologica Italiana 21 (2012) 750–2. Florenzano A., Marignani M., Rosati L., Fascetti S. and Mercuri A. M. (2015) “Are Cichorieae an indicator of open habitats and pastoralism in current and past vegetation studies?”, Plant Biosystems 149 (2015) 154–65. Follieri M., Magri D. and Sadori L. (1988) “250,000-year pollen record from Valle di Castiglione (Roma)”, Pollen et Spores 30 (1988) 329–56. Francke A., Wagner B., Leng M. J. and Rethemeyer J. (2013) “A Late Glacial to Holocene record of environmental change from Lake Dojran (Macedonia, Greece)”, Climate of the Past 9 (2013) 481–98. Frank A. H. E. (1969) “Pollen stratigraphy of the Lake of Vico (central Italy)”, Palaeogeography, Palaeoclimatology Palaeoecology 6 (1969) 67–85. Gogou A., Triantaphyllou M., Xoplaki E., Izdebski A., Parinos C., Dimiza M., Bouloubassi I., Luterbacher J., Kouli K., Martrat B., Toreti A., Fleitmann D., Rousakis G., Kaberi H., Athanasiou M. and Lykousis V. (2016) “Climate variability and socio-environmental changes in the northern Aegean (NE Mediterranean) during the last 1500 years”, Quaternary Science Reviews 136 (2016) 209–28. Giardini M. (2007) “Late Quaternary vegetation history at Stracciacappa (Rome, central Italy)”, Vegetation History and Archaeobotany 16 (2007) 301–16. Gregory T. E and Wharton A. J. (1991) “Kastoria”, in The Oxford Dictionary of Byzantium, ed. A Kazhdam (Oxford-New York 1991) 1110–11. Hall V. A. (1989) “A comparison of grass foliage, moss polsters and soil surfaces as pollen traps in modern pollen studies”, Circaea 6 (1989) 63–69. Hjelle K. L. (1999) “Modern pollen assemblages from mown and grazed vegetation types in western Norway”, Review Palaeobotany Palynology 107 (1999) 55–81.
Regional Vegetation Histories Jahns S. (1993) “On the Holocene vegetation history of the Argive Plain (Peloponnese, southern Greece)”, Vegetation History Archaeobotany 2 (1993) 187–203. Karkanas P., Pavlopoulos K., Kouli K., Ntinou M., Tsartsidou G., Facorellis Y. and Tsourou T. (2011) “Palaeoenvironments and site formation processes at the Neolithic lakeside settlement of Dispilio, Kastoria, northern Greece”, Geoarchaeology 26 (2011) 83–117. Kotthoff U., Müller U. C., Pross J., Schmiedl G., Lawson I. T., Van De Schootbrugge B. and Schulz H. (2008) “Lateglacial and Holocene vegetation dynamics in the Aegean region: an integrated view based on pollen data from marine and terrestrial archives”, Holocene 18 (2008) 1019–32. Kouli K. (2015) “Plant landscape and land use at the Neolithic lake settlement of Dispilió (Macedonia, northern Greece)”, Plant Biosystems 149 (2015) 195–204. Kouli K. (2012) “Vegetation development and human Activities in Attiki (SE Greece) during the last 5,000 Years”, Vegetation History and Archaeobotany 21 (2012) 267–78. Kouli K. (2007) “Palaeoenvironmental and palaeoecological reconstruction of the area of the Dispilio Neolithic lake settlement, Lake Kastoria”, Gaia 17 (2007) 1–150. Kouli K. and Dermitzakis M. D. (2010) “Lake Orestiás (Kastoria, northern Greece)”, Grana 49 (2010) 154–56. Lawson I. T., Al-Omari S., Tzedakis P. C., Bryant C. L. and Christanis K. (2005) “Lateglacial and Holocene vegetation history at Nisi Fen and the Boras mountains, northern Greece”, Holocene 15 (2005) 873–87. Leroy S., Giralt S., Frankus P. and Seret G. (1996) “The high sensitivity of the palynological record in the Vico Maar lacustrine sequence (Latium, Italy) highlights the climatic gradient through Europe for the last 90 ka”, Quaternary Science Reviews 15 (1996) 189–201. Lionello P., Malanotte-Rizzoli P., Boscolo R., Alpert P., Artale V., Li L., Luterbacher J., May W., Trigo R., Simplis M. T., Ulbrich U. and Xoplaki E. (2006). “The Mediterranean climate: an overview of the main characteristics and issues”, in Mediterranean Climate Variability, edd. P. Lionello, P. Malanotte-Rizzoli, R. Boscolo (Amsterdam 2006) 1–26. Lowe J. J., Accorsi C. A., Bandini Mazzanti M., Bishop A., Forlani L., Van der Kaars S., Mercuri A. M., Rivalenti C., Torri P. and Watson C. (1996) “Pollen stratigraphy of sediment sequences from crater lakes (Lago Albano and Lago Nemi) and the central Adriatic spanning the interval from Oxygen isotope stage 2 to present day”, Memorie Istituto Italiano Idrobiologia 55 (1996) 71–98. Magri D. and Sadori L. (1999) “Late Pleistocene and Holocene pollen stratigraphy at Lago di Vico (central Italy)”, Vegetation History and Archaeobotany 8 (1999) 247–60. Mariotti Lippi M., Bellini C., Trinci C., Benvenuti M., Pallecchi P. and Sagri M. (2007) “Pollen analysis of the ship site of Pisa San Rossore, Tuscany, Italy: the implications for catastrophic hydrological events and climatic change during the Late Holocene”, Vegetation History and Archaeobotany 16 (2007) 453–65. Masi A., Francke A., Pepe C., Thienemann M., Wagner B. and Sadori L. (2018) “Vegetation history and paleoclimate at Lake Dojran (FYROM/Greece) during the Late Glacial and Holocene”, Climate of the Past 14 (2018) 351–367. Mazier F. (2007) Modélisation de la relation entre pluie pollinique actuelle, végétation et pratiques pastorales en moyenne montagne (Pyrénées et Jura) application pour l’interprétation des données polliniques fossiles (Ph.D. diss., Univ. de Franche Comté 2007). Mazier F., Gaillard M. J., Kunes P., Sugita S., Trondman A.-K. and Brostrom A. (2012) “Testing the effect of site selection and parameter setting on REVEALS-model estimates of plant abundance using the Czech Quaternary Palynological database”, Review of Palaeobotany and Palynology 187 (2012) 38–49.
81 Mazier F., Galop D., Gaillard M. J., Rendu C., Cugny C., Legaz A., Peyron O. and Buttler A. (2009) “Multidisciplinary approach to reconstructing local pastoral activities: an example from the Pyrenean Mountains (Pays Basque)”, Holocene 19 (2009) 171–88. Mazier F., Galop D., Brun C. and Buttler A. (2006) “Modern pollen assemblages from grazed vegetation in the western Pyrenees, France: a numerical tool for more precise reconstruction of past cultural landscapes”, The Holocene 161 (2006) 91–103. Mercuri A. M. (2014) “Genesis and evolution of the cultural landscape in the central Mediterranean: the “where, when and how” through the palynological approach”, Landscape Ecology 29 (2014) 1799–810. 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 (2008) 1950–67. Mercuri A. M., Montecchi M. C., Florenzano A., Rattighieri E., Torri P. and Vaccaro E. (2017) “The late antique plant landscape in Sicily: the pollen data from the agro-pastoral Villa del Casale— Philosophiana system”, Quaternary International: doi:10.1016/j .quaint.2017.09.036 (last accessed 24 August 2018). Mercuri A. M. and Sadori L. (2014). “Mediterranean culture and climatic change: past patterns and future trends”, in The Mediterranean Sea: its History and Present Challenges, edd. S. Goffredo and D. Dubinsky (Dordrecht-Heidelberg-NewYorkLondon 2014) 183–211. Mercuri A. M., Bandini Mazzanti M., Florenzano A., Montecchi M. C. and Rattighieri E. (2013a) “Olea, Juglans and Castanea: the OJC group as pollen evidence of the development of human-induced environments in the Italian peninsula”, Quaternary International 303 (2013) 24–42. Mercuri A. M., Bandini Mazzanti M., Florenzano A., Montecchi M. C., Rattighieri E. and Torri P. (2013b) “Anthropogenic Pollen Indicators (API) from archaeological sites as local evidence of human-induced environments in the Italian peninsula”, Annali di Botanica 3 (2013) 143–53. Mercuri A. M., Bandini Mazzanti M., Torri P., Vigliotti L., Bosi G., Florenzano A., Olmi L., Massamba N’siala I. (2012) “A marine/ terrestrial integration for Mid-Late Holocene vegetation history and the development of the cultural landscape in the Po valley as a result of human impact and climate change”, Vegetation History and Archaeobotany 21 (2012) 353–72. Mercuri A. M., Sadori L., Uzquiano Ollero P. (2011) “Mediterranean and North African cultural adaptations to Mid-Holocene environmental and climatic changes”, Holocene 21 (2011) 189–206. Mercuri A. M., Florenzano A., Massamba N’siala I., Olmi L., Roubis D. and Sogliani F. (2010) “Pollen from archaeological layers and cultural landscape reconstruction: case studies from the Bradano valley (Basilicata, southern Italy)”, Plant Biosystems 144 (2010) 888–901. Mercuri A. M., Accorsi C. A., Bandini Mazzanti M., Bosi G., Cardarelli A., Labate D., Marchesini M. and Trevisan Grandi G. (2006) “Economy and environment of Bronze Age settlements— Terramaras—on the Po Plain (northern Italy): first results of the archaeobotanical research at the Terramara di Montale”, Vegetation History and Archaeobotany 16 (2006) 43–60. Mercuri A. M., Accorsi C. A. and Bandini Mazzanti M. (2002) “The long history of cannabis and its cultivation by the Romans in central Italy, shown by the pollen records from Lago Albano and Lago di Nemi”, Vegetation History and Arhaeobotany 11 (2002) 263–76. Messora R., Florenzano A., Torri P., Mercuri A. M. Muzzalupo I. and Arru L. (2017) “Morphology and discrimination features of pollen from Italian olive cultivars (Olea europaea L.)”, Grana 56 (2017) 204–14.
82 Montecchi M. C. and Mercuri A. M. (2018) “When palynology meets classical archaeology: the Roman and medieval landscapes at the Villa del Casale di Piazza Armerina, UNESCO site in Sicily”, Archaeological and Anthropological Science 10 (2018) 749–57. Panagiotopoulos K., Aufgebauer A., Schäbitz F. and Wagner B. (2013) “Vegetation and climate history of the Lake Prespa region since the Lateglacial”, Quaternary International 293 (2013) 157–69. Pepe C., Giardini M., Giraudi C., Masi A., Mazzini I. and Sadori L. (2013) “Plant landscape and environmental changes recorded in marginal marine environments: the ancient Roman harbour of Portus (Rome, Italy)”, Quaternary International 303 (2013) 73–81. Rattighieri E., Rinaldi R., Mercuri A. M., Bowes K. (2013) “Land use from seasonal archaeological sites: the archaeobotanical evidence of small Roman farmhouses in Cinigiano, south-eastern Tuscany—central Italy”, Annali di Botanica Roma 3 (2013) 207–15. Roberts N., Brayshaw D., Kuzucuoğlu C., Perez R. and Sadori L. (2011) “The Mid-Holocene climatic transition in the Mediterranean: causes and consequences”, Holocene 21 (2011) 3–13. Rohling E. J., Cane T. R., Cooke S., Sprovieri M., Bouloubassi I., Emeis K. C., Schiebel R., Kroon D., Jorissen F. J., Lorre A. and Kemp A. E. S. (2002) “African monsoon variability during the previous inter-glacial maximum”, Earth Planetary Science Letters 202 (2002) 61–75. Russo Ermolli E., Romano P., Ruello M. R., Barone-Lumaga M. R. (2014) “The natural and cultural landscape of Naples (southern Italy) during the Graeco-Roman and late antique periods”, Journal of Archaeological Science 42 (2014) 399–411. Sabato D., Masi A., Pepe C., Ucchesu M., Peña-Chocarro L., Usai A., Giachi G., Capretti C., Bacchetta G. (2015) “Archaeobotanical analysis of a Bronze Age well in Sardinia: a wealth of knowledge”, Plant Biosystems 149 (2015) 205–15. Sadori L., Giraudi C., Masi A., Magny M., Ortu E., Zanchetta G. and Izdebski A. (2016) “Climate, environment and society in southern Italy during the last 2000 years. A review of the environmental, historical and archaeological evidence”, Quaternary Science Reviews 136 (2016) 173–88. Sadori L., Giardini M., Gliozzi E., Mazzini I., Sulpizio R., Welden A. V. and Zanchetta G. (2015) “Vegetation, climate and environmental history of the last 4500 years at Lake Shkodra (Albania/ Montenegro)”, Holocene 25 (2015) 435–44. Sadori L., Bertini A., Combourieu-Nebout N., Kouli K., Mariotti Lippi M., Roberts N. and Mercuri A. M. (2013a) “Palynology and Mediterranean vegetation history”, Flora Mediterranea 23 (2013) 141–56. Sadori L., Ortu E., Peyron O., Zanchetta G., Vannière B., Desmet M. and Magny M. (2013b) “The last 7 millennia of vegetation and climate changes at Lago di Pergusa (central Sicily, Italy)”, Climate of the Past 9 (2013) 1969–84. Sadori L., Jahns S. and Peyron O. (2011) “Mid-Holocene vegetation history of the central Mediterranean”, Holocene 21 (2011) 117–29. Sadori L., Giardini M., Giraudi C. and Mazzini I. (2010a) “The plant landscape of the imperial harbour of Rome”, JAS 37 (2010) 3294–305. Sadori L., Mercuri A. M. and Mariotti Lippi M. (2010b) “Reconstructing past cultural landscape and human impact using pollen and plant macroremains”, Plant Biosystems 144 (2010) 940–51. Sadori L., Zanchetta G. and Giardini M. (2008) “Last Glacial to Holocene palaeoenvironmental evolution at Lago di Pergusa (Sicily,
Kouli et al. southern Italy) as inferred by pollen, microcharcoal, and stable isotopes”, Quaternary International 181 (2008) 4–14. Sadori L. and Giardini M. (2008) “Environmental history in the Mediterranean basin: microcharcoal as a tool to disentangle human impact and climate change”, in Charcoals from the Past: Cultural and Palaeoenvironmental Implications, edd. G. Fiorentino and D. Magri (BAR-IS 1807) (Oxford 2008) 229–36. Sadori L. and Giardini M. (2007) “Charcoal analysis, a method to study vegetation and climate of the Holocene: the case of Lago di Pergusa, Sicily (Italy)”, Geobios 40 (2007) 173–80. Sadori L. and Narcisi B. (2001) “The postglacial record of environmental history from Lago di Pergusa (Sicily)” Holocene 11 (2001) 655–671. Sevink J., van Bergen M., van der Plicht J., Feiken H., Anastasia C. and Huizinga A. (2011) “Robust date for the Bronze Age Avellino eruption (Somma-Vesuvius): 3945 ± 10 cal BP (1995 ± 10 cal BC)”, Quaternary Science Reviews 30 (2011) 1035–46. Sulpizio R., Bonasia R., Dellino P., di Vito M. A., la Volpe L., Mele D., Zanchetta G. and Sadori L. (2008) “Discriminating the long distance dispersal of fine ash from sustained columns or near ground ash clouds: the example of the Pomici di Avellino eruption (Somma-Vesuvius, Italy)”, Journal of Volcanology and Geothermal Research 17 (2008) 263–76. Thienemann M., Masi A., Kusch S., Sadori L., John S., Francke A., Wagner B. and Rethemeyer J. (2017) “Organic geochemical and palynological evidence for Holocene natural and anthropogenic environmental change at Lake Dojran (Macedonia/Greece)”, Holocene
(2017): doi: 10.1177/0959683616683261 (last accessed 2 July 2017). Triantaphyllou M. V., Pavlopoulos K. P., Kouli K. et al. (2016) “Μultiproxy paleoenvironmental reconstruction: the Piraeus coastal plain case study”, Bulletin of the Geological Society of Greece 50 (2016) 478–87. Triantaphyllou M. V., Gogou A., Bouloubassi I. et al. (2014) “Evidence for a warm and humid Mid-Holocene episode in the Aegean and northern Levantine seas (Greece, NE Mediterranean)”, Regional Environmental Change 14 (2014) 1697–712. Triantaphyllou M. V., Kouli K., Tsourou T., Koukousioura O., Pavlopoulos K. and Dermitzakis M. D. (2010) “Paleoenvironmental changes since 3000 BC in the coastal marsh of Vravron (Attica, SE Greece)”, Quaternary International 216 (2010) 14–22. Vaccaro E. (2013) “Patterning the late antique economies of inland Sicily in a Mediterranean context”, in Local Economies? Production and Exchange of Inland Regions in Late Antiquity, ed. L. Lavan (Late Antique Archaeology 10) (Leiden-Boston 2013) 259–313. Vaccaro E., Ghisleni M., Arnoldus-Huyzendveld A. et al. (2013) “Excavating the Roman peasant II: Excavations at Case Nuove, Cinigiano (GR)”, Papers of the British School at Rome 81 (2013) 129–79. Weiberg E., Unkel I., Kouli K. et al. (2016) “The socio-environmental history of the Peloponnese during the Holocene: towards an integrated understanding of the past”, Quaternary Science Reviews 136 (2016) 40–65. Zanchetta G., Van Welden A., Baneschi I. et al. (2012) “Multiproxy record for the last 4500 years from Lake Shkodra (Albania/ Montenegro)”, Journal of Quaternary Science 27 (2012) 780–89. Zohary D., Hopf M. and Weiss E. (2012) Domestication of Plants in the Old World (New York, 4th edn. 2000).
A Late Antique Vegetation History of the Western Mediterranean in Context José Antonio López-Sáez, Sebastián Pérez-Díaz, Didier Galop, Francisca Alba-Sánchez and Daniel Abel-Schaad Abstract Fossil pollen records from 70 sites with reliable chronologies and high-resolution data in the western Mediterranean, were synthesised to document Late Holocene vegetation and climate change. The key elements of vegetation dynamics and landscape construction during Late Antiquity are clear in the light of the fossil pollen records. These are: fire events (natural or anthropogenically induced); grazing activities in high-mountain areas; agriculture; arboriculture; and human settlement in the lowlands. In terms of anthropogenic pressure, the differences recorded between highlands and lowlands suggest an imbalance in land use. Such practices were related to three main types of activities: wood exploitation and management, cultivation, and pastoralism. In lowland areas there seems to be some synchronism in vegetation dynamics during the late antique period, since most of the territories of the western Mediterranean had been deforested by the Early Roman period. However, in mountainous regions, pollen records document a clear asynchrony.
Introduction Knowledge of the impact of climate change on social vicissitudes, and associated human responses, can provide an historical reference point for dealing with current and future global climate change. The impact of climate change is the result of interactions among a variety of factors in both natural and social systems, and at multiple spatial and temporal scales.1 During recent decades, the western Mediterranean region has aroused great interest in the study of Late Quaternary landscapes, as this region is very sensitive to climate change. It has also witnessed strong human influence, through polyculture, intensive grazing and deliberate use of fire, for millennia. Fossil pollen data from lakes and peat bog deposits have been frequently used in vegetation and climate reconstructions, as they tend to reflect vegetation and climate changes at a regional scale. In this sense, reconstruction of past vegetation has become important in order to understand environmental and land use changes. 1 Gil-Romera et al. (2010); Dearing et al. (2015).
This article will thus review the existing environmental evidence for the western Mediterranean from 300 BC to AD 1000, and will compare it with relevant archaeological and historical data. The western Mediterranean is an enormously complex territory from an orographic, biogeographical and historical point of view, and it is therefore extremely difficult to present a detailed synthesis without taking into account the uniqueness of each of the concerned territories. Therefore, in this article, we have decided to consider certain geographical areas separately, and ultimately discern possible asynchronism or synchronism between vegetation evolution and anthropic dynamics.2 Through land-use change, human activity could have brought about synchronous or asynchronous changes in vegetation composition, depending on the particularities and possibilities of exploitation of each territory. However, available pollen records from the western Mediterranean have been primarily focused on the Pleistocene-Holocene transition, or on the potential impact of Prehistoric communities, or rapid climate changes, on the environment during the Mid Holocene.3 In contrast, high-resolution and continuous pollen series, which extend over the last two millennia, are limited. In this review, we used 70 fossil pollen records from the western Mediterranean (fig. 1). Our review and synthesis differ from earlier published reviews,4 as we applied a uniform set of site-selection criteria, and mostly used recent pollen data. The objectives of this paper will be to assess regional vegetation patterns by synthesising fossil pollen records, and to evaluate and understand vegetation response to climate change and/ or human impact.
2 De Beaulieu et al. (2005); Giesecke et al. (2011); Luterbacher et al. (2012). 3 Carrión et al. (2010); Carrión et al. (2012); Fletcher and Zielhofer (2013); Carrión et al. (2015). 4 Hernández-Beloqui (2011); Hernández-Beloqui et al. (2015). The synthetic nature of this review means that only the most useful and relevant pollen data will be shown and discussed, so some collected types will be omitted.
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_007 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 83–104
84
figure 1
López-Sáez et al.
Pollen records mentioned in the text: 1, Cruz do Bocelo; 2, Pena da Cadela; 3, Borralleiras da Cal Grande; 4, Penido Vello; 5, Braña Rubia; 6, Alto de la Espina; 7, Monte Areo; 8, Sotombo; 9, Culazón; 10, Zalama; 11, Fuente del Vaquero; 12, Randulanda; 13, Arreo; 14, Almenara de Adaja; 15, Villafáfila; 16, Lanzahíta; 17, Serranillos; 18, Pozo de la Nieve; 19, Redondo; 20, Tiemblo; 21, Ojos Albos; 22, Peña Negra; 23, Fuente de la Leche; 24, Fuente del Pino Blanco; 25, Labradillos; 26, Charco da Candieira; 27, El Payo; 28, Maíllo; 29, La Meseguera; 30, Rascafría; 31, Peñalara; 32, Somolinos; 33, Tablas de Daimiel; 34, Patateros; 35, Siles; 36, Ojos de Villaverde; 37, Gádor; 38, Cañada de la Cruz; 39, Borreguil de la Caldera; 40, Zóñar; 41, Castelló; 42, Sobrestany; 43, Portlligat; 44, Montcortès; 45, Pla de Soulcem; 46, Pla de Labinas; 47, Estanya; 48, Prats de Vila; 49; Estanilles; 50, Redon; 51, Basa de la Mora; 52, Pélissié; 53, Col d’Ech; 54, Pla de l’Orri; 55, Pradell; 56, Gabarn; 57, Quinto Real; 58, Sourzay; 59, Artxilondo; 60, Orri de Théo; 61, Palavas; 62, La Calade; 63, Fifi; 64, Aanasser; 65, Maison Forestière; 66, Dayet M’Had; 67, Sidi Ali; 68, Tifounassine; 69, Taguelmam n’Harcha; 70, Tigalmamine.
Methodological Considerations The geographical area considered in this review covers the Iberian Peninsula (Spain and Portugal), southwestern and south-eastern France, and north-western Africa (Morocco). Abundant pollen records of various time spans and data quality are available for this area, however, many records tend to have low temporal resolution and can be discontinuous. In this study, we selected pollen sites based on three criteria: (i) a reliable chronology with a minimum of three dating control points over the last 2,300 years; (ii) high sampling resolution with a minimum of 50 years per sample; and (iii) a continuous record covering most of the Late Holocene
without a documented depositional hiatus. A total of 70 lake and peat sites are included as they fulfill these criteria. There are no satisfactory records for south-western Iberia and south-western France, and for the northern and southern Iberian plateaus there are only a few records. Therefore, we have included two sequences with a lower resolution when we discuss the general vegetation and climate history. Pollen data were obtained from the original authors or digitised from pollen diagrams in published material. Radiocarbon dating was the geochronological technique used for all pollen records in this synthesis. All dates for the reviewed sequences have been calibrated or recalibrated to calendar years BC (before Christ)
A Late Antique Vegetation History of the Western Mediterranean in Context
85
Cantabrian Range (sites 6–9) Four high-resolution pollen records are available in this region. In the lowlands, the pollen sequences of Monte Areo (200 masl), Culazón (582 masl) and Alto
de la Espina-La Molina (650 masl)7 peat bogs, show the permanence of dense deciduous forests (Quercus robur type, Corylus, Betula, Alnus) and heathlands (Erica) during the Late Iron Age (fig. 2–top). At higher altitudes, the Sotombo pollen record (1290 masl)8 shows a similar situation, namely dominant mixed forests of hazel, oak and beech, heaths and grasslands, below a belt likely consisting of high-mountain pine forests (Pinus sylvestris type). Coprophilous fungi and Cerealia are attested at this time in Sotombo, suggesting transterminant-grazing activities (fig. 2–bottom). During the Early Roman period (ca. 20 BC–AD 240), deciduous Quercus and Corylus are still the most abundant trees in the lowlands, although cultivated ones such as Castanea sativa, Olea europaea and Juglans regia show a more regular presence. In the highlands, forests remain stable, except for some development of heathlands and the expansion of beech to the detriment of hazel in Sotombo. This period documents a very significant change, characterised by high values of hydro-hygrophytes (Cyperaceae, Myriophyllum, Potamogetonaceae, Typha) in the Alto de la Espina record (fig. 2–top), suggesting an important rise in the water-table level from ca. AD 20, likely related to its Roman use as a water-deposit associated with mining activities. This is followed by several phases of a drawdown ca. AD 150 and 200 (probably drainage). Additionally, agricultural practices close to the wetland, as the high values of cereals suggest, could be related to the mining period during the early Roman empire. During the Late Roman period (ca. AD 240–450) the forests remain relatively stable in the lowlands, with less cereal and olive cultivation, and an increasing pastoral pressure (high values of coprophilous fungi). In the highlands, beech becomes more important, like grasslands, while heathlands are significantly reduced. This situation remains stable during the Germanic period (ca. AD 450–700) in the lowlands, when permanent grazing (recurrence of coprophilous fungi) is the most prominent feature, while Cerealia is almost absent. In contrast, in the highlands the situation is quite different, with declining high-mountain pine and oak forests, while hazel and grassland spread, and cereals and walnuts grow again in the Sotombo pollen record. Tree cover (Quercus robur type, Corylus, Betula) gradually diminishes through the Middle Ages (ca. AD 700–1300), whereas Cerealia, Castanea and Juglans
5 Blaauw (2010); Reimer et al. (2013). 6 Törnqvist et al. (1989); Muñoz-Sobrino et al. (2005); Mighall et al. (2006); Silva-Sánchez et al. (2014).
7 López-Merino et al. (2010); López-Merino et al. (2011); López-Sáez et al. (2013); González-Pellejero et al. (2014); López-Merino et al. (2014). 8 Pérez-Díaz et al. (2016a).
or AD (anno Domini), using the latest IntCal13 calibration dataset.5 An age-depth model was produced for each site using Clam 2.2 software. The best fit was obtained by applying a smoothing spline to the available radiocarbon dates. The confidence intervals of the calibrations and the age-depth models were calculated at 95% (2σ), with 1,000 iterations. Pollen percentages were calculated from the count data, and were based on the sum of all terrestrial pollen excluding aquatic taxa, spores and non-pollen palynomorphs. For discursive convenience, we organised the sites into seven geographical areas, as described below. We present and discuss 13 summary pollen diagrams, that are representative of the vegetation changes in these regions of different vegetation types. Northern Iberia Galician Mountains (sites 1–5) In the north-western part of the Iberian Peninsula, several pollen records of peaty deposits (Cruz do Bocelo, Pena da Cadela, Borralleiras da Cal Grande, Penido Vello and Braña Rubia6), on the Galician Mountains (600–900 masl), show the development of mesophilous forests (Alnus, Corylus, deciduous Quercus) during the Iron Age. However, evidence of human impact is detected ca. 350–270 BC by the increase of coprophilous fungi and nitrophilous taxa, indicative of animal husbandry. The most dramatic environmental change occurred at the transition from the Iron Age to the Roman period ca. 200 BC, when mesophilous forest suffers a large decline in favour of grasslands, while coprophilous fungi and nitrophilous taxa strongly increase. Although the forest recovers ca. 170 BC, thereafter, deforestation was more or less continuous from Roman times until AD 1300. This was characterised by two major phases of forest clearance: the first one during the Roman period (ca. 160 BC–AD 200); and the second during the Germanic era (ca. AD 450–700). Both phases are synchronous with grassland expansion, an increase in agricultural and tree crops (Castanea, Olea), and livestock activity.
86
figure 2
López-Sáez et al.
Percentage pollen diagrams of Alto de la Espina (top) and Sotombo (bottom) peat bogs for selected pollen and non-pollen palynomorph taxa. The grey bands represent the different phases described in the text.
A Late Antique Vegetation History of the Western Mediterranean in Context
percentages increase again, and heathlands and grasslands extend into the lowlands. Fagus sylvatica becomes the most important tree species in the mountainous areas of the Cantabrian Range ca. AD 750. Basque Mountains (sites 10–13) In the lowlands of the Basque Country at the end of the Roman period, the data show a landscape characterised by important fluctuations of tree cover ca. AD 0–450. It is essentially dominated by deciduous and evergreen oaks, beech and hazel, although evidence for anthropogenic activities is documented in the pollen records of Arreo lake (655 masl) and Randulanda peat bog (807 masl; fig. 3-top),9 suggesting a non-intensive exploitation of valleys and forested areas in a humid and warm context, related to the Ibero-Roman humid period (ca. 250 BC–AD 450). The high percentages of charcoal in the Arreo sequence indicate a high frequency of fires, most likely caused by salt exploitation of the Añana diapir in Roman times. At higher altitude, the Zalama pollen record (1330 masl)10 also documents the great abundance of forests, particularly those of deciduous Quercus and Corylus, as well as the development of heathlands and presence of Cerealia pollen (fig. 3-bottom). The lowlands of this region were intensively exploited ca. AD 450–790, especially by agricultural and livestock activities, shown by the development of grasslands and heathlands. We also see an increase in coprophilous fungi, suggesting intensive land use during the onset of the Early Middle Ages. The end of Roman domination in this territory caused important changes, linked with the disarticulation of hierarchies. By the mid 5th c. AD, many Late Roman settlements were abandoned, while other, previously secondary, ones seem to become new centres. At the same time, previously untouched, marginal areas are now populated by new villages, understood as hierarchical centres with new power structures. All these large-scale transformations affected the way of life of villagers, which, judging from the pollen records of Randulanda (fig. 3-top) and Fuente del Vaquero (722 masl), translate to a somewhat more intensive exploitation of forest spaces.11 In the Zalama sequence, this period is also characterised by an increase in pastoral pressure and the development of ferns (Pteridium aquilinum). In contrast, the decrease in pastures and
9 Pérez-Díaz and López-Sáez (2012); Corella et al. (2013); PérezDíaz and López-Sáez (2014). 10 Pérez-Díaz et al. (2016b). 11 Pérez-Díaz and López-Sáez (2014); Pérez-Díaz and López-Sáez (2015).
87
coprophilous fungi seen in the Arreo record during the Early Middle Ages, was probably linked to a decrease in salt production and the progressive depopulation of the area during an arid phase, related to the Early Medieval Cold Episode (AD 500–900). During the High Middle Ages (ca. AD 790–1200), the forests undergo a very important regression that affects both evergreen and deciduous oaks, as well as beech and hazel, in both lowlands and highlands. All taxa related to anthropogenic activities reach their maximum value at this time, such as: Cerealia, the Erica arborea type, anthropozoogenous and nitrophilous herbs, and coprophilous fungi. This period tallies with the onset of the Medieval Climate Anomaly (AD 900–1350), the most intense arid phase recorded in the pollen records, which, in the Arreo lake, is reflected by increased salinity and lower lake levels. In fact, the decreasing trend in deciduous Quercus and Fagus from the 8th c. AD in the Arreo, Fuente del Vaquero and Randulanda sequences, could be interpreted as both a progressive increase in arid conditions and a sign of increased human impact. From the 8th c. AD a dense network of medieval villages was created, in a context characterised by population growth and the expansion of cultivated areas. Central Iberia Northern Plateau (sites 14–15) Pollen records from the Almenara de Adaja (784 masl) and Villafáfila (680 masl) lagoons show in great detail the vegetation dynamics during the Late Holocene, in relation to both climatic variability and anthropic activity.12 During the Early Iron Age (ca. 1000–500 BC) the climate was humid, as shown by the development of hydro-hygrophytic elements (Cyperaceae, Phragmites, Callitriche) and non-pollen palynomorphs,13 indicative of eu-mesotrophic waters, under high human impact. This is evidenced by progressive degradation of evergreen oak woodlands, high values of both anthropozoogenous and nitrophilous herbs as well as coprophilous fungi, and cereal cultivation (fig. 4). Later, an arid phase begins, extending until ca. 300 BC, with progressing xerophytic (Artemisia) and xerohalophytic (Chenopodiaceae) communities. The period from the Late Iron Age to the end of the Roman period (ca. 300 BC–AD 450) shows a slight recovery of regional Scots pine forests (Pinus sylvestris
12 López-Merino et al. (2009a); López-Sáez et al. (2017a). 13 Hereafter referred as NPPs.
88
figure 3 Percentage pollen diagrams of Randulanda (top) and Zalama (bottom) peat bogs for selected pollen and non-pollen palynomorph taxa. The grey bands represent the different phases described in the text.
López-Sáez et al.
A Late Antique Vegetation History of the Western Mediterranean in Context
89
figure 4 Percentage pollen diagram of Almenara de Adaja lagoon (selected pollen and non-pollen palynomorph taxa). The grey bands represent the different phases described in the text.
type) and a decrease of Pinus pinaster, although with a significant advance of anthropozoogenic nitrophilous herbs, and the presence of Cerealia pollen and coprophilous fungi at the end of the Iron Age. These findings suggest a humid climate and a temperature increase, that is the so-called Ibero-Roman humid period (ca. 250 BC–AD 450). This wet and warm period is confirmed in the Almenara de Adaja and Villafáfila pollen records by increasing values of Cyperaceae. However, the Roman period shows the progressive decline of Quercus ilex type percentages, suggesting woodland clearings accompanied by the spread of Poaceae. Climatic conditions changed markedly with the onset of the Early Medieval Cold Episode (ca. AD 450/500– 900/950), a period marked by lower temperatures and greater aridity. This climatic change clearly influenced the vegetation of the northern Iberian plateau in the Early Middle Ages, as evidenced by the percentage increase of Pinus pinaster and Juniperus, and the decrease of Cyperaceae and NPPs, indicative of eu-mesotrophic waters (fig. 4). An important change in vegetation composition occurred ca. AD 400, that is a rapid decline of Quercus ilex type and a marked rise in Poaceae. This likely indicates the definitive replacement of evergreen woodlands by grasslands and agricultural crops, so conforming to the current landscape of the greater part of the northern Iberian plateau.
Iberian Central System (sites 16–32) The region of the western Mediterranean in which the most high-resolution pollen records are available is the Iberian Central System, particularly the Gredos Range in its central-western part, with ten palaeoenvironmental sequences (sites 16–25).14 Four more are available for the Francia (sites 28–29), Gata (site 27) and Estrela (site 26) ranges in the western part,15 and three others can be found for the Guadarrama (sites 30–31) and Pela (site 32) ranges in the eastern end of the Iberian Central System.16 By the time the Romans got to central Iberia, different Celtic ethnic groups inhabited the Iberian Central System ca. 400–50 BC. The western part was included in the Vettonia area, whereas the eastern highlands constituted the Carpetania region. In the Gredos and Gata ranges, within the Vettonian region, pollen records from 14 Blanco-González et al. (2009); López-Merino et al. (2009b); López-Sáez et al. (2009b); López-Sáez et al. (2010); Abel-Schaad and López-Sáez (2013); López-Sáez et al. (2014); López-Sáez et al. (2016a); López-Sáez et al. (2016b); López-Sáez et al. (2017b); Robles-López et al. (2017a); Robles-López et al. (2017b). 15 Van der Knaap and van Leeuwen (1995); Abel-Schaad et al. (2009); Morales-Molino et al. (2013); Abel-Schaad et al. (2014); Silva-Sánchez et al. (2016). 16 Ruiz-Zapata et al. (2007); Gómez-González et al. (2009); Currás et al. (2012).
90 low (Lanzahíta, Peña Negra and El Payo: 590–1,000 masl) and mid-high-altitude (Serranillos, Redondo, Fuente de la Leche and Pozo de la Nieve: 1,380–1,800 masl) peat bogs show the presence of a well-developed pine (Pinus sylvestris type) or pine-birch forest in the highlands. In the lowlands, the record indicates Pinus pinaster (central Gredos Range), deciduous oaks-hazel (western Gredos Range), or alder-birch (Gata Range) forests (figs. 5a and 5b). Despite the dominance of forests, there is some evidence for anthropogenic activities linked to grazing and agriculture. Towards the westernmost part of the massif (Estrela Range), the Charco da Candieira record (1,400 masl) shows an area of deciduous oak woodlands and heath formations in a highly anthropised environment, with a presence of cereal pollen. Within the Carpetanian region (Guadarrama Range), the high-altitude pollen record from Peñalara (1,940 masl) also shows a well-developed Scots pine forest in the highlands, and evergreen and deciduous oaks in the lowlands, with low percentages of anthropogenic pollen indicators. However, in the Lozoya-Paular valley (Guadarrama Range), the Rascafría sequence (1,113 masl) is indicative of a deforested landscape, with low tree cover and the proliferation of ruderal and anthropozoogenous taxa and shrublands. The easternmost pollen sequence, from Somolinos lake (Pela Range: 1,280 masl), shows the presence of a well-developed Scots pine forest, accompanied by evergreen and deciduous oaks, with a low representation of herbs and an increase in anthropogenic pollen indicators ca. 400 BC. The identification of Cerealia and Secale is evidence for economically productive activities in this area, and coprophilous fungi ascospores also indicate the presence of crops and some local grazing activities. Pollen diagrams during the Roman period show a higher intensity of human impact in low-altitude deposits around the western-central areas of the massif, such as Maíllo peat bog on the northern slopes (1,100 masl), or Lanzahíta, El Payo and Peña Negra on the southern (fig. 5-top). Arboreal pollen reaches its minimum values eastwards in Rascafría and Somolinos, pointing to pine wood deforestation, probably due to the use of timber for construction or as fuel. The high levels of Cerealia and Secale in Somolinos are also noteworthy. This is related to the existence of large urban centres and many small villages and farming communities, that are especially abundant in the eastern massif, but much rarer to the west. In the Gredos and Estrela ranges, the landscape appears scarcely altered by human activity during the period ca. 50 BC–AD 450, apparent in high-altitude deposits, such as those from Serranillos, Charco da Candieira, Redondo or Pozo de la Nieve (fig. 5-bottom).
López-Sáez et al.
However, at Fuente de la Leche, the progressive decline of Pinus sylvestris type suggests the existence of woodland clearings, which are accompanied by an increase in coprophilous fungi and macrocharcoal maximum values. During the Roman period, the Gredos Range was only used as an access route to other regions; the area was marginal territory and sparsely populated due to its inhospitable character. Yet, the situation is completely different in the Guadarrama Range, judging by some of the high-altitude pollen records from there, such as at Peñalara, where, ca. 100 BC–AD 450 the pine forests are largely absent, and anthropogenic pollen indicators show high values. However, all the high-altitude pollen records show the above-mentioned farming intensification during the Late Roman period, resulting in broader livestock pressure and a rise in pollen percentages of different crops. Castanea and Olea curves become continuous during this time in almost all cases, with a wider appearance of sweet chestnut westwards and olive tree eastwards. Beyond the spread of grasslands, to supply the increasing livestock density, the Iberian Central System acted as an area of transhumance through the communication network developed by the Romans, who provide the first evidence for long-distance livestock movement. The slight expansion of both evergreen and deciduous Quercus forests, favoured by the increasing temperatures and the repeated use of fire, is also significant, as is the increase in charcoal accumulation rates in most of the deposits during this period. The Tiemblo (1,250 masl) pollen record provides the first Iberian pollen sequence undertaken within a Castanea sativa-dominated woodland. This sequence is the irrefutable proof not only of the naturalness of sweet chestnut trees in central Iberia, but also of the existence of well-characterised forests of this species, during the last 3,000 years at least. The Visigothic kingdom (AD 450–711) that evolved in Hispania after the period of Roman occupation, was based on subsistence farming, and the fall of international trade eased the pressure on forest resources. However, climatic conditions changed strikingly, with the onset of the Early Medieval Cold Episode (AD 450– 950), marked by lower temperatures and greater aridity. This climatic change clearly influenced the vegetation of the Iberian Central System in the Early Middle Ages, as evidenced in Somolinos, which reports an arid phase extending from the final decades of the Late Roman period until AD 700. Birch forests increase their representation in almost all the pollen records, favoured by the colder temperatures, possibly indicating a drop in the tree line in high-altitude deposits. However, lower records located at the westernmost area also display this spread in birch, like at El Payo or Maíllo. Forests also recover in
A Late Antique Vegetation History of the Western Mediterranean in Context
figure 5 Percentage pollen diagrams of Lanzahíta (top) and Pozo de la Nieve (bottom) peat bogs for selected pollen and non-pollen palynomorph taxa, and charcoal accumulation rate-CHAR. The grey bands represent the different phases described in the text.
91
92 the highest-altitude deposits of the Guadarrama Range, such as at Peñalara. Nevertheless, the Visigothic period was a phase of large deforestation processes, within a clear livestock-oriented pattern, although cropping continues. Forests were cleared to obtain new pastures, and new lands were brought into cultivation, in a period of rural habitat proliferation and land exploitation led by peasantry. These forest clearances were mainly carried out by means of fire. This is reported in low-altitude deposits like Maíllo, Lanzahíta (fig. 5-top), Peña Negra and Rascafría, following the trend set by the Romans, but is apparent more strongly in higher altitude records, like Serranillos, Redondo, Fuente del Pino Blanco, Labradillos, Ojos Albos, Fuente de la Leche (fig. 5bottom), Pozo de la Nieve and Somolinos. This is likely related to grazing activities (high values of anthropozoogenous taxa and coprophilous fungi), in order to preserve upland pastures, but also to bring new lands into cultivation, as part of new, expansive and dynamic farming strategies. Nonetheless, some extension of olive and sweet chestnut groves is detected, especially on southern slopes. The presence of cereal pollen declines. However, rye, a cereal better adapted to low temperatures, appears on several sites during this period, like Ojos Albos or Peña Negra, pointing to increasing human pressure on the environment. Until the effective political possession and colonisation of the Iberian Central System by the northern Christian kingdoms ca. AD 1080, this massif acted as a borderline between Muslim and Christian powers for several centuries (ca. AD 711–1100). This was also a time of climatic transition, between the Early Medieval Cold Episode and the Late Medieval Warm Episode (ca. AD 950–1350), the latter characterised by similar temperatures to the current regime, moderate rainfall, but a marked dry period from the 9th to 11th c. AD At high-altitudes sites in western areas of the massif, an initial phase of forest recovery (birch, alder and deciduous oaks) is recorded ca. AD 700–950 in Charco da Candieira and Redondo, as a result of a decline in livestock and farming activities. At this time, heathlands and grasslands decreased, as well as anthropozoogenous taxa, and there is no evidence for olive or sweet chestnut cultivation. The same situation is documented on the southern slopes of the Gata Range in the El Payo sequence. Conversely, at low-altitude sites in the western Gredos Range (Peña Negra), grasslands spread again by means of fire, in order to sustain the increasing livestock population. Differences between both slopes are evidenced in the Francia Range, where the Maíllo northern record shows the dominance of birch forests, while heathlands are widely spread on the southern slopes, as recorded in the La Meseguera deposit (900 masl).
López-Sáez et al.
A similar picture is documented in the central Gredos Range, at high-altitude sites such as Ojos Albos, Fuente de la Leche, Pozo de la Nieve and Serranillos, ca. AD 680– 1010, where high-mountain pine forests recover thanks to a cooler and drier climate. There is little evidence of anthropozoogenous taxa, coprophilous fungi, or olive cultivation, and there are low values of charcoal. In contrast, in the low-altitude pollen records of the southern slopes of the Gredos Range, such as at Lanzahíta, Pinus pinaster increase considerably, as well as Cistus ladanifer, while anthropogenic taxa (Olea, pyrophilous fungi and macrocharcoal) levels decrease ca. AD 710–1120. High-altitude pollen records from the Guadarrama Range, such as Peñalara, show a very clear human impact, through both the increase in anthropozoogenous taxa and decreasing tree cover. In the Lozoya-Paular valley, the Rascafría record shows a significant increase of Plantago lanceolata type ca. AD 800–1000, as well as a decline in tree cover ca. AD 895–1170. The historical data for this period (AD 711–1100) fully corroborate the presented palynological data. They describe an economic model based on livestock grazing and small subsistence crops, managed by relatively self-sufficient peasants, in central and western areas of the Iberian Central System, in a sparsely populated land, denser on mid-altitudes than on valleys or high-altitudes.17 In the Guadarrama Range, in contrast, the settlements were mostly located in the river valleys, while in the mountains a very large number of watchtowers were built in order to guard the mountain passes.18 After this initial phase of forest recovery, forest clearance is continuous around high-altitude deposits ca. AD 950. This could be related to a seasonal exploitation of these areas, within a transterminant-type model. Deposits located at lower altitudes show a different pattern. Livestock intensification does not involve a broad clearance of forested areas, neither the use of fire, but a recovery of pine woods and oak forests in fact. This is possible thanks to pasture management, especially by means of irrigation in previously deforested areas, like Peña Negra or Rascafría. Fire and grazing indicators provide a bigger signal too, especially ca. AD 1000–1100, pointing to slightly more intense human pressure on the environment in La Meseguera in the Francia Range. Crops also improve slightly during this phase, helped by the milder climatic conditions of the Late Medieval Warm Episode. Olive and, to a lesser extent, sweet chestnut groves, grew in almost all sites, as did cereal crops, especially at the transition to the new millennium, as
17 Blanco-González et al. (2009). 18 Jiménez and Rollón (1987).
A Late Antique Vegetation History of the Western Mediterranean in Context
93
figure 6 Percentage pollen diagram of Patateros peat bog (selected pollen and non-pollen palynomorph taxa). The grey bands represent the different phases described in the text.
detected in Ojos Albos, where Cerealia and Secale increase their percentages. Southern Plateau (sites 33–34) Located in the La Mancha Plain, the Tablas de Daimiel pollen record19 comes from a core taken from a wetland area at 605 masl. Between 300 BC–AD 1000 the landscape appears to be dominated by grasslands and hydro-hygrophytic taxa, while forest formations (evergreen oak woodlands and sparse Mediterranean pines) are always very scarce, with the latter even declining from ca. AD 950. Evidence of anthropisation is confirmed throughout the sequence, especially ca. 300 BC– AD 250, whereas pastoral pressure is more evident ca. AD 250–1000. The Patateros peat bog pollen record (700 masl) is located at the northern edge of the Toledo Mountains.20 This sequence is unique because throughout its temporal development, the local vegetation has always been dominated by a shrub community of: Erica arborea type, Cistus ladanifer, Myrtus, Lonicera and Daphne gnidium type (fig. 6). There is little arboreal pollen, with low values of Quercus pyrenaica type, Q. ilex type and Pinus pinea type; while Corylus and Betula are only present
19 Gil-García et al. (2007). 20 Dorado-Valiño et al. (2014).
ca. 400–50 BC. Between 50 BC–AD 1200, local vegetation shows a greater representation of Cistus ladanifer, Poaceae, and nitrophilous and anthropozoogenous herbs; Olea europaea, Cerealia and coprophilous fungi show higher values ca. AD 500. This suggests the development of agricultural and livestock activities in the Early Middle Ages. Southern Iberia Baetic Mountains (sites 35–39) Within this study area, available palaeoenvironmental data come mainly from the south-eastern part: the Ojos de Villaverde wetland; Cañada de la Cruz and Borreguil de la Caldera peat bogs; and Siles and Gádor lakes.21 Between 400–100 BC the overall pattern of the regional landscape corresponds to high-mountain pine forests accompanied, at lower altitudes, by deciduous and evergreen Quercus, Pinus pinaster and mesophilous taxa (Betula, Corylus). In the Ojos de Villaverde (870 masl) and Siles (1,320 masl) pollen records (Segura Mountains), mountain pine forests (Pinus nigra type) decrease abruptly
21 Carrión et al. (2001a); Carrión et al. (2001b); Carrión (2002); Carrión et al. (2003); Ramos-Román et al. (2016).
94
López-Sáez et al.
figure 7 Percentage pollen diagram of Siles lake (selected pollen and non-pollen palynomorph taxa, charcoal accumulation rate-CHAR). The grey bands represent the different phases described in the text.
ca. 100 BC–AD 150, and Poaceae becomes dominant during the Roman period. Significantly, maximum values of Poaceae coincide with a remarkable increase in coprophilous fungi (Sordariaceae) percentages, suggesting an increase in grazing in the lakes’ surroundings. These findings are associated with a recurrence of anthropogenic fires (an increase in the charcoal accumulation rate) that leads to a gradual disappearance of the forest cover (fig. 7). Later, the period AD 150–1100 is characterised by rapid alternances of Pinus nigra type and Poaceae, related to frequent fire events, anthropogenic disturbance (mainly grazing pressure) and regional aridity. The Ojos de Villaverde record points to local deforestation and the expansion of pasture plants and arable agriculture, at least from AD 350, while the Siles record shows the existence of arable agriculture (Cerealia, Vitis) from AD 550. Documentary evidence is scarce, but it is clear that agricultural practices existed during the Roman period, and an important urban expansion had already taken place by the Islamic period, when the flanks of the Segura Mountains were dotted with a string of villages practising irrigation, separated by large areas of pasture.22 In the same mountain range, the Cañada de la Cruz (1,595 masl) record only begins to show evidence of human impact ca. AD 1000, probably
22 Carrión et al. (2001a).
indicating that economic activities were more important at lower altitudes. Further south, in the Gádor Range, the pollen record from Gádor lake (1,530 masl) also shows the dominance of a pine forest here until AD 330, in a context of progressive aridity. Later, Pinus decreases, while Poaceae, Juniperus and Lygeum become more frequent, as well as Sordariaceae, Polygonum aviculare type and Riccia, suggesting a substantial increase in pastoral pressure. Microcharcoal particles are especially abundant, reaching maximum values ca. 150 BC–AD 350 during the Roman period. In the Sierra Nevada Range, the pollen record from Borreguil de la Caldera (2,992 masl), unlike previous examples, and due to its location at a much higher altitude, documents the co-dominance of grassland communities and pine groves (probably Pinus sylvestris) in highland mountain areas, until 200 BC. Between 200 BC–AD 1000, pine forests are progressively reduced as a result of the pastoral pressure (an increase of Sporormiella and Sordaria values), while in the valleys olive and chestnut cultivation begins. Guadalquivir Basin (site 40) In the south-central area of southern Iberia, a key pollen record is available from the Zoñar lake,23 located at
23 Valero-Garcés et al. (2006); Martín-Puertas et al. (2008); Martín-Puertas et al. (2011).
A Late Antique Vegetation History of the Western Mediterranean in Context
300 masl in the southern margin of the Guadalquivir river basin. From AD 180–935, the regional vegetation comprised evergreen Quercus, Mediterranean pines, junipers, and some thermophilous elements, such as Olea europaea var. sylvestris, Pistacia lentiscus, Rhamnus, Lycium and Ceratonia siliqua; while local vegetation around the lake was dominated by riparian woodlands (Alnus, Fraxinus, Populus, Ulmus). High values of Myriophyllum spicatum, Botryococcus, Pediastrum and Zygnemataceae suggests high lake levels and mesoeuthrophic fresh waters. Agricultural (Cerealia) and livestock activities (Plantago lanceolata type) are attested at this time. Later, until the 11th c. AD, a growing trend in Olea indicates increasing cultivation at a regional scale, as well as around the lake; while cereal recurrence suggests a period of more intense farming during the Muslim period (AD 600–1100). Eastern Iberia Three high-resolution pollen records are available from the north-eastern regions of the Iberian Peninsula. The palaeoenvironmental study of the Castelló lagoon (site 41),24 located at the northern border of the Empordà basin, offers significant data that characterise three phases of vegetation evolution and human impact in the Mediterranean coastal area. Between 1500–800 BC, Quercus suber and Q. ilex type percentages are fairly high, probably representing mesomediterranean oak forests with some pines in the vicinity of the lagoon system. This is additionally well-discriminated by the presence of thermophilous shrubs (Cistus, Erica, Olea, Pistacia) and halophytes (Chenopodiaceae) in a flooding phase of the lagoon by the sea, when archaeological settlements located in the floodplain were abandoned. A period of increasing human impact is recorded ca. 800–150 BC by a new development of grazing activities (maximum values of coprophilous fungi) and the recording of Cerealia and Secale. These data suggest the local presence of arable fields and moderate human pressure during the Late Iron Age, concomitant with the clearance of evergreen and cork oak woodlands, and the exploitation of littoral wetlands under lower salinity conditions. This also coincides with increasing settlement, and the onset of cultural and commercial contacts between local populations and Phoenician and Greek cultures. The Roman period (150 BC–AD 750) appears to be a threshold phase in the landscape shaping of the coastal areas, in a gradual process of reduced marine influence 24 Ejarque et al. (2016).
95
with the replacement of salt vegetation by freshwater marshes, reeds and wet meadows (Cyperaceae, Poaceae). The progressive decline of tree percentages suggests woodland clearings, which are accompanied by the spread of anthropozoogenic nitrophilous grasslands and coprophilous fungi, and a more regular presence of crops (Cerealia, Secale, Castanea). This confirms increasing agropastoral exploitation around the lagoon, as a result of the intense rural occupation of this territory following Roman colonisation after the 2nd c. BC.25 Finally, ca. AD 750–1250, agropastoral activities and the cultivation of olive trees are still documented. An important change in the vegetation composition of the Castelló record occurred ca. AD 750, namely a large-scale and long-lasting clearance of littoral woodlands and the expansion of pasturelands. This likely indicates a replacement of cork oak forests by shrublands, dominated by Erica arborea type and Cistus. This major change in land use practices correlates with the Carolingian conquest and the pioneer settling of Christian monastic communities in the area. The above-mentioned colonisation of the frontier zone was incited by Carolingians through the aprisio landholding system, granting ownership of undeveloped lands to new settlers through their clearance and cultivation. The implementation of this system largely impacted on the Empordà basin by putting in place an generally open coastal landscape type.26 Regionally, this change is confirmed by the Sobrestany lagoon’s pollen record in the lower Empordà basin (site 42), where oak woodlands were cleared after the mid 7th c. AD.27 In north-eastern Iberia, on the Mediterranean coast (Cape Creus), a very interesting pollen sequence is that of Portlligat Bay (site 43).28 This sedimentary record is considered to be quite unusual, as it is a peat-like deposit derived from the accumulation of the below-ground dead parts of the marine phanerogam Posidonia oceanica through the last 4,500 years. Between AD 800–1000, the combined presence of Quercus pyrenaica type and Q. suber suggests a relatively open oak wood with cork oaks (fig. 8). The low percentages of Pinus halepensis type could be a sign of a short period of development of the coastal mesomediterranean pine forest in the area. At this time, chestnut tree (Castanea sativa) was cultivated. Regionally, high values of Pinus sylvestris type point to a wide development of high-mountain pinewoods around
25 Palet et al. (2014). 26 Chandler (2002); Ejarque et al. (2016). 27 Parra et al. (2005). 28 López-Sáez et al. (2008); López-Sáez et al. (2009a); Serrano et al. (2011); López-Merino et al. (2015).
96
López-Sáez et al.
figure 8 Percentage pollen diagram of Portlligat Bay peat-like deposit (selected pollen and non-pollen palynomorph taxa). The grey bands represent the different phases described in the text.
the central-eastern foothills of the Pyrenees. Shrub elements show a remarkable importance, with Erica arborea type and Cistus ladanifer as main taxa, as well as a xerothermophilous macchia composed of myrtles (Myrtus), Phillyrea and wild olive tree pollen seen. The presence of anthropozoogenous herbs (Urtica, Plantago lanceolata type) and coprophilous fungi (Sordaria) is indicative of significant local livestock pressure. Unlike the two above-mentioned pollen records, in the Portlligat sequence, oak woodlands remained untouched until the 11th c. AD, stressing the complexity of landscape management in north-east Iberia. During the 11th–12th c. (AD 1000–1150), a drastic deforestation of high-mountain pine forests in the eastern Pyrenees occurred; chestnut cultivation ceased, attended by a local deforestation of oak woodlands, with a resulting proliferation of heathers. These findings are related to an increase in medieval settlement in the eastern Pyrenees, probably benefiting from the mild and humid climate, associated with the Late Medieval Warm Episode. Interestingly, this population shift towards the high areas of the Pyrenees causes local pastoral pressure to disappear.
extend back at least 7,000 years.29 It is located on the southern flank of the central Pyrenees (1,027 masl). Its pollen record shows that the pre-feudal period (8th– 9th c. AD) were characterised by the presence of welldeveloped pine forests (Pinus nigra type), while crops and weeds were almost absent. An intensification of burning practices at the beginning of the feudal period, ca. AD 900, led to intensive clearance of woodlands and to their progressive replacement by rye (Secale) and hemp (Cannabis) cultivation, as well as by meadows and pastures. In the nearby pollen records of the Prats de Vila (1,150 masl) and Estanilles (2,247 masl) peat bogs and Redon lake (2,240 masl) (sites 48–50), a dense forest of firs (Abies alba) and pines dominated the landscape ca. 250–50 BC.30 Later, the fir forest suffered a strong regression from 50 BC, derived from the selective use of its wood in metallurgical activities, particularly during the 3rd and 4th c. AD, coinciding with the occasional presence of Juglans and Cerealia, although the pine forest remained relatively stable. Between AD 650–1150, pine and fir forests suffered a strong regression, related to the extension of chestnut and cereal crops, as well as a high pastoral pressure.
Pyrenees The small lake of Montcortès (site 44) is especially wellsuited for high-resolution palaeoecological reconstruction due to its annually laminated sediments, which
29 Corella et al. (2011); Rull et al. (2011); Rull and Vegas-Villarrúbia (2014); Rull and Vegas-Villarrúbia (2015). 30 Pèlachs et al. (2009); Pérez-Obiol et al. (2012); GonzálezSampériz et al. (2017).
A Late Antique Vegetation History of the Western Mediterranean in Context
Further south, in the Pre-Pyrenees, the palaeoenvironmental record (site 47) from Estanya lake31 (670 masl) shows a relatively forested and well-preserved landscape of deciduous and evergreen Quercus, where Mediterranean pines were also an important element ca. 500 BC–AD 750 (Late Roman and Visigothic era). Cultivated taxa (Olea europaea, Vitis, Cerealia, Cannabis) are abundant, although the presence of nitrophilous plants is still very small, suggesting that stockbreeding and grazing were less developed. The highest values of Cerealia pollen are found ca. 50 BC, coinciding with the Roman period. In contrast, a lower-degree of anthropogenic impact is recorded at the highland locations of the central Pyrenees. Not even the Redon or Basa de la Mora32 (1,914 masl) lacustrine sequences (sites 50–51) show cultivation, stockbreeding or deforestation signals during this period, as at Estanilles (site 49). Forests were slightly reduced by means of fire ca. AD 750–1000 in the Estanya record (site 47), coinciding with the introduction of new crops (Juglans, Secale) during the Muslim period. Human deforestation processes have been also recognised at higher altitude in the above-mentioned Estanilles and Redon records from AD 750, although this phase is not clearly identified in the Basa de la Mora sequence (sites 49–51). At high-altitude sites on the northern flank of the central Pyrenees,33 the pollen records from the Pla de Labinas (1,800 masl), Orri de Théo (1,640 masl) and Pla de Soulcem (1,520 masl) peat bogs (sites 45–46, 60) show a well-developed pine-fir forest during the Iron Age, accompanied by birches, hazel and beech. Although the anthropic use of fire is attested, its incidence in the forests was scarce. From the Gallo-Roman period (ca. 50 BC–AD 400), an increase in human impact is recorded by a pronounced decline in pine cover and the increase of heliophilous species, such as birch and beech, and grasslands. This process is even more evident during the second half of the Early Middle Ages ca. AD 400–1000, when the presence of cereal pollen is attested. The same dynamics is observed at lower altitude in the Pélissié (312 masl) and Col d’Ech (710 masl) pollen records (sites 52–53),34 although with a greater amount of deciduous oak and beech forests; whereas the cultivation of cereals is attested from AD 250 and that of rye from AD 700. In the eastern Pyrenees, the Pla de l’Orri (2,105 masl) and Pradell (1,975 masl) sequences (sites 54–55) follow the
31 Riera et al. (2004); Riera et al. (2006); Morellón et al. (2008); Morellón et al. (2009); Morellón et al. (2011); GonzálezSampériz et al. (2017). 32 Pérez-Sánz et al. (2011); Pérez-Sánz et al. (2013). 33 Galop and Jalut (1994). 34 Rius et al. (2012); Cunill et al. (2015).
97
same pattern mentioned above, for sites located at high altitude in the central Pyrenees.35 In the western Pyrenees, pollen records from the Gabarn (310 masl), Quinto Real (910 masl), Artxilondo (950 masl) and Sourzay (1130 masl) peat bogs (sites 56–59),36 show the existence of well-preserved oak and beech forests during much of the Iron Age, witnessing the cultivation of cereal in the valley areas. The end of the Iron Age and antiquity (ca. 150 BC–AD 400) is reflected in a spatial extension of anthropogenic activities, including woodland clearances, grazing pressure, pastureland development and anthropic fire episodes, as well as metallurgical activities (fig. 9). The steady level of Cerealia pollen suggests that this increase in human impact occurred at a greater scale. During the Middle Ages (ca. AD 400–1000) the fire regime is characterised by high fire frequencies, and a diversification in cultivated crops (the first occurrences of Secale), and by the expansion of husbandry activities to their greatest extent. South-Eastern France Only two high-resolution pollen records are available from this region. The first one comes from the Palavas lagoon (site 61), located in the Languedoc under both Mediterranean and Atlantic climatic influences.37 Between ca. 500–100 BC, trees dominate the pollen spectra, mainly comprising deciduous and evergreen Quercus, associated with Pinus; Juglans, Vitis and Cerealia are recorded sporadically during the Late Iron Age (fig. 10). Later, ca. 100 BC–AD 600, forest formations were gradually degraded, while bushes (Ericaceae) increased, with a continuation of olive cultivation throughout the period. These findings can be explained by taking into account settlement migration from the hinterland to the coastal areas; most villae were located in coastal plains. Finally, ca. AD 650–750, deciduous Quercus forest recovery appears to be occurring in the transition to the Early Middle Ages, coinciding with a gap in the archaeological record. Although later, ca. AD 800–1000, such forest is reduced again, coinciding with greater agricultural and livestock activity (maxima of Olea and Plantago; presence of Juglans, Castanea, Vitis and Cerealia) and an evergreen shrubland extension, pointing to a strong human influence on the environment during the High Middle Ages.
35 Vannière et al. (2001); Ejarque et al. (2009). 36 Galop et al. (2001); Galop et al. (2002); Monna et al. (2004); Galop (2005); Jouffroy-Bapicot et al. (2007); Mazier et al. (2009); Rius et al. (2009); Pérez-Díaz et al. (2015). 37 Azuara et al. (2015).
98
López-Sáez et al.
figure 9 Percentage pollen diagram of Quinto Real peat bog (selected pollen taxa) with Pb/Sc and 206Pb/207Pb ratios reflecting historical mining activities. The grey bands represent the different phases described in the text.
Rif Mountains (sites 63–66) Pollen records from this region show the dominance of cedar (Cedrus atlantica) and mirbeck’s oak (Quercus canariensis, Q. pyrenaica) forests ca. 1000 BC–AD 400, likely related to increasing moisture, suggested by the general sharp increase of hydro-hygrophytic taxa. This picture is confirmed by several pollen records in the western
Rif Mountains (fig. 11):39 Maison Forestière (1,602 masl) and Dayet M’Had (754 masl) in the Jbel Bou Hachen (sites 65–66); Aanasser (1,342 masl) in the Jbel Tizirène (site 64); and Fifi (1,200 masl) in the Jbel Khesana (site 63). At this time, agricultural activities (Cerealia, Vicia faba type) are attested. It seems that anthropogenic activities (fire and grazing) during the Phoenician period ca. 1000 BC–AD 50 did not impact greatly on a welldeveloped mirbeck’s oak-cedar forest. The further clearance of Quercus canariensis forests ca. AD 50 coincides with the development of cork oak and holm oak woodlands (Quercus suber, Q. ilex) at lower altitudes, and with the beginning of cedar decline. This period corresponds to the Roman colonisation of northern Morocco, and to the foundation of the Roman province of Mauretania Tingitana (AD 40). The area of modern day Morocco remained a part of the Roman empire until AD 429, when Vandals displaced Romans. After ca. AD 500, the Rif remained occupied by the indigenous Berber inhabitants and the Byzantines, until the arrival of the Arabs in AD 681. This set of events gave rise to a remarkable phase of deforestation, which took place in the western Rif Mountains ca. AD 400–700, carried out by fire, leading forest levels to their minima, while grasslands and shrublands spread. Later, ca. AD 700–1000, humid conditions, shown by a rise in aquatic taxa, allowed forests to recover, especially Quercus canariensis/pyrenaica and Q. suber ones, while Cedrus always remained at a significant percentage. The absence of coprophilous fungi points to scarce grazing activity. Cereal and bean cultivation continued, with the first occurrences of Linum (AD 400) and Cannabis
38 Andrieu-Ponel et al. (2000).
39 Cheddadi et al. (2015); Muller et al. (2015); Abel-Schaad et al. (2016).
To the east of the Palavas lagoon, the La Calade record (site 62) comes from a wetland located in the lower Rhône floodplain (4 masl).38 In a first phase, corresponding to the period of Greek presence at Marseille and to the Roman empire in Provence, the sequence shows a similar vegetation pattern to that seen in Palavas ca. 500 BC–AD 600. That is a dominance of oak woodland with Mediterranean pines, which are gradually reduced; and a cultivation of walnut, chestnut, olive, grape and cereals including rye. Cattle breeding was also practised locally, as inferred by the presence of anthropozoogenous taxa (Chenopodiaceae, Plantago lanceolata type, Polygonum aviculare type). During a second phase, forest cover recovers in the Merovingian period; although it quickly degrades again in a third phase. This is as a result of a very high pastoral pressure during the High Middle Ages, while cereals and rye are still cultivated, but with greater intensity. This phase could be related to the establishment of a monastic community (Montmajour Abbey, mid-9th c.) and to an increasing number of farmers in the surrounding area. North-Western Africa
A Late Antique Vegetation History of the Western Mediterranean in Context
figure 10 Percentage pollen diagram of Palavas lagoon (selected pollen taxa). The grey bands represent the different phases described in the text.
figure 11
Percentage pollen diagram of Fifi wetland (selected pollen and non-pollen palynomorph taxa). The grey bands represent the different phases described in the text.
99
100 (AD 750). At later dates (AD 900–1300), the cedar progressively declined, until it disappears in the Maison Forestière sequence ca. AD 1050. Middle Atlas Mountains (sites 67–70) Four high-resolution pollen records are available from lakes located in this area,40 in the present-day cedar-holm oak woodland. These are Taguelmam n’Harcha (1,640 masl), Tigalmamine (1,626 masl), Tifounassine (1,913 masl) and Sidi Ali (2,080 masl). At Tigalmamine, before 200 BC, pollen spectra are dominated by the Quercus ilex type, Cedrus atlantica and Poaceae. However, from that date on, a slow decline in forest formations began, parallel to a greater abundance of anthropozoogenous herbs, showing that pastoral pressure was responsible for such a decline. At Taguelmam n’Harcha, forest clearance and grazing became more intense ca. AD 300, when the presence of cultivated cereals is detected. At Sidi Ali and Tifounassine there is some weak evidence for human impact ca. 1000 BC, but it seems that anthropogenic processes became evident only ca. 400–200 BC. This evidence for landscape anthropisation results from the Roman expansion into the Middle Atlas, as Roman populations settled the town of Volubilis, near Meknes. Discussion and Conclusions Palynological investigations of human impact in the mountains of north-western Morocco suggest that human use of forest resources was less drastic prior to the expansion of the Roman empire. However, pollen data show a regression of forest taxa ca. 200 BC in the Middle Atlas and ca. AD 50 in the Rif Mountains, which suggests an increase in the human use of high-altitude forests from the Roman period. The contemporaneous vegetation disturbances recorded in the Middle Atlas and the Rif Mountains resulted from large-scale anthropogenic forest disturbances, which nevertheless appear very weak in comparison to those occurring at the same time in south-western Europe. Indeed, the activities of Roman/Romanised populations triggered, almost everywhere in south-western Europe, dramatic deforestations, accompanied by severe ecosystem modifications.41 In the south-east of Iberia, several records from the Baetic Mountains demonstrate the millennial persistence of pine forests until 100 BC, as well as a drastic deforestation of these during the Roman period ca. 100
40 Lamb et al. (1989); Lamb et al. (1991); Lamb and van der Kaars (1995); Lamb et al. (1999); Cheddadi et al. (2015). 41 López-García (1994); De Beaulieu et al. (2005); Blondel (2006).
López-Sáez et al.
BC–AD 150, linked to an increase in pastoral pressure and the occurrence of anthropic fires. A progressive decline in tree cover during the Roman period has also been documented in the northern Iberian plateau and south-eastern France. In the Iberian Central System, at the time of the Late Iron Age (ca. 400–50 BC), anthropic dynamics had different impacts on each mountain range considered here. Within the Vettonian region, forests are well-developed both in highland and lowland areas. Within the Carpetanian region, high-mountain forests are also welldeveloped, but the lowlands, especially easternmost ones, evidence a phase of socio-cultural change. This is shown by an increase in the number of settlements, territorial reorganisation and new land use systems, resulting in a deforested landscape with evidence for both grazing and farming activities. During the Roman period (ca. 50 BC–AD 450), large-scale deforestations resulted in open landscapes characterising lowland areas, as a consequence of the intensification of human activities, like cropping, grazing, mining or forestry. However, anthropogenic dynamics in the highlands was different in each area of the massif, and was connected to mining (western areas) or greater settlement (eastern areas), while the Gredos Range (central area) was hardly altered at all, due to its inaccessibility. The decline of western Mediterranean forests, although initiated during the Roman era, really began several centuries later. In the Baetic Mountains, from the Roman period until the 10th c. AD, the vegetation shows an alternating dynamic, related to livestock activities and fire events. Evidence of agriculture is initially detected ca. AD 350 at low altitude, and then at higher altitude ca. AD 550. At higher altitudes, human impact is only evident from ca. AD 1000. In the north-eastern regions of Iberia, from the Iron Age to the end of the ancient period (1500 BC–AD 750) human settlement and agropastoral exploitation of wetland ecosystems responded to natural coastal dynamics. During this period, human land uses produced a limited impact on the lagoon landscapes, that were dominated by evergreen and cork oak woodlands. In contrast, societies become a major agent of landscape transformation ca. AD 750–1250, actively controlling the lagoon dynamics and resources. Evidence for this can be seen in the degradation of coastal forests and the expansion of agricultural and livestock activities from the 8th c. AD. In the lowlands of the southern Pyrenees, pollen records show a strong deforestation phase of the pine forest from AD 900, linked to the cultivation of hemp and rye. Pollen data from south-eastern France show a period of widespread deforestation at the beginning of the Middle Ages ca. AD 800–1000, after an earlier,
A Late Antique Vegetation History of the Western Mediterranean in Context
limited recovery of the forest ca. AD 650–750. In the Rif Mountains, pollen records suggest significant deforestation ca. AD 400–700 during the period of Byzantine dominance. The disappearance of evergreen oak woodlands on the northern Iberian plateau is attested ca. AD 400. Deforestation during the Germanic era, also represented a tipping point in the history of mesophilous forests in the northern Iberian mountains ca. AD 450–700. In the lowlands of the Cantabrian Range, the Middle Ages led to the onset of a gradual but permanent deforestation. Previously forested areas in northern Iberia were occupied by heathlands and grasslands. At the beginning of the Middle Ages, the intensity of human impact on the environment increased notably ca. AD 450–700 after the collapse of the Roman system in the Iberian Central System and the Toledo Mountains. This is documented by large deforestation processes at high altitudes, and by a rise in anthropogenic pollen indicators. This is within a clear livestock-oriented pattern, although new lands were brought into cultivation with the onset of the Early Medieval Cold Episode (AD 450–950). An initial phase of forest recovery (birch, alder and deciduous oaks in western areas; pine in central and eastern areas) is recorded in high-mountain environments from AD 700–950, as a result of a decline in livestock and farming activities, while agropastoral pressure increases at low altitudes. Then a new phase of forest clearance takes place ca. AD 950–1100, related to transterminant movements and the milder weather conditions of the Late Medieval Warm Episode (AD 950–1350). In the lowlands, livestock intensification did not involve a broad clearance of forested areas, only some diversification in crops. Landscape clearance after ca. 150 BC–AD 300 does not have to be primarily determined by increased human impact; a climatic explanation is also plausible. In fact, pollen data from the Baetic Range document an increasing trend in dryness from the Mid Holocene, as shown by the progressive decrease in tree species (mainly pines) and the increase in xerophytes, such as Artemisia. The semi-desert expansion and high-mountain forest decline during the Late Holocene in southern Iberia could be also explained by decreasing summer insolation.42 One of the most conspicuous features of the Zoñar pollen record is the succession of wet and dry conditions at a centennial-scale, generally synchronous with rapid climate changes that occurred in the Mediterranean basin during the Late Holocene. Arid phases are identified prior to 950 BC, in the Roman period from 150
42 Jiménez-Moreno et al. (2015).
101
BC–AD 150, and AD 800–1300.43 The second phase coincides with documented deforestation processes in the Gádor, Segura and Sierra Nevada ranges, mediated by a higher incidence of forest fires. More humid conditions occurred ca. AD 150–350 during the later Roman empire. This period coincides with a recovery in high mountain pine forests throughout the whole of the Baetic Mountains, as well as with the first evidence for agriculture there. It is interesting to note that the most important recorded landscape changes in the western Mediterranean region coincide with the lower temperatures and greater aridity of the Early Medieval Cold Episode (AD 500–900); although such changes are even more evident with the onset of the Medieval Climate Anomaly (MCA) (AD 900– 1350).44 In Mediterranean Iberia, the MCA was a period characterised by decreased lake levels, more xerophytic and heliophytic vegetation (Juniperus, Artemisia), and a low frequency of floods, suggesting warmer yet still arid conditions.45 Acknowledgements The research presented in this paper was funded by the following projects: Desirè-HAR2013–43701-P (Plan Nacional I+D+I, Spanish Ministry of Economy and Competitiveness); and Relic-Flora-P11-RNM-7033 (Excellence Research Projects Program from the Andalusian Government). Bibliography Abel-Schaad D. and López-Sáez J. A. (2013) “Vegetation changes in relation to fire history and human activities at the Peña Negra mire (Bejar Range, Iberian Central Mountain System, Spain) during the past 4.000 years”, Vegetation History and Archaeobotany 22 (2013) 199–214. Abel-Schaad D. et al. (2016) “Declive de los bosques de cedro (Cedrus atlantica) durante el Holoceno reciente en el sector occidental de la cordillera del Rif (Marruecos)”, in Avances en Biogeografía: áreas de distribución: entre puentes y barreras, edd. J. Gómez, J. Arias, J. A. Olmedo and J. L. Serrano (Granada 2016) 592–600. Abel-Schaad D. et al. (2014) “Heathlands, fire and grazing. A paleoenvironmental view of Las Hurdes (Cáceres, Spain) history during the last 1200 years”, Forest Systems 23 (2014) 247–58. Abel-Schaad D. et al. (2009) “Evolución de la vegetación en la Sierra de Gata (Cáceres-Salamanca, España) durante el Holoceno
43 Martín-Puertas et al. (2010); Martín-Puertas et al. (2011). 44 Morellón et al. (2011); Morellón et al. (2012); Blanco-González et al. (2015). 45 Morellón et al. (2012); Moreno et al. (2012); Roberts et al. (2012); González-Sampériz et al. (2017).
102 reciente. Implicaciones biogeográficas”, Revista Española de Micropaleontología 41 (2009) 91–105. Andrieu-Ponel V. et al. (2000) “Palaeoenvironments and cultural landscapes of the last 2000 years reconstructed from pollen and coleopteran records in the Lower Rhône Valley, southern France”, The Holocene 10 (2000) 341–55. Azuara J. et al. (2015) “Late Holocene vegetation changes in relation with climate fluctuations and human activity in Languedoc (southern France)”, Climate of the Past 11 (2015) 1769–84. Blaauw M. (2010) “Methods and code for classical age-modelling of radiocarbon sequences”, Quaternary Geochronology 5 (2010) 512–18. Blanco-González A. et al. (2015) “Medieval landscapes in the Spanish Central System (450–1350): a palaeoenvironmental and historical perspective”, Journal of Medieval Iberian Studies 7 (2015) 1–17. Blanco-González A. et al. (2009) “Ocupación y uso del territorio en el sector centromeridional de la Cuenca del Duero entre la Antigüedad y la Alta Edad Media (siglos I–XI d.C.)”, Archivo Español de Arqueología 82 (2009) 275–300. Blondel J. (2006) “The ‘design’ of Mediterranean landscapes: a millennial story of humans and ecological systems during the historic period”, Human Ecology 34 (2006) 713–29. Carrión J. S. (2002) “Patterns and processes of Late Quaternary environmental change in a montane region of southwestern Europe”, Quaternary Science Reviews 21 (2002) 2047–66. Carrión J. S. et al. (2015) Cinco millones de años de cambio florístico y vegetal en la Península Ibérica e Islas Baleares (Murcia 2015). Carrión J. S. et al. (2012) Paleoflora y paleovegetación de la Península Ibérica e Islas Baleares: Plioceno-Cuaternario (Madrid 2012). Carrión J. S. et al. (2010) “Expected trends and surprises in the Late Glacial and Holocene vegetation history of the Iberian Peninsula and Balearic Islands”, Review of Palaeobotany and Palynology 162 (2010) 458–75. Carrión J. S. et al. (2003) “Holocene vegetation dynamics, fire and grazing in the Sierra de Gádor, southern Spain”, The Holocene 13 (2003) 839–49. Carrión J. S. et al. (2001a) “Abrupt vegetation changes in the Segura Mountains of southern Spain throughout the Holocene”, Journal of Ecology 89 (2001) 783–97. Carrión J. S. et al. (2001b) “Crossing forest thresholds: inertia and collapse in a Holocene sequence from south-central Spain”, The Holocene 11 (2001) 635–53. Chandler C. J. (2002) “Between court and counts: Carolingian Catalonia and the aprisio grant, 778–897”, Early Medieval Europe 11 (2002) 19–44. Cheddadi R. et al. (2015) “A history of human impact of Moroccan mountain landscapes”, African Archaeological Review 32 (2015) 233–48. Corella J. P. et al. (2013) “A 2500-year multi-proxy reconstruction of climate change and human activities in northern Spain: the Lake Arreo record”, Palaeogeography, Palaeoclimatology, Palaeoecology 386 (2013) 555–68. Corella J. P. et al. (2011) “Climate and human impact on a meromictic lake during the last 6,000 years (Montcortès lake, central Pyrenees, Spain)”, Journal of Paleolimnology 46 (2011) 351–67. Cunill R. et al. (2015) “Paleoecological study of Pyrenean lowland fir forests: exploring Mid-Late Holocene history of Abies alba in Montbrun (Ariège, France)”, Quaternary International 366 (2015) 37–50. Currás A. et al. (2012) “Climate change and human impact in central Spain during Roman times: high-resolution multi-proxy analysis of a tufa lake record (Somolinos, 1280 masl)”, Catena 89 (2012) 31–53. Dearing J. A. et al. (2015) “Socio-ecological systems in the Anthropocene: the need for integrating social and biophysical
López-Sáez et al. records at regional scales”, The Anthropocene Review 2 (2015) 220–46. De Beaulieu J. L. et al. (2005) “Vegetation dynamics in north-western Mediterranean regions: instability of the Mediterranean bioclimate”, Plant Biosystems 139 (2005) 114–26. Dorado-Valiño M. et al. (2014) “Contributions to the European Pollen Database. 21. Patateros, Toledo Mountains (central Spain)”, Grana 53 (2014) 171–73. Ejarque A. et al. (2016) “Coastal evolution in a Mediterranean microtidal zone: Mid to Late Holocene natural dynamics and human management of the Castelló Lagoon, NE Spain”, PLoS ONE 11 (2016) e0155446: http://dx.doi.org/10.1371/jour nal.pone.0155446 (last accessed March 2017). Ejarque A. et al. (2009) “Tracing the history of highland human management in the eastern Pre-Pyrenees: an interdisciplinary palaeoenvironmental study at the Pradell fen, Spain”, The Holocene 19 (2009) 1241–55. Fletcher W. J. and Zielhofer C. (2013) “Fragility of western Mediterranean landscapes during Holocene rapid climate changes”, Catena 103 (2013) 16–29. Galop D. (2005) “La conquête de la montagne pyrénéenne au Néolithique. Chronologie, rythmes et transformations des paysages à partir des données polliniques”, in Populations néolithiques et environnement, ed. J. Guilaine (Paris 2005) 279–85. Galop D. and Jalut G. (1994) “Differential human impact and vegetation history in two adjacent Pyrenean valleys in the Ariège basin, southern France, from 3000 B.P. to the present”, Vegetation History and Archaeobotany 3 (1994) 225–44. Galop D. et al. (2002) “Métallurgie et histoire de l’environnement au cours des cinq derniers millénaires en pays Basque nord (Vallée de Baigorri, Pyrénées Atlantiques, France): résultats préliminaires d’une approche interdisciplinaire”, Archeologia Postmedievale 6 (2002) 155–69. Galop D. et al. (2001) “Cinq millénaires de métallurgie en montagne Basque. Les apports d’une démarche intégrée alliant palynologie et géochimie isotopique du plomb”, Sud-Ouest Européen 11 (2001) 3–15. Giesecke T. et al. (2011) “The pace of Holocene vegetation change— testing for synchronous developments”, Quaternary Science Reviews 30 (2011) 2805–14. Gil-García M. J. et al. (2007) “Late Holocene environments in Las Tablas de Daimiel (south-central Iberian península, Spain)”, Vegetation History and Archaeobotany 16 (2007) 241–50. Gil-Romera G. et al. (2010) “Interpreting resilience through long-term ecology: potential insights in western Mediterranean landscapes”, The Open Ecology Journal 3 (2010) 43–53. Gómez-González C. et al. (2009) “Evolución del paisaje vegetal durante los últimos 1.680 años BP en el Macizo de Peñalara (Sierra de Guadarrama, Madrid)”, Revista Española de Micropaleontología 41 (2009) 75–89. González-Pellejero R. et al. (2014) “Dinámicas naturales y antrópicas en los paisajes vegetales de los valles internos de Cantabria occidental (norte de España)”, Boletín de la Asociación de Geógrafos Españoles 65 (2014) 139–65. González-Sampériz P. et al. (2017) “Environmental and climate change in the southern central Pyrenees since the Last Glacial Maximum: a view from the lake records”, Catena 149 (2017) 668–88. Hernández-Beloqui B. (2011) “La palinología aplicada al estudio de contextos de cronología medieval en la península Ibérica: estado de la cuestión”, Revista Arkeogazte 1 (2011) 97–124. Hernández-Beloqui B. et al. (2015) “The Late Holocene in the western Pyrenees: a critical review of the current situation of palaeopalynological research”, Quaternary International 364 (2015) 78–85. Jiménez J. and Rollón A. (1987) Guía de los castillos de Madrid (Madrid 1987).
A Late Antique Vegetation History of the Western Mediterranean in Context Jiménez-Moreno G. et al. (2015) “Impact of Late-Holocene aridification trend, climate variability and geodynamic control on the environment from a coastal area in SW Spain”, The Holocene 25 (2015) 607–17. Jouffroy-Bapicot et al. (2007) “Environmental impact of early palaeometallurgy: pollen and geochemical analysis”, Vegetation History and Archaeobotany 16 (2007) 251–58. Lamb H. F. and van der Kaars S. (1995) “Vegetation response to Holocene climatic change: pollen and palaeolimnological data from the Middle Atlas, Morocco”, The Holocene 5 (1995) 400–408. Lamb H. F. et al. (1999) “Lake evolution in a semi-arid montane environment: responses to catchment change and hydroclimatic variation”, Journal of Paleolimnology 21 (1999) 325–43. Lamb H. F. et al. (1991) “Human impact on the vegetation of the Middle Atlas, Morocco, during the last 5000 years”, Journal of Biogeography 18 (1991) 519–32. Lamb H. F. et al. (1989) “An 18,000-year record of vegetation, lake-level and climatic change from Tigalmamine, Middle Atlas, Morocco”, Journal of Biogeography 16 (1989) 65–74. López-García P. (1994) “Forest, forest clearance and open land during the time of the Roman empire in Spain”, European Palaeoclimate and Man 5 (1994) 23–35. López-Merino L. et al. (2015) “Glomalin accumulated in seagrass sediments reveals past alterations in soil quality due to land-use change”, Global and Planetary Change 133 (2015) 87–95. López-Merino L. et al. (2014) “Reconstructing the impact of human activities in a NW Iberian Roman mining landscape for the last 2500 years”, JAS 50 (2014) 208–18. López-Merino L. et al. (2011) “Human-induced changes on wetlands: a study case from NW Iberia”, Quaternary Science Reviews 30 (2011) 2745–54. López-Merino L. et al. (2010) “Early agriculture and palaeoenvironmental history in the north of the Iberian Peninsula: a multiproxy analysis of the Monte Areo mire (Asturias, Spain)”, JAS 37 (2010) 1978–88. López-Merino L. et al. (2009a) “Estudio polínico de una laguna endorreica en Almenara de Adaja (Valladolid, Meseta Norte): cambios ambientales y actividad humana durante los últimos 2.800 años”, Revista Española de Micropaleontología 41 (2009) 333–47. López-Merino L. et al. (2009b) “2000 years of pastoralism and fire shaping high-altitude vegetation of Sierra de Gredos in central Spain”, Review of Palaeobotany and Palynology 158 (2009) 42–51. López-Sáez J. A. et al. (2017a) “Una perspectiva paleoambiental de la explotación de la sal en las Lagunas de Villafáfila (Tierra de Campos, Zamora)”, Cuaternario y Geomorfología (forthcoming). López-Sáez J. A. et al. (2017b) “Unraveling the naturalness of sweet chestnut forests (Castanea sativa Mill.) in central Spain”, Vegetation History and Archaeobotany 26 (2017) 167–82. López-Sáez J. A. et al. (2016a) “Exploring seven hundred years of transhumance, climate dynamic, fire and human activity through a historical mountain pass in central Spain”, Journal of Mountain Science 13 (2016) 1139–53. López-Sáez J. A. et al. (2016b) “Landscape dynamics and human impact on high-mountain woodlands in the western Spanish Central System during the last three millennia”, JAS: Reports 9 (2016) 203–18. López-Sáez J. A. et al. (2014) “Vegetation history, climate and human impact in the Spanish Central System over the last 9,000 years”, Quaternary International 353 (2014) 98–122. López-Sáez J. A. et al. (2013) “Contributions to the European Pollen Database. Culazón, Cantabrian mountains (northern Spain)”, Grana 52 (2013) 316–18. López-Sáez J. A. et al. (2010) “Late Holocene ecological history of Pinus pinaster forests in the Sierra de Gredos of central Spain”, Plant Ecology 206 (2010) 195–209. López-Sáez J. A. et al. (2009a) “Palaeoecological potential of the marine organic deposits of Posidonia oceanica: a case study in
103
the NE Iberian Peninsula”, Palaeogeography, Palaeoclimatology, Palaeoecology 271 (2009) 215–24. López-Sáez J. A. et al. (2009b) “Contribución paleoambiental al estudio de la trashumancia en el sector abulense de la Sierra de Gredos”, Hispania. Revista Española de Historia 231 (2009) 9–38. López-Sáez J. A. et al. (2008) “Historia de la vegetación en el litoral norte de Girona entre los siglos VIII y XX d.C.: cambios climáticos y socioeconómicos desde una perspectiva paleoambiental”, Arqueología y Territorio Medieval 15 (2008) 13–28. Luterbacher J. et al. (2012) “A review of 2000 years of paleoclimatic evidence in the Mediterranean”, in The Climate of the Mediterranean Region: from the Past to the Future, ed. P. Lionello (Amsterdam 2012) 87–185. Martín-Puertas C. et al. (2011) “Geochemical processes in a Mediterranean lake: a high-resolution study of the last 4,000 years in Zoñar lake, southern Spain”, Journal of Paleolimnology 46 (2011) 405–21. Martín-Puertas C. et al. (2010) “Late Holocene climate variability in the southwestern Mediterranean region: an integrated marine and terrestrial geochemical approach”, Climate of the Past 6 (2010) 807–16. Martín-Puertas C. et al. (2008) “Arid and humid phases in southern Spain during the last 4000 years: the Zoñar lake record, Córdoba”, The Holocene 18 (2008) 907–21. Mazier F. et al. (2009) “Multidisciplinary approach to reconstructing local pastoral activities: an example from the Pyrenean Mountains (Pays Basque)”, The Holocene 19 (2009) 171–88. Mighall T. M. et al. (2006) “Proxy climate and vegetation changes during the last five millennia in NW Iberia: pollen and nonpollen palynomorph data from two ombrotrophic peat bogs in the north western Iberian Peninsula”, Review of Palaeobotany and Palynology 141 (2006) 203–23. Monna F. et al. (2004) “Environmental impact of early Basque mining and smelting recorded in a high ash minerogenic peat deposit”, Science of the Total Environment 327 (2004) 197–214. Morales-Molino C. et al. (2013) “Holocene vegetation, fire and climate interactions on the westernmost fringe of the Mediterranean basin”, Quaternary Science Reviews 59 (2013) 5–17. Morellón M. et al. (2012) “A multi-proxy perspective on millenniumlong climate variability in the southern Pyrenees”, Climate of the Past 8 (2012) 683–700. Morellón M. et al. (2011) “Climate changes and human activities recorded in the sediments of lake Estanya (NE Spain) during the Medieval Warm Period and Little Ice Age”, Journal of Paleolimnology 46 (2011) 423–52. Morellón M. et al. (2009) “Lateglacial and Holocene palaeohydrology in the western Mediterranean region: the lake Estanya record (NE Spain)”, Quaternary Science Reviews 28 (2009) 2582–99. Morellón M. et al. (2008) “Holocene palaeohydrology and climate variability in northeastern Spain: the sedimentary record of lake Estanya (Pre-Pyrenean Range)”, Quaternary International 181 (2008) 15–31. Moreno A. et al. (2012) “The Medieval Climate Anomaly in the Iberian Peninsula reconstructred from marine and lake records”, Quaternary Science Reviews 43 (2012) 16–32. Muller S. D. et al. (2015) “Vegetation history of the western Rif Mountains (NW Morocco): origin, Late-Holocene dynamics and human impact”, Vegetation History and Archaeobotany 24 (2015) 487–501. Muñoz-Sobrino C. et al. (2005) “Palynological data on major Holocene climatic events in NW Iberia”, Boreas 34 (2005) 381–400. Palet J. M. et al. (2014) “Landscape dynamics and territorial organization in the Empordà littoral plain (north-eastern Catalonia) from the Antiquity to the Medieval period: an integrated landscape analysis”, in Proceedings of the 18th International Congress of Classical Archaeology: Centre and Periphery in the Ancient World, edd. J. M. Álvarez, T. Nogales and I. Rodà (Mérida 2013) 311–16.
104 Parra I. et al. (2005) “Análisis palinológico y radiométrico del sondeo Sobrestany. Nueve milenios de historia natural e impactos humanos sobre la vegetación del Alt Empordà”, Empúries 54 (2005) 33–44. Pèlachs A. et al. (2009) “Landscape dynamics of Abies and Fagus in the southern Pyrenees during the last 2200 years as a result of anthropogenic impacts”, Review of Palaeobotany and Palynology 156 (2009) 337–49. Pérez Díaz S. and López-Sáez J. A. (2015) “Contributions to the European Pollen Database. Fuente del Vaquero peat bog (Basque Country, Northern Iberian Peninsula, Spain)”, Grana 54 (2025) 82–84. Pérez Díaz S. and López-Sáez J. A. (2014) “Contributions to the European Pollen Database. Prados de Randulanda peat bog (Basque Country, Northern Iberian Peninsula, Spain)”, Grana 53 (2014) 252–54. Pérez Díaz S. and López-Sáez J. A. (2012) “Paisajes medievales: paleoambiente y antropización en Treviño en los últimos 1800 años”, in Viaje a Íbita. Estudios históricos del Condado de Treviño, edd. R. González de Viñaspre and R. Garay (Treviño 2011) 377–90. Pérez-Díaz et al. (2016a) “Seis mil años de gestión y dinámica antrópica en el entorno del Parque Natural de los Collados del Asón (Cordillera Cantábrica Oriental)”, Cuaternario y Geomorfología 30 (2016) 49–74. Pérez-Díaz et al. (2016b) “8000 years of vegetation history in the northern Iberian Peninsula inferred from the palaeoenvironmental study of the Zalama ombrotrophic bog (Basque-Cantabrian Mountains, Spain)”, Boreas 45 (2016) 658–72. Pérez-Díaz et al. (2015) “Vegetation dynamics and human activity in the western Pyrenean region during the Holocene”, Quaternary International 364 (2015) 65–77. Pérez-Obiol R. et al. (2012) “Vegetation dynamics and anthropogenically forced changes in the Estanilles peat bog (southern Pyrenees) during the last seven millennia”, Vegetation History and Archaeobotany 21 (2012) 385–96. Pérez-Sanz A. et al. (2013) “Holocene climate variability, vegetation dynamics and fire regime in the central Pyrenees: the Basa de la Mora sequence (NE Spain)”, Quaternary Science Reviews 73 (2013) 149–69. Pérez-Sanz A. et al. (2011) “Clima y actividades humana en la dinámica de la vegetación durante los últimos 2000 años en el Pirineo central: el registro palinológico de la Basa de la Mora (Macizo de Cotiella)”, Zubía. Monográfico 23 (2011) 17–38. Ramos-Román M. J. et al. (2016) “Centennial-scale vegetation and North Atlantic Oscillation changes during the Late Holocene in the southern Iberia”, Quaternary Science Reviews 143 (2016) 84–95. Reimer P. J. et al. (2013) “Intcal13 and marine13 radiocarbon age calibration curves 0–50,000 years cal yr BP”, Radiocarbon 55 (2013) 1869–87. Riera S. et al. (2006) “Lake responses to historical land use changes in northern Spain: the contribution of non-pollen palynomorphs in a multiproxy study”, Review of Palaeobotany and Palynology 141 (2006) 127–37. Riera S. et al. (2004) “2000-year environmental history of a karstic lake in the Mediterranean Pre-Pyrenees: the Estanya lakes (Spain)”, Catena 55 (2004) 293–324.
López-Sáez et al. Rius D. et al. (2012) “Holocene history of fire, vegetation and land use from the central Pyrenees (France)”, Quaternary Research 77 (2012) 54–64. Riera S. et al. (2009) “Fire frequency and landscape management in the northwestern Pyrenean piedmont, France, since the Early Neolithic (8000 cal. BP)”, The Holocene 19 (2009) 847–59. Roberts N. et al. (2012) “Paleolimnological evidence for an east-west climate see-saw in the Mediterranean since AD 900”, Global and Planetary Change 84–85 (2012) 23–34. Robles-López S. et al. (2017a) “The dialectic between deciduous and coniferous forests in central Iberia: a palaeoenvironmental perspective during the Late Holocene in the Gredos Range”, Quaternary International (forthcoming). Robles-López S. et al. (2017b). “Contributions to the European Pollen Database. Labradillos mire, Gredos Range (central Spain)”, Grana (forthcoming). Ruiz Zapata M. B. et al. (2007) “Cambios en la vegetación durante el Holoceno reciente en el Valle del Lozoya (Sierra de Guadarrama, Madrid)”, Revista Española de Paleontología 22 (2007) 95–102. Rull V. and Vegas-Villarrúbia T. (2015) “Crops and weeds from the Estany de Montcortès catchement, central Pyrenees, during the last millennium: a comparison of palynological and historical records”, Vegetation History and Archaeobotany 24 (2015) 699–710. Rull V. and Vegas-Villarrúbia T. (2014) “Preliminary report on a mid19th century Cannabis pollen peak in NE Spain: historical context and potential chronological significance”, The Holocene 24 (2014) 1378–83. Rull V. et al. (2011) “Vegetation changes in the southern Pyrenean flank during the last millennium in relation to climate and human activities: the Montcortès lacustrine record”, Journal of Paleolimnology 46 (2011) 387–404. Serrano O. et al. (2011) “The Posidonia oceanica marine sedimentary record: a Holocene archive of heavy metal pollution”, Science of the Total Environment 409 (2011) 4831–40. Silva-Sánchez N. et al. (2016) “Influence of climate change and human activities in the organic and inorganic composition of peat during the Little Ice Age (El Payo mire, Gata Range, W Spain)”, The Holocene 26 (2016) 1290–1303. Silva-Sánchez N. et al. (2014) “Linking forest cover, soil erosion and mire hydrology to Late-Holocene human activity and climate in NW Spain”, The Holocene 24 (2014) 714–25. Törnqvist T. E. et al. (1989) “Degradación antropogénica de la vegetación en el noroeste de Galicia durante los últimos 2500 años”, Cuadernos de Estudios Gallegos 103 (1989) 175–98. Valero-Garcés B. L. et al. (2006) “Human impact since medieval times and recent ecological restoration in a Mediterranean lake: the Laguna Zoñar, southern Spain”, Journal of Paleolimnology 35 (2006) 441–65. Van der Knaap W. O. and van Leeuwen J. F. N. (1995) “Holocene vegetation and degradation as responses to climatic change and human activity in the Serra da Estrela, Portugal”, Review of Palaeobotany and Palynology 89 (1995) 153–211. Vannière B. et al. (2001) “Feu et pratiques agro-pastorales dans les Pyrénées Orientales: le cas de la montagne d’Enveitg (Cerdagne, Pyrénées-Orientales, France)”, Sud-Ouest Européen 11 (2001) 29–42.
Vegetation and Land-Use Change in Northern Europe during Late Antiquity: A Regional-Scale Pollen-Based Reconstruction Jessie Woodbridge, Neil Roberts and Ralph Fyfe Abstract This chapter presents an overview of land cover and land use change in northern Europe, particularly during Late Antiquity (ca. 3rd–8th c. AD) based on fossil pollen preserved in sediments. We have transformed fossil pollen datasets from 462 sites into eight major land-cover classes using the pseudobiomisation method (PBM). Through using pollen-vegetation evidence, we show that north-central Europe, lying outside the Roman frontier (the so-called ‘Barbaricum’ region), remained predominantly forested until Medieval times, with the main clearance phase only starting from ca. AD 750. This stands in contrast to north-west Europe, both inside (France/ England) and outside (Scotland/Ireland) the Roman imperial frontier; here a majority of forested land was already cleared prior to antiquity. The implications of this are that Roman expansion into the periphery of the empire largely took over existing intensive agrarian regions in the case of ‘Gaul’ (France) and ‘Britannia’ (England and Wales). Pre-existing land-use systems and levels of landscape openness may have played a role in directing the expansion of the Roman empire northwards into Gaul and Britannia, rather than eastwards into Germania. After the period of Roman occupation, partial reforestation is evident in some areas.
Introduction Regional landscape characteristics may have been important in influencing human land use patterns and the advancement of the Roman frontier across Europe. The aim of this study is to explore land use change on a regional scale, within and beyond the Roman frontier, in north-west and north-central Europe. For this, we use multiple pollen records in order to compare the timing of shifts in land use patterns in relation to historical and archaeological evidence of changing human activity during Late Antiquity (ca. 3rd to 8th c. AD), within the longer-term context of the Mid-Late Holocene (4000 BC until present). Previous pollen-based syntheses of land-cover change for central and northern Europe have highlighted the importance of human land
use in influencing patterns of Mid-Late Holocene vegetation change.1 The European continent has undergone significant landscape alterations throughout the Holocene as a result of climatic fluctuations and human land-use changes. Fossil pollen, preserved in lake and peat sediments, has provided a valuable tool in understanding past vegetation change, and various approaches have been developed to turn pollen data into records of quantified vegetation cover and land use change.2 These methods can allow us to consider the time at which human land use began to alter vegetation at both local and regional scales. Synthesis of pollen records from northern Europe has the potential to explore human land use and vegetation change. For example, initial forest loss at the regional and continental scale is recorded from ca. 4000 BC, as a result of Neolithic land use activities, such as tree clearance for agriculture.3 Comparison with radiocarbon-based demographic data at the national scale, has confirmed the relationship between population and land cover change from prehistoric Britain, and a relationship between secondary woodland regeneration and population in southern Germany.4 Approaches that quantify vegetation using pollen data, have indicated that there was an increase in open land between 4000 and 1000 BC in temperate Europe.5 Furthermore, an important intensification of land use, reflected in greater forest clearance and the establishment of pastoral and arable land use, has been identified beginning in the Late Iron Age (ca. 250 BC).6 Thus, by the start of the late antique period, the landscape of temperate Europe was
1 E.g. Behre (1988); Berglund et al. (1996); Berglund (2003). 2 Quantified vegetation cover: e.g. Marquer et al. (2014); Trondman et al. (2015). Land-use change: e.g. Fyfe et al. (2015). 3 Behre (1988); Woodbridge et al. (2014a); Fyfe et al. (2015); Gaillard et al. (2015). 4 Prehistoric Britain: Woodbridge et al. (2014a). Southern Germany: Lechterbeck et al. (2014). 5 Marquer et al. (2014) and Trondman et al. (2015). 6 Fyfe et al. (2015).
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_008 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 105–118
106
Woodbridge, Roberts and Fyfe
already notably transformed. Continent-scale reconstruction indicates that forest cover was relatively stable during Late Antiquity. Following the ‘Migration Period’ (from around AD 400), there was a downward trend in forest cover, as arable and pasture/grassland increased in extent from ca. AD 750,7 and it was during Medieval times that Europe’s pre-industrial landscape was established. These changes primarily reflect patterns of human land use change, both in mode of production and intensity of land use. The Romans conquered the non-Mediterranean area of ‘Gaul’ under Julius Caesar during the 1st c. BC. From there he went on to lead the first invasion of Britain around 55 BC,8 but Britain was not incorporated into the empire until ca. AD 43, with the full extent of the western Roman empire established by ca. AD 100. Randsborg has described the development of the villa system of large agricultural estates under the empire, which required abundant resources and workers.9 In the Roman period in north-west Europe, settlements were mainly concentrated on good grain-growing soils with intensive cultivation of cereals, along with long-distance trading.10 The centuries of the Roman empire’s rise coincided with general climatic favourability (the Roman optimum: 100 BC to AD 200) and warm mean summer temperatures.11 There has been much debate over the relative importance of climate changes and human land use pressure on the vegetation of Europe; for example, Huntley argued that climate was more important until around AD 1000,12 although this is hard to reconcile against the abundant evidence for intensive human land use of the continent from archaeological and historical sources. Equally, Berglund demonstrated that human impact events across north-west Europe were not concordant with periods of Holocene climate changes.13 In north-central Europe Germanic resistance and military conflicts—such as the battle of Teutoburg Forest in AD 9, in which three Roman legions were ambushed and destroyed—restricted Roman expansion to the natural frontier of the river Rhine. Similarly, the Romans made no concerted attempt to occupy Ireland, and although Scotland was invaded, the northern defensive frontier of the empire was eventually established at Hadrian’s Wall, in close proximity to the present-day border between England and Scotland. The western .
7 Fyfe et al. (2015). 8 Russel (2005). 9 Randsborg (1991). 10 Randsborg (1991). 11 Roman optimum: McCormick et al. (2012). Warm summer temperatures: Luterbacher et al. (2016). 12 Huntley (1990). 13 Berglund (2003).
Roman empire eventually declined during the 5th c. AD, with far reaching social impacts, and led to the collapse of trade networks, the decline of cities, and the advent of the so-called post-Classical ‘Dark Ages’ across Europe. The amalgamation of pollen datasets across broad subcontinental areas, and sub-regions within Northern Europe, allows differences in the timing of vegetation shifts to be identified in relation to these historicalcultural changes. Methods Pollen Analysis and Pseudobiomisation (PBM) Fossil pollen preserved in peat and lake sediments provides a proxy for vegetation and landscape change through time. Numerous radiocarbon-dated pollen records from the European Pollen Database (EPD)14 have been synthesised and transformed for a selection of regions (fig. 1) to reflect patterns in different regions encompassing areas that would have been within and outside the margins of the Roman empire. The Pseudobiomisation Method (PBM) has been used in order to identify regional-scale vegetation changes. This includes 190 sites in north-west Europe and 225 from north-central Europe, which are split into a number of sub-regions. ‘Germania’ covers the region of north-central Europe from the Rhine in the west to Vistula (present day Poland) in the east, the Danube in the south to the Baltic Sea in the north. ‘Gaul’ covers France, Belgium, small parts of south-west Germany and the southern Netherlands. Although this is an extensive dataset, there is spatial bias in the distribution of sites, with limited coverage of EPD sites in some regions, such as lowland England. The PBM provides a relatively simple and easily applied approach for turning complex pollen data into Land-Cover Classes (LCCs) related to human land-use change at 200-year time intervals.15 The key strengths of the PBM are that it can be easily applied to large datasets, is informative about broad-scale vegetation change through time, and has been developed to indicate anthropogenically-induced land-cover change by incorporating taxa that are indicators of land uses, such as agriculture and grazing. The approach is less complex than some other methods for the transformation of pollen to vegetation cover data, which rely on a detailed understanding of the differences in pollen productivity and dispersal between plant taxa. This is the case with the REVEALS (Regional Estimates
14 Leydet (2007–16). 15 For more detail of this see Fyfe et al. (2010) and Woodbridge et al. (2014a).
Vegetation and Land-Use Change in Northern Europe during Late Antiquity
figure 1
107
Northern Europe case study areas with fossil pollen site locations (numbers reflect sites presented in fig. 4). Symbols indicate sites grouped for analyses. The shaded area represents the Roman imperial frontier and the grey box highlights the area of the historic forest map: Schlüter (1952); Thomas (1970).
of VEgetation Abundance from Large Sites) approach,16 which incorporates differences in pollen production based on PPEs (Pollen Productivity Estimates), albeit for a more restricted number of taxa. The correlation and correct interpretation of pollen records is dependent on good chronological control from radiocarbon-dated core sequences. Chronologies constructed in recent years have been used to sum the pollen count data from each site into 200 year time windows for the records analysed in this study.17 The PBM approach is based on assigning each pollen taxon to one of a series of LCCs (table 1)—such as Quercus to deciduous woodland, or Secale cereale to arable landcover)—and identifying which LCC has the majority of pollen grains assigned to it. This becomes the ‘dominant’ land-cover type for that particular pollen sample level within a core sequence, which can include one of
eight classes (table 2). Additionally, broad-leaf forest was down-weighted (× 0.6) and arable/disturbed land was up-weighted (× 1.3) in the dataset. This partly compensates for the over—and under-production of pollen in these broad groups, and is based on the results of a comparison of modern pollen and remotely-sensed vegetation.18 Semi-open and mixed vegetation LCCs are defined using a threshold value between the sum of closed and open land-cover,19 which is a proxy for relative change in vegetation openness ranging between −100 and +100. This has been presented stratigraphically as symbols, indicating the assigned LCC for a selection of sites representing each case study region. The proportion of each LCC represented in each 200 year time window has been calculated to reconstruct spatially aggregated records of land-cover change for different regions of north-west and north-central Europe (fig. 1).
16 Sugita (2007); Gaillard et al. (2010). 17 Giesecke et al. (2014) and Leydet (2007–16).
18 Woodbridge et al. (2014b). 19 Woodbridge et al. (2014b).
108 table 1
Woodbridge, Roberts and Fyfe Land Cover Classes (LCCs) defined for the pseudobiomisation (PBM) method with assigned pollen taxa. Land Cover Class
Pollen taxa assigned for PBM
LCC1
Needle-leaf forest
Abies, Cupressaceae, Larix, Picea, Pinaceae, Taxus
LCC2
Broad-leaf forest
Acer, Aesculus, Alnus, Amygdalus, Anacardiaceae, Arceuthobium, Betula, Buxus, Carpinus, Castanea, Celtis, Cercis siliquastrum, Corylus, Crataegus, Daphne, deciduous Quercus, Euonymus, Fagus, Frangula, Fraxinus, Hedera, Ilex, Juglans, Liquidambar, Liriodendron, Lonicera, Malus, Morus, Myrica, Myrtaceae, Nyssa, Olea, Parrotia, Platanus, Populus, Prunus, Pulmonaria, Pyrus, Rhamnaceae, Robinia, Salix, Sambucus nigra, Sorbus, Staphylea, Tilia, Ulmaceae, Viscum, Vitaceae
LCC3
Mixed forest (dominated by a mix of needleleaf and broad-leaf forest)
Defined by thresholds between different forest types
Closed LCCs
Semi-open LCC Semi-open vegetation (mixed land cover including both forest and open vegetation)
Defined by thresholds between other classes
LCC5
Heath/scrubland
Aellenia, Alkanna, Anagallis, Arctostaphylos, Argania spinosa, Asphodelus, Astragalus, Ballota, Beta, Bongardia, Calluna, Carex, Casuarina, Centaurea sclerophyllus, Ceratonia, Chaenorhinum, Chamaerops, Chrozophora, Cistaceae, Citrus, Colchicum, Colutea, Convolvulaceae, Coriaria, Cynoglossum, Cyperaceae, Cytisus, Echium, Empetrum, Ephedra, Erica, Fagonia, Genista, Hippophae, Hypericum pulchrum, Hyphaene, Juniperus, Laurus, Lotus, Melampyrum pratense, Moltkia, Narthecium, Onosma, Parnassia, Phlomis, Pistacia, Polygonum heath types, evergreen Quercus, Rhynchospora, Rosmarinus, Rutaceae, Salvia, Sapotaceae, Sarcobatus, Scheuchzeria, Tamarix, Trachycarpus, Ulex, Valerianaceae heath types, Verbena
LCC6
Pasture/natural grassland
Actaea, Agrimonia, Anemone, Anethum, Anthyllis, Apiaceae, Asperula, Asteraceae (subfamily Lactucoides), Asteraceae undifferentiated, Bunium, Bupleurum, Calendula, Caltha, Campanulaceae, Centaurea grassland types, Coronilla, Dipsacaceae, Epilobium, Filipendula, Galega, Galium, Gentianaceae, Geum, Knautia, Lithospermum officinale, Ornithopus, Paliurus spina christi, Papaveraceae, Plantago alpina, Plantago coronopus, Plantago lanceolata, Plantago tenuiflora, Poaceae, Polygala, Potentilla, Ranunculaceae, Rhinanthus, Rumex acetosa, Rumex acetosella, Ruppia, Sanguisorba, Senecio, Sherardia arvensis, Sisyrinchium, Stachys, Succisa, Thalictrum, Trifolium, Urticaceae, Valerianaceae grassland types, Veronica, Viciaceae
LCC4
Open LCCs
Vegetation and Land-Use Change in Northern Europe during Late Antiquity table 1
109
Land Cover Classes (LCCs) defined for the pseudobiomisation (PBM) method with assigned pollen taxa (cont.). Land Cover Class
Pollen taxa assigned for PBM
LCC7
Arable/disturbed land
Allium, Anagallis arvensis, Anchusa, Aphanes, Arctium, Artemisia, Asparagaceae, Asteraceae (subfamily Asteroideae), Borago, Brassicaceae, Calligonum, Cannabaceae, Carduus, Caryophyllaceae, Cassia, Centaurea arable types, Centaurea undifferentiated, Cerealia, Chenopodiaceae, Cirsium, Cucurbitaceae, Fagopyrum, Galeopsis, Humulus, Hypericum calycinum, Jurinea, Lamiaceae, Lappula, Lathyrus, Leguminosae, Linaceae, Lithospermum arvense, Medicago, Melilotus, Noaea, Ononis, Pastinaca sativa, Phaseolus, Pisum, Plantaginaceae, Plantago albicans, Plantago atrata, Plantago cylindrica, Plantago major, Plantago maritima, Plantago media, Plantago montana, Plantago ovata, Plantago psyllium, Polygonaceae arable types, Prosopis, Punica, Rheum, Ribes, Ricinus, Rubus arable types, Rumex type, Scorzonera, Sideritis, Silenaceae, Solanaceae, Spinacia, Valerianaceae arable types, Zea
LCC8
Mixed open vegetation (dominated by a combination of heath/scrubland, pastures/natural grassland and arable land)
Defined by thresholds between open classes
table 2
Land Cover Classes (LCCs) defined for the pseudobiomisation (PBM) method. Land Cover Class
Description
Closed LCCs
LCC1
Needle-leaf forest
Defined by a small number of needle-leaf trees (typically high pollen producers)
LCC2
Broad-leaf forest
Defined by a large number of broad-leaf trees (e.g. deciduous Quercus), including a small number of epiphytes (e.g. Hedera), and some fruit trees (e.g. Olea)
LCC3
Mixed forest
Dominated by a mix of needle-leaf and broad-leaf forest Defined by thresholds between different forest types
Semi-open LCC LCC4
Semi-open vegetation
Mixed land cover including both forest and open vegetation Defined by thresholds between other classes
110
Woodbridge, Roberts and Fyfe
table 2
Land Cover Classes (LCCs) defined for the pseudobiomisation (PBM) method (cont.).
Open LCCs
LCC5
Heath/scrubland
Defined by a mixture of heath/scrub type taxa, including evergreen Quercus.
LCC6
Pasture/natural grassland
Predominantly defined by a large mixture of grassland herbs including a number of taxa associated with pasture
LCC7
Arable/disturbed land
Defined by a large number of herbaceous taxa indicative of arable and disturbed land (excluding fruit trees)
LCC8
Mixed open vegetation
Dominated by a combination of heath/scrubland, pastures/natural grassland and arable land Defined by thresholds between open classes
The results are presented as the percentage of samples assigned to each LCC through time, both for multiple records within each region, in order to create a regional average, and also for individual pollen records, in order to show more local trends. Historic Geographic Forest Maps/Domesday Records Historic records of forest cover have been compared against the pollen-inferred land-cover change reconstructions, to test their ability to accurately reconstruct vegetation. The percentage of forest, for AD 700–900 and AD 1900, covering central-northern Europe (highlighted by the grey box in fig. 1) was calculated from a digitised historic geographic forest map.20 Values for woodland percentage were also obtained from Rackham, who provides details of land cover in AD 1086 and 1895, and is partly based on the historic Domesday record of English counties.21 The use of these historic sources is limited by the data collection approaches used (e.g. place name evidence), and consequently the area mapped as forest may not represent true forest cover, particularly for the early Medieval period. Results Relative change in the abundance of different landcover types throughout Late Antiquity is illustrated in figures 2 and 3, which is placed within the context of the Mid-Late Holocene. Regional variation in the timing and amplitude of forest loss and increasing open land from around 4000 BC is demonstrated in figure 2,
20 Schlüter (1952); Darby and Terrett (1954); Thomas (1970). 21 Rackham (2003).
which summarises the percentage of pollen samples assigned to closed, semi-open, and open vegetation LCCs throughout the Mid-Late Holocene. The percentage of sites assigned to forest classes in north-west Europe was as low as 20% by ca. 450 BC. However, sites in north-central Europe remained forested for longer, with high forest values (above 40%) persisting until around AD 950. Separate analyses for sites in England and Wales, and Scotland and Ireland, indicate that these regions became more open earlier in the records in comparison with sites from France and Belgium. Separate analyses for parts of the central European region show a similar trend to the pattern shown for the amalgamated central European dataset, although the records from southern Sweden and Denmark show an initial decline in forest cover earlier (around AD 300) than the other sub-regions. Individual open LCCs are shown in figure 3, which illustrates how different open vegetation landcover types dominated in different areas throughout Late Antiquity. For example, pasture/natural grassland was the principal LCC in England and Wales, with values reaching almost 80% by ca. AD 1350, whereas heath/ scrubland was more abundant in records from Scotland and Ireland. Arable land was relatively more significant in continental Europe, especially inside the Roman frontier (i.e. in ‘Gaul’) (fig. 3). Over the Mid-Late Holocene timescale (fig. 2) the different regions analysed vary in the timing of changes in the main land-cover types. When analysed separately, records indicate that England was significantly more open earlier in the record in comparison with other regions, whereas the dominant vegetation type in Scotland and Ireland is semi-open land-cover. When the central European dataset is split into smaller sub-regions, all areas show similar patterns, with longer persistence of forest classes until around AD 500,
Vegetation and Land-Use Change in Northern Europe during Late Antiquity
111
figure 2 Total forest, semi-open and open vegetation percentage for sites grouped for north-west Europe: England + Wales, Scotland + Ireland and France + Belgium, and north-central Europe: Germany + west Czech Republic, Poland + east Czech Republic and southern Sweden + Denmark. Values based on the percentage of samples assigned to each land-cover class (LCC) using the pseudobiomisation (PBM) approach: 7000 cal. CAL. BP (5050 BC) to modern. The grey box highlights the time period covered by Late Antiquity.
although the semi-open class increases earlier in the records from southern Sweden. A greater percentage of samples have been assigned to mixed open vegetation for the combined records from north-west and northcentral Europe, but values never reach above 20% for any region for this mixed LCC. Therefore, pollen samples dominated by open vegetation indicators appear to typically fall into distinct classes, such as heathland or pasture/grassland in these regions. Figure 4 shows a representative selection of individual pollen records from different regions of northern Europe, in order to highlight variations in vegetation
history across the wider landscape. The pollen sites show relative change in landscape openness and the assigned LCC for each sample, and indicate similar trends to those recognised at the amalgamated regional level (site location numbers are provided on figure 1). The graphs provide an indication of vegetation openness, but do not reflect fully quantified vegetation. The LCC symbols plotted to the right in each graph, with a score between 0 and 100, indicate more closed landscapes; symbols plotted within the ± 20 range of 0 represent semi-open vegetation, and symbols plotted to the left of each plot (−100 to 0) represent open vegetation types.
112
Woodbridge, Roberts and Fyfe
figure 3 Percentage of samples assigned to each land-cover class (LCC) using the pseudobiomisation (PBM) approach for LCC5 (heath/ scrubland), LCC6 (pasture/natural grassland), LCC7 (arable/disturbed land) and LCC8 (mixed open vegetation) based on sites grouped for north-west Europe: England + Wales, Scotland + Ireland and France + Belgium, and north-central Europe: Germany + west Czech Republic, Poland + east Czech Republic and southern Sweden + Denmark: 7000 cal. CAL. BP (5050 BC) to modern. The grey box highlights the time period covered by Late Antiquity.
Figure 4 shows how north-west Europe was an open landscape earlier in the Holocene, whereas many areas of north-central Europe (particularly Germany, Poland and the Czech Republic) remained forested for longer, but there is variation in the types of open/closed LCCs assigned, and the timing of changes varies between sites. Some records from Germany (Steerenmoos and Grosser Treppelsee) were dominated by mixed forest, whereas other records from the same region were assigned to broad-leaf forest (e.g. Lüttersee). Holzmaar, located inside the Roman frontier, was dominated by
semi-open vegetation prior to antiquity according to the PBM results. Within these records, the main shift from closed to semi-open LCCs in Germany and Poland seems to have occurred around AD 1000, whereas semi-open LCCs define the record from Denmark (Lake Solsø) from 2500 BC. Differences in the timing of shifts and the types of forest LCCs are also shown in the two records from Sweden in figure 4: Ageröds Mosse is defined as deciduous forest, whereas Holtjärnen is mixed forest, and both remain semi-open, with no open LCCs assigned later in their
figure 4 PBM score: closed forest sum minus open vegetation sum (proxy for vegetation openness) for pollen samples summed into 200 year time windows for a selection of sites from a) north-central Europe (all sites are located outside of the Roman frontier, and some are positioned close to the border (Holzmaar and Steerenmoos); and b) north-west Europe, which includes sites both within (France, England and Wales) and outside (Scotland and Ireland) the Roman frontier. The symbols represent the assigned Land Cover Class (LCC). Symbols plotted to the left indicate open land-cover types, and to the right indicate closed land-cover types. The grey boxes highlight the period 7000 cal. CAL. BP (5050 BC) to 600 cal. CAL. BP (AD 1350).
Vegetation and Land-Use Change in Northern Europe during Late Antiquity
113
114
Woodbridge, Roberts and Fyfe
figure 5 a) Woodland percentage by English county, according to Rackham (2003), with pollen-inferred closed sum based on the pseudobiomisation (PBM) approach and historic woodland percentage for Devon and averaged for England, from Rackham (2003); b) comparison of a historic forest map (Schlüter (1952); Thomas (1970)) for north central Europe, and the pollen-inferred PBM forest sum for the same region.
records. Similarly, the records from north-west Europe show variation in the dominant LCCs and the timing of changes in their records. For example, Dallican Water from Scotland is assigned to heath/scrubland from 4000 BC, whereas these open LCCs do not appear until later in the other records from Scotland and Ireland (ca. AD 600). However, as Dallican Water is located on an island off the north coast of Scotland, the different environmental characteristics and degree of exposure to prevailing winds may have influenced this record, which has been recognised as relating to decreased tree abundance in pollen records from Scottish islands.22 The forest LCC types assigned to sites in north-west Europe (shown in figure 4) are all broad-leaf, however, the open LCCs vary regionally. Sites in England and Wales are defined by pasture/natural grassland, whereas sites in France and Switzerland include more semi-open land, arable/disturbed land and mixed open vegetation. The woodland percentages obtained from Rackham (fig. 5) indicate that tree cover in England decreased
somewhat between AD 1086 and 1895.23 However, the pollen-inferred PBM closed sum, which provides an indication of forest cover, shows higher values than those provided by Rackham (fig. 5a). Differences in the exact time periods covered by these datasets, and the limitations of inferring past forest cover from historic sources, may account for the differences in these values. The comparison between an historic forest map for north-central Europe (mainly covering Germany) (fig. 1) and the PBM closed sum (fig. 5b) indicates similar values around AD 700–900 and AD 1850–1900, showing a decrease in forest cover between these dates, but also shows that values remained high (above 30%) in the more recent time period. The forest map for AD 1850–1900 is likely to be a more reliable historic source, whereas the map for AD 700–900 may show significant over/under estimates of forest cover. However, when considering relative changes, both the pollen data and the historic maps show a decline in forest cover between these dates.
22 Lowe and Walker (1986).
23 Rackham (2003).
Vegetation and Land-Use Change in Northern Europe during Late Antiquity
Discussion Regional and Local Vegetation Changes Inside and Outside the Western Roman Empire The sub-continental and regional-scale changes evident in the PBM-transformed pollen records for north-central and north-west Europe, highlight the extent of spatial and temporal variability that existed in land use and forest loss patterns throughout the Mid-Late Holocene. The greater percentage of samples assigned to forest LCCs in north-central Europe (Germany, Poland, Czech Republic, southern Sweden and Denmark), indicates that a more densely forested landscape persisted for longer in all areas (excluding parts of southern Sweden and Denmark) in comparison with north-west Europe (England, Wales, Scotland, Ireland, France and Belgium), where open vegetation types are abundant much earlier in the records. Thus, prior to antiquity, it appears that there was already greater intensity of human land use for agriculture in these north-western areas. These patterns are similarly reflected in other broadscale vegetation reconstructions for northern Europe.24 The patterns of landscape change identified in the PBM results (figs. 2, 3 and 4) are also reflected in numerous individual pollen sequences from temperate Europe. For example, in a record from northern England (Gormire Lake), a pattern indicating forest loss was identified from 200 BC that was contemporaneous with increased grass pollen and various other agricultural indicators, which was followed by forest recovery from ca. AD 500.25 Similarly, another record from northern England showed three Late Neolithic—Bronze Age woodland clearance phases from ca. 2290 BC, with Late Iron Age clearance and agricultural intensification from ca. 300 BC, and a peak period of clearance ca. 90 BC until AD 50.26 This highlights that forest removal was extensive in England prior to Late Antiquity, and the patterns described are consistent with the amalgamated PBM results for England.27 In a study of long-term human impacts on landscape and vegetal biodiversity in the Auvergne region of France, using pollen data, a decline in forest cover from around the 2nd c. AD was identified, along with a progressive period of widespread woodland exploitation for agro-pastoral purposes.28 Woodland clearance was identified as already evident prior to this period, with pastoral and arable agriculture apparent at the 24 E.g. Trondman et al. (2015); Fyfe et al. (2015). 25 Oldfield et al. (2003). 26 Yeloff et al. (2007). 27 See Rippon and Fyfe (this volume) for a discussion of pollen records from the British Isles covering Late Antiquity. 28 Miras et al. (2015).
115
local and regional scales. These kinds of patterns reflect the overall regional trends identified in the PBM results for north-west Europe. Individual pollen records from north-central Europe indicate similar patterns of land use change to those in the records from northwest Europe, but with differing impacts on forest cover. For example, in a pollen record from the Eifel region of Germany close to the Roman frontier (Holzmaar) (fig. 4: site 9), a high percentage of trees and shrubs was identified, but with variations in the main tree types throughout Late Antiquity.29 However, this record also shows persistent grasses and agricultural pollen types throughout this period, followed by forest recovery from ca. AD 200. The PBM-assigned LCCs for this record (fig. 4) indicate that the forest type remained broadleaved, with a shift to semi-open vegetation through Late Antiquity, and a subsequent return to forest LCCs around AD 500. In a pollen record from Poland, continuous human land use was identified between the 7th c. BC and the 10th c. AD, linked to agriculture.30 Therefore, human land use appears to have been significant throughout Late Antiquity in some parts of north-central Europe (Germany and Poland in particular) and the records mentioned above provide evidence for considerable agriculture in north-central Europe throughout this time period. This demonstrates the added value of a synthetic approach, using multiple pollen records covering large areas, and the sub-regions within these areas, which allows interpretations between regions. The inter-regional comparisons provided by the PBM analyses can help us to explore where and when some areas experienced greater forest loss than others. Unlike the individual pollen records discussed above, the amalgamated PBM-based vegetation reconstruction shows that while there is evidence of continuous agriculture in north-central Europe, its impact on forest cover was not as significant as in parts of north-west Europe. Using the REVEALS approach, it was found that the degree of anthropogenic transformation of forests to cultivated and grazing land in northern Europe was significantly higher than inferences based on pollen percentages alone.31 However, the REVEALS results still show high values of forest cover for the AD 1250–1600 (350–700 CAL. BP) time period in northern Germany, with values above 40% for summer-green trees.32 The division of the PBM pollen dataset into regions inside and outside the Roman frontier indicates that
29 Litt et al. (2009). 30 Szal et al. (2014). 31 Trondman et al. (2015). 32 Trondman et al., (2015).
116 the vegetation of Scotland and Ireland remained partly open for longer, and that open land was the most important land-cover type in England and Wales throughout Late Antiquity. Both regions had significantly more open land in comparison with north-central Europe at this time (fig. 2). The main difference between these two regions was the type of open land-cover: namely, heath/scrubland in Scotland and Ireland, whereas pasture/grassland was the main LCC in the available EPD records from England and Wales. Heath/scrubland is evidently of lower economic value in comparison with pastureland. In an analysis of 73 pollen records from the British Isles using the REVEALS approach, it has been shown that Britain had a greater degree of landscape openness at the regional scale than areas across mainland Europe.33 This study also highlighted the considerable spatial bias in the dataset towards wetland areas (such as fens) and uplands in the British Isles, and this may partly explain the higher estimates of landscape openness for the region. The patterns in the PBM results for different sub-regions of northern Europe may be explained by differences in certain landscape characteristics. For example, flat expanses on clay-rich soils have been described as areas with the most intensive agricultural impact, with hilly areas remaining rich in forest, and poor sandy soils becoming dominated by heathland, for example, in Denmark and southern Sweden.34 Semi-open vegetation is the most abundant class in the PBM results for southern Sweden and Denmark (fig. 2), and the patterns identified in the PBM results for southern Sweden are different to other parts of the central European study area. The quantified land-cover reconstruction based on the REVEALS approach suggests that by the Late Bronze Age/Early Iron Age, significant areas of southern Sweden were under human use for crop cultivation and pastures,35 which includes areas that the PBM approach has defined as semi-open vegetation. In some western parts of Europe the most pronounced human impacts on vegetation took place during the Roman period; however, in Ireland renewed agricultural activity occurred later, after around AD 300.36 In continental Europe, regional forest regeneration has been identified during the ‘Migration Period’ (AD 400–550) in marginal areas of Denmark, southern Sweden, the Baltic Sea coast, north-east Germany, and northern Poland.37 When multiple fossil pollen sequences are analysed over broad regions, the results support the hypothesis 33 Fyfe et al. (2013). 34 Berglund et al. (2003); Gaillard et al. (2015). 35 Gaillard et al. (2015). 36 Behre (1988). 37 Gaillard et al. (2015).
Woodbridge, Roberts and Fyfe
that landscapes differed within and outside the Roman frontier, with greater open land inside the frontier prior to Roman advancement into this region, and continued persistence of forested land for longer outside of the frontier in continental Europe, particularly in what is now Germany, Poland and the Czech Republic. Figure 4, which is based on a selection of individual pollen sequences, highlights that within each region there are differences between individual sites. Three out of the six selected sites in Germany show evidence of forest loss prior to ca. 8th c. BC, while three others show this only occurring in the Medieval period (from around the 5th c. AD). However, in comparison with England, none of the six German sites show predominantly open landscapes in the Roman period; even those that had experienced some clearance were still partly forested. Temporal and spatial changes in forest cover will have also been influenced by climatic variability, but this is likely to have been more significant over longer multimillennial timescales than during the 1st millennium AD. Associations have been made between climatic conditions and the advance and retreat of the Roman empire.38 Palaeoclimate records also reveal regional variability in climate trends through Late Antiquity across Europe.39 Vegetation Change and the Expansion/Retreat of the Cultural Landscape The last 2500 years (from around 550 BC) is described as a period characterised by the widespread transformation of Europe’s land cover associated with human land use.40 This is reflected in shifts in pollen records from closed forest to open land-cover,41 and also coincides with increased human population levels.42 Considerable forest loss had therefore already occurred in north-west Europe before the arrival of the Romans, as a result of land-use and land-cover change during later prehistory, which accelerated during antiquity.43 There are numerous driving forces that would have influenced the pattern of Roman expansion and the regions conquered and incorporated into the empire. For example, as demonstrated in the pollen record, lowland Britain was already a largely cleared agricultural landscape in 55 BC. ‘Britannia’ may therefore have appeared economically valuable, whereas much of ‘Germania’ was still densely forested at this time.
38 E.g. McCormick et al. (2012). 39 E.g. Luterbacher et al. (2016); Labuhn et al. (this volume). 40 E.g. Fyfe et al. (2015). 41 E.g. Marquer et al. (2014). 42 E.g. Collard et al. (2010). 43 E.g. Woodbridge et al. (2014a).
Vegetation and Land-Use Change in Northern Europe during Late Antiquity
The causality behind expansions or regressions of the cultural landscape, and land-use changes, is complex.44 However, one possibility is that Britain was invaded and occupied by the Romans due to the economic gains that were offered by an already cleared agricultural landscape, with many existing rural settlements being incorporated into the Roman villa system.45 The Roman army also experienced setbacks in Germania, such as the Battle of the Teutoburg Forest in AD 9, one of the Roman army’s greatest military defeats, and which discouraged further expansion into this region. The dense forests that still existed east of the Rhine were likely to have been an impediment to the Roman armies and an asset to Germanic tribes; the latter may have had a military advantage thanks to a lack of battles in open terrain. The pollen evidence therefore implies that differences in regional landscapes, and particularly in the economic value of agricultural and forested land, may have influenced the spatial pattern of Roman expansion. At the northern European scale, it has been shown that there was a reduction in the pace of forest clearance between ca. 50 BC and ca. AD 600,’.46 This was followed by renewed forest clearance during the Medieval period, particularly in central Europe, which had remained largely forested throughout Late Antiquity. The PBM results show evidence of reversals in the total forest cover for north-central and north-west Europe, but the changes in the total forest cover were small during and after Late Antiquity, according to the records analysed at an amalgamated regional scale, even if individual sites show periods of reforestation.47 When the Rhine frontier collapsed in AD 406 the Romans began to withdraw from Britain and never returned.48 The post-Roman decline of agriculture and the gradual appearance of new land-cover types leading to reforestation, would have been spatially and temporally variable. During the period covering the fall of the Roman empire, woodland levels stabilised in the pollen records for north-west and north-central Europe, until a further more dramatic decline in forest cover over the last 1100 years (since ca. AD 950). A comparison with historic records using the forest map for central Europe and Rackham’s land cover records for Domesday England, implies that there was still a tangible difference in forest cover between England and Germany in Medieval times.49
44 Berglund (2003). 45 There was a large degree of regional variation in RomanoBritish agriculture, with significant diversity of Roman rural settlement across the landscape: Smith et al. (2016). 46 Fyfe et al. (2015). 47 E.g. Oldfield et al. (2003). 48 Nicholas (2014). 49 Schlüter (1952); Thomas (1970); Rackham (2003).
117
Conclusion The contrasting vegetation reconstructions for northwest and north-central Europe indicate significant regional variation in the timing and extent of forest loss between these regions. Many forests in England and Wales (‘Britannia’) and France (‘Gaul’) were converted to arable or pasture land by existing Celtic populations in the Iron/Bronze Ages. By the 1st c. BC, both of these regions had dense rural populations and productive agrarian economies that would have made them appealing targets for Roman imperial expansion. These areas also may have been easier to conquer and to control militarily by Roman invaders than the still-densely forested lands east of the Rhine. The PBM-inferred land-cover reconstruction indicates that much of north-central Europe remained densely forested until ca. AD 950, whereas north-west Europe, particularly Britain, was already an open landscape prior to the Roman arrival. During the centuries after the end of Roman imperial rule, the decline in forest cover was halted or reversed in many but not all regions; in most of northern Europe there is evidence of forest recovery in the period before the main onset of Medieval economic and demographic growth. Forests, or the lack of them, may therefore have been both a cause and a consequence of the nature of Roman imperial expansion, and its post-antique retreat and the societal reorganisation that followed. Acknowledgements This research was funded by Leverhulme Trust grant F00568W. The authors would like to thank a number of individuals who have supported or contributed towards this project, including contributors to the European Pollen Database (EPD) for sharing their data, Shaun Lewin for assistance with data manipulation, and Basil Davis for useful discussions. We would also like to thank Thomas Giesecke and others involved in the formulation of new chronologies for the EPD pollen records, and Michelle Leydet (the EPD manager). Bibliography Behre K. E. (1988) “The role of man in European vegetation history”, in Vegetation History, edd. B. Huntley and T. Webb III (DordrechtBoston-London 1988) 633–72. Berglund B. E. (2003) “Human impact and climate changes— synchronous events and a causal link?”, Quaternary International 105 (2003) 7–12. Berglund B. E., Birks H. J. B., Ralska-Jasiewiczowa M. and Wright H. E. (1996) Palaeoecological Events During the Last 15000 Years:
118 Regional Synthesis of Palaeoecological Studies of Lakes and Mires in Europe (Chichester 1996). Collard M., Edinborough K., Shennan S. and Thomas M. G. (2010) “Radiocarbon evidence indicates that migrants introduced farming to Britain”, JAS 37 (2010) 866–70. Darby H. C. and Terrett I. B. (1954) edd. The Domesday Geography of Midland England (New York 1954). Fyfe R. M., Woodbridge J. and Roberts N. (2015) “From forest to farmland: pollen inferred land-cover change across Europe using the pseudobiomization approach”, Global Change Biology 21 (2015) 1197–212. Fyfe R. M., Twiddle C., Sugita S. et al. (2013) “The Holocene vegetation cover of Britain and Ireland: overcoming problems of scale and discerning patterns of openness”, Quaternary Science Reviews 73 (2013) 132–48. Fyfe R. M., Roberts N. and Woodbridge J. (2010) “A pollen-based pseudobiomisation approach to anthropogenic land-cover change”, The Holocene 20 (2010) 1165–71. Gaillard M. J., Kleinen T., Samuelsson P. et al. (2015) “Causes of regional change—land cover”, in Second Assessment of Climate Change for the Baltic Sea Basin (BACC II), edd. H. J. Bolle, M. Menenti, S. Sebastiano al Vesuvio and S. Ichtiaque Rasool (ChamHeidelberg-New York-Dordrecht-London 2015). Gaillard M. J., Sugita S., Mazier F. et al. (2010) “Holocene land-cover reconstructions for studies on land-cover-climate feedbacks”, Climate of the Past 6 (2010) 483–99. Giesecke T., Davis B., Brewer B. et al. (2014) “Towards mapping the Late Quaternary vegetation change of Europe”, Vegetation History and Archaeobotany 23 (2014) 75–86. Huntley B. (1990) “European vegetation history: palaeovegetation maps from pollen data—13 000 yr cal. CAL. BP to present”, Journal of Quaternary Science 5 (1990) 103–22. Leydet M. (2007–16) The European Pollen Database: http://www .europeanpollendatabase.net/ (accessed Sept 2013). Lechterbeck J., Edinborough K., Kerig T., Fyfe R., Roberts N. and Shennan S. (2014) “Is Neolithic land-use correlated with demography? An evaluation of pollen derived land-cover and radiocarbon inferred demographic change from central Europe”, Holocene 24 (2014) 1297–1307. Litt T., Schölzel C., Kühl N. and Brauer A. (2009) “Vegetation and climate history in the Westeifel Volcanic Field (Germany) during the past 11,000 years based on annually laminated lacustrine maar sediments”, Boreas 38 (2009) 679–90. Lowe J. J. and Walker M. J. C. (1986) “Flandrian environmental history of the Isle of Mull, Scotland II: pollen analytical data from sites in western and northern Mull”, New Phytologist 103 (1986) 417–36. Luterbacher J., Werner J. P., Smerdon J. E. et al. (2016) “European summer temperatures since Roman times”, Environmental Research Letters 11 (2016): 024001 doi:10.1088/1748-9326/11/ 2/024001 (accessed May 2016). McCormick M., Büntgen U., Cane M. A. et al. (2012) “Climate change during and after the Roman empire: reconstructing the past from scientific and historical evidence”, Journal of Interdisciplinary History 43 (2012) 169–220.
Woodbridge, Roberts and Fyfe Marquer L., Gaillard M. J. and Sugita S. et al. (2014) “Holocene changes in vegetation composition in northern Europe: why pollenbased quantitative reconstructions matter?”, Quaternary Science Reviews 90 (2014) 199–216. Miras Y., Beauger A., Lavrieux M. et al. (2015) “Tracking long-term human impacts on landscape, vegetal biodiversity and water quality in the Lake Aydat catchment (Auvergne, France) using pollen, non-pollen palynomorphs and diatom assemblages”, Palaeogeography, Palaeoclimatology, Palaeoecology 424 (2015) 76–90. Nicholas D. M. (2014) The Growth of the Medieval City: From Late Antiquity to the Early Fourteenth Century (Oxford-New York 2014). Oldfield F., Wake R., Boyle J. et al. (2003) “The Late-Holocene history of Gormire Lake (NE England) and its catchment: a multiproxy reconstruction of past human impact”, Holocene 13 (2003) 677–90. Rackham O. (2003) Ancient Woodland. Its History, Vegetation and Uses in England (Colvend, 2nd edn. 2003). Randsborg K. (1991) The First Millennium AD in Europe and the Mediterranean: An Archaeological Essay (New York 1991). Russel M. (2005) “Ruling Britannia”, History Today 55 (2005) 5–6. Schlüter O. (1952) Die Siedlungsräume Mitteleuropas in frühgeschichtlicher Zeit, Part 1 (Forschungen zur Deutschen Landeskunde 63) (Hamburg 1952). Smith A., Allen M., Brindle T. and Fulford M. (2016) The Rural Settlement of Roman Britain. (Roman Society Publications) (Reading 2016). Sugita S. (2007) “Theory of quantitative reconstruction of vegetation I. Pollen from large sites REVEALS regional vegetation”, Holocene 17 (2007) 229–41. Szal M., Kupryjanowicz M., Wyczółkowski M. and Tylmann W. (2014) “The Iron Age in the Mrągowo Lake District, Masuria, NE Poland: the Salęt settlement microregion as an example of longlasting human impact on vegetation”, Vegetation History and Archaeobotany 23 (2014) 419–37. Thomas W. L. (1970) Man’s Role in Changing the Face of the Earth (Chicago-London 1970). Trondman A. K., Gaillard M. J. Sugita S. et al. (2015) “Pollen-based land-cover reconstructions for the study of past vegetation-climate interactions in NW Europe at 0.2 k, 0.5 k, 3 k and 6 k years before present”, Global Change Biology 21 (2015) 676–97. Woodbridge J., Fyfe R. M., Roberts N., Downey S., Edinborough K. and Shennan S. (2014a) “The impact of the Neolithic agricultural transition in Britain: a comparison of pollen-based land-cover and archaeological 14C date-inferred population change”, JAS 15 (2014) 216–24. Woodbridge J., Fyfe R. M. and Roberts C. N. (2014b) “A comparison of remotely sensed and pollen-based approaches to mapping Europe’s land-cover”, Journal of Biogeography 41 (2014) 2080–92. Yeloff D., van Geel B., Broekens P., Bakker J., and Mauquoy D. (2007) “Mid- to Late-Holocene vegetation and land-use history in the Hadrian’s Wall region of northern England: the record from Butterburn Flow”, Holocene 17 (2007) 527–38.
Case Studies – West
Hadrian’s Wall in Context: A Multi-Proxy Palaeoenvironmental Perspective from Lakes Petra Dark Abstract The Hadrian’s Wall area has more pollen sequences spanning Late Antiquity than any other part of the British Isles, but most are from peat bogs, posing problems of distinguishing between changes in the local wetland vegetation and events in the wider landscape. Here, an alternative perspective is offered by multi-proxy analyses of sediments from two lakes— Crag Lough and Grindon Lough—adjacent to the central sector of Hadrian’s Wall and the Stanegate, respectively. These demonstrate that at least the central sector of the Hadrianic frontier was constructed in a landscape already shaped by two millennia of woodland clearance, burning, farming and soil erosion. Roman military presence led to changes in agricultural and settlement patterns, but the overall impact, from an environmental perspective, was minor compared to that of prehistoric peoples. Roman withdrawal led to a relaxation in land use intensity, resulting in woodland regeneration on areas least favourable to agriculture, probably encouraged by climatic deterioration. The landscape, overall, remained predominantly open and agricultural, however, resembling that of the Late Iron Age. A multi-proxy multi-site approach offers the greatest prospect of understanding environmental and landscape changes connected with Roman military presence and withdrawal, and the varied spatial and temporal scales on which they occurred.
Introduction Hadrian’s Wall spans the north of England from the estuary of the River Tyne to the Solway Firth (fig. 1),1 a landscape rich in lakes and mires that have allowed the production of the highest concentration of pollen sequences spanning Late Antiquity from any part of the British Isles.2 These sequences have been used to explore questions concerning the nature of local vegetation and land use at the time of the construction of Hadrian’s Wall, which began in the AD 120s. These sequences have also been valuable in ascertaining what,
1 Breeze and Dobson (2000). 2 See overviews in Dark (1996), (1999), (2000); and Dark and Dark (1996), (1997).
if any, impact the withdrawal of the Roman military in the 5th c. had on the environment, especially in terms of the extent of woodland. Key difficulties in addressing these issues arise from: the relatively poor chronological control of the pollen sequences, in comparison with the local Roman-period archaeological record; uncertainties over the sources of pollen reaching peat bogs (especially from in situ wetland vegetation); and problems of disentangling the effects of human vs ‘natural’ (e.g., climate, vegetational succession) factors in the pollen record. Some of these difficulties can be addressed by examination of sediment sequences from lakes, which lack extensive growth of in situ vegetation, and which act as ‘sinks’ for eroded soil from their catchments, analysis of which can provide important information on land use changes complementary to the pollen record. Combined with analysis of microscopic charcoal particles, the prospect of disentangling the effects of human and natural agencies on vegetation and land use change is considerably enhanced, compared with the use of pollen data alone. This being the case, it is perhaps surprising that these complementary analyses have barely been used in studies of the environmental history of the Hadrian’s Wall area. This paper presents multi-proxy palaeoenvironmental data from two lakes close to the central sector of Hadrian’s Wall: Crag Lough and Grindon Lough. It will use these to shed new light on questions of the environmental context, and impact, of the establishment of the Hadrianic frontier, and the nature of vegetational and land use changes at the end of the Roman period. The results are discussed in the light of the peat-bog and onsite palaeoenvironmental data, highlighting the varying, and complementary, perspectives provided by different types of sampling site. Background to the Research on the Environment of Hadrian’s Wall The building of Hadrian’s Wall in the AD 120s to 130s was preceded by the construction of the Stanegate road,3 3 By how long is a matter of debate: Hodgson (2009).
© koninklijke brill nv, leiden, 2019 | doi:10.1163/9789004392083_009 Adam Izdebski and Michael Mulryan (eds) Environment and Society in the Long Late Antiquity (Late Antique Archaeology 11–12) (Leiden 2018), pp. 121–134
122
Dark
0
20 km
Steng Moss
Fellend Moss
Vin do lan da
sw ald
e egat Stan
Bir do
Glasson Moss
Walton Moss
Ca rlis le
ay Firth Solw
ut
Hadrian
's Wall
Grindon Lough
Quick Moss
figure 1
So
Fozy Moss
Co rbr idg e
Butterburn Flow H ou se ste Midgeholme ad Bolton Fell Moss Moss Crag Lough s
Thornbrough
h
Sh i
eld
s
yne R. T
Hallowell Moss
The Hadrian’s Wall area, showing sites mentioned in the text (triangles = pollen sequences, squares = forts, circle = farmstead).
which runs slightly further south of the Wall on a line from Corbridge to Carlisle (fig. 1). Much research and discussion has focused on the timing and extent of the clearance of the original prehistoric woodland covering the area,4 exploring the degree of landscape modification required for the establishment of the Hadrianic frontier, in terms of provision of timber, agricultural resources, and so on. In most of the relevant pollen sequences5 significant clearance is not registered until the Late Iron Age or later. This is perhaps linked to the establishment of the frontier, to make space for the construction of the Wall, roads, forts, and so on, and also perhaps the result of the military demand for timber, fuel, and agricultural produce. A problem with attempting to establish links between the relatively well-dated sequence of events surrounding the establishment of the Hadrianic frontier and changes in the local environment, lies in the reliance of the environmental chronology on radiocarbon dating. Even where a pollen sequence has a radiocarbon date directly associated with a clearance phase, which might be linked to the construction of Hadrian’s Wall, the inherent imprecision of radiocarbon dating—which typically provides an age range spanning around 200
4 E.g. Turner (1979); Dumayne (1993), (1994); Dumayne and Barber (1994); McCarthy (1995); Hanson (1996), (1997); Tipping (1997); Dumayne-Peaty (1998); Huntley (1999); Dark (2000) 58–63, 100– 108; Dark (2005). 5 Including some very close to Hadrian’s Wall, such as Fozy Moss: Dumayne (1993), (1994); Dumayne and Barber (1994).
years after calibration—means that it is usually impossible to prove a direct link.6 Similar chronological problems attend debates on the fate of the Hadrian’s Wall landscape at the end of the Roman period, and the question of whether there was any continuity of agricultural activity, or land abandonment leading to woodland regeneration.7Although most adequately-dated pollen sequences show evidence for at least some woodland regeneration (although not total land abandonment) at around the end of the Roman period in the region,8 some archaeologists have invoked pollen evidence to argue for an interlude of continuity before this regeneration occurred.9 This is in the context of indications from the archaeological evidence for post-Roman activity at sites such as Birdoswald and Housesteads,10 yet the chronological limitations of the pollen evidence makes such an assertion problematic.11 Recent research at Butterburn 6 This problem has been discussed in relation to the pollen sequence from Fozy Moss and other sites in the Hadrian’s Wall area: Dumayne et al. (1995). 7 Turner (1979); Dark and Dark (1996); Dumayne-Peaty (1999); Dark (2000) 150–54; Huntley (2000). 8 As detailed in Dark and Dark (1996), including (from west to east): Walton Moss: Dumayne and Barber (1994); Fellend Moss: Davies and Turner (1979); Fozy Moss: Dumayne and Barber (1994); Quick Moss: Rowell and Turner (1985); Steng Moss: Davies and Turner (1979); and Hallowell Moss: Donaldson and Turner (1977). 9 E.g. Casey (1993); Collins (2004); Collins (2012) 134–37; Wilmott (1997) 225. 10 Birdoswald: Wilmott (1997). Housesteads: Rushworth (2014) 197–99. 11 For example, Collins (2004) 15 asserts that at Fellend Moss “maximum levels of clearance established in the Roman
Hadrian ’ s Wall in Context
Flow, 10 km north of the central sector of the Wall, has demonstrated the possibility of obtaining more precise (decadal) chronologies by ‘wiggle-matching’ a large set of AMS (Accelerator Mass Spectrometry) dates to the radiocarbon calibration curve, but the number of dates required makes this expensive.12 Further challenges lie in identifying the extent of the vegetation around a site represented by its pollen assemblages. It has often been assumed that the peat bogs providing the source of most ‘off-site’ pollen sequences from the Hadrian’s Wall area have regional pollen source areas extending for distances of at least several hundred metres around the site (following Jacobson and Bradshaw’s theoretical model), and perhaps, as argued by Huntley, for as much as 10–30 km (based on Prentice’s model).13 This, and the chronological problems noted above, has led to understandable scepticism by some archaeologists as to the value of palynological evidence for shedding any light on the ‘transition period’.14 Fortunately, there has been a significant increase in research into pollen source areas over the last 20 years, including on sites close to Hadrian’s Wall,15 indicating that, in fact, most of the pollen in such peat-bog sequences is more local in origin, from vegetation within a few hundred metres of the sampling site.16 This makes their spatial resolution more relevant to understanding the environs of adjacent archaeological sites. Unfortunately, however, peat-bog sequences suffer from
period are continued and maintained up to c. 620, determined through numerous calibrated radiocarbon dates (Davies and Turner 1979)”. In fact there are two radiocarbon determinations for the part of the sequence from the 1st millennium AD, 1948+45 B.P. (SRR-876) and 1330+40 B.P. (SRR-875). The latter corresponds to Davies and Turner’s ‘maximum forest regeneration’, which they placed at ‘AD 620’, the uncorrected centre point of the determination, although Turner (1979) later gave a ‘tree-ring range’ for this date of AD 620–50 (calibration against the latest IntCal13 calibration curve: Reimer et al. (2013)), using the OxCal 4.2 program (Bronk Ramsey (2009)), gives a calibrated range (rounded) of 640–770 cal. AD). As the date relates to the ‘maximum forest regeneration’, the trigger to this regeneration must have been earlier, given that it takes decades for trees to colonise an abandoned site and to mature to flowering. The ribwort plantain (Plantago lanceolata) curve suggests that this change was some 20 cm lower down the sequence which, at Davies and Turner’s estimated accumulation rate of ca. 7 cm/100 yr, implies that regeneration actually began some 300 years before the dated arboreal pollen peak, i.e. around the 4th–5th c. AD 12 Yeloff et al. (2007) obtained 42 for their ca. 5000-year long sequence. 13 Jacobson and Bradshaw (1981); Huntley (2000); Prentice (1988). 14 E.g. Wilmott (2000) 15–16. 15 Notably Bolton Fell Moss and Walton Moss: Dumayne-Peaty and Barber (1998). 16 Bunting (2002).
123 the additional problem that a significant component of the pollen incorporated into the peat—including that of the heather family (Ericaceae, of which the commonest species is usually Calluna vulgaris), sedges (Cyperaceae) and grasses (Poaceae)—may derive from plants growing on the peat surface itself, potentially from within a few metres of the sampling site.17 This presents difficulties of differentiating between natural successional/climate-induced changes in this in situ vegetation and anthropogenic impacts on the wider landscape, as these taxa are also important in many non-wetland plant communities. In reality, no single source area can be defined for pollen reaching a site, of whatever type, because of differences in pollen production and dispersal characteristics of different plant groups.18 For example, the pollen of cereals—which, with the exception of rye (Secale cereale), are mostly self-pollinated and release very little, poorly-dispersed pollen—is strongly under-represented in the pollen record, to the extent that its mere presence is likely to indicate local cultivation. On the other hand, the pollen of most north-west European trees—which are, with some exceptions, wind-pollinated and produce abundant pollen—is widely dispersed and much better represented in pollen sequences. The interpretation of the predominant sources of pollen to a deposit must, therefore, be tailored to a consideration of the pollination biology of the individual taxa concerned, as well as to the type and extent of the depositional context. Some of the problems of chronology and source area can be addressed by the examination of on-site pollen sequences directly linked with Hadrian’s Wall and associated structures.19 Such evidence can provide a ‘snapshot’ view of local vegetation at a time defined by the stratigraphic relationship of the samples to the archaeological sequence, but this may be hampered by poor preservation and disturbance by on-site activities. At Vindolanda, for example, deposits from the ditches of the first fort contained a high proportion (over 40% in most samples) of deteriorated pollen,20 perhaps reworked from erosion of the ditch sides. This potentially biases the assemblages, thus limiting their value for environmental reconstruction. Although also often affected by poor preservation, pollen assemblages from the original land surface sealed beneath Hadrian’s Wall have been particularly instructive in illustrating spatial variability in the immediate pre-Wall environment, as at Birdoswald, situated at a point where Hadrian’s
17 Bunting (2003). 18 See, for example, Broström et al. (2008). 19 Usefully summarised by Huntley (1999); Huntley et al. (2009). 20 Manning et al. (1997).
124 Wall was originally built of turf.21 Here, pollen samples were analysed from the buried surface beneath the Turf Wall,22 which was constructed in the 120s and later partially demolished when the stone fort was built. These suggested that the site was wooded immediately before the local section of Wall was built, pointing to a direct link between clearance and construction, although the availability of turf also suggests local open areas, unless it was brought in from further afield. In contrast (pollen analysis of the buried surface below) the Turf Wall at Appletree, 2 km to the west, indicated an open moorland environment before construction.23 On-site samples have yet to contribute to the post-Roman environmental change debate, but have the potential to do so where deposition continued across this crucial period, as may have been the case for ditch fills, wells, and so on. Multi-Proxy Palaeoenvironmental Analyses from Lakes An alternative approach to improving the link between the archaeological and palaeoenvironmental records from the Hadrianic frontier is to use lake sediments (which lack the complicating factor of extensive in situ vegetation), and to combine pollen analysis with analysis of other palaeoenvironmental indicators which can help disentangle ‘natural’ from anthropogenic changes. This approach has been adopted for long sediment sequences from two lakes, Crag Lough and Grindon Lough, adjacent to the central sector of Hadrian’s Wall and the Stanegate respectively (fig. 1). Crag Lough Crag Lough lies immediately north of the central sector of Hadrian’s Wall, at the foot of the dramatic escarpment of the Whin Sill at Highshield Crags (fig. 2). Two turrets (38A and 38B) overlooked the lake in the Roman period, and milecastles 38 and 39 lay close to the eastern and western ends of the lake respectively. The remains of the fort at Housesteads (Vercovicium),24 with its vicus (civilian settlement) and field systems, lie on the line of the Wall 2 km east of the lake, while the fort and vicus at Vindolanda are a similar distance to the south, by the Stanegate (fig. 3).25 Just 350 m from the eastern end of the lake, between the Wall and the vallum, is the site of the Milking Gap farmstead,26 which appears, on the basis of pottery recovered during excavations in 1937, to 21 Wilmott (1997). 22 Wiltshire (1997). 23 Wiltshire (1997); Wells (2009). 24 Crow (2004); Rushworth (2014). 25 Birley (2009). 26 Kilbride-Jones (1938).
Dark
date from the 1st/2nd c. AD, and is generally assumed to have been abandoned when the Wall was built. A 7.5 m long sediment core was obtained from the western end of the lake, and analysed for pollen, charcoal particles, macroscopic plant remains and sedimentology (loss-on-ignition and magnetic susceptibility). A pollen percentage diagram showing the main taxa and crops present is presented in Figure 4,27 accompanied by data on the minerogenic component of the sediment determined by loss-on-ignition analysis, and its micro-charcoal content, estimated by point counting.28 Full details of the pollen and associated analyses, and programme of radiocarbon dating, have been published elsewhere.29 The pollen diagram has been divided into zones by statistical analysis of the pollen assemblages, with pollen zone boundaries assigned an approximate calendar date using an age-depth model. This was derived from the series of radiocarbon accelerator dates on terrestrial plant remains picked from the sediment matrix.30 The Crag Lough sequence extends back to the Late Neolithic period, ca. 3000 cal. BC, at which time the landscape was substantially wooded. Clearance of oak (Quercus) and hazel (Corylus avellana) woodland began at a level dated 4030+50 BP (AA-28172) (2860–2460 cal. BC), in the Late Neolithic-Early Bronze Age, associated with a major increase of minerogenic inwash and charcoal deposition, reflecting soil disturbance and possible use of fire to assist in clearance. Sporadic records of Hordeum-type pollen appear soon after, probably representing local cultivation of barley (Hordeum vulgare), although wetland grasses are another possible source as the pollen type includes flote-grass (Glyceria fluitans). By the later Bronze Age, ca. 1200 BC, the more consistent presence of Hordeum-type pollen, accompanied by a first record for Avena-type,31 suggests sustained local cereal cultivation, in the context of a further decline in
27 Nomenclature of pollen and spore types follows Bennett et al. (1994). Pollen percentage calculations are based on a sum including all identifiable pollen and spores, excluding those of obligate aquatics and Sphagnum moss. The pollen data have been zoned, and the pollen diagram produced, using the ANSI C program psimpoll: Bennett (1994). 28 Clark (1982). 29 Dark (2005). 30 The dates are here calibrated using OxCal 4.2 (Bronk Ramsey (2009)) with the IntCal13 datasets (Reimer et al. (2013)), the results of which vary in some cases by a decade when rounded from those previously published in Dark (2005). Linear interpolation between adjacent determinations has been used to estimate the date of the pollen zone boundaries. The error terms for the radiocarbon dates, which are displayed on the pollen diagram, range between +40 at one standard deviation for the upper 300 cm of the sequence (i.e. deposits