The Routledge Companion to Big History 2019015132, 2019980202, 9781138905818, 9780429299322

939 130 10MB

English Pages [507] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

The Routledge Companion to Big History
 2019015132, 2019980202, 9781138905818, 9780429299322

Table of contents :
Cover
Half Title
Title Page
Copyright Page
Dedication
Contents
List of figures
List of tables
List of contributors
Introductory chapters
Introduction to The Routledge Companion to Big History
1 What is big history?
PART I: Big history and science
2 Big history and the study of time: the underlying temporalities of big history
3 Big history and astronomy – space is big: the Fermi paradox: its relevance to big history and the human race
4 Big history and macro-evolution: evolutionary principles and mechanisms at biological and social phases of the big history
PART II: Big history, social science and the humanities
5 Big history and anthropology: our place in the multiverse: anthropology, civilization and big history
6 Big history and archaeology: archaeology is big history
7 Big history and philosophy: philosophical foundations of big history: why big history makes sense
8 Big history and political science: science, the deep past, and the political
9 Big history and historiography: deep tides and swirling foam: the influence of macro-historical trends on micro-historical events
10 Big history and critical theory: science, history and why theory matters
11 Big history, morality and religion
PART III: Little big histories
12 A case for little big histories
13 The little big history of the Nalón River, Asturias, Spain
14 Sketch of a little big history of Private E.E. Benjamin and the Great War
PART IV: Teaching big history
15 The Big History Project in Australia
16 Big history teaching in Korea
17 Crossing thresholds: using big history to meet challenges in teaching and learning in the United States
PART V: Big history and the future
18 Big history and the future of technology
19 Big history and the Singularity
20 Underground metro systems: a durable geological proxy of rapid urban population growth and energy consumption during the Anthropocene
21 The coming energy transition: what comes after fossil-fueled civilization?
Index

Citation preview

THE ROUTLEDGE COMPANION TO BIG HISTORY

The Routledge Companion to Big History guides readers though the variety of themes and concepts that structure contemporary scholarship in the field of big history. The volume is divided into five parts, each representing current and evolving areas of interest to the community, including big history’s relationship to science, social science, the humanities, and the future, as well as teaching big history and ‘little big ­histories’. Considering an ever-expanding range of theoretical, pedagogical, and ­research topics, the book addresses such questions as what is the relationship between big history and scientific research, how are big historians working with ­philosophers and religious thinkers to help construct ‘meaning’, how are leading theoreticians making sense of big history and its relationship to other creation narratives and ­paradigms, what is ‘little big history’, and how does big history impact on thinking about the future? The book highlights the place of big history in historiographical traditions and the ways in which it can be used in education and public discourse across disciplines and at all levels. A timely collection with contributions from leading proponents in the field, it is the ideal guide for those wanting to engage with the theories and concepts behind big history. Craig Benjamin is a Professor of History at Grand Valley State University in ­Michigan, ­ urasian history. USA, where he researches and teaches big history and ancient E ­Recent books include Big History: Between Nothing and Everything (­co-­authored with David Christian and Cynthia Stokes Brown, 2014); and Empires of Ancient ­Eurasia.The First Silk Roads Era 100 BCE – 250 CE (2018). Esther Quaedackers is a Lecturer in Big History at the University of ­Amsterdam, the Netherlands, where she has been developing, coordinating, and teaching big ­history courses for over a decade. She is the inventor of the ‘little big history’

approach, which is a research and teaching method in which small subjects are connected to aspects of big history in order to generate creative new ideas about how these ­subjects came to be the way they are. David Baker is Lecturer of Big History at Macquarie University, Australia, and is the first scholar worldwide to achieve a PhD in that field. He is co-designer of Big ­History School (three K-12 curricula), Big History: Connecting Knowledge on Coursera, and contributor to the Big History Project. He was scriptwriter for two seasons of Crash Course: Big History with over 12 million views on Youtube. He has produced numerous research articles, chapters, and edited volumes, including works in the Journal of World History and Proceedings of the National Academy of Sciences.

THE ROUTLEDGE COMPANION TO BIG HISTORY

Edited by Craig Benjamin, Esther Quaedackers and David Baker

First published 2020 by Routledge 2 Park Square, Milton Park, Abingdon, Oxon OX14 4RN and by Routledge 52 Vanderbilt Avenue, New York, NY 10017 Routledge is an imprint of the Taylor & Francis Group, an informa business © 2020 selection and editorial matter, Craig Benjamin, Esther Quaedackers and David Baker; individual chapters, the contributors The right of Craig Benjamin, Esther Quaedackers and David Baker to be identified as the authors of the editorial material, and of the authors for their individual chapters, has been asserted in accordance with sections 77 and 78 of the Copyright, Designs and Patents Act 1988. All rights reserved. No part of this book may be reprinted or reproduced or utilised in any form or by any electronic, mechanical, or other means, now known or hereafter invented, including photocopying and recording, or in any information storage or retrieval system, without permission in writing from the publishers. Trademark notice: Product or corporate names may be trademarks or registered trademarks, and are used only for identification and explanation without intent to infringe. British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library Library of Congress Cataloging-in-Publication Data Names: Benjamin, Craig, editor. | Quaedackers, Esther, editor. | Baker, David, 1986- editor. Title: The Routledge companion to big history / edited by Craig Benjamin, Esther Quaedackers and David Baker. Description: Abingdon, Oxon; New York, NY: Routledge, 2020. | Includes bibliographical references and index. | Summary: “The Routledge Companion to Big History guides readers though the variety of themes and concepts that structure contemporary scholarship in the field of big history. The book is divided into five parts, each representing current and evolving areas of interest to the community, including big history’s relationship to science, social science, the humanities, and the future, as well as teaching big history and ‘little big histories’. A timely collection with contributions from leading proponents in the field, it is the ideal guide for those wanting to engage with the theories and concepts behind big history”— Provided by publisher. Identifiers: LCCN 2019015132 (print) | LCCN 2019980202 (ebook) | ISBN 9781138905818 (hardback: alk. paper) | ISBN 9780429299322 (ebk) Subjects: LCSH: Civilization—Philosophy. | World history. | History—Study and teaching. | Science and civilization. Classification: LCC CB19 .R66 2020 (print) | LCC CB19 (ebook) | DDC 909—dc23 LC record available at https://lccn.loc.gov/2019015132 LC ebook record available at https://lccn.loc.gov/2019980202 ISBN: 978-1-138-90581-8 (hbk) ISBN: 978-0-429-29932-2 (ebk) Typeset in Bembo by codeMantra

This book is dedicated to the memory of two dear departed colleagues, both important pioneers in the field of Big History: Cynthia Stokes Brown and Akop Nazaretyan.

CONTENTS

List of figures x List of tables xiii List of contributors xiv Introductory chapters

1

Introduction to The Routledge Companion to Big History 3 Craig Benjamin, Esther Quaedackers and David Baker 1 What is big history? 16 David Christian PART I

Big history and science 35 2 Big history and the study of time: the underlying temporalities of big history 37 Barry Wood 3 Big history and astronomy – space is big: the Fermi paradox: its relevance to big history and the human race 57 Jonathan Markley 4 Big history and macro-evolution: evolutionary principles and mechanisms at biological and social phases of the big history 72 Leonid E. Grinin, Andrey Korotayev and Alexander Markov vii

Contents PART II

Big history, social science and the humanities 107 5 Big history and anthropology: our place in the multiverse: anthropology, civilization and big history 109 Barry H. Rodrigue 6 Big history and archaeology: archaeology is big history 156 Brian Fagan 7 Big history and philosophy: philosophical foundations of big history: why big history makes sense 170 Armando Menéndez Viso 8 Big history and political science: science, the deep past, and the political 180 Lowell Gustafson 9 Big history and historiography: deep tides and swirling foam: the influence of macro-historical trends on micro-historical events 202 David Baker 10 Big history and critical theory: science, history and why theory matters 233 David Blanks 11 Big history, morality and religion 251 Cynthia Stokes Brown PART III

Little big histories 277 279

12 A case for little big histories Esther Quaedackers

13 The little big history of the Nalón River, Asturias, Spain 300 Olga García-Moreno, Diego Álvarez-Laó, Miguel Arbizu, Eduardo Dopico, Eva García-Vázquez, Joaquín García Sansegundo, Montserrat Jiménez-Sánchez, Laura Miralles, Ícaro Obeso, Ángel Rodríguez-Rey, Marco de la Rasilla Vives, Luis Vicente Sánchez Fernández, Luis Rodríguez Terente, Luigi Toffolatti and Pablo Turrero viii

Contents

14 Sketch of a little big history of Private E.E. Benjamin and the Great War 320 Craig Benjamin PART IV

Teaching big history 337 15 The Big History Project in Australia 339 Tracy Sullivan 16 Big history teaching in Korea 361 Seohyung Kim 17 Crossing thresholds: using big history to meet challenges in teaching and learning in the United States 372 Robert B. Bain PART V

Big history and the future 395 18 Big history and the future of technology 397 Leonid E. Grinin and Anton L. Grinin 19 Big history and the Singularity 420 Akop P. Nazaretyan 20 Underground metro systems: a durable geological proxy of rapid urban population growth and energy consumption during the Anthropocene 434 Mark Williams, Matt Edgeworth, Jan Zalasiewicz, Colin N. Waters, Will Steffen, Alexander P. Wolfe, Nicholas J. Minter, Alejandro Cearreta, Agnieszka Gałuszka, Peter Haff, John McNeill, Andrew Revkin, Daniel deB. Richter, Simon Price, and Colin Summerhayes 21 The coming energy transition: what comes after fossil-fueled civilization? Joseph Voros

456

Index481

ix

FIGURES

5.1 5.2 5.3 5.4 5.5 5.6 5.7 5.8 5.9 5.10 5.11 5.12 5.13

Recently discovered cuneiform Tablet V of the Gilgamesh Epic, c 1800 bce, which provides fresh insights about ethno-geographic encounters in the epic’s Cedar Forest110 Petroglyphs and a Picenean inscription found in the vicinity of Mt. Conero in Italy111 Bisotun inscription. John Quackenbos, Illustrated History of Ancient Literature, Oriental and Classical, New York: Harper & Brothers, 1882, p. 65111 Antler carving of faces interpreted as Dorset (below) and European (above), c fourteenth century ce, Baffin Island, Nunavut (Canada)113 Sultan Firoz Shah Tughlaq of the Delhi Sultanate had this Ashokan pillar (third century bce) removed from Topra Kalan (Haryana) to Firozabad as part of his antiquarian collections in 1356 ce114 Hathor shrine (fifteenth century bce), Deir el-Bahari, Egypt116 Excavation of the Ōmori shell mound, Jōmon culture, Shinagawa (Tokyo), Japan, c 1877117 French postcard from Buffalo Bill’s Wild West Show, 1903118 Anthropologist Irawati Karve conducting a field interview, Maharashtra, India c 1960120 Green Dragon Bridge, near Nankau Pass, Great Wall of China121 Paul Gauguin, Where Do We Come From? What Are We? Where Are We Going? D’où Venons Nous/Que Sommes Nous/Où Allons Nous, Tahiti, 1897122 Cover of the Bangla edition of the book by Nigel Hughes and Rati Basu, Monishar Pathorer Bon [Monisha and the Stone Forest], Kolkata: Monfakira Press, 2012125 Anthropology and big history students from Symbiosis International University doing field work at a megalith in Lohegaon, Maharashtra, India on 10 May 2018125 x

Figures

NASA’s ‘Plaque of Humanity’, etching made for the Pioneer 10 space probe in 1972129 9.1 Relationship between population growth and carrying capacity (Richerson, Boyd, Bettinger, 219)222 9.2 Relationship between s-curves and carrying capacity. The asterisk (*) marks a period of severe population decline where learning is lost (Richerson, Boyd, Bettinger, 219)222 12.1 A possible layered structure of a little big history284 13.1 Iberian Peninsula. Inset: Asturias and the Nalón River basin301 13.2 Anthropic changes through time in landscapes in the Nalón River basin304 13.3 Recreation of Carboniferous period forests309 13.4 Woolly rhinoceros lived in the Asturias area during cold phases in the Quaternary311 13.5 Peña de Candamo Cave. Niche: Horses and aurochs312 13.6 Spark Plasma Sintering equipment (left) in the Nanomaterials and Nanotechnology Research Centre (upper right) and nanostructured materials made therein (lower right)316 15.1 Disciplinary continuum342 15.2 Big history pedagogical framework352 15.3 Mezirow’s progression of autonomous thinking355 15.4 Big history pedagogical framework and transformative learning theory357 17.1 Cause & Consequence Activity 1378 17.2 Cause & Consequence Activity 2378 18.1 The phases of the Cybernetic Revolution400 18.2 The relationship between citation frequency in scientific publications and the technologies forming MANBRIC, according to the Web of Science, 2010–2015402 18.3 Dynamics of the global combined share of four technologies with the highest share of patent applications in 1985 (electrical machinery, measurement, machine tools, and other special machines) in comparison with the dynamics of the global combined share of patent applications in four top categories (medical, pharmaceutical, computer, and biotechnologies), 1985–2014403 19.1 Scaling law in the phase transitions426 20.1 Linear correlation between the number of operational metros (1859–2010) and global urban population (data from Table 20.1)435 20.2 Development of the London Underground System from 1863 to present, showing a broad reduction in age of construction and transfer from subsurface to surface lines and stations towards the network periphery436 20.3 The relationships of metros with surface anthroturbation, the archaeosphere and deep geology437 5.14

xi

Figures

20.4

Cutaway of London’s five levels of traffic at Charing Cross Station (now Embankment Station), taken from the Popular Science Magazine, January 1921, pp. 44–45, drawing by S.W. Clatworthy445 20.5 (A,  B) Urban population growth and energy use 1850–2010 (based on figures in Table 20.1); (C) number of metro systems plotted against time (see also ­Gonzalez-Navarro & Turner, 2016); (D) urban population as a percentage of ­total population plotted against number of metro systems.The increase in number of metro systems in the post 1950 period is evident and approximates ­economic changes associated with the Great Acceleration (Steffen et al., 2015)446 20.6  Global spread of major metro systems, 1863 to present, with inset maps for Europe and China447

xii

TABLES

5.1 Rendition of Eric Chaisson’s table of average energy rate densities128 9.1 Predictions for the theory of secular cycles, in any given phase217 9.2 Amount of free energy running through a gram per second, and the australopithecine and human free energy rate density is determined from the average energy consumption of an individual, Chaisson 2010: 28 & 36223 18.1 The phases of the Agrarian Revolution398 18.2 The phases of the Industrial Revolution399 20.1 Urban primary energy use: 1850–2010435

xiii

CONTRIBUTORS

Diego Álvarez-Laó is Profesor Contratado Doctor of Paleontology at the Geology Department, Universidad de Oviedo, Spain. Miguel Arbizu is Honorary Professor, Universidad de Oviedo, Spain. Robert B. Bain is an Associate Professor in the Departments of Educational Studies and History at the University of Michigan, USA, and co-faculty lead on the Big History Project and the new On World History Projects. A PhD in history, Bob taught more than 25 years in a high school and has been teaching history and history education at University of Michigan for the past 21 years. He studies the practice of teaching and learning history in multiple contexts including elementary, secondary, and university classrooms as well as in museums. His current work centers on ways to improve the quality of world and big history instruction. David Baker is Lecturer of Big History at Macquarie University, Australia, and is the first scholar worldwide to achieve a PhD in that field. He is co-designer of Big ­History School (three K-12 curricula), Big History: Connecting Knowledge on Coursera, and contributor to the Big History Project. He was scriptwriter for two seasons of Crash Course: Big History with over 12 million views on Youtube. He has produced numerous research articles, chapters, and edited volumes, including works in the Journal of World History and Proceedings of the National Academy of Sciences. Craig Benjamin is a Professor of History at Grand Valley State University in ­Michigan, USA, where he researches and teaches big history and ancient ­Eurasian history. ­Recent books include Big History: Between Nothing and Everything (­co-­authored with David Christian and Cynthia Stokes Brown, 2014); and Empires of Ancient ­Eurasia.The First Silk Roads Era 100 BCE – 250 CE (2018). David Blanks  is a Professor of History at Arkansas Tech University, USA. His ­research interests include the philosophy of Big History, the history of European attitudes toward Islam, and the social history of medieval and early modern Europe. xiv

Contributors

Alejandro Cearreta  is in the Departamento de Estratigrafía y Paleontología, ­Facultad de Ciencia y Tecnología, Universidad del País Vasco UPV/EHU, Spain. David Christian is a Professor of Modern History and Director of the Big History ­Institute at Macquarie University, Sydney, Australia. His research Interests include Big ­History and Russian and Soviet history. Recent publications include Origin Story: A Big History of Everything (2018), A History of Russia, Central Asia and Mongolia: Vol 2: Inner ­Eurasia from the Mongol Empire to today: 1260–2000 (2018); and Big History: Between ­Nothing and Everything (co-authored with Cynthia Stokes Brown and Craig Benjamin, 2013). Marco de la Rasilla Vives  is a Faculty Member in the History Department at ­Universidad de Oviedo, Spain. Daniel deB. Richter is a Professor in the Division of Environmental Science and Policy in the Nicholas School of the Environment, Duke University, USA. Eduardo Dopico is Associate Professor of Pedagogy at the Department of ­Education Sciences, Universidad de Oviedo, Spain. Matt Edgeworth is a field archaeologist. He is also Honorary Visiting Research Fellow at the School of Archaeology and Ancient History, University of Leicester, UK. Brian Fagan is Distinguished Emeritus Professor of Anthropology at the University of California, Santa Barbara, USA. He has written several books on ancient ­climate change, including The Little Ice Age and The Great Warming, and recently Fishing (2018). Agnieszka Gałuszka  is in the Geochemistry and the Environment Division, ­Institute of Chemistry, Jan Kochanowski University, Poland. Olga García-Moreno  is a professor in the Geology Department of Universidad de  Oviedo, Spain. Coordinator of the Gran Historia (big history) course in that university, she has interests in little big histories and the link between geological ­processes and evolution of life and humanity. Joaquín García Sansegundo is Professor Titular, in the Department of Geology, Universidad de Oviedo, Spain. Eva García-Vázquez  is Professor in the Department of Functional Biology, ­Universidad de Oviedo, Spain. Anton L. Grinin  has a PhD in Biological Sciences and is Senior Research Fellow of the International Center for Education and Social and Humanitarian Studies as well as leading Research of Volgograd Centre for Social Research, Russia. His main research interests include Big History, evolution, biotechnologies, global ­technological transformations and forecasts. Anton L. Grinin is the author of over 50 ­scholarly publications, including 2 monographs. He is the co-author of the xv

Contributors

monograph ‘From ­Biface to Nanorobots: The World on the Way to the Epoch of Self-Regulating Systems’ and a number of articles including ‘Technological Forecasting and Social Change’ ‘Macroevolution of Technology’ and ‘Global Technological Transformations’. Leonid E. Grinin is a Senior Research Professor at the Institute for Oriental Studies of the Russian Academy of Sciences in Moscow, Russia, and Senior Research Professor at the Laboratory for Destabilization Risk Monitoring at the National Research University Higher School of Economics. His research interests include evolution, the philosophy of history, and historical trends and future studies. He is the Editor-in-Chief of the journal Age of Globalization (in Russian), as well as a co-editor of the international journals Social Evolution & History and the Journal of Globalization Studies. Dr. Grinin is the author of over 500 scholarly publications in Russian and English, including 33 monographs. Lowell Gustafson  is a Professor of Political Science at Villanova University, ­Philadelphia, USA. Peter Haff is a Professor in the Nicholas School of the Environment, Duke University, USA. Montserrat Jiménez-Sánchez is a Professor at Universidad de Oviedo, Spain. Seohyung Kim is Chief Director of the CHO Big History Academy in Korea; a Research Professor at the Eurasian Center for Big History and Systems Forecasting, and Russia; and was a Research Professor at the Institute of Global History, Ewha Womans University, Seoul, Korea. Recent publications include Big History for Primary Students (2017); and Big History: Origin of Human History (2018). Andrey Korotayev  is Head of the Laboratory for Monitoring of Destabilization Risks in the National Research University Higher School of Economics, Moscow, Russia. He is also a Senior Research Professor in the System Forecasting Center, Oriental Institute, Russian Academy of Science; and a Senior Research Professor at the International Laboratory of Political Demography, Russian Presidential Academy for National Economy and Public Administration. Jonathan Markley  is a professor in the History Department at California State University Fullerton, USA. Alexander Markov  is a biologist, DSc, head of the department of Evolutionary Biology of Moscow State University and a leading researcher at the Paleontological Institute of the Russian Academy of Sciences. His works are devoted to evolutionary theory, quantitative analysis of the paleontological record, and major evolutionary transitions in the history of life on Earth. John McNeill is at Georgetown University, USA. xvi

Contributors

Armando Menéndez Viso graduated in Philosophy (Oviedo) and in Economics (UNED), he received his PhD in 2002 from the Complutense University of ­Madrid, after four years as a research fellow at the Institute for Philosophy of the CSIC (­Spanish national Higher Council for Scientific Research). He has worked as a lecturer at the European University of Madrid and as a post-doctoral research fellow at the ­University of Exeter (UK). He is currently Associate Professor of Philosophy. His main research interests are science and values, philosophy of social sciences (particularly philosophy of economics), ethics of science and technology, and sustainability. He has been visiting scholar at the Delft University of Technology, the University of Amsterdam (The ­Netherlands), the Universidad Autónoma de Santo Domingo (Dominican Republic), and the University of Guadalajara (Mexico). Currently he is a member of the STS Group and the Big History Group of the University of Oviedo, Spain. Nicholas J. Minter is in the School of Earth and Environmental Sciences, ­University of Portsmouth, UK. Laura Miralles is working at the Department of Functional Biology, Universidad de Oviedo, Spain. Akop P. Nazaretyan  was Director of the Eurasian Center for Big History & ­System Forecasting; a Senior Research Professor of the Institute of Oriental Studies of ­Russian Academy of Sciences; a Full Professor at Moscow State University; and ­Editor of the academic journal Historical Psychology & Sociology (in Russian). He is the author of over 300 scholarly publications, including books: Intelligence in the Universe: Sources, Formation and Perspectives (1991, in Russian); Aggression, Morals and the Crises in World Cultural Development (1995, 1996, in Russian); Aggressive Crowds, Mass Panic, and Rumors: Lectures in Social and Political Psychology (2001, 2003, 2005, in Russian); Civilization Crises within the Context of Big History: Self-Organization, Psychology, and Forecasts (2001, 2004, in Russian); Anthropology of Violence and Culture of Self-Organization. Essays on Evolutionary Historical Psychology (2007, 2008, in Russian); and Evolution of Non-violence: Studies in Big History, Self-Organization and Historical Psychology (2010, in English). Akop passed away in 2019. Ícaro Obeso is a Predoctoral researcher at Universidad de Oviedo, Spain. Simon Price is in the Department of Geography, University of Cambridge, UK. Esther Quaedackers is a Lecturer in Big History at the University of ­Amsterdam, the Netherlands, where she has been developing, coordinating, and teaching big ­history courses for over a decade. She is the inventor of the ‘Little Big History’ approach, which is a research and teaching method in which small subjects are connected to aspects of Big History in order to generate creative new ideas about how these ­subjects came to be the way they are. Andrew Revkin is at Pace University, Dyson College Institute for Sustainability and the Environment, USA. xvii

Contributors

Barry H. Rodrigue is a geographer, archaeologist, and professor of anthropology at Symbiosis International University, Pune, Maharashtra, India. He engages in the study of human ecology around the north Pacific Rim, the eastern borderlands of Canada, and the Eurasian Highlands. He serves on the editorial board of History, Archaeology and Ethnography of the Caucasus, a journal of the Daghestan Scientific Centre, Russian Academy of Sciences, in Makhachkala. He sees the use of Big History’s micro/ macro lens as an important tool for humanity’s understanding of its place in the cosmos, as well as a key to our survival as a species on Earth. As editor-in-chief of the three-­ volume collection, From Big Bang to Galactic Civilizations: A Big History Anthology (2015–2017), he compiled chapters by 100 scholars from two dozen nations. Ángel Rodríguez-Rey is Professor of Petrology and Geochemistry at Universidad de Oviedo, Spain. Luis Rodríguez Terente is Conservator of the Museum of Geology at Universidad de Oviedo, Spain. Luis Vicente Sánchez Fernández is Associate Professor at Universidad de Oviedo, Spain. Will Steffen is at the Australian National University, Australia. Cynthia Stokes Brown was a Professor Emerita of Education and History at the Dominican University of California. She is the author of Big History: From the Big Bang to the Present (2007) and co-author (with David Christian and Craig Benjamin) of Big History: Between Nothing and Everything (2015); and Big History, Small World: From the Big Bang to You (2016). Cynthia passed away in 2017. Tracy Sullivan  is Education Leader for the Big History Institute at Macquarie ­University, Sydney, Australia. A former classroom teacher and development team member for the Big History Project and Big History School, her research focuses on interdisciplinary curriculum and related learning outcomes for adolescent and adult learners. Colin Summerhayes is at the Scott Polar Research Institute, Cambridge University, UK. Luigi Toffolatti  is Professor of Astronomy and Astrophysics at Universidad de Oviedo, Spain. Pablo Turrero is at the Universidad Nacional de Educación a Distancia, Spain. Joseph Voros is a Senior Lecturer in Strategic Foresight in the Faculty of Business and Law at Swinburne University of Technology in Melbourne, Australia, where he teaches courses in Big History and Foresight/Futures Studies. His recent research is focused on the coming ‘civilizational transition’, using Big History as the framing perspective, and on the potential futures for humankind which lie in and beyond it. xviii

Contributors

Colin N. Waters  is in the School of Geography, Geology and Environment, ­University of Leicester, UK. Mark Williams is in the School of Geography, Geology and Environment, ­University of Leicester, UK. Alexander P. Wolfe  is in the Department of Biological Sciences, University of ­Alberta, Canada. Barry Wood  is a Professor of English in the College of Liberal Arts and Social ­Sciences at the University of Houston, Texas, USA. Jan Zalasiewicz  is in the School of Geography, Geology and Environment, ­University of Leicester, UK.

xix

Introductory chapters

INTRODUCTION TO THE ROUTLEDGE COMPANION TO BIG HISTORY Craig Benjamin, Esther Quaedackers and David Baker

Welcome to The Routledge Companion to Big History, a collection of 21 chapters authored by leading proponents in the field on a variety of current topics relevant to big history research and education. The field of big history has been in existence for around a quarter of a century and is practiced today as a coherent form of research and teaching by historians, physicists, geologists, biologists, anthropologists, and other disciplinary specialists around the world. Over the past decade in particular big history has been increasing its academic profile through a variety of new publications, pedagogical innovations, more serious thinking about a research agenda, and through the establishment of the International Big History Association (IBHA). At the same time, big history has also increased its public presence through various high-profile media and web-based outlets, including big history courses produced for The Teaching Company’s Great Courses collection, big history programs produced by the History Channel and Discovery Channel, TED talks, and the involvement of Bill Gates in the creation of the Big History Project (BHP), which has led to more than a thousand schools and hundreds of thousands of high school students around the world taking courses in big history. All these developments mean that the publication of this Routledge Companion to big history comes at a most opportune moment in the evolution of the field. This introduction to the volume is intended to offer a brief overview of the field of big history, and also of some of the most important of these recent developments, in order to provide readers of the volume with some context for the chapters it contains, and of the current state of the field. Despite this increasing profile, many educators and researchers remain unaware of big history, its place in the historiographical traditions of history writing, and how the approaches of big history might be easily and usefully incorporated into a wide range of foundational classes in a variety of disciplines, as the chapters in this collection attempt to demonstrate.

Historiographical origins of big history Big history did not spring out of some historical vacuum. It is a continuation of the great historiographical tradition of universal history, which in its written form dates 3

Craig Benjamin et al.

back to Classical Greece and Han China, and in its oral form to the earliest human communities.The first century BCE Sicilian Greek historian Diodorus Siculus made explicit connections between the cosmos and events on Earth, in his attempt to recount the ‘common affairs of the inhabited world’ in a single, unified narrative.1 This idea stemmed from the much more ancient oral creation myths that have been devised and told by every human society since our species first appeared on the planet some 250,000 years ago. Following the invention of writing, attempts to create a written unified reckoning of the past quickly followed. For the Chinese Early Han Dynasty historian Sima Qian, this meant using history to recreate the universe in microcosm, with all its contradictions, inconsistencies, and hidden connections.2 For early Christian historians like Orosius and Augustine, human and cosmic history were seen as part of the same connected narrative focused on the unfolding of God’s divine plan.3 Islamic historian Ibn Khaldun largely rejected this, and in his great work the Muqaddimah focused instead on connections between the environment and the various types of societies humans have constructed over time.4 European Enlightenment historians like Voltaire and de Condorcet, while acknowledging the vagaries and impact of geology and biology, preferred to unfold a universal narrative that was progressive and secular, and that culminated in the triumph of civilization and reason over both barbarism and the physical world in their own ‘enlightened’ times.5 Immanuel Kant made explicit connections between the heavens and the inner workings of human history both in his philosophy and on his tombstone, which reads: ‘The starry skies above and the moral law within’. But for his successor Hegel, the single reckoning of universal history could be reduced to one human-sized maxim: ‘the history of the world is none other than the progress of the consciousness of freedom’.6 For Marx, large-scale history demonstrated nothing more than the rejection of natural human relationships with the environment and the ultimate enslavement of the human species to the mindless and ‘unnatural’ productive forces of capitalism.7 Despite these implicit intentions to explore the universality of the human experience and its connection to the physical world, most of these historians and philosophers remained narrowly focused on written documents as their primary source of evidence, and thus on human history. An exception to this was ­eighteenth century naturalist, Alexander von Humbolt, who in his intended (but unfinished) multi-­volume publication Kosmos planned to summarize and connect all the knowledge that had thus far been accumulated about both the physical world and the realm of human history.8 Early in the twentieth century, the science fiction novelist H.G. Wells enjoyed great success with his book The Outline of History, which was another attempt to expound in a unified way the latest knowledge about the solar system, planet, life, and the essential elements of human history.9 In the first half of the twentieth century, attempts to relate the ‘big’ story fell on hard times. The work of Toynbee, Spengler, and others was seen as too vague and generalized to be of much use to the smaller scale, more specialized historians who now dominated the discipline. In the 1960s and 1970s, it was the scientific community that re-embraced the meta-narrative, despite the critiques of post-modernists. Physicists began to recount the history of the entire universe from the big bang to the present in audacious, chronological narratives; while biologists and geologists used 4

The Routledge Companion to Big History

evolutionary theory and stratigraphical studies to offer their own ‘big histories’ of the planet and all life upon it. Even an abstract entity like ‘time’ was seen as fair game for a universal, historical account, with the 1988 publication of Stephen Hawking’s A Brief History of Time an obvious example.10 One of the implications of this historicization of science is that, when large-scale history began to make a historiographical comeback later in the twentieth century, it often did so with the help of the historical sciences. These various elements then – the oral creation myth, attempts to write a ‘single reckoning of past events’, the identification of key themes that run like threads through the confusing morass of natural and human history, and the historicization of science – are at the intellectual heart of big history today, the latest development in this historiographical and scientific tradition. Like its predecessors, big history uses intensive interdisciplinary research and the most advanced historical and scientific knowledge to unfold the story of the evolution of the cosmos, and of the place of humans within it. Because of the extraordinary scientific breakthroughs that have occurred since the 1950s, particularly the discovery of evidence for the big bang theory, the solar nebula theory of the formation of stars and solar systems, the principles of plate tectonics, genetic evidence for evolution, and the techniques of radiometric dating, big historians are now equipped with the knowledge and tools to write the most accurate creation story ever devised. David Christian, one of the founders of the modern field of big history (and the coiner of the term ‘big history’) has argued that it was only after these discoveries, particularly of the principles of radiometric dating which facilitated a profound ‘chronometric revolution’, that the modern big history account became possible.11 These techniques and scientific breakthroughs have allowed big historians to dramatically expand the breadth and depth of their enquiries, and to push the start date of their narrative back to that point in time before which we can say nothing with any certainty – the big bang some 13.8 billion years ago. So, as extraordinary as this sometimes sounds to traditional historians, most big history courses cover billions of years of cosmic and human history in a single semester, from the origins to the possible ultimate fate of the universe. Here are some of the questions that students in a big history course, either at the high school or college level, might reasonably be expected to provide coherent, accurate answers to by the end of a semester: • • • • • • • • • • •

How and when does modern science suggest the universe was created? How and when were our sun and solar system created? How and when was the earth created? What forces created and shaped the earth’s lands, seas, and atmosphere? How and when did life first appear on earth, according to modern science? How did life assume the variety of forms we see today? How and when did modern humans evolve? How did the earliest human communities live? When does human history begin, and why? How and when did agriculture first appear? How did humans live in the first millennia following the adoption of agriculture? 5

Craig Benjamin et al.

• • • • • • • • •

How and when did the first cities and states evolve? Why did some of these evolve into agrarian civilizations, while others did not? What are some of the defining characteristics of agrarian civilizations? How did agrarian civilizations change over 4,000 years of history? What are the origins of modern industrial society? Why did the ‘modern revolution’ take a European form? Is the twentieth century different from all previous centuries in human history? Does a study of history on this scale help us predict the future? What is the future trajectory of complexity over billions of years?

Perhaps most importantly, by the end of the semester, students should also be able to understand how the answers to all of these questions are related. This means they should also be able to answer other types of questions such as: • • •

What creates, sustains, and increases forms of complexity throughout big history? Are there any unifying patterns and trends in human history? How can big history contribute to a better understanding of different aspects of the world around us?

Readers unfamiliar with big history would be excused for thinking that these questions could surely be addressed only in a very superficial and generalized way in a single course, but this is not the case. Students learn the essential elements of all the relevant scientific, social, and historical theories, along with details of the evidence that supports them. They also learn to connect the essential of cosmology, astronomy, physics, chemistry, geology, biology, anthropology, archaeology, economics, sociology, and history, which adds to their understanding of the world in ways learning about individual disciplines does not. As Fred Spier, another important pioneer of big history, has noted, ‘although all the knowledge taught in big history courses is readily available in academic settings, only rarely is it presented in the form of one single historical account’.12 It is this interdisciplinarity and connectivity that makes big history so relevant and exciting to students and members of the public alike, as seen by its increasing profile and popularity. When compared to other genres within the discipline of history, big history is clearly audacious in scope, content, and methodology. Yet over the past quarter of a century, it has proven to be an extraordinarily popular and useful approach for students of all levels, who are continuously encouraged to seek ways to connect various fields of human knowledge, and to find a context for their own careers and interests. It is not unusual to find students using course evaluations to describe a semester of big history as literally a life-changing experience. Partly this is because big history also demands something of a moral commitment from students, as they are encouraged to discover their place in the connected global village of the twenty-first century, and how they might better contribute to the future of that society and the biosphere that supports it. In the tradition of the aboriginal Dreamtime stories, and of Diodorus, Khaldun, Voltaire, Kant, Marx, and Hawking to name but a few, big historians have constructed a new ‘single reckoning of the past’ that begins with the big bang and ends with the 6

The Routledge Companion to Big History

fate of the universe billions of years in the future. This is world history writ large across a vast cosmic canvas. As Bruce Mazlish so neatly puts it, big history is a ‘testament to the human desire to know the whole of the past, envisioned in one sweeping vision, overleaping the limited and limiting boundaries humans have sought to place on the earth’.13

Twenty-first century developments in the field There can be no doubt that the expansion of the field of world history in the late twentieth century enriched the entire discipline of history. The examination of human affairs from an enhanced perspective has allowed for the identification of processes, themes, and patterns that have provided context and meaning for all the seemingly chaotic details of human interaction. World history has not attempted to replace small-scale historical research, which provides the details and substance upon which these broader patterns are constructed, but rather to complement that work. Over recent years, it is often world historians that have been at the forefront of arguments that big history does precisely the same thing, but at much greater scales again. While the world history lens is wider than the lens more familiar to the national or biographical historian, the big history lens examines human, planetary, and cosmic history at the widest angle thus far possible. Marnie Hughes-Warrington, a pioneer in both world and big history, argues that for far too long historians have limited the array of lenses available to them:‘It is as if the lens through which we view the past has got stuck at a certain magnification – the viewing of individual actions lens – and over time we have forgotten that other lenses are available’.14 By opening up the entire bag of lenses, from the most detailed close-up to the widest of wide angles, the big historian reveals a whole new set of themes and patterns invisible even to the world historian. Big history provides the framework, not just for an understanding of the evolution of different human communities, but of the entire human species, and of the biospheric, geological, and cosmic stage upon which the story of this species has played out. Other disciplines beyond history have also become increasingly aware of the potential for big history to take interdisciplinarity to a much higher level of intensity, as many of the chapters in this Companion demonstrate. As Fred Spier has noted, even today ‘interdisciplinary studies in the form of theoretically integrated approaches are still rare’.15 Big history is thus at the forefront of efforts to offer an interdisciplinary approach to the past, present, and future as part of a genuine attempt to unify all human knowledge, something that Edward O. Wilson described as ‘consilience’.16 This is now recognized by a range of physical, life, and social scientists, many of who have embraced the approach and presented big history papers from their own disciplinary perspectives at academic conferences, and have also contributed disciplinary-related chapters to this volume. There are increasingly insistent voices from within many these disciplines arguing that some form of consilience is desperately needed in the twenty-first century, because the problems facing humans and planet Earth are much too big to be even considered by narrow disciplinary thinking, let alone solved. One example of this from within the field of world history was the argument made in a keynote address 7

Craig Benjamin et al.

titled ‘The Future of World History’, presented at the 2009 World History Association Conference by Alfred W. Crosby, Professor Emeritus of History, Geography, and American Studies at the University of Texas at Austin.17 Crosby argued that big history is uniquely positioned to not only address potentially catastrophic issues like global warming, but also to offer some reassurance that humanity might just be equal to the challenges posed by them. Crosby noted that world history’s greatest contribution to the discipline has been to emphasize the common historical experience of life on earth, particularly the processes that have affected human adaptation to environmental change. Big history not only addresses these same issues, Crosby suggested, but also offers historical evidence that humanity might perhaps be equal to the contemporary challenges posed by them. Between 100,000 and 10,000 years ago, he pointed out, in the face of wildly fluctuating global climates, human beings were able to undertake extraordinary migrations and settle every continent on the planet with the exception of Antarctica. Within the last 15,000 years, human migrants into the American world zone adapted to the ice-bound North American continent, the deserts and forests of Central and upper South America, and then found ways of re-adapting to the ice-bound fiords of Patagonia, all within a couple of thousand years. What, Crosby asked, was the challenge of global warming compared to this history of human adaptation? He concluded that those historians most aware of this long history of human migration and adaptation have not only acquired a deeper understanding of the human past, but perhaps also of its future, and this might ultimately prove to be big history’s most important contribution to human knowledge. A number of publishers, media outlets, and technology innovators have come to similar conclusions over recent years, given the degree of interest they have shown in big history. David Christian’s 48-lecture big history course for The Great Courses has been one of the best sellers in that company’s Great Courses series, and is the medium through which Bill Gates discovered big history. The Great Courses then commissioned a follow-up course entitled The Big History of Civilization, written and presented by Craig Benjamin. David Christian was also featured in a 2011 TED talk where he was introduced by Bill Gates personally; and he was an invited delegate to the February 2012 World Economic Forum in Davos. Both Christian and Fred Spier were also invitees to a 2011 international environmental symposium convened by Al Gore; and members of the Board of the International Big History Association were featured presenters at the 2012 Global Futures 2045 Conference held in Moscow in March 2012. The History Channel has also developed an interest in big history, initially producing a program titled ‘A History of the World in Two Hours’, which featured several well-known big historians, and following this up with a 17-episode series on big history. In addition, the media giant YouTube currently hosts two seasons of Crash Course Big History written by David Baker, and hosted by John Green, Hank Green, and Emily Graslie that has garnered over 10 million views. Publishers have also been showing an increasing interest in big history. Important books published in the past few years include Berkshire’s brief This Fleeting World by David Christian (second edition 2013)18; Wiley-Blackwell’s 2010 book Big History and the Future of Humanity by Fred Spier19; McGraw-Hill’s 2014 publication of the 8

The Routledge Companion to Big History

first ever text book in the field, Big History: From Nothing to Everything authored by David Christian, Cynthia Brown, and Craig Benjamin20; and Cynthia Brown’s 2007 publication Big History: From the Big Bang to the Present. Further relatively recent developments in the field of big history have come through innovative approaches to the teaching of big history, and the establishment of the International Big History Association in 2011.

Recent developments in big history pedagogy Big history continues to expand as a coherent undergraduate course at universities around the world. One example of the increasing relevance of big history at university level has been the establishment of big history as a mandatory general education course for incoming first-year students at the Dominican University of California, because in many ways big history offers the quintessential introduction to liberal ­education. Big history is also being used more frequently to study more specialized subjects, for instance, in academic little big history projects and local big history courses. At the same time, big history has established itself as a viable course of instruction in secondary schools, thanks largely to the work of the Bill Gates – ­supported BHP. Most recently, Macquarie University in Sydney, Australia, released Big History School that provides a big history curriculum for K-12.

The Big History Project The Big History Project is a Seattle-based organization that, in collaboration with Macquarie University in Australia and contributions from big historians around the world, has brought high-quality big history instruction to high school students. The course is designed primarily as a social studies/history course for 9th or 10th grade students, although it is taken by students of various ages. The aims of this high school initiative are the same as those of big history educators at every level – to link different areas of knowledge into one unified story that provides students with a deeper awareness of our past, and thus hopefully better prepares them to help shape the future of our fragile planet in a positive way.21 The BHP has made all course content and curriculum freely available online. By using a web-based instructional model, the developers ensure that content remains up-to-date; that the online platform relieves schools of the need for expensive textbooks; and BHP managers offer teachers all manner of guidance in utilizing media-rich materials to enhance student learning. BHP content has been devised by course designers in partnership with many leading big historians and educators, including several members of the IBHA Board. In December 2018, Macquarie University released its preview launch of the next generation big history curriculum for primary and secondary schools, for students K-12. Big History Junior is a project-based course that teaches the grand narrative to primary school students. Big History Core deepens this knowledge for middle school students. Big History Senior focuses on what the trends of big history indicate about the many perils facing humanity in the Anthropocene. Different versions of Big History School are being tailor-made to the national curriculum requirements of 9

Craig Benjamin et al.

every country in the world in order to make it easier for teachers to adopt the course around the world. Big History School is also being translated into ten major global languages. The goal is to double the number of schools taking big history within the next five years.22

Little and local big histories Big history has been used at various universities to study more specialized subjects for over a decade now. A few examples of such efforts, that have been dubbed little big histories, can be found in Part 3 of this book. More recently, such a little big history approach has been applied to local histories, for instance, at the Universities of Amsterdam, Milan and Oviedo. Doing so allows students to connect the local history they identify with most to the big history that all people in the world share and that can stimulate the development of a global identity. It aims to teach students to appreciate what has made certain people and places special while also reflecting on what they all have in common, and to prepare them for working in our current world, as global citizens that can also relate to what is important on the local level.

First year big history experience at the Dominican University of California Since 2010, instructors at the Dominican University of California have been helping first-year students explore the origins and evolution of the universe, Earth, life, and humanity in a variety of interactive big history course. Not only is this the first institution anywhere in the world to make big history a required course, but Dominican’s program is also the first to be offered as a cohesive sequence rather than a single course. It uses big history as the contextual framework to further the study of the arts, humanities, and social sciences. As original Director of General Education and the First Year Experience program Mojgan Behmand once explained it: “We have expanded on the big history concept in order to develop a course sequence that emphasizes the students’ critical and creative thinking and helps students think about the future of humanity as a species on our planet.”23 Dominican’s path to big history began in 2008 when the University overhauled its general education curriculum and began looking for innovative first-year seminars that would be both foundational and global in focus and content. Eminent Big Historian and Dominican faculty member, the late Cynthia Stokes Brown, author of Big History: From the Big Bang to the Present,24 suggested that big history could provide precisely the sort of global, foundational, and interdisciplinary experience Dominican was looking for, and in 2009 the faculty voted to adopt big history as the focus of the seminars. The course sequence was launched in 2010, and big history courses have been taught every semester since then by a variety of Dominican faculty from across the disciplines, including art, business, English, history, mathematics, music, occupational therapy, political science, psychology, religion, and humanities. Dominican continues to show the way for a deeper integration of big history into general education in the future. To quote Behmand again, “Students study the past, 10

The Routledge Companion to Big History

make projections for the future, and then enter an integrative second-year course sequence that addresses essential learning outcomes – specifically personal and social responsibility – through … engaging with ‘real world’ problems. We are providing students with the forum to develop knowledge and practical skills needed to act to transform issues of vital importance for our shared future, and to do so in a manner that is realistic but positively empowered.”25 It would be difficult to find a clearer description of the pedagogical aims and potentials of big history at all levels of education.

Establishment of the IBHA Another important development in the spread and long-term sustainability of the field has been the establishment of the IBHA. This grew out a meeting in August 2010 at the Geological Observatory at Coldigioco in Italy. The Big Historians who met at Coldigioco were David Christian of Macquarie University in Sydney (­Australia), Fred Spier of the University of Amsterdam (Netherlands), Walter Alvarez of the University of California at Berkeley (USA), Craig Benjamin of Grand Valley State University in Michigan (USA), Cynthia Brown of Dominican University in California (USA), Lowell Gustafson of Villanova University in Pennsylvania (USA), and Barry Rodrigue of the University of Southern Maine (USA). Also in attendance were Pamela Benjamin, Gina Giandomenico, and Penelope Markle who constituted an advisory committee; representatives of the BHP and the Microsoft ChronoZoom Project; and graduate geology students from the University of California, Berkeley. The big historians in attendance formally constituted themselves as a provisional executive committee and voted in favor of the following initiatives: (1) (2) (3) (4)

To establish the International Big History Association. To establish an International Big History Journal. To establish an International Big History Website. To hold the first international Big History conference in 2012 (convened at Grand Valley University, Michigan, USA).

The embryonic IBHA was fortunate to receive a start-up grant from Microsoft External Research, which facilitated the convening of the first formal meeting of the IBHA Board at Grand Valley State University (GVSU) in January 2011. At this meeting the By-Laws and Articles of Association were constituted and accepted; and an offer from the administration of GVSU to provide an office and assistance for the IBHA to be based at GVSU was gratefully accepted. The IBHA was headquartered at GVSU from 2011 to 2018. The IBHA has successfully held four conferences: at GVSU in 2012; The Dominican University of California in 2014; the University of Amsterdam in 2016; and at Villanova University in Pennsylvania in 2018. The IBHA has published a formal mission statement: The International Big History Association (IBHA) exists to promote the unified and interdisciplinary study and teaching of the history of Cosmos, Earth, Life, and Humanity. 11

Craig Benjamin et al.

The organization of chapters within this volume With this background in mind, the editors are delighted to present The Routledge ­Companion to Big History, which, following this Introduction and David Christian’s ­chapter ‘What is big history’, is divided into five thematic parts. The themes of these parts represent current and evolving areas of great interest to the big history community. Individual ­chapters in the thematic parts engage with the relevant secondary literature and develop the part’s theme in the context of each author’s own particular interests.

Introductory chapters Introduction to The Routledge Companion to Big History: Craig Benjamin, Esther Quaedackers, David Baker (1) What is big history? – David Christian

Part I: Big history and science In this part, big historians with an interest in the physical and life sciences discuss the application of the big history approach to research and teaching within their disciplines of astrophysics, astronomy, and evolution. This part might be useful to other physical and life scientists, and students and instructors, who want ideas on how to incorporate big history perspectives and methodologies into their own research and teaching. (2) Big history and the study of time: the underlying temporalities of big history: Barry Wood (3) Big history and astronomy – space is big: the Fermi paradox: its relevance to big history and the human race: Jonathan Markley (4) Big history and macro-evolution: evolutionary principles and mechanisms at biological and social phases of the big history: Leonid E. Grinin, Andrey Korotayev and Alexander Markov

Part II: Big history, social science and the humanities In Part II, leading big historians from the social sciences and humanities discuss the application of the big history approach to research and teaching within their disciplines, including anthropology, archaeology, philosophy, political science, historiography, post-modernist critical theory, and moral and religious studies. This part might be useful to other social science and humanities students and instructors who want ideas on how to incorporate big history perspectives and methodologies into their own research and teaching. (5) Big history and anthropology: our place in the multiverse: anthropology, ­civilization and big history: Barry H. Rodrigue (6) Big history and archaeology: archaeology is big history: Brian Fagan (7) Big history and philosophy: philosophical foundations of big history: why big history makes sense: Armando Menéndez Viso 12

The Routledge Companion to Big History

(8) Big history and political science: science, the deep past, and the political: ­Lowell Gustafson (9) Big history and historiography: deep tides and swirling foam: the influence of macro-historical trends on micro-historical events: David Baker (10) Big history and critical theory: science, history and why theory matters: David Blanks (11) Big history, morality and religion: Cynthia Stokes Brown

Part III: Little big histories In this part, individual big historians and a team of interdisciplinary researchers contribute their own ‘little big histories’, chapters focused on some particular process, place, event, or idea that demonstrate how these phenomena take on new meaning when traced back through deep time and space. This is a growing area of interest within the field of big history at the moment and is proving to be a tremendously interesting approach for students and instructors at schools and universities all over the world, particularly those involved with the BHP. (12)  A case for little big histories: Esther Quaedackers (13)   The little big history of the Nalón River, Asturias, Spain: Olga GarcíaMoreno, Diego Álvarez-Laó, Miguel Arbizu, Eduardo Dopico, Eva GarcíaVázquez, ­Joaquín García Sansegundo, Montserrat Jiménez-Sánchez, Laura Miralles, Ícaro Obeso, Ángel Rodríguez-Rey, Marco de la Rasilla Vives, Luis Vicente Sánchez Fernández, Luis Rodríguez Terente, Luigi Toffolatti and Pablo Turrero (14)  Sketch of a little big history of Private E.E.  Benjamin and the Great War: Craig Benjamin

Part IV: Teaching big history In Part IV, three experienced big history educators contribute chapters on the teaching of big history at various educational levels in Korea, Australia, and the USA. The growth of big history education in the K-12 sector, as well as undergraduate and graduate college courses, demonstrates how relevant the big history approach is to students of all ages. (15) The Big History Project in Australia: Tracy Sullivan (16)  Big history teaching in Korea: Seohyung Kim (17) Crossing thresholds: using big history to meet challenges in teaching and learning in the United States: Robert B. Bain

Part V: Big history and the future Big history has always provided an excellent fit with futurology in its various guises. Examining the past and present on the largest possible temporal and spatial scales means that big historians are very well equipped to say something useful 13

Craig Benjamin et al.

and important about the future. In this part, leading Russian, European, ­American, and Australian big historians contribute chapters about the future on various scales. (18)  Big history and the future of technology: Leonid E. Grinin and Anton L. Grinin (19)  Big history and the Singularity: Akop P. Nazaretyan Underground metro systems: a durable geological proxy of rapid urban (20)  ­population growth and energy consumption during the ­Anthropocene: Mark Williams, Matt Edgeworth, Jan Zalasiewicz, Colin N. Waters, Will Steffen, ­Alexander P. Wolfe, Nicholas J. Minter, Alejandro Cearreta, Agnieszka Gałuszka, Peter Haff, John McNeill, Andrew Revkin, Daniel deB. Richter, Simon Price, and Colin Summerhayes (21)  The coming energy transition: what comes after fossil-fueled civilization? Joseph Vorros

Conclusion This brief introduction to the field of big history, and to this volume, suggests that big history is well positioned for continued expansion as a research and teaching field and has the potential to make a significant contribution to education and public discourse at all levels. Those of us deeply committed to the field are hopeful of the eventual widespread introduction of big history into high school and university programs around the world. We are particularly convinced of the genuine pedagogical and societal gains to be made by introducing big history as the cornerstone of university general education programs. The evidence collected over the past 25 years of teaching has shown that by exposing the leaders of the future to big history, students learn to use the tools of interdisciplinarity and critical thinking on a macro scale to conceptualize and think about real solutions to the great problems of our times. The future of humanity might well depend on facilitating the acquisition of these skills and perspectives through widespread big history education and research. January 2019

Notes 1 Diodorus Siculus, The Library of History, in C.H. Oldfather, Diodorus of Siculus Vol. 1, Cambridge, MA: Harvard University Press, 1933. 2 Sima Qian, Shi Ji (trans. B. Watson), Records of the Grand Historian by Sima Qian – Han Dynasty II, New York: Columbia University Press, 1993. 3 See, for example, Paulus Orosius, Seven Books of History against the Pagans, Washington, DC: Catholic University of America Press, 1964. 4 Ibn Khaldun, The Muqadimah, Princeton, NJ: Princeton University Press, 1958. 5 See, for example, Voltaire, The Philosophy of History, Brookline, MA: Vision Press, 1965; and Nicolas de Condorcet, ‘The Progress of the Human Mind’, in Patrick Gardiner, ed., Theories of History, New York: Free Press, 1959. 6 George Hegel, Philosophy of History, New York: Dover Publications, 1956, 19. 7 Karl Marx, ‘The German Ideology’, in K. Marx and F. Engels, eds., Collected Works, V, London, 1976. 14

The Routledge Companion to Big History

8 Alexander von Humbolt, Cosmos, Volume 1 (Foundations of natural History), Baltimore, MD: Johns Hopkins University Press, 1845. 9 H.G. Wells, The Outline of History: Being a Plain History of Life and Mankind, New York: Garden City Publishing Company, 1920. 10 Stephen Hawking, A Brief History of Time, New York: Bantam Dell Publishing Group, 1988. 11 David Christian, ‘The Evolutionary Epic and the Chronometric revolution’, in C. Genet et al., eds., The Evolutionary Epic. Science’s Story and Humanity’s Response, Santa Margarita: Collins Foundation Press, 2009, 91–99. 12 Fred Spier, Big History and the Future of Humanity, Chichester: Wiley-Blackwell, 2010. 13 Bruce Mazlish, ‘Terms’, in M. Hughes-Warrington, ed., World Histories, London and New York: Palgrave/MacMillan, 2005, 18–43. 14 M.H. Hughes-Warrington (2005). Big History. Social Evolution & History. Vol. 4, no. 1 (Spring 2005), ed. Graeme Donald Snooks, 7–21. 15 Fred Spier, ‘Big History: The Emergence of an Interdisciplinary Science?’, Interdisciplinary Science Reviews, 2008, Vol. 33, no. 2 © 2008 Institute of Materials, Minerals and Mining. Published by Maney on behalf of the Institute. 16 E.O. Wilson, Consilience, New York: Alfred A. Knopf Inc., 1998. 17 Alfred W. Crosby, plenary panel address, The Future of World History, WHA Conference, Salem MA, June 25–28, 2009. 18 David Christian, This Fleeting World, Great Barrington, NH: Berkshire Publishing, 2008. 19 Spier, Big History and the Future of Humanity, 2010. 2 0 David Christian, Cynthia Brown and Craig Benjamin, Big History: From Nothing to Everything, New York: McGraw-Hill, custom edition published in 2010. 21 Big History Project website: www.bighistoryproject.com/. 22 See the Macquarie Big History School website: bighistoryschool.org. 23 Mojgan Behmand, Dominican University of California First Year Big History Experience: www.dominican.edu/dominicannews/first-year-experience-focuses-on-big-history. 24 Cynthia Stokes Brown, Big History: From the Big Bang to the Present, New York: The New Press, 2007. 25 Mojgan Behmand, Dominican University of California First Year Big History Experience.

15

1 WHAT IS BIG HISTORY?1 David Christian

Big history is based on a very ancient idea, which sixteenth century French historian Jean Bodin captured nicely: “As they err who study the maps of regions before they have learned accurately the relation of the whole universe and the separate parts of it to each other and to the whole, so they are not less mistaken who think they can understand particular histories before they have judged the order and sequence of universal history and of all times, set forth as it were in a table.”2 Big history represents an attempt at what E.O. Wilson has called “consilience,” a return to the goal of a unified understanding of reality, in place of the fragmented visions that dominate modern education and scholarship.3 Though it may seem new, the goal of consilience is very old. And even in its modern forms, big history has been around for at least a quarter of a century. So the publication of the first issue of the Journal of Big History provides the ideal opportunity for a stock take. This article is a personal account of the field. It sees big history as the modern form of an ancient project. I am a historian by training, so my account focuses on the relationship of big history to the discipline of history. It reflects the perspective of a historian trained in the English-speaking world, and it focuses on big history’s relationship to Anglophone historical scholarship. But not just to Anglophone historical scholarship, because the debates I discuss had their counterparts and echoes in many other traditions of historical scholarship. Nor do I focus just on historical scholarship as it is normally understood within the academy, because big history sees human history as part of a much larger past that includes the pasts studied by biologists, paleontologists, geologists, and cosmologists. By linking different perspectives and scales, and many different scholarly disciplines, all of which try to understand the deep roots of today’s world, big history can transform our understanding of “history.” However, to fully capture the richness and range of this vibrant new field of research, scholarship, and teaching, we will eventually need the perspectives of big historians trained in many other disciplines. I hope this essay may encourage such scholars to offer their distinctive perspectives on big history. 16

What is big history?

The evolution of historical scholarship in the twentieth century Historians will recognize that my title comes from a classic essay on history, studied by most Anglophone history graduates. It was written in 1961 by E.H. Carr, an English historian of the Soviet Union. Carr’s book began as a lecture series given at ­Cambridge in 1961 in honor of George Macauley Trevelyan, a historian who, unlike Carr, saw history as a literary discipline, and quite distinct from the sciences. As a historian of Russia and the Soviet Union, Carr took seriously the Marxist insistence that history should be regarded as a branch of science, and that idea influenced my own thinking about history as I, too, entered the field of Russian history as a graduate student in the early 1970s. In “What is History?” Carr tracks the evolution of the history discipline in ­England in the early twentieth century. At one level, his story is of a sustained trend away from the confident realism, positivism, and even universalism of many nineteenth century historical thinkers, towards increasing fragmentation and skepticism. He begins by citing Lord Acton’s confident vision of historical scholarship from the 1890s, as ­Acton presided over the first edition of the Cambridge Modern History. Acton saw the Cambridge Modern History as “a unique opportunity of recording, … the fullness of the knowledge which the nineteenth century is about to bequeath….” He added: “Ultimate history we cannot have in this generation [but] … all information is now within reach, and every problem has become capable of solution.”4 Acton’s view of history is confident, positivist, and optimistic, and it assumes that history is part of the larger project of increasing human knowledge in general. His vision of history is also broad. He assumed that historians should aim at some kind of “universal history,” though he seems to have understood that phrase to mean, not an early form of big history, but something closer to modern “world history” or “global history.” Acton defined universal history as “that which is distinct from the combined history of all countries.”5 In the early twentieth century, English historical scholarship underwent a profound transformation, and when Carr wrote, the discipline was more fractured and less sure of itself. These shifts were part of a sea-change that affected most scholarly disciplines, from the humanities to the natural sciences, as specialization and professionalization broke scholarship into ever-smaller compartments, each offering its own pin-hole view of the world. Specialization proved a powerful research strategy, but it was achieved by severing ancient links among fields of knowledge, leaving them increasingly isolated from each other. The idea of a single world of knowledge, whether united by religious cosmologies, such as that of Christianity, or by scientific scholarship—the vision that lay behind Alexander von Humboldt’s attempt to write a scientific universal history in his Kosmos— was abandoned.6 In humanities disciplines such as history, which lacked the sort of unifying paradigm ideas characteristic of the natural sciences in the era of ­Darwin, of Maxwell and of Einstein, specialization also undermined Acton’s confident epistemological realism.7 Carr captures some of these changes by citing the introduction to the second ­edition of the Cambridge Modern History, written by George Clark in 1957, more than 17

David Christian

half a century after Acton’s confident pronouncements. After citing Acton’s hopes for an “ultimate history,” Clark writes: Historians of a later generation do not look forward to any such prospect. They expect their work to be superseded again and again. …The exploration seems to be endless, and some impatient scholars take refuge in skepticism, or at least in the doctrine that, since all historical judgements involve persons and points of view, one is as good as another and there is no ‘objective’ historical truth.8 The loss of confidence in a realist or naturalist epistemology in disciplines, such as history, widened the gulf between the “two cultures” of the sciences and humanities that so worried C.P. Snow in a famous lecture delivered in 1959.9 The gulf was particularly wide in the English-speaking world, because English, unlike most other scholarly languages, confined the word, “science,” to the natural sciences. In English, the very idea of “historical science” began to seem absurd. By Carr’s time, historical scholarship had lost confidence both in the “scientific” nature of historical ­scholarship, and in the realist epistemology that still underpinned research in the natural sciences. Skepticism and intellectual fragmentation sapped confidence in the value of ­historical research, and undermined the ancient hope that history could empower us by helping us better understand the present. As historians became increasingly isolated from other disciplines and even from each other, they were left with increasingly fragmented visions of the past, and of the nature and goals of history. This growing sense of fragmentation was the scholarly counterpart of what Durkheim called anomie, the loss of a sense of coherence and meaning, an idea that Carr himself glosses in a footnote as “the condition of the individual isolated from … society.”10 Scholarly anomie arose from the growing isolation of scholars both from each other and from a unified world of knowledge.The one force that partially mitigated the growing sense of scholarly isolation was nationalism. Though tribal by their very nature, national histories, which had flourished since the nineteenth century, provided some sense of cohesion for historians working within national historiographical traditions. Carr’s own position falls between the robust scientific realism of Acton and the hesitant relativism of Clark. He explores brilliantly the complex dialectic between history as truth and history as stories we tell about the past. He takes truth and science seriously, because he believes that history, like science, and like truth in general, has a purpose: it can empower us. It empowers us by improving our understanding of the present, and it does that by mapping the present on to the past: “The function of the historian is neither to love the past nor to emancipate himself from the past, but to master and understand it as the key to the understanding of the present.”11 It followed that the maps of the past created by historians had to be good maps. Like good science, they had to give us a better grip on the real world. So Carr, like Marx, was a philosophical realist and saw no fundamental chasm between the humanities and the natural sciences. Scientists, social scientists, and historians are all engaged in different branches of the same study: the study of man and his environment, of the effects of 18

What is big history?

man on his environment and of his environment on man. The object of the study is the same: to increase man’s understanding of, and mastery over, his environment.12 On the other hand, Carr understood more clearly than Acton that the past is not simply waiting to be discovered, “like fish on a fishmonger’s slab.”13 History consists of stories about the past constructed by historians, and how we construct those stories changes as our world and our purposes change. We need empirical rigor to get at the truth about the past, but when telling stories about the past we will need the skills of storytellers, including what Carr calls “imaginative understanding,” the ability to understand and empathize with those who lived in the past.14 In this, Carr was influenced by one of the great English philosophers of history, R.G. Collingwood, though he warned that Collingwood’s emphasis on the empathetic role of the historian, if taken too far, could lead to extreme skepticism.15 Particularly, influential on Carr’s thinking was Marx’s dialectical balance between science and activism. Marx insisted that there is an objective past. But making something of that past is a creative task, and how we approach it depends on who we are and the particular present in which we write and study.This is the dialectic that Marx described in a famous passage from the “18th Brumaire of Louis Napoleon.” Men make their own history, but they do not make it just as they please; they do not make it under circumstances chosen by themselves, but under circumstances directly found, given and transmitted from the past. The tradition of all the dead generations weighs like a nightmare on the brain of the living.16 Historians, too,“make their own history,” but they do so “under circumstances d­ irectly found, given and transmitted from the past.” What they make of the past depends on the time and place in which they write. But the stories they construct about the past may, in their turn, influence the pasts studied by future historians. As an activist, Marx understood well that how we describe the past matters, because our accounts may shape the future. Indeed, he hoped that his own account of the evolution of capitalism would have a profound impact on the future, as indeed, it did. Like Marx, then, Carr understood the complex and delicate balance between history as truth and history as story. History is, Carr wrote, in a passage familiar to many a graduate student in history: “a continuous process of interaction between the historian and [the] facts, an unending dialogue between the present and the past.”17 Like memory, history does not recall the past; it re-creates it. But what past? Carr was even more committed than Acton to broadening the scope of historical research. He was, after all, a historian of Russia, and keen to demonstrate the significance of histories that had been neglected by English-­speaking historians. As an admirer of Joseph Needham, he also insisted on the importance of Chinese history and the histories of many other parts of the world beyond Europe. But, though Carr’s past is broad, it is not deep. He shows little interest in human prehistory or in the histories of the biosphere and the universe. And that is surprising, given his interest in Marx, who saw history as part of a knowledge continuum that included all the sciences. Indeed, Marx, like von Humboldt, was a big historian 19

David Christian

before his time. But Carr wrote in an era of scholarly fragmentation, and the idea of universal history was not on his radar, or on the radar of any English-language historians of his generation. Strangely, though, it was on the radar of historians in the Soviet Union, the country whose history Carr wrote most about, because the Soviet Union’s Marxist heritage ensured that the idea of “universal” or “general” history never entirely lost its inclusive Marxist sense. That is one reason why, today, there is a flourishing Russian school of big history research led by scholars such as Andrey Korotayev and Leonid Grinin. In 2001, David Cannadine edited a collection of essays called What is History Now? based on a conference held to mark the 40th anniversary of Carr’s book.18 Much had changed since Carr wrote.The history discipline had become even more f­ragmented, in both content and epistemology, and even less sure of itself. The universalist ­vision of Marx or von Humboldt or H.G.Wells seemed to have vanished completely, surviving only in the cut-down version of national histories. Many of the changes evident in Cannadine’s collection reflect the post-war proliferation of universities, university students, historians, and historical sub-disciplines. This was a worldwide phenomenon, so similar trends can be found, with variations, in many different historiographical traditions. Since Cannadine’s book was no longer about a single history discipline, it was appropriate that it had multiple authors. More historians and more students seemed to mean more diverse ideas on the content, the meaning, and the purpose of historical scholarship. Each chapter is about a different type of history, so there are ­chapters called: “What is Social History now?” “What is intellectual History Now?” and “What is Cultural History Now?” The absence of “What is Women’s History Now?” or “What is Environmental History Now?” is striking, though Cannadine insists that his book reflects just a small number of the sub-disciplines into which history was then divided. Fragmentation was accompanied by increasing skepticism about the objectivity and the scientific nature of the discipline.True, most historians continued to approach the details of their research with a robust, realist empiricism, so much so, that many caricatured the discipline as just a catalogue of facts. But, as the circle of questions widened, the confidence of historians seemed to dwindle, and few were comfortable with the idea of historical scholarship as part of a larger system of knowledge or meaning. Historians became increasingly isolated from other disciplines (the decline of economic history is a striking example of this process), and even from each other, and any consensus about the nature and goals of history seemed to evaporate. In an introductory essay to Cannadine’s book, Richard Evans noted the increasing focus in a postmodernist era on the creative and subjective role of the historian and on the historian’s role as storyteller. This approach had been epitomized in Hayden White’s 1973 classic, Metahistory:The Historical Imagination in Nineteenth Century Europe, which focused almost entirely on the literary aspects of historical scholarship, rather than on the truth claims it made. Historical scholarship seemed to have splintered into multiple, incommensurable, stories about the past, each representing a particular perspective, and none confident about its claims on historical truth. Historians seemed to have taken on the deep skepticism towards grand narratives or meta-narratives that Jean-François Lyotard saw as a defining feature of postmodern thought.19 20

What is big history?

And yet, though the tremors barely registered on the seismograph of Cannadine’s volume, by the year 2000, the idea of a new form of universal history was already rattling the margins of historical scholarship. World history was flourishing in the USA, had a well-established scholarly organization and a successful journal (The Journal of World History), and was taught in an increasing number of universities and schools. But several scholars now ventured far beyond world history. They began to explore the possibility of a truly universal history that would embrace the whole of the past, including the pasts of the biosphere and the entire universe. By 2001, I had been teaching big history for 12 years, but I was just one member of a small but vigorous community of scholars moving in the same direction. Eric Chaisson had been teaching astronomer’s versions of big history for more than 20 years, and big history was being taught in Amsterdam by Fred Spier and Johan Goudsblom, in Dallas by John Mears, in San Rafael by Cynthia Stokes Brown, in Melbourne by Tom Griffiths and Graeme Davidson, and elsewhere. Big history snuck up on a history discipline that was looking in the opposite direction. Today, 15 years after Cannadine’s volume, big history remains marginal, but it is beginning to shake up the history discipline.20 There is an emerging scholarly ­literature that proves big history can be written with rigor and precision and can yield new, sometimes transformative, insights into the past.21 Big history is being taught successfully in several universities, mostly in the English-speaking world, and even those history departments that do not teach it often include discussions of big history in their historiography seminars. There are several MOOCs (Massive Open Online Courses) on big history. There is a scholarly association (the IBHA [International Big History Association]), which has held three major conferences, and now there is a journal of big history. Macquarie University has established a Big History Institute, which has organized two research conferences. Big history is even being taught in hundreds of high schools, mostly in the USA and Australia, through the Big History Project, a free, online high school syllabus in big history, launched in 2011 and funded by Bill Gates. In 2018, M ­ acquarie ­University’s Big History Institute released its own free online courses in big history. These were adapted for different educational environments, and also included, for the first time, a course for primary students and one for students about to enter University. What seemed just decades ago an archaic, unrealistic, and perverse approach to historical scholarship is now beginning to look like a powerful, rigorous, and even transformative form of modern scholarship, which can re-connect historical scholarship and teaching to other disciplines in both the humanities and the sciences.

Why the return to universal history? What happened? Some of the crucial changes occurred within the history discipline itself. There had always been a few scholars, such as H.G. Wells or Arnold Toynbee, who kept alive the vision of a more capacious understanding of the past. But specialist research also laid the foundations for a broader view of the past, by generating a colossal amount of new historical scholarship and tackling subjects and regions and epochs that had been ignored 21

David Christian

by earlier generations of historians. Felipe Fernandez-Armesto, a world historian with extraordinarily broad interests, puts it nicely in a chapter in Cannadine’s volume: Historians dig ever deeper, narrower furrows in ever more desiccated soil until the furrows collapse and they are buried under their own aridity.Yet on the other hand, whenever one climbs out of one’s furrow, there is now so much more of the field to survey, so much enriching new work, which can change one’s perspective or broaden one’s framework of comparison.22 However, many of the changes that allowed a return to universal history occurred beyond the history discipline, and particularly within the natural sciences, which had always been more friendly than the humanities to the idea of consilience.23 The quantum physicist, Erwin Schrödinger, had already anticipated new forms of scholarly unification in a book he wrote just after World War II on the nature of life. We have inherited from our forefathers the keen longing for unified, all-­ embracing knowledge. The very name given to the highest institutions of learning reminds us that from antiquity and throughout many centuries the universal aspect has been the only one to be given full credit. … We feel clearly that we are only now beginning to acquire reliable material for welding together the sum total of all that is known into a whole; …24 In the natural sciences, as in the humanities, specialized scholarship over many decades yielded a huge bounty of new information and ideas. Equally important was the emergence of new unifying paradigm ideas. The most important were Big Bang cosmology, plate tectonics, and the modern Darwinian synthesis. The new paradigms were barely visible when Carr wrote. DNA had been discovered in Carr’s own University of Cambridge, in 1953, but the full significance of that discovery would only become apparent over the next decade or two. The discoveries that clinched plate tectonics and Big Bang cosmology still lay a few years in the future. By 1970, though, the new paradigms were already encouraging hopes of a new unification of knowledge, at least in the natural sciences. Some scientists began to talk of “Grand Unified Theories.” Particularly striking is the fact that the new scientific paradigms were historical in nature. Gone was the static universe of Newton, replaced by a universe that operated according to historical and evolutionary rules. E.H. Carr was aware of the “historical turn” in the natural sciences, and its significance for history, though his insights would be ignored by most historians over the next 50 years or so. Science, he wrote: had undergone a profound revolution….What Lyell did for geology and Darwin for biology has now been done for astronomy, which has become a science of how the universe came to be what it is…. The historian has some excuse for feeling himself more at home in the world of science today than he could have done a hundred years ago.25 In the English-speaking world, Big Bang cosmology encouraged astronomers such as Carl Sagan to recount the history of the universe, while plate tectonics encouraged 22

What is big history?

geologists such as Preston Cloud to write new histories of planet Earth.26 It turned out that many natural scientists were in the same messy business as historians—that of trying to reconstruct a vanished past from the random clues it had left to the present. The historical turn in the natural sciences brought the methods of scientists closer to those of historians. Controlled experiments on the origins of life on Earth or the Russian Revolution were out of the question. Instead, it turned out that many scientific disciplines faced the same methodological challenge as historians: that of collecting as many clues to the past as they could—from ancient starlight, to zircon crystals, to fossil trilobites—and using them to reconstruct plausible and even meaningful accounts of the past. This was territory familiar to historians. The knockdown disproofs favored by Karl Popper were rarely available, and other, fuzzier, skills familiar to historians, such as pattern-recognition or hunches bases on prolonged familiarity with a given field, acquired increasing salience in the natural sciences.27 Particularly important for the emergence of modern forms of universal history was the development of radiometric dating techniques that could provide a firm chronological skeleton for histories of the deep past.28 When H.G. Wells attempted a universal history just after World War I, the early parts of his story sagged because, as Wells admitted, all his absolute dates depended on written records, so he could provide none before the First Olympiad (776 BCE).29 Nineteenth century geologists had learned how to construct relative chronologies by studying the layering of ancient rocks, but none could tell when the Cambrian explosion occurred or when Earth formed. This all changed with the emergence of radiometric dating techniques in the 1950s. In 1953, Claire Paterson used the half-life of uranium in meteorites to determine that Earth is 4.56 billion years old. His date stands to this day. When Carr wrote in 1961, radiometric dating was just beginning to transform the thinking of ­archaeologists and pre-historians. In 1962, at Kenniff Cave in South Queensland, John ­Mulvaney used radiometric techniques to show that humans had lived in ­Australia since before the end of the last ice age, and over the next few decades, the earliest dates for human settlement in Australia would be pushed back to between 50,000 and perhaps 60,000 years.30 As Colin Renfrew writes: … the development of radiometric dating methods, … allowed the construction of a chronology for prehistory in every part of the world. It was, moreover, a chronology free of any assumptions about cultural developments or relationships, and it could be applied as well to nonliterate societies as to those with written records.To be prehistoric no longer meant to be a historic in a chronological sense.31 Eventually, radiometric and other dating techniques made it possible to construct rigorous chronologies reaching back to the origins of the universe. For the first time, it is now possible to tell a universal history based on a robust universal chronology. Some of these changes did just register in David Cannadine’s collection of essays. In the last chapter of that book, Felipe Fernandez-Armesto argued that history had widened its scope, specialization by specialization, and now needed to embrace the natural sciences: “history can no longer remain encamped in one of ‘two cultures’. Human beings are obviously part of the animal continuum.”32 In 1998, the great 23

David Christian

world historian,William H. McNeill, argued that historians needed to embed the history of humanity within the history of the biosphere and even the universe as a whole: Human beings, it appears, do indeed belong in the universe and share its unstable, evolving character. … [W]hat happens among human beings and what happens among the stars looks to be part of a grand, evolving story featuring spontaneous emergence of complexity that generates new sorts of behavior at every level of organization from the minutest quarks and leptons to the galaxies, from long carbon chains to living organisms and the biosphere, and from the biosphere to the symbolic universes of meaning within which human beings live and labor, …33 In his last years, McNeill became increasingly interested in the idea of Big History, seeing it as a natural extension of his own broad vision of history. It was, as his son, John, has written: “the thing that excited him most (aside from grandchildren).”34

What is big history? So, what is big history? In the final part of this essay I would like to explore several, overlapping descriptions of what big history is and what it could be. These are personal thoughts, and some are speculative. But I hope they may interest even those who are less persuaded by them than I am. And I hope they may encourage a broad discussion about big history and its future. My thoughts are organized, loosely, along a spectrum running from the “truth” end of Carr’s dialectic of history towards the “storytelling” end. The goal of big history, like that of all good knowledge, is to empower us by helping us understand the world we live in. Big history empowers us by helping us understand our world. Like all forms of history, big history empowers us primarily by mapping the present onto the past, so as to help us better understand how today’s world came to be as it is.This claim about the purpose of history assumes a realist or naturalist understanding of knowledge. As evolved creatures, we interact with our surroundings with some degree of success, and that success presupposes that we (like all living organisms) can attain a limited but real understanding of our surroundings. Though aware of the limits to knowledge, big history, like science in general, resists extreme forms of skepticism or relativism. It builds on the same realist and naturalist foundations as good science, and has the same ultimate goal, of empowerment.

Big history is universal But if understanding the past can empower us, shouldn’t we try to understand the whole of the past? What distinguishes big history most decisively from other forms of historical scholarship is its attempt to understand the past as a whole. It aspires to a universal understanding of history. Big history is not hostile to specialist historical scholarship. On the contrary, it is utterly dependent on the rich scholarship of specialists. But it tries to link the findings of specialist scholarship into a larger unifying vision, just as millions of local maps can be connected to form a single 24

What is big history?

world map. These ambitious goals mean that big history swims against the tide of intellectual fragmentation that structured so much scholarship in the twentieth century. Big history aims at consilience, at what Alexander von Humboldt once called the “Mad Frenzy … of representing in a single work the whole material world.”35 Many interesting consequences flow from big history’s ambitious universalism. Big history recognizes no disciplinary barriers to historical knowledge. It presumes the existence of a whole range of historically oriented disciplines, all of them linked by the same goal: that of reconstructing how our world came to be as it is. Indeed, I often wonder if we may not see, sometime in the future, a re-arrangement of university campuses, so that, instead of putting the sciences at one end and the humanities at the other, you would find a zone devoted to “the historical sciences,” in which astronomers, geologists, evolutionary biologists, neuroscientists, and historians would all be working together. The universal aspirations of big history mean that it will embrace all areas of knowledge that have generated plausible, rigorous, evidence-based accounts of the past, and any discipline whose insights can illuminate the past. This means that, at present, it makes sense to draw a line between everything that happened just after the big bang—a past that can be reconstructed with oodles of evidence—and anything that preceded the big bang, territory where there is plenty of interesting speculation, but not, as yet, a taut, evidence-based story. This may change, of course, in which case, the big history story itself will expand to incorporate, perhaps, evidence for a multiverse or for string theory. Similar changes may occur in other parts of the big history story, as biologists probe the origins of life on Earth, or astronomers look for life around other star systems, or as neuroscientists and psychologists begin to get a grip on the “hard” problem of consciousness, or historians get a better understanding of the role of religion and science in human history at multiple scales.36 With these qualifications, big history aims at a comprehensive understanding of history, the intellectual equivalent of a world map of the past. Like a world map, the big history story can help us see not just the major nations and oceans of the past, but also the links and synergies that connect different scholarly continents, regions, and islands into a single knowledge world. The broad perspective of big history also encourages us to move among multiple scales, from those of the universe itself, to those of humans, to those of individual cells, within which millions of precisely calibrated reactions occur every second. Big history encourages us to connect the dots in time and space, to look for the synergies between disparate entities, disciplines, and scales. Russian and Russian-trained scholars such as Andrey Korotayev and Peter Turchin have been particularly active in the important task of looking for mathematical patterns in the evolution of complexity at multiple scales.37 By focusing on the ideas that link disciplines, big history can help us overcome the more extreme forms of skepticism characteristic of much twentieth century scholarship, particularly in the humanities. In Durkheim’s hands, the idea of “anomie” referred to the absence of a clear sense of place or meaning, a condition of intellectual homelessness in which the world itself made little sense and individuals could feel isolated enough to contemplate suicide. The extreme fragmentation of twentieth century scholarship allowed great intellectual progress, discipline by discipline. 25

David Christian

But it did so at the cost of isolating disciplines from each other, which limited the possibilities both for a larger, unifying vision, and for truth-checking between disciplines. ­Particularly in the humanities, intellectual isolation generated scholarly forms of anomie that sapped confidence in claims to generate meaning or to achieve a more general grasp of reality. The postmodernist skepticism shared by so many scholars in the humanities in the late twentieth century was a useful corrective to over-confident forms of positivism. But, when taken to extremes, it created a splintered sense of reality that could be profoundly dis-empowering, both intellectually and ethically. Some saw it as the scholarly equivalent of suicide. Big history returns, with due scientific modesty, to the ancient project of trying to assemble unified maps of reality. By removing the partitions between disciplines, big history can help re-establish a more balanced relationship between specialist scholarship and large, paradigm ideas.

Big history is collaborative and collective The big history story is being assembled, like a vast mosaic, using tiles from many different countries, epochs, and scholarly disciplines. All scholarship is collaborative. But the extraordinary range of big history puts collaboration at the heart of the new discipline. A rich and reliable big history story will not be the product of individual scholarly minds, but the joint creation of millions of minds. The extreme scholarly collaboration required to write big history should encourage a re-think of what we mean by expertise. Specialization encouraged the notion that, if you narrowed the field of enquiry enough, individual scholars could achieve total mastery of a field. They became experts. This view was always naïve because even the narrowest of experts drew on insights and paradigms from outside their fields of expertise. But the extraordinary breadth of big history means that, though it will build on the insights of experts, it will also require many other scholarly skills, not all of which are valued in today’s fragmented knowledge world. Big history requires, above all, an ability to grasp and then link scholarship from many different disciplines. It demands breadth as much as depth, and a sharp eye for unexpected synergies among disciplines. And it requires an ability to tune into the different intellectual frequencies of multiple disciplines. Big historians will have to be interdisciplinary translators, sensitive to subtle nuances in the way different disciplines use similar concepts, words, and methods. And they will also ask deep interdisciplinary questions. Are there ideas that work well across multiple disciplines, from cosmology to biology and history, ideas such as the “regimes” and “Goldilocks conditions” described by Fred Spier, or the “free energy density” rates that lie at the heart of Eric Chaisson’s work? Can the idea of entropy, which plays such a powerful role in physics, illuminate our understanding of human history? Can the atomic level molecular machines being explored today by nano-biologists suggest new ways of managing energy flows in today’s world?38 Are there universal mechanisms (perhaps some form of universal Darwinism?) that explain the appearance of increasingly complex entities despite the second law of thermodynamics? By focusing not just on the individual islands and continents of modern scholarship, but also on the many links between them, big history can provide a new framework for interdisciplinary thinking and research. Researchers familiar with big 26

What is big history?

history’s world map of the past will naturally seek out useful ideas and methods from beyond their own specialist disciplines. Transdisciplinary research will become particularly important as more and more problems, from climate change to the study of cancer or financial crises, begin to depend on findings and insights from multiple disciplines. Indeed, the very success of research within disciplines explains why more and more interesting and important problems now lie between disciplines. As interdisciplinary research becomes increasingly important, big history can offer a new model of scholarly expertise that demands breadth of knowledge and an alertness to unexpected interdisciplinary synergies. The young discipline of big history has also shown that intellectual collaboration is a distinctive feature of our species, Homo sapiens. Though many evolutionary features define us as a species, our technological creativity seems to have been clinched by the evolution of an exceptionally powerful form of language that allows us to exchange ideas and insights with such precision and in such volume that they can accumulate in the collective memory. We know of no other species in which learned knowledge accumulates across multiple generations so that later generations know, not just different things, but more things than earlier generations. And this difference has proved transformative. The accumulation of learned information by millions of individuals across multiple generations explains our increasing control over the resources and energy flows of the biosphere.This accelerating trend has shaped much of human history, and has culminated today in making us the single most powerful force for change in the biosphere. In my own work, I have described our unique capacity for sharing and accumulating information as “collective learning.” It has given us humans not only increasing control over flows of energy and resources through the environment, but also increasing insight into the world and the universe we inhabit. Modern science as well as modern religions and literatures are all the creations of millions of individuals, working within shared networks of knowledge. In just one century, the sphere of human mind, or the “Noösphere,” as Vernadsky called it, has become a planet-changing force.39 My personal conviction is that the idea of “collective learning” offers a paradigm idea that can frame our understanding of human history and of the distinctive nature of our own species. Human history is driven by collective learning just as the history of living organisms is driven by natural selection. If this idea is broadly correct, it illustrates the capacity of big history to clarify deep problems by helping us see them against an exceptionally broad background, as part of the “world map” of modern knowledge.

Big history is a story So far, I have discussed the nature of the truth-claims that can be made by big history, and its capacity to synergize collaborative, interdisciplinary research. But of course, big history also tells a story. It arises, as Carr wrote of all history, from “an unending dialogue between the present and the past.” Its two poles are the past as a whole and the historians who view that past from a particular vantage point in the present. Like history in general, big history is very much a product of the historians who are constructing the big history story. That means, of course, that big history is evolving and will evolve, like all stories, as it is told by different tellers, writing in different contexts and with different preoccupations. 27

David Christian

Big history is an origin story But because of its universalist ambitions, big history is not just another story about the past. Its universal ambitions mean that big history shares much with traditional origin stories.40 As far as we know, all human communities have tried to construct unified accounts of the origins of everything that surrounds us. This is the sense in which I will use the idea of “origin stories.” Origin stories attempt to hold together and pass on all that is known in a given community about how our world came to be as it is. They are extraordinarily powerful if they are believed, if they ring true to those who hear and re-tell them, whether we are talking about foraging communities of the Paleolithic world, or the great philo­ onfucianism sophical and religious traditions of major world civilizations, from C to Buddhism to the traditions of the Aztec world, of Christianity and of Islam. They are also powerful because they are shared by most members of a given community, who learn the rudiments of their origin stories as children, and then internalize those stories in the course of many years of education, with increasing detail and sophistication. As far as we know, origin stories can be found at the core of all forms of education. They have provided foundational knowledge in seminaries and universities, as well as in the rich oral traditions passed on by elders in all foraging communities. In the light of this discussion it is apparent that Durkheim’s notion of “anomie” can also be understood as the state of mind of those who lack access to a credible, rich, and authoritative origin story. Intellectual anomie is a state of map-lessness and meaninglessness. Curiously, it is the intellectual state that became the norm in the twentieth century, as globalization and modern science battered confidence in traditional origin stories, both in the metropolitan centers of the world and at its colonial margins. Everywhere, modern secular educational systems ceased to teach within shared traditions of foundational knowledge. Some found the decline of traditional origin stories exhilarating and liberating, and glorified in the multiple, free-floating perspectives of a world without a shared origin story. But many, both in the colonial world and in the metropolitan heartlands, experienced, and continue to experience, a deep sense of loss. Today, we are so used to a world without universal framing ideas (particularly in the humanities), that it is easy to forget how painful it was to lose the sense of intellectual coherence that goes with trust in an origin story. But that sense of loss is apparent in much of the literature, philosophy, and art of the late nineteenth and early twentieth centuries. Here are just two, more or less random, examples of what I mean. In his 1851 poem, “Dover Beach,” Matthew Arnold writes: The Sea of Faith Was once, too, at the full, and round earth’s shore Lay like the folds of a bright girdle furled. But now I only hear Its melancholy, long, withdrawing roar, Retreating, to the breath Of the night-wind, down the vast edges drear And naked shingles of the world. 28

What is big history?

The poem continues with a terrifying vision of a future without coherence or meaning: Ah, love, let us be true To one another! for the world, which seems To lie before us like a land of dreams, So various, so beautiful, so new, Hath really neither joy, nor love, nor light, Nor certitude, nor peace, nor help for pain; And we are here as on a darkling plain Swept with confused alarms of struggle and flight, Where ignorant armies clash by night. W.B. Yeats’ “The Second Coming,” was written in 1919, just after the Great War seemed to realize Arnold’s haunting vision of the future. Turning and turning in the widening gyre The falcon cannot hear the falconer; Things fall apart; the centre cannot hold; Mere anarchy is loosed upon the world, The blood-dimmed tide is loosed, and everywhere The ceremony of innocence is drowned; The poem ends with a famous and terrifying image: what rough beast, its hour come round at last, Slouches towards Bethlehem to be born? Specialization and the loss of traditional unifying narratives were symptomatic of the chaotic and incoherent world described in so much twentieth century literature, art, and philosophy. Indeed, it has often been assumed that this world of isolated, even incommensurable disciplines and perspectives is characteristic of modernity in general. The modern world threw together peoples, cultures, religions, and traditions so violently that it created a growing sense of a single humanity, while undermining confidence in traditional visions of the world. In the Communist Manifesto, we read that, in the bourgeois era of human history: “All fixed, fast-frozen relations, with their train of ancient and venerable prejudices and opinions, are swept away, all new-formed ones become antiquated before they can ossify. All that is solid melts into air, all that is holy is profaned, …” In a book on modernity that takes its title from this passage, Marshall Berman writes that the modern world has created: a paradoxical unity, a unity of disunity; it pours us all into a maelstrom of perpetual disintegration and renewal, of struggle and contradiction, of ambiguity and anguish. To be modern is to be part of a universe in which, as Marx said: ‘all that is solid melts into air.’41 But a different interpretation is also possible. Perhaps for much of the twentieth ­century, we have lived in a sort of intellectual building site, surrounded by the debris 29

David Christian

of older origin stories, while a new origin story was being constructed all around us, a story for humanity as a whole. The best evidence for this idea is the re-emergence of new unifying stories in the last 50 years. Seen from this perspective, big history is the project of trying to tease out and build a modern, global origin story.

Big history is an origin story for the Anthropocene Epoch Perhaps, then, we can think of big history as an origin story for the twenty-first century. Big history builds on the intellectual achievements of modern science, but it is also the product of an increasingly globalized world, that is very different from the world of E.H. Carr. Scientific knowledge has advanced faster than he could have imagined, and new technologies such as the Internet have created a much more intertwined world. But perhaps the most important changes arise from the “Great Acceleration,” the astonishing increase in human numbers, human energy use, human control over the environment, and human inter-connectedness, in the 60 years since Carr wrote.42 In that brief period, we humans have collectively become the single most important force for change in the biosphere, the first single species to play such a role in the 4 billion year history of life on Earth. That is an outcome that Carr could not have imagined in 1961. These spectacular changes mean that questions about the nature and source of the astonishing power wielded collectively by 7.4 billion humans loom much larger today than they did in Carr’s time. In this sense, big history can be thought of as an origin story for the Anthropocene Epoch of human history. We will need the broad scale of big history to see the Anthropocene clearly, because it is not just a turning point in modern world history, but a significant threshold within human history as a whole, and even in the history of planet Earth. Most contemporary historical scholarship studies the last 500 years. The danger of this foreshortened perspective is that it can normalize recent history, making the technologically and economically dynamic societies of recent centuries seem typical of human history in general. They are not. Their dynamism is extraordinary and exceptional. The very idea of history, of long-term change, is modern and, as John McNeill has shown, the scale of change in the modern era, and particularly since the mid twentieth century, really is “something new under the sun.”43 In contrast, most people in most human societies over the last 200,000 years lived lives whose structures and surroundings seemed relatively stable, because change was so slow that it could not be observed at the scale of a few generations. Only within the capacious scales of big history is it possible to see clearly that the Anthropocene Epoch is strange not just on human scales, but also on those of the history of planet Earth.This is perhaps why, in a recent article, a group of paleontologists suggest that the Anthropocene Epoch counts as one of the three most important turning points in the history of the biosphere, along with the emergence of life, almost 4 billion years ago, and of multicellular life 600 million years ago.44 Never before has a single species dominated change in the biosphere as we humans do today, and never before has the near future depended as it does today, on the decisions, insights, 30

What is big history?

and whims, of a single species. Appreciating the strangeness of modern society is vital if we are to tackle the global challenges it poses for the near future. Understanding how strange today’s world is may also give us a renewed appreciation for the insights and understanding of our ancestors, who maintained over many millennia a much more stable relationship the biosphere as a whole.

Big history is the first origin story for all humans If big history is an origin story, it is also the first origin story for humanity as a whole. Emerging as it does in a densely interconnected world, it is the first origin story constructed by, and available to, all human beings.While traditional origin stories tried to sum over knowledge from particular communities or regions or cultural traditions, this is the first origin story that tries to sum over accumulated knowledge from all parts of the world. That alone suggests the wealth of information and the astonishing richness of detail of a modern origin story. Traditional origin stories provided a unifying vision for particular communities, by highlighting the ideas that different people shared, just as modern national histories provided a unifying vision for nation states despite internal differences of language, culture, religion, and ethnicity. In a similar way, the big history story can start to provide a unifying vision for humanity as a whole, despite the many differences between regions, classes, nations, and cultural traditions. The construction and dissemination of a global origin story can help generate the sense of human unity that will be needed as human societies navigate collectively through the global challenges of the next few decades. Though the national and cultural tribalisms that dominated Carr’s world are still very much present today, he would have been astonished to see, emerging alongside them, an origin story for humanity as a whole. So interconnected is today’s world that the idea of a unified humanity with a history of its own has a salience that it lacked in Carr’s time, when the most significant human communities seemed to be either nation states or culturally cohesive regions such as “the West” or the Muslim world, or the zones dominated by great traditional empires such as China or India.Today, a sense of global citizenship, of belonging to the global community of humanity, is not just a matter of scientific precision. (­Generically speaking we are, after all, a remarkably homogenous species, so that the category, Homo sapiens has a scientific precision that the category of “Chinese human being” or “American human being” lacks.) Awareness of what all humans share is increasingly a matter of self-preservation, particularly in a world with nuclear weapons. E.H. Carr wrote “What is History?” one year before the Cuban missile crisis, when, according to President Kennedy, the odds of an all-out nuclear war lay “between one out of three and even.”45 H.G. Wells’ attempt to write a universal history in 1919, when the horrors of the Great War were still vivid in his mind, was driven by a similar sense of human unity. Peace, he argued, required new ways of thinking. It required: …common historical ideas.Without such ideas to hold them together in harmonious co-operation, with nothing but narrow, selfish, and conflicting nationalist traditions, races and peoples are bound to drift towards conflict and destruction. 31

David Christian

This truth, which was apparent to that great philosopher Kant a century or more ago … is now plain to the man in the street.46 More recently, the great American world historian, William McNeill, has made the point with equal eloquence: Humanity entire possess a commonality which historians may hope to understand just as firmly as they can comprehend what unites any lesser group. Instead of enhancing conflicts, as parochial historiography inevitably does, an intelligible world history might be expected to diminish the lethality of group encounters by cultivating a sense of individual identification with the triumphs and tribulations of humanity as a whole. This, indeed, strikes me as the moral duty of the historical profession in our time.We need to develop an ecumenical history, with plenty of room for human diversity in all its complexity.47 As Wells understood, a universal history is the natural vehicle for a unified history of humanity, because, unlike national histories, big history first encounters humans not as warring tribes, but as a single, and remarkably homogenous, species. And it is a story that can now be told with increasing precision and confidence, and can help us understand the place of our species not just in the recent past, but in the history of the biosphere, and of the entire universe.

Notes 1 Our thanks to Lowell Gustafson, the editor of the Journal of Big History, for permission to reprint this article, which first appeared in the Journal of Big History, 1, no. 1 (2017), 4–19. Here, the article appears with minor changes, and additional references to more recent scholarship. 2 Jean Bodin, sixteenth century, cited from Craig Benjamin “Beginnings and Endings,” in Marnie Hughes-Warrington, ed., Palgrave Advances in World Histories (New York: Palgrave Macmillan, 2005), 95. 3 E. O. Wilson, Consilience: The Unity of Knowledge (London: Abacus, 1998). 4 E. H. Carr, What Is History? (Harmondsworth: Penguin, 1964), 7. 1st published in 1961, based on the George Macaulay Trevelyan Lectures, delivered in 1961 in Cambridge. 5 Carr, What Is History? 150. 6 On Humboldt as a big historian before his time, see Fred Spier, Big History and the Future of Humanity, 2nd ed. (Malden, MA: Wiley Blackwell, 2015), 18–21, and Andrea Wulf, The Invention of Nature: The Adventures of Alexander von Humboldt, the Lost Hero of Science (London: John Murray, 2015). 7 The distinction between paradigm and pre-paradigm disciplines was introduced by a book whose first edition appeared in 1962, just a year after Carr’s book: Thomas Kuhn, The Structure of Scientific Revolutions, 2nd ed. (Chicago: University of Chicago Press, 1970). 8 Carr, What Is History? 7–8. 9 C. P. Snow, The Two Cultures and the Scientific Revolution (Cambridge: Cambridge ­University Press, 1959). 10 Carr, What Is History? 32. 11 Carr, What Is History? 26. 12 Carr, What Is History? 84. 32

What is big history?

1 3 Carr, What Is History? 23. 14 Carr, What Is History? 24. 15 Collingwood’s work, like Carr’s, was staple fare for graduates of my generation. His most important work was R. G. Collingwood, The Idea of History, rev. ed., Jan Van der Dussen (Oxford and New York: Oxford University Press, 1994). 16 Cited from Robert C. Tucker, ed., The Marx-Engels Reader, 2nd ed. (New York and London: W.W. Norton & Co., 1978), 595. 17 Carr, What Is History? 30. 18 David Cannadine, ed., What Is History Now? (Basingstoke: Palgrave/Macmillan, 2002). 19 Jean-François Lyotard, The Postmodern Condition: A Report on Knowledge, trans. Geoff Bennington and Brian Massumi (Minneapolis: University of Minnesota Press, 1984). 20 One interesting example is The History Manifesto, by Jo Guldi and David Armitage (Cambridge: CUP, 2014), which offers an aggressive critique of short-termism in contemporary historical scholarship. 21 A start up list might include Eric Chaisson, Cosmic Evolution: The Rise of Complexity in Nature (Cambridge, MA: Harvard University Press, 2001); David Christian, Maps of Time: An Introduction to Big History, 2nd ed. (Berkeley, CA: University of California Press, 2011), and Origin Story: A Big History of Everything (Little, Brown and ­Penguin, 2018); Fred Spier, Big History and the Future of Humanity, 2nd ed. (Malden, MA: Wiley/Blackwell, 2015); Cynthia Stokes Brown, Big History: From the Big Bang to the Present, 2nd ed. (New York: New Press, 2012); a university text, David Christian, ­Cynthia Stokes Brown, and Craig Benjamin, Big History: Between Nothing and Everything (New York: McGraw-Hill, 2014); anthologies of essays, such as Barry Rodrique, Leonid ­Grinin, and Andrey ­Korotayev, eds., From Big Bang to Galactic Civilizations: A Big ­History Anthology, Vol. 1, Our Place in the Universe (Delhi: Primus Books, 2015); and a beautifully illustrated overview, Macquarie University Big History Institute, Big History (London: DK Books, 2016). 22 Cannadine, ed., What Is History Now? 149. 23 This section summarizes and adds to arguments I have presented in “The Return of Universal History,” History and Theory, Theme Issue, 49 (December, 2010), 5–26. 24 Erwin Schrödinger, What Is Life? (Cambridge: CUP, 2000), 1 [first pub. 1944]; Schrödinger was also acutely aware of the barriers that specialization placed in the way of such ambitions. 25 Carr, What Is History? 57. 26 Carl Sagan’s television series, Cosmos, was first broadcast in 1980; Preston Cloud’s Cosmos, Earth, and Man: A Short History of the Universe (New Haven: Yale University Press, 1978) was published just two years earlier; the Soviet Union already had a flourishing tradition of “biosphere” history, pioneered by the great geologist, Vladimir Vernadsky in works such as V. I. Vernadsky, The Biosphere (New York: Springer-Verlag, 1998). 27 There is a fine account of the real, as opposed to the idealized, methodologies of modern science in John Ziman, Real Science: What It Is, and What It Means (Cambridge: CUP, 2000). 28 See David Christian, “Historia, complejidad y revolución cronométrica” [“History, Complexity and the Chronometric Revolution”], Revista de Occidente, no. 323 (Abril 2008), 27–57, and David Christian, “History and Science after the Chronometric Revolution,” in Steven J. Dick and Mark L. Lupisella, eds., Cosmos & Culture: Cultural Evolution in a Cosmic Context (NASA, 2009), 441–462; and see Doug Macdougall, ­Natures’ Clocks: How Scientists Measure the Age of Almost Everything (Berkeley: University of ­California Press, 2008). 29 H. G. Wells, Outline of History: Being a Plain History of Life and Mankind, 3rd ed. (New York: Macmillan, 1921), 1102. 33

David Christian

30 John Mulvaney and Johan Kamminga, Prehistory of Australia (Sydney: Allen & Unwin, 1999), 1–2; a fine recent history of Australian archaeology is Billy Griffiths, Deep Time Dreaming: Uncovering Ancient Australia (Carlton, VIC: Black, Inc., 2018). 31 Colin Renfrew, Prehistory: The Making of the Human Mind (London: Weidenfeld and Nicolson, 2007), 41. 32 Cannadine, What Is History Now? 153. 33 William H. McNeill, “History and the Scientific Worldview,” History and Theory, 37, no. 1 (1998): 12–13. 34 Origins (Newsletter of the International Big History Association), VI.08 (2016), 7. 35 Wulf, The Invention of Nature, Chapter 18, “Humboldt’s Cosmos.” 36 Currently, new techniques for deciphering the genomes of ancient humans are transforming our understanding of human evolution and prehistory: see David Reich, Who We Are and How We Got Here: Ancient DNA and the New Science of the Human Past (­Oxford: Oxford University Press, 2018). 37 See, for example, Andrey V. Korotayev and Alexander V. Markov, “Mathematical Modeling of Biological and Social Phases of Big History,” in Leonid Grinin, et al., eds., Teaching and Researching Big History: Exploring a new Scholarly Field (Volgograd: ‘Uchitel’ Publishing House, 2014), 188–219; and Peter Turchin and Sergey A. Nefedov, Secular Cycles (Princeton: Princeton University Press, 2009). Turchin is the editor of Cliodynamics: The Journal of Quantitative History. 38 Peter M. Hoffmann, Life’s Ratchet: How Molecular Machines Extract Order from Chaos (New York: Basic Books, 2012), is a wonderful exploration of how molecular machines exploit the “molecular storm” created by the random energy of individual molecules to power the chemistry of cells; and why doing so does not breach the second law of thermodynamics, because it depends on additional sources of free energy, mostly supplied by the battery molecule, ATP. 39 On the idea of a Noösphere, see David Christian, “The Noösphere,” forthcoming ( ­January 2017) on www.edge.org/, 2017, Annual Question. 40 David Christian, Origin Story: A Big History of Everything, explores the idea of big history as a modern, global, science-based origin story. 41 Marshall Berman, All that Is Solid Melts into Air: the Experience of Modernity (New York: Penguin, 1988, 1st published 1982), 15. 42 See John McNeill, The Great Acceleration: An Environmental History of the Anthropocene since 1945 (Cambridge, MA: Harvard University Press, 2014). 43 For more on these claims, see David Christian, “History and Time,” Australian Journal of Politics and History 57, no. 3 (2011): 353–365, and John McNeill, Something New under the Sun: An Environmental History of the Twentieth-Century World (New York: W. W. Norton, 2000). 44 Mark Williams, Jan Zalasiewicz, et al., “The Anthropocene Biosphere,” The Anthropocene Review (2015): 1–24. 45 Graham Allison and Philip Zelikow, Essence of Decision: Explaining the Cuban Missile Crisis, 2nd ed. (New York: Longman, 1999), 271. 46 Wells, Outline of History, vi. 47 William H. McNeill, “Mythistory, or Truth, Myth, History, and Historians,” The American Historical Review 91, no. 1 (Feb. 1986), 7.

34

PART I

Big history and science

2 BIG HISTORY AND THE STUDY OF TIME The underlying temporalities of big history Barry Wood Thirty-six years ago the philosopher–physicist T. J. Fraser, founder of the International Society for the Study of Time (ISST), summarized the evolutionary levels of the cosmos, from the particulate to the stars and galaxies, earth and life, and finally humans and their social institutions, adding that “the stable integrative levels created by these steps survive and coexist today.” They coexist in a complex nested reality where laws of lower levels provide parameters for higher levels while higher levels exhibit emergent properties. Fraser then established a foundation for his own philosophy of temporalities; “it follows that time itself has evolved along a path corresponding to the evolutionary complexification of matter” (Fraser 1982, 35). Over the past half century, the sciences have provided a chronology for the history of humanity, life, earth, the Milky Way, and the Universe. The age of Homo sapiens is now been extended to 300,000 years, life to 3.8 billion years (byr), the earth to 4.56 byr, the Milky Way to 12.6 byr, and the universe to 13.8 byr. These dates were established following 300 years of guesswork dating including various ingenious calculations for the age of the earth (Gorst 2001), but we now have good reason to expect that these dates will stand with only minor adjustments. All such dates, however, emanate from a single idea of time that is intensely geocentric and uniformly regular like the marks on a meter stick or the tick-tock of a clock.We use units derived from earth’s orbit around the sun, the size and mass of the earth, and the behavior of water at the earth’s surface. Our second began as a fraction of a mean solar day; our meter began as a fraction of the distance around the earth; our kilogram began as a measure of the mass of a liter of water at sea level; the liter began as a specific volume derived from the meter; the centigrade scale arbitrarily divides temperature between the freezing and boiling points of water—at sea level—into 100 degrees (Angier 2007, 71–86). Older measuring systems (inches, feet, yards) contrarily retained by the United States are Anglocentric, but the rest of the world has graduated from England to earth, except for our base-ten mathematical system which is homocentrically derived from our accidental inheritance of ten digits on our hands from our primate ancestors. 37

Barry Wood

Like our usual confusion of measurements with reality, we take the clock tick-tock as real time, regular, unchanging, proceeding invariably since The Initial Moment of Emergence (TIME), commonly known as the Big Bang. We feel comfortable, ­Norman Berrill wrote, with an idea of time as “something like an infinitely long ribbon marked off into empty blank spaces all calling for some kind of entry” (Berrill 1955, 164). Erich Jantsch used a similar metaphor in describing time as an a priori container, “an absolute timescale into which evolution pours like water into a driedout river bed” ( Jantsch 1980, 84–85). Eric Chaisson, who utilizes an arrow-of-time image, assumes clock time as the framework for cosmic evolution: “However time flows and for how long, we take it to be a linear phenomenon, to unfold at a steady pace from its fiery origins to the here and now of the present” (Chaisson 2001, 6). Virtually all writers find it easier to follow this ribbon-river flow than wrestle with conceptual conundrums of temporality. It suits ordinary life on earth but it is illsuited to cosmic history.Years in the millions and billions soon get out of hand; thus various metaphors have been developed to make it understandable in human terms. Cynthia Brown noted the value of “compressing” cosmic time to human dimensions, each page of her 288-page book, Big History, representing 15 million years (myr) with Homo sapiens arriving on the scene in the last third of the last line (Brown 2007, 39–40). Nobel winner Francis Crick likened his 192-page book, Life Itself, to the 600-myr history of life since the Cambrian period, each letter representing 1,500 years and an individual life less than the final period on the last page (Crick 1981, 21). Carl Sagan equated cosmic history to a one-year calendar with recorded history occupying the last ten seconds of December 31 (Sagan 1977, 20–22). David Christian equated the history of the universe to 13 years; modern industrial society occupies the last six seconds (Christian 2003, 440). Nigel Calder equated it to a walk the length of Manhattan Island; the life of a single human is confined to the paint on the railing of Battery Park (Calder 1983, 73–74). The Hewlett-Packard walk-throughtime exhibit telescopes the history of the world to a one-mile walk, each 30-inch stride representing more than 200,000 years (Liebes et al. 1998). While all such time compressions aid the imagination, they all start with uniform clock time. Albert Einstein upset the clock cart when he theorized that time might move more slowly at super speeds and stop entirely at the speed of light. Measurements have since proved the former; the latter is beyond comprehension. In everyday experience, we are all aware of how irrelevant clock time can be, completely overrun by psychological time that may drag, pass us by, or rush ahead—interesting physical and spatial metaphors.This variation in the passage of time suggests that there may be several different times in different realms—atomic, chemical, organic, social—which big historians treat as an imaginative challenge: learning to think across multiple “scales” of time (Christian 2004). Einstein’s linking of time and space forces us to think of time as intimately connected with the primary levels of organization that have dominated at different stages of cosmic history. The idea of reality as being comprised of different levels has a long history that extends into arcane models of the universe in ancient and medieval times. For our purposes, versions from the last half century are the most relevant. Preston Cloud presented one of the first comprehensive histories of the universe in Cosmos, Earth, and Man (1978); the three levels of the universe in his title were expanded in the 38

Big history and the study of time

text to four which coincidently anticipated the taxonomy in the definition of big history: “the attempt to understand, in a unified, interdisciplinary way, the history of Cosmos, Earth, Life, and Humanity” (Christian 2011, 20). A decade before the term “big history” was coined, Chaisson divided cosmic history into seven levels he called “epochs,” which he altered slightly a quarter century later to the “seven ages of the cosmos”: particulate, galactic, stellar, planetary, chemical, biological, and cultural (Chaisson 1981, 2006). Fraser’s hierarchy emphasizes six “major integrative levels” based on physics: (1) The world of particles with zero restmass, always on the move at the speed of light; (2) The world of particles with nonzero restmass, always on the move but at speeds below that of light; (3) The world of massive, ponderable masses gathered into stars, galaxies, and groups of galaxies; (4) The world of living organisms; (5) Man as a species and as an individual member of the species; (6) The collective institutions of human societies to the extent that they function on semiautonomous structures. (Fraser 1982, 28–29) The world of physics is daunting for most of us, and the logic of three out of six levels given over to the pre-organic may not be immediately obvious but Fraser’s “major integrative levels” provide a precise physical foundation for his theory of various temporalities that correspond to these physical levels. The entire second chapter of The Genesis and Evolution of Time (1982) is devoted to “the principle of temporal levels” (Fraser 1982, 18–36). In his essay “The Study of Time” which appeared at the end of his comprehensive anthology, The Voices of Time, Fraser combined the Greek roots chronos and sophia to coin the word “chronosophy.” This word does not recur in his later writings, nor has it entered general usage, but one of the purposes he proposed for chronosophy was “to promote communication between the humanities and the sciences using time as the common theme” (Fraser 1966, 590–592). The form time takes in the humanities is narrative which is basic to literature and history and often proves useful for the presentation of the sciences and social sciences. Elsewhere this writer has argued that narrative provides a transdisciplinary template for bridging what C. P. Snow called the “two cultures”—the sciences and the humanities (Wood 2013). Narrative as our basic cognitive endowment provides more than Fraser’s “common theme”; as the entry point for understanding the world from grade school to grad school, narrative organization is the first axiom of chronosophy. Based on Fraser’s chronosophical insights, the present study looks at the temporalities underlying various cosmic narratives that together add up to the grand narrative of big history. Fred Spier (1996) has presented the “structure” of big history as a series of “regimes”; David Christian (2004) has organized it as a sequence of “thresholds”; Tyler Volk has organized it around sequential events of “cosmogenesis.” Fraser’s lasting contribution remains his argument that different temporalities prevail 39

Barry Wood

at various integrative levels of reality. He postulated six kinds of time—­atemporality, prototermporality, eotemporality, biotemporality, nootemporality, and s­ociotemporality— which correspond precisely with his six integrative levels (Fraser 1982, 29–31). Here we ­argue that atemporality stands as the first underlying structure for big history but the others, which we here term petrotemporality, magnetotemporality, and genotemporality, are needed to construct the full story. The resulting big history fulfills Fraser’s goal for chronosophy by joining the sciences and humanities within a unifying narrative. Fraser’s approach to various temporalities derived from Jakob von Uexkull’s concept of Umwelt, a term that has been naturalized into English. A creature’s umwelt is the universe made possible but limited by a creature’s sensory receptors and effectors (Uexkull 1957). An ant’s umwelt does not extend to a human standing over it or even bending close to observe it but a sand crab’s umwelt includes a nearby human; thus it scurries away. Defined psychologically, a creature’s umwelt is “the circumscribed portion of the environment which is meaningful and effective for a given species” (­English 1964). Fraser’s addition to the umwelt idea is the human use of instruments and formulas to open up and explore non-biological umwelts. “We can use mathematics and theories to explore aspects of the universe that are not otherwise accessible to any living creature, such as the world of a traveling photon, the motion of distant galaxies, the implosion of a star, or the explosion of the Big Bang.”This is what Fraser calls the “extended umwelt principle” (Fraser 1982, 28–29). Fraser’s use of umwelt seems anthropocentric because he applies the extended umwelt principle to pre-organic and non-organic entities. It is however best understood as a thought experiment, as is clear in his approach to the “proper framework of the proton … one enters its umwelt in imagination” (Fraser 1982, 38; italics added).We know the spatial nature of the photon’s umwelt; it has no mass and occupies no space. Its temporal umwelt is a primal mystery—not an isolated mystery from some ancient world far removed from us but constantly present. The world we know is awash with photons conveying images not only of loyal pets and faces of friends before us but also everything in the universe; in Fraser’s words, they are “the surviving ancestors of everything that evolved” (Fraser 1982, 39). Through this evolution, lesser velocities, complex levels of organization, and new temporalities have been put in place. Imaginatively, we must look into their umwelts and variant temporalities because they underlie the grand narrative of big history.

Atemporality and big history Fraser’s discussion of time is firmly based on the work of Albert Einstein, specifically his special theory of relativity. Everything we know about the universe beyond earth is gleaned from radiation—a variety of waves reaching earth from everything in the sky, from the nearby moon to the most distant galaxies. Until the mid-twentieth century, we saw only visible light collected by optical telescopes; in the past half century, this has been extended through viewing devices capable of picking up radiation above and below the visible spectrum. The Hubble Telescope, launched in 1991, provides limited images from both the infrared and ultraviolent ends of the spectrum. The Cosmic Background Explorer (COBE) launched in 1989, and its successor, the 40

Big history and the study of time

Wilkinson Microwave Anistropy Probe (WMAP) launched in 2001, observed the microwave portion of the spectrum so finely that temperature differences of two ten-thousands of a degree were mapped which made sense of the later distribution of galaxies and galaxy clusters. Radiation photons belong to a family of pre-particles (Max Planck’s quanta) that includes gravitons and neutrinos. These never exist at rest but are always moving at a constant speed (186,000 mps; 300,000 kps). Recognizing that the absolute velocity of light was more reliable and real than the putative absolute rest assumed in Newtonian theory, Einstein tied matter and energy to the speed of light and thus rewrote the fundamentals of physics. In this new framework, time itself is redefined; as Fraser notes, “It is known from special relativity theory that in the proper framework of the photon no meaning can be attached to ideas of futurity, pastness, and presentness” (Fraser 1982, 31). While few people understand the intricacies of relativity, one paradox has gained popularity: time runs slower as velocity increases and stops at the speed of light. The cessation of time at light speed defines Fraser’s atemporality. This paradoxical fact about the speed of light stretches the imagination until we see it in operation. Let us suppose a star exploded 6,500 light years from earth. At this instant, a change occurred within untold trillions of atoms at the site of the explosion: in our current understanding, atomic particulates called electrons dropped from higher to lower orbits, signifying a loss of energy; this energy was then imparted to non-particulates called photons which rushed away in all directions at the speed of light. At this velocity time stopped; thus they carried a precise unchanging, unaltered “photograph” of the explosion that arrived at the earth 6,500 years later. This arrival is precisely what happened on July 4, 1054, when Chinese observers and probably Native Americans recorded an exploding star where the Crab Nebula is now located. What they saw was an exact image generated 6,500 years earlier, identical because time and change did not and do not exist at the speed of light. This is different from a photograph sent from one cell phone to another: such an image is in fact separated into a temporal stream of information bits which are reassembled into an image by the receiving phone. The supernova image was delivered to earth through time, 6,500 years, by an atemporal medium. While this phenomenon allows us to “look back in time” (an anthropocentric view of what happens) what really occurs is that an instant in past time is delivered atemporally, but through time, to us here and now. Atemporality at the speed of light has allowed us to reconstruct a big history of the universe. When we focus our instruments at a distance of 65 million light years in the direction of the Virgo Constellation, we are able to discern the Virgo Cluster of 1,300 to 2,000 galaxies, with the full array of 92 elements in evidence. Refocusing at 250 million light years, we discover a powerful region known as the Great Attractor which has recently been established as the center of the Laniakea Supercluster that includes 300 to 500 clusters containing approximately 100,000 galaxies (Tully et al. 2014). Refocusing at a distance of 12 billion light years we see that globular galaxies are far more numerous than spirals with hydrogen and helium predominating and elements 3 (Lithium) to 92 (Uranium) almost entirely missing. From this preservation of images through the atemporal mediation of light, we construct a narrative history of the universe which includes the unorganized clustering of stars into globules in 41

Barry Wood

the early universe, the progressive reorganizing of these into spirals arranged in superclusters strung out along immense filaments, and the progressive enrichment of the universe through nuclear fusion in stars and their end-game supernovas and kilonovas which scatter element-loaded stardust into space. A final refocusing at wavelengths far longer than visible light allows us to see the Cosmic Background Radiation (CBR), a nearly uniform surround of microwaves, the redshifted remnants of the first light emerging from the primordial fireball when the universe was just 380,000 years old. This redshift can be subtracted—an involved calculation which must account for our motion around our own galaxy—to reveal the miniscule variations in temperature (and thus density) which correspond to galaxies and galaxy clusters that began to emerge 13.2 byr BP. By arranging all of these atemporal images along a 13.8-byr timeline, we are able to translate atemporality into our own geotemporal frame to construct a big history.

Petrotemporality and big history In Fraser’s definition, the level of reality mediated by photons across the electromagnetic spectrum is the astronomical world of galaxy clusters, galaxies, stars, and their accompanying planets, all of which he describes a “ponderable masses” (Fraser 1982, 28). His treatment of this realm is minimal though it harbors vast processes of galaxy formation, stellar evolution, element fusion, and the formation of planetary systems. Its operative force is gravity; it is dominated by deterministic cycles—the slow turning of galaxies, spinning stars, orbiting planets, “directionless time … pure succession” which he terms eotemporality, meaning first time, earliest time, or dawn (Eos) time (Fraser 1982, 30–31). But eotemporality in the context of the earth, one of the four rocky planets of the inner solar system, includes various processes associated with rock: its formation, endurance, preservation, and destruction through various petrotemporal events. Petrotemporality or time as resident in rocks is as complex as the history of the earth that we now know includes multiple geological processes. One was the formation of the earth, a process of accretion and continuing bombardment over several 100 myr now called the Hadean period, referring to the hellish conditions of Hades, the Greek underworld. Temperature was so high that little has survived, but bombardment continues today, though greatly attenuated, bringing to earth a surprising number of meteors from the earliest time of the solar system (Dalrymple 1991, 257–304). Second, numerous igneous and metamorphic rocks have survived from the Archean Era (4.0 to 2.5 byr), generally found within the most long-­lasting planetary cratons (Dalrymple 1991, 125–192). Third, once oceans formed and the water cycle was in place, erosion created sedimentary layers on ocean bottoms, pieces of which have survived folding, fracturing, faulting, uplift, and continental collisions. Fourth, the rise of life, especially shell-enclosed creatures and vertebrates, has produced fossils which, when embedded in sedimentary rock are mineralized or fossilized. Each of these processes creates a geological record at a specific place which can be located on a three-dimensional grid—longitude, latitude, and depth or elevation. In addition, they may be described in terms of time, giving rise to various kinds of petrotemporality. 42

Big history and the study of time

Petrotemporality and fossils A simple approach to petrotemporality is sequential according to the history of discovery, which began in the seventeenth century. The Danish anatomist ­Nicholas Steno (1638–1686) gained fame for his discovery of saliva ducts, tear glands, and mechanisms of the heart that established him as one of the most prominent scientists of the time and brought him an offer to serve as physician to Ferdinand II, the Grand Duke of Tuscany (Montgomery 2012, 56–81). Steno’s contributions to geology began in 1666 when a gigantic shark was caught off the Tuscan coast and Ferdinand ordered the head sent to Florence so that Steno could explore its anatomy. His 1667 report, Canis Carcharia Dissectum Caput (Dissection of a Shark’s Head), ranks as the earliest geological treatise and a stimulus for Steno’s subsequent formulation of basic laws of sedimentation. Steno noticed that the shark’s teeth were identical to stoneswithin-rocks that had puzzled naturalists all the way back to Pliny the Elder (23–79 CE), author of the 37-volume Natural History, who thought they must have fallen to the earth during lunar eclipses. Later mythology held that they were tongues of serpents turned to stone by the Apostle Paul, a legend that probably traces to a viper bite on the island of Malta from which he recovered (Acts 28:3–6); hence they came to be known as glossopetrae, “tongue stones.” Displaying down-to-earth common sense, Steno recognized that “bodies dug out of the ground and looking like parts of animals should be considered parts of animals” (Steno 1667, 35). Applying this insight to tongue stones, he wrote, “That they are sharks’ teeth (Canis Carchariae dentes) is proved by their shape, since they are quite alike, planes to planes, sides to sides, base to base” (Steno 1667, 42–43)—his language sounding like Euclid’s tests for geometric congruency. In an observation anticipating much later understandings of geology, Steno wrote that “new islands have emerged from the sea … Perhaps formerly when this place [Malta] was submerged in the sea it was the haunt of sharks, whose teeth in times past were buried in the muddy sea-bed” (Steno 1667,  45). Steno included several conjectures (conjectura) for how soil might mix with water, then harden into rock, but he had no explanation for the process we now call fossilization or mineralization. The fossilization of an organism constitutes a reversion of living material cell by cell to non-organic molecules. The biological is transformed into a lower but much more stable level of reality and with it biotemporality is replaced with petrotemporality, which includes an interesting ambiguity. In Fraser’s framework, the umwelt of a mineralized shark’s tooth, a petrified tree, or a fossil of any kind exists in a frozen present, thus mimicking atemporality, though in fact a fossil hovers in an uncertain eotemporality of material processes we now know extend beyond the earth to other planets and presumably to the far reaches of the universe: “the eotemporal world does not contain structures capable of protecting and maintaining their identities” (Fraser 1982, 30). Fossils are thus subject to various forces that operate on irregular timescales we might term geotemporal—erosion, plate tectonics, vulcanism, glaciation, and the like—adding various kinds of temporal uncertainty. A fossil’s umwelt is always in danger of catastrophic alteration by a variety of geophysical processes. Our interest, however, is in petrotemporality where geophysical alteration has occurred very slowly or not at all. 43

Barry Wood

The fossil record of the earth shows petrotemporal lifespans of 520 myr for p­ roto-vertebrate fossils of Pikaia found in the Burgess shale of British Columbia (Palmer 2006, 66–67). Extensive beds of fossilized stromatolites aged 1.9 byr have been found on Blanchett Island, Great Slave Lake, Canada (Luyles and Miali 2007, 118). Much older and more complex stromatolites aged 3.4 byr are found in rocks of the Fig Tree Group in Southern Africa. The oldest stromatalites, aged 3.5 byr, are small, rare, and scattered through rocks from the Warrawoona Group in Western Australia (Derenne et al. 2008). Chemical signatures in rock have provided evidence of microscopic life 3.7 byr (Ohtomo 2014); published timetables suggest that the first living cells probably formed around 4 byr ( Jantsch 1980, 121; Swimme and Tucker 2011, 120). If undisturbed, petrotemporality suggests that fossils would endure as long as their enclosing rock.

Petrotemporality and stratigraphy In his Prodromus (1669), Steno puzzled again over the mystery of the tongue stone, “a solid body enclosed by process of nature within a solid” (solido intra solidum naturaliter contento). His new concern, however, is evident in systemized observations on the strata of the earth that were visible in the Alps and Tuscan Hills. Steno’s presentation is set forth most often as numbered propositions which reveal the kind of scientific observation-and-conclusion methodology evident in early scientific writing—in ­Francis Bacon’s Novum Organum (1620), for instance. The method suited his recognition that rock layers were created successively by sequences of sedimentation so that the oldest layer is located at the bottom, the youngest at the top. This is today known as the principle of superposition; moreover, it gives rise to two others: the principle of original horizontality, meaning that sedimentary rocks first form in horizontal layers, and the principle of lateral continuity, which states that layers found at one location extend laterally underground in all directions until they meet a geographical obstacle—a valley, mountain, or body of water (Steno 1669, 227–231). The recognition of lateral continuity triggered the life work of William Smith (1769–1839) whose explorations led to his landmark geological map of England (Winchester 2001). While Steno’s principles allowed for relative dating of sedimentary layers, their real importance resided in materials and forms found within successive layers, allowing for what we now call relative dating. Known as the founder of modern geology, Steno laid a foundation for stratigraphical processes and for petrotemporality which governs both the rocks of the earth and the fossilized solids they enclose. In an interesting paradox, long-lasting petrotemporality for fossils depends on stability within sedimentary rock but not necessarily for the rock itself. Each may be described in terms of different umwelts defined by different material surroundings. Folding, faulting, and fracturing may damage sedimentary rock but miss enclosed fossils and leave them intact, in which case the disturbed rock and its fossilized contents are subject to different temporalities. This paradox is what allows us to find fossils in uplifted rock that has been severely fractured and boulders called “floats” carried by glaciers and deposited hundreds of miles from their points of origin. In a remarkable sequence of observations, Steno recognized that mountains include level strata, inclined strata, exposed edges of strata, fragments of broken strata at the base of mountains (scree), and evidence of volcanism which he described as “clear 44

Big history and the study of time

traces of subterranean fire” (Steno 1669, 232). These observations provided evidence that mountains represented severe disruptions of the earth but Steno was unable to systematize what he observed. At one point he confesses “that the extension of crests of mountains, or chains, as some prefer to call them, along the lines of certain definite zones of the earth, accords with neither reason nor experience” (Steno 1669, 234). Despite his awareness of geological processes we know occurred over many millions of years, Steno could not overcome religious assumptions based on the Bible, including the kind of chronology Bishop Ussher had published in Annals of the World (1651) which placed the creation of the world in the year 4004 BCE. Steno believed, for instance, that the stratified layers of rock he had studied were laid down during the Biblical flood and all tilting, faulting, and fracturing he observed had occurred since then. Sounder geological explanations were delayed for a century. “He burst the boundaries of time,” wrote Stephen Jay Gould of James Hutton (1785–1788), the Scotsman who explained the processes by which earth alteration could occur and systematized what is now known as the rock cycle (Repchek 2003, 1). Residing on inherited farms known as Slighhouses and Nether Monynut in Scotland near the English border, Hutton observed significant erosion on the slopes and, well ahead of his time, set about reducing it with strategic stone walls, drainage ditches, and crop rotation (Repchek 2003, 103–113). Extending his observations beyond his landholdings, Hutton reasoned that the continuous erosion of land and mountains should have long ago moved all the earth’s land into the oceans; subsequently he developed a theory to explain the regeneration of the land. In 1785, Hutton presented his ideas to the recently formed (1783) Royal Society of Edinburgh, arguing that there must be a compensating process whereby rocks formed by sedimentation on sea bottoms were uplifted to form new land (Hutton 1785). In the following years, Hutton continued his relentless search for evidence, including expeditions into the highlands of Scotland, though his own farmlands triggered his most important discovery. Nether Monynut, his upland farm, was located on schistus; Slighhouses had a red sandstone foundation. A north-running schistus-­sandstone boundary was minimally visible along the eastern boundary of Nether Monynut. ­Hutton may have anticipated an exposed outcrop of this boundary at the coast. His boating expedition with John Playfair and Sir James Hall led to its discovery at Siccar Point along with decisive proof for his theory of the earth—a striking narrative of the earth’s past made visible (Wood 2019). Standing in the boat while he and his friends gazed at the formation, Hutton described how the vertical layers were originally laid down on the ocean bottom, then subjected to lateral pressure that folded them into vertical waves the tops of which, after uplift above sea level, had then been eroded away; subsequent submergence had then allowed an accumulation of new horizontal layers which now, following another uplift, stood high above the water. The disjunction between the vertical and horizontal layers which provided Hutton with the evidence he sought is now known as an unconformity. Nothing Hutton wrote conveys the excitement of this moment; his Theory of the Earth (1794) was so turgid that his insight received little notice. John Hall who had doubted Hutton’s theory was instantly converted and John Playfair later wrote of that moment at Siccar Point, “We felt ourselves necessarily carried back…. The mind seemed to grow giddy by looking so far into the abyss of time” (Playfair 1956, xiv). 45

Barry Wood

While hosting the young geologist Charles Lyell (1797–1875) in 1824, more than a quarter century after Hutton had died, John Hall repeated that moment: they rowed down the coast to see Siccar Point as he, Playfair, and Hutton had seen it 36 years earlier. Lyell’s words reflects his appropriation of Hutton’s vision:“Worlds are seen beyond worlds immeasurably distant from each other, and beyond them all innumerable other systems are faintly traced on the confines of the visible universe” (Lyell 1990, I, 63). In his three-volume Principles of Geology (1930–1932) Lyell built on Hutton’s insights to become the greatest geologist of the century. Hutton’s recognition of what is now called the rock cycle demonstrates that petrotemporality, while exhibiting remarkable stability if rocks are undisturbed, may be subject to gradual disjunctions. Hutton’s descriptions tend toward a steady process, gradual change over vast periods of time, implying that all highlands might erode away and all sea bottoms might eventually rise to form new land. In terms of umwelt, Hutton’s rock cycle must continue; there appears to be no room for exceptions, “no vestige of a beginning—no prospect of an end” (Hutton 1785, 304). This slow inexorable cycling of the earth’s materials is a defining feature of the sedimentary-rock umwelt. Hutton would not have known how to explain ancient non-sedimentary rocks or fossils which have somehow escaped this cycle of erosion and sedimentary reconstruction. Hutton’s theories were part of the geological story known as uniformitarianism, which posited slow and gradual change that extended the timeline of geohistory deep into prehistory. The opposing story was catastrophism, which argued that sudden cataclysmic events could account for changes in the earth. The latter, whose advocates have drifted to the backwaters of historical geology, were influenced by belief in the biblical deluge, a desire to preserve it, and retain the belief in a recent creation. In setting biblical chronology aside, Hutton was clearly in the vanguard of geological thinking. Now we know that the full narrative of geohistory is indeed complex: long periods of uniform and virtually imperceptible change punctuated by sudden ­disjunctions—volcanic eruptions, landslides, earthquakes, tsunamis, and the occasional arrival of an asteroid that keeps seismologists, climatologists, and asteroid watchers busy. Hutton’s theory of the earth provided the first long-term geohistory; it formed the basis for the far more extensive explorations and theories that emerged over the next two centuries. The full richness of petrotemporality would have to await the dramatic narrative of geological process that came to the fore in the ­twentieth century with plate tectonics.

Petrotemporality and radioactive isotopes A third approach to petrotemporality derives from the formation of atomic matter itself, first in the Big Beginning commonly called the Big Bang, then in nucleosynthesis within stars, supernovas when stars self-destruct (Weinberg 1977; Silk 1989; Harpaz 1994; Gribbin 2000; Chown 2001) and most ultimately, we now realize, in kilonovas (Wood 2018). Most presentations of the evolution of the elements focus on the naturally occurring 92 as if the enormous energies of nuclear fusion produce none other than those portrayed in the Periodic Table of Elements. Stellar cooking was understood by the mid-twentieth century (Hoyle 1946; Burbidge et al. 1957) 46

Big history and the study of time

but the possibility that a sample of an element might contain more than one kind of atom was suggested as early as 1871 with the term isotope applied in 1911 to all variants. Now, a century later, isotopes of all but 9 out of 92 have been found, element number 50 (tin) setting the record with over a dozen. Isotope differences reflect extra neutrons in the nucleus. Counting all 92 elements plus their known isotopes adds up to hundreds; scores are radioactive; approximately 20 are used for radiometric dating (Dalrymple 1991, 92) to which radiocarbon dating has been added (Libby 1955). These go through a process of decay, shedding excess neutrons on a rigorous and reliable timetable until they achieve a stable form, either a standard element or a stable isotope further down the Periodic Table. The discovery of radioactive decay which dates to the work of Madame Currie (1867–1934) provides a third version of petrotemporality the regularity and certainty of which exceeds both mineralize fossils or fossil-bearing sedimentary rocks. Once isotopic decay was discovered and the concept of the half-life worked out, the value of radioactivity (Madame Curie’s term) as a temporal measurement became clear; by 1905 its value for dating mineral deposits was recognized; it was a short step to begin its application to the oldest rocks on earth and thus to determine the age of the earth. Work on measuring the half-lives of numerous radioactive isotopes occupied dozens of physicists over the next 40 years. Igneous rocks from volcanos are particularly valuable for radiometric dating because their molten state in effect “sets the clock” so that the proportion of the original “mother” element to the derived “daughter” element provides an accurate dating scale. Decades of searching turned up several regions where unaltered samples of the world’s oldest rocks could be dated: various lavas from the Swaziland Supergroup in South Africa dated at 3.53 byr; basalt formations in the Pilbara region of Western Australia dated at 3.57 byr; granitoid gneises years near Granite Falls, Minnesota dated at 3.68 byr; supercrustals near Isua in west Greenland dated at 3.81 byr (Dalrymple 1991, 141, 169, 180). The most useful components of ancient rocks are long-lived zircons which look enough like diamonds that they are used extensively in jewelry. Their isotopic content is exceedingly long lasting, typically including trace amounts of Uranium 238 and Uranium 235 with half-lives of 4.47 byr and 704 myr. This makes them subject to accurate radiometric dating over the whole history of the earth. Like fossils, zircons display paradoxical petrotemporality, their surrounding rock—whether metamorphic or sedimentary—may erode away, leading to zircon recontainment millions of years later. In terms of umwelt, such breakdown of containing rock would lead to the destruction of fossils, even the most firmly mineralized versions but zircons are virtually indestructible with various combinations of zirconium and silicon arranged in exceeding hard tetragonal crystals. These endure not only the typical hazards of their containing rock but long distance transport. The unique structural fingerprint of a zircon can often connect it to a source hundreds or thousands of miles away, which allows tracking of plate rifts in early supercontinents and consequent continent drift. The oldest zircons have been dated to 4.4 byr, just 100 myr after the earth’s formation (Wilde et al. 2001). While movement, breakdown of enclosing rock, and successive rehabilitation in later rock formations makes a precise history difficult to construct, water content in the most ancient zircons indicates oceans were beginning to accumulate on the earth this early in the Hadean period. 47

Barry Wood

An important application of radioactive isotopic decay was developed in the 1940s with the work of Willard Libby. Recognizing that cosmic rays reacted with ­Nitrogen-14 in the atmosphere to produce Carbon-14, a mildly radioactive isotope of Carbon-12, Libby set out to discover its absorption by living tissue. His preliminary work (Libby 1946) eventually led to a systematic method of radiocarbon dating (Libby 1952).With a measured half-life of 5730 years, Carbon-14 provides a temporal scale for well-preserved organic materials (bones, plant seeds, wooden carvings, cave paint pigments) reliable to about 40,000 years with declining reliability over longer timespans. Radiocarbon dating provides a biotemporal scale for life forms prior to the process of mineral replacement resulting in petrification. Carbon-14 temporality is also useful as an adjunct to genotemporality to be discussed shortly.

Petrotemporality and meteors While the virtual indestructability of zircon provides a petrotemporal scale for earth-origin rocks deep into the Hadean Era, another form of petrotemporal stability is evident from age measurements of extra-planetary rocks—meteor fragments containing chondrites, achondrites, and iron that current theory suggests were forming from interstellar debris at the same time as the solar system. Verification depends on similar dates for meteor fragments from other planets. The Martian meteorite Northwest Africa (NWA) 7533, which pairs with NWA 7034 (nick-named “Black Beauty”), is thought to originate in the crust of the southern highlands region of Mars that is estimated to have taken 100 myr to form. Recent dating has established its age as 4.4 byr (Marchi et al. 2012), suggesting that the planet Mars formed perhaps 4.5 byr ago, the same date calculated for the age of the earth. The recent Dawn spacecraft exploration of Vesta, one of the largest asteroids in the solar system, dated it within a few million years of the formation of the first solar system solids (Humayun et al. 2013). Dalrymple summarizes the evidence of meteor dating as a “convincing body of evidence showing that the solid bodies of the Solar System formed 4.5–4.6 byr ago” (Dalrymple, 1991, 404). Applying radiometric techniques to extra-planetary rocks has thus led to consistent evidence verifying the value of petrotemporality for reconstructing a big history of the earth and the solar system.

Magnetotemporality and plate tectonics Alongside the recognition of various petrotemporalities based on sequencing of fossils and radiometric dating of rock strata, zircons, and meteorites, another form of geochronology has emerged that may be usefully termed magnetotemporality. At the dawn of the twentieth century, the French geophysicists Bernard Brunhes (1857–1910) recognized an opposite magnetic alignment from today’s in sedimentary rock (Brunhes and David 1901). Exploring an extinct volcano near his observatory in central France, he discovered a similar reversed alignment in ancient volcanic rock (Brunhes 1906). From these observations he speculated that there must have been a reversal of the earth’s magnetic poles at some time in the past. Brunhes did not live to appreciate the significance of this discovery, nor did it receive attention from geologists for several decades. However, between 1927 and 1929 the Japanese geophysicist Motonori 48

Big history and the study of time

Matuyama (1884–1958) undertook a gravity survey in J­apan, including Korea and Manchuria (under ­Japanese occupation at the time). This led to corroborating evidence of magnetic reversals along with the recognition that these could be correlated with the established ages of sedimentary rock strata (Matuyama 1929). Such reversals subsequently became one of the most useful “earth cycles” for investigating geotemporality (Hoffman 1988; Olroyd 2006). Magnetic reversals in minerals that crystalized during rock formation (notably the iron mineral magnetite) occurred on a variable schedule and thus could not by themselves establish an absolute geochronology; moreover, even a relative chronology was difficult to ascertain. Much like dendrochronology, which depends on an unbroken continuity of tree rings that correlate recent with progressively earlier timbers, a reliable scale of magnetotemporality depends on an unbroken sequence of rock depositions. In a meticulously argued paper a century ago, Joseph Barrell (1917) contended that sedimentation is most often arrhythmic or “pulsatory”; for this reason, sedimentary discontinuities make reliable chronology difficult and often impossible. Earth’s “deep history” and the recognition of paleomagnetism constitute a pivotal   orld War II chapter in geological discovery (Sullivan 1974; Rudwick 2014). Following W and the development of submarines capable of deeper descent, military-­motivated sea-bottom measurements occurred as needed during random surveys with magnetometers. Piecemeal evidence of magnetic reversals was predictably difficult to analyze. However, more systematic investigation during the International Geophysical Year (1 July 1957—31 December 1958) revealed a sequence of magnetic reversals on either side of the Mid-Atlantic ridge and similar reversals on either side of the East Pacific Rise which parallels the west coast of Central and South America. It was soon clear that ocean floors were being created at these mid-ocean ridges; upwelling lava was progressively pushing its way into a divide in the ocean floor and forcing sections of the seafloor apart. This discovery of seafloor spreading provided the first sound evidence for what had hitherto been called “continental drift,” a term associated with Arthur Wegener decades earlier from his book The Origin of Continents and Oceans (1915). Wegener’s theory was intuitively based on the apparent “fit” between coastlines on either side of the Atlantic which he illustrated with maps of today’s continents assembled into a single landmass which eventually came to be known as Pangea (“all land”). His evidence was also geological, with coastal rock formations on both coasts of the Atlantic clearly aligned; additionally, it was paleontological, based on similar fossils of flora and fauna separated by 1,000 of miles of ocean. However, neither Wegener nor anyone else had come up with a theory to explain how the massive structures of continents could have separated and moved 1,000 of miles apart. The mechanism was made clear when Princeton geologist Harry Hess (1962) set forth his “history of ocean basins,” theorizing that upwelling lava at mid-ocean ridges and the movements of continents were caused by convection currents in the earth’s mantle. Persuasive evidence resided in parallel “stripes” on each side of mid-ocean ridges that recorded successive magnetic reversals. Unlike the discontinuities of sedimentation, seafloor lava sequences turned out to be continuous, thus lending a degree of certainty to the dating of magnetic reversals. The most recent magnetic reversal was thus dated reliably to 780,000 years ago—fairly rapidly according to the record of 49

Barry Wood

the rocks. Earlier reversals—183 over the past 83 myr—reveal a cycle though not cyclical regularity; as John McPhee describes it, “a distinct arrhythmic yardstick through time” (McPhee 21). Increasing ages of volcanic rock on each side of the divide thus provide a relative rather than absolute magnetotemporality. The formulation of a mechanism for seafloor spreading provided a physical explanation; continental drift thus quickly acquired precision and a new descriptive term: plate tectonics. Seafloor spreading occurring in the Pacific and Atlantic Oceans has been the cause of longitudinal change in continental locations: in general, the “sidewise” drift of continents has been the most obvious. However, evidence of latitudinal movement had already accumulated decades earlier with fossil finds of life forms far from their expected habitat though this had at first remained a puzzle for geographers and biologists. North America had an extensive record of tropical jungle around an ancient inland seaway that covered most of the land between the Rockies and the ­Appalachians— the origin of extensive Carboniferous coal deposits—along with remains of gigantic dinosaur species that could not have survived on today’s temperate forests. Evidence of different locations for most land masses in past times was extensive. It was soon recognized that continental movement could be tracked independently of a continuous rock record. A periodic sampling of continental magnetic alignment at intervals of a few million years combined with evidence of climate alterations and fossil changes was enough to establish latitudinal movement. Kenneth Deffeyes of Princeton University has summarized how it works: “The earth’s magnetic field is such that a compass needle at the equator will lie flat, while a compass needle at the poles will want to stand straight up on end—with all possible gradations of that in the latitudes between.” Thus “you can tell not only whether the magnetic pole was in the north or south when the rock formed but also—from the more subtle positions of the needles—the latitude of the rock at the time it formed” (McPhee 115). Geologists thus select rock samples from stable continental cratons insulated from tectonic distortion, erosional alteration, and marginal terrane collisions to assure that any twisting or rotation of land masses detected are authentic cratonic movements. The assembling of paleomagnetic evidence has revealed extensive latitudinal movement of continents. The progressive tilt of magnetic alignment as one traces North America back in time indicates that it has rotated to its present “vertical” location from a nearly “horizontal” position on the Equator at the time the Americas were joined to Africa in a single landmass. Similar latitudinal movement accounts for evidence of ancient glaciation in tropical Africa, temperate forests in Antarctica, and fossils of tropical vegetation in Scandinavia and other areas now far from equatorial regions. Analysis has now achieved a remarkable degree of precision; to quote McPhee again, “In the argot of geology, paleomagnetic specialists are sometimes called paleomagicians” (McPhee 22). Magnetotemporality thus turns out to be of surpassing value for tracking continental drift and terrane migration even where the rock record has extensive discontinuities.

Genotemporality and big history Following Fraser’s idea of umwelt as a way of formulating temporality at various levels of reality, biotemporality refers to the many species-specific worlds available to animals 50

Big history and the study of time

or humans according to their endowments of sensors and effectors. The increasing complexity of organisms achieved through the evolutionary process thus results in a succession of umwelts and their associated biotemporalities. A new dimension of biotemporality appeared with the sequencing of the human genome; as Frank Ryan puts it, “nothing quite prepared us for the strange glory that was unveiled in ­February 2001, when, for the first time in history, two rival consortia revealed the genetic makeup of our human genome” (Ryan 2009, 100). The startling discovery was that human endogenous retroviruses or HERVs make up a huge proportion of the human genome (Ryan 2009, 102). We tend to regard viruses as the cause of sicknesses from the common cold and flu to lethal diseases like AIDS and ebola but the fact is that the human genome includes numerous recognizable viral sequences which were incorporated into the biological tree millions or billions of years ago. Lynn Margulis (1998; Margulis and Sagan 2002) was a pioneer in this field with research on bacterial and viral symbiosis as a driver of evolutionary change. Carl Zimmer describes the incorporation of viral components in the cells of higher species as “almost philosophical in its weirdness … a mishmash of genes, cobbled together by evolution … as if the world was filled with hybrid monsters, with clear lines of identity blurred away” (Zimmer 2011, 47). Despite this initial sense of weirdness, its fundamental logic may be illustrated by an industrial analogy: if a manufacturing company wishes to incorporate a new process, the most difficult procedure is expensive R&D to reinvent the process; the easier way is to headhunt—find people who have already developed the process and hire them into the company. Essentially, this appears to be how the evolution of higher organisms has proceeded: useful viral processes were incorporated into higher organisms’ DNA. In this way, complexity of function was assembled from pre-existing skills. As we have seen, temporalities, with the exception of atemporality, incorporate the past into the present. Petrotemporality incorporates and preserves the past history of ancient life in mineralized fossils, the past history of geological processes in sedimentary layers, deformations, and unconformities; and former isotopic instabilities provide evidence for constructing chronologies according to specific radioactivity schedules. Magnetotemporality records past movements of tectonic plates in alternating and tilted magnetic alignments. In similar fashion, the genomes of higher organism record processes developed in a distant past in earlier organisms, including earth’s earliest bacteria. A specific genotemporality thus governs genetic processes, bringing the past into the present and laying down a blueprint and recipe for a species’ future. With living organisms, we enter a far more complex world. The stability of fossils over millions of years or zircons over billions of years is replaced with fluid continuity difficult to conceptualize. Living organisms are characterized by a continual flowthrough of energy and cellular replacement. This makes genotemporality different from the petrotemperality of rock-encased fossils. How can we define the umwelt of a gene when it does not survive from one organism to the next? Matt Ridley provides a suggestive description: “The genes in the cells of your little finger are the direct descendants of the first replicator molecules through an unbroken chain of tens of billions of copyings; they come to us today still bearing a digital message that has traces of those earlier struggles of life” (Ridley 2000, 22). While an individual gene has a limited lifespan, the recipe for its duplication and the blueprint for its 51

Barry Wood

offspring outlive it and both the recipe and the blueprint are passed on so that the process can be repeated. In a curious way the umwelt is inherited too, in the sense that the offspring cell will find a place in an identical organelle-niche with identical opportunities and limitations. The symbiosis of this relationship guarantees that both the offspring gene and its inherited umwelt will be the same. This underlying genotemporality is what makes the larger context of biotemporality work. Like all higher-level temporalities, genotemporality involves uncertainties which have been long recognized. Genetic mixing through sexual reproduction is one; genetic drift is another; and genetic mutations is a third. These add stress to the genotemporal situation but this is precisely the stress of Darwinian variation out of which those that adapt most effectively will survive. A fourth uncertainty traces to virolutionary changes within the incorporated DNA fragments themselves that may result in adjustments all around. The most familiar version of this is the annual adaptations of the flu virus to last year’s vaccine, along with the occasional mutation such as the one that occurred in 2014 that rendered the seasonal flu vaccine less effective. Less familiar is the deep cellular co-evolution of cellular host and viral inhabitant which guarantees biogenetic complexity over the long run but short-term genotemporal uncertainty. The long-term survival of genetic material has resulted in hitherto unimagined discoveries. Rutherford (2016) summarizes the 2008 discovery of a fingerbone and later a tooth in Densova Cave in Siberia that has led to recognition of a new species inhabiting Asia cotemporally with Neanderthals. Now known as Denisovans, traces of their genome are now recognized in contemporary Melanesians of Fiji, Papuan New Guinea and islands off the northeast coast of Australia.The genotemporal record may be subject to hazards of decay or loss, but fascinating events have survived for tens of 1,000 of years in the human genome as current accounts have documented (Harari 2015; Reich 2018). The genome of present Homo sapiens includes 2% or 3% ­Neanderthal DNA testifying to more frequent matings than formerly imagined; moreover, Homo sapiens/ Neanderthal mating encounters appear to have occurred in separate episodes 1,000 of years apart, with the earliest occurring between male Neanderthals and Homo sapiens females (Rutherford, 51–54). Perhaps the most astonishing discovery (Slon et al. 2018) is genomic sequencing from a 90,000-year-old bone fragment that has revealed a first-generation offspring of a Denisovan male and a Neanderthal female. With discoveries of this accuracy and antiquity, genotemporality has demonstrated itself as the most sophisticated structure for reconstructing prehistoric human history. Conventional Darwinian theory posits a narrow-focus view of evolutionary change based on offspring variation with survival keyed to those variations that maximize survival and promote coupling and reproduction. The framework for evolutionary change is biotemporal. The larger framework introduced by the complex assemblage of DNA from earlier species—including a multitude of primeval viruses, bacteria, protozoans, and fungi—suggests a wide-angle genotemporal view as appropriate for the full complexities of speciation, coevolution, and the emergence of higher levels in the human story: psychological and sociological. These involve new kinds of time, psychotemporality and sociotemporality, which govern the final though as yet not fully written chapters of the historical narrative. 52

Big history and the study of time

Summary Prior to the big history perspective, available temporalities were limited and anthropocentric. In Fraser’s terms, the range of umwelts was largely undeveloped. Bishop James Ussher’s chronology, which placed the creation of the earth, sun, moon, and animals as stage props for humans, reduced cosmic time to sociotemporality. The ­Judeo-Christian association of human origins with Hebrew genealogy and the history of Israel reduced human history to ethnotemporality. Georges-Louis Leclerc, Comte de Buffon’s assumption that the earth tore away from the sun and his attempt to determine its age by calculating the cooling rate of various metal and stone globes moved geohistory into workshop temporality (Dalrymple 1991, 30–31). His conclusion that 74,832 years—the cooling time of an 8,000 mile diameter heated ball of rock and metal—was the age of the earth was ingenious but lacking essential data (Haber 1959, 118). Apart from his uniformed guesswork about the rock-and-mineral composition of the earth, what Buffon lacked was a framework of petrotemporality, specifically the effect of continuous radioactive decay deep in the earth producing additional heat that would necessarily alter and lengthen the cooling time of the planet. Other prescientific time schemes were ingenious but obviously anthropocentric, psychocentric, sociocentric, or ethnocentric. The precise calculations required for big history required new temporalities aligned with the multiple levels of reality, their distinctive complexities, and unique emergent properties. The atemporality of the preparticulate universe has allowed for the delivery of the past, unchanged, to the present, mediated by the spectrum of electromagnetic radiation. Interpreting this past based on variations produced by redshift, emission lines of radiation spectra, and enhancements of gravitational lensing has allowed us to view past states of the universe and thus construct a 13.8-billion-year big history of the universe. The petrotemporality of mineralized fossils, sedimentary rocks, and established decay-rates of radioactive isotopes has allowed us to develop a chronology for past events of the solar system and construct a 4.5-byr big history of the earth. With the coupling of isotopic decay with paleomagnetic reversals recorded in iron-­ bearing rocks, an effective magnetotemporality has produced a record of seafloor creation and continental movement to fill in additional chapters of the terrestrial story. The genotemporality of the human genome has allowed us to view the incorporation of pre-vertebrate and subsequent DNA into our own, allowing for a big-history perspective on the story of life and humanity. Without these various temporalities, the big-history perspective would be impossible. We would be confined to a reality lacking an extended past with no chronological perspective on the future—a world of mere succession, one damn thing after another. This would plunge us back centuries to a deterministic view of the world and ourselves that would confine us to the realm of Fraser’s eotemporality, the dawn time of ancient mythology.

References Angier, Natalie. 2007. The Canon: A Whirligig Tour of the Beautiful Basics of Science. New York: Houghton Mifflin. Barrell, Joseph. 1917. “Rhythms and Measurement of Geological Time.” Bulletin of the Geological Society of America 28, 745–904 and plates. 53

Barry Wood

Berrill, N. J. 1955. Man’s Emerging Mind. New York: Dodd, Mead & Company. Brown, Cynthia. 2007. Big History: From the Big Bang to the Present. New York: The New Press. Brunhes, Bernard. 1906. “Recherches sur la direction d’aimentation des roches volcaniques.” Journal des physique théorique et appliqué 8 (Ser.4), 705–724. Brunhes, Bernard, and Pierre David. 1901. “Sur la direction d’aimantation permanante dans des couches d’argille transformée en brique par des coulées de lave.” Comptes rendu de l’Académie des Sciences, Paris 133, 155–157. Burbidge, E. Margaret, G. R. Burbidge, William A. Fowler, and Fred Hoyle. 1957. “­Synthesis of the Elements in Stars.” Review of Modern Physics 29, 547–650. Calder, Nigel. 1983. Timescale: An Atlas of the Third Dimension. New York: Viking Press. Chaisson, Eric. 1981. Cosmic Dawn: The Origins of Matter and Life. Boston, MA: Little, Brown and Company. Chaisson, Eric. 2001. Cosmic Evolution: The Rise of Complexity in Nature. Cambridge, MA: Harvard University Press. Chaisson, Eric. 2006. Epic of Evolution: Seven Ages of the Cosmos. New York: Columbia University Press. Chown, Marcus. 2001. The Magic Furnace: The Search for the Origin of Atoms. Oxford: ­Oxford University Press. Christian, David. 2003. “World History in Context.” Journal of World History 14, 437–488. Christian, David. 2004. Maps of Time: An Introduction to Big History. Berkeley: University of California Press. Christian, David. 2011. “The Evolution of Big History: A Short Introduction.” Evolution: A Big History Perspective. Eds. Leonid E. Grinin, Andrey V. Korotayev, Barry H. ­Rodrigue. Volgograd: ‘Uchitel’ Publishing House, 20–25. Cloud, Preston. 1978. Cosmos, Earth, and Man: A Short History of the Universe. New Haven, CT: Yale University Press. Crick, Francis. 1981. Life Itself: Its Origin and Nature. New York: Simon and Schuster. Dalrymple, G. Brent. 1991. The Age of the Earth. Stanford, CA: Stanford University Press. Derenne, Sylvie, et al. 2008. “Molecular Evidence for Life in the 3.5 Billion Year Old Warrawoona Chert.” Earth and Planetary Science Letters 272 (1–2), 476–480. English, Horace B., and Ava Champney English. 1964. A Comprehensive Dictionary of Psychological and Psychoanalytic Terms. New York: Davis McKay. Fraser, T. J. 1966. “The Study of Time.” The Voices of Time: A Cooperative Survey of Man’s Views of Time as Expressed by the Sciences and by the Humanities. Ed. T. J. Fraser. New York: George Braziller, 582–592. Fraser, T. J. 1982. The Genesis and Evolution of Time. Amherst: University of Massachusetts Press. Gorst, Martin. 2001. Measuring Eternity: The Search for the Beginning of Time. New York: Broadway Books. Gribbin, John. 2000. Stardust: Supernovae and Life—The Cosmic Connection. New Haven, CT: Yale University Press. Haber, F. C. 1959. The Age of the Earth: Moses to Darwin. Baltimore, MD: Johns Hopkins Press. Harari, Yuval Noah. 2015. Sapiens: A Brief History of Humankind. New York: Harper Collins. Harpaz, Amos. 1994. Stellar Evolution. Wellesley, MA: A. K. Peters.

54

Big history and the study of time

Hess, Harry. 1962. “History of Ocean Basins.” Petrologic Studies: A Volume to Honor A. F. Buddington. Eds. Albert Engel, Harold James, and Benjamin Leonard. New York: Geological Society of America, 599–620. Hoffman, Kenneth. 1988. “Ancient Magnetic Reversals: Clues to the Geodynamo.” Scientific American 258 (May), 50–59. Hoyle, Fred. 1946. “The Synthesis of the Elements from Hydrogen.” Monthly Notices of the Royal Astronomical Society 106, 343. Humayun, M., et al. 2013. “Origin and Age of the Earliest Martian Crust from Meteorite NWA 7533.” Nature 503, 513–516. Hutton, James. 1788. “Theory of the Earth.” Royal Society of Edinburgh Transactions 1, 209–304. Jantsch, Erich. 1980. The Self-Organizing Universe: Scientific and Human Implications of the Emerging Paradigm of Evolution. Elmsford, NY: Pergamon Press. Libby, Willard F. 1946. “Atmospheric Helium Three and Radiocarbon from Cosmic ­Radiation.” Physical Review 69 (11–12): 671–672. Libby, Willard F. 1952. Radiocarbon Dating. Chicago: Phoenix. Liebes, Sidney, Elisabeth Sahtouris, and Brian Swimme. 1998. A Walk through Time: From Stardust to Us. New York: John Wiley & Sons. Luyles, Nick, and Andrew Miali, 2007. Canada Rocks: The Geologic Journey. Markham, ON: Fitzhenry & Whiteside. Lyell, Charles. 1990. Principles of Geology. 3 vols. Chicago: University of Chicago Press. Marchi, S., et al. 2012. The Violent Collisional History of Asteroid 4 Vesta. Science 336 (11 May), 690–694. Margulis, Lynn. 1998. Symbiotic Planet: A New View of Evolution. New York: Basic Books. Margulis, Lynn, and Dorion Sagan. 2002. Acquiring Genomes: A Theory of the Origin of Species. New York: Basic Books. Matuyama, Mononori. 1929. “On the Direction of Magnetisation of Basalt in Japan, Tyȏsen and Manchuria.” Proceedings of the Imperial Academy of Japan 5, 203–205. Montgomery, David R. 2012. The Rocks Don’t Lie. New York: W. W. Norton. Ohtomo, Yoko. 2014. “Evidence for Biogenic Graphite in Early Archaen Isua Metasedimentary Rocks.” Nature Geoscience 7, 25–28. Olroyd, David. 2006. Earth Cycles: A Historical Perspective. Westport, CT: Greenwood Press. Palmer, Douglas, ed. 2006. The Illustrated Encyclopedia of the Prehistoric World. Edison, NJ: Chartwell Books. Playfair, John. 1956. Illustrations of the Huttonian Theory of the Earth. New York: Dover. Reich, David. 2018. Who We Are and How We Got Here: Ancient DNA and the New Science of the Human Past. New York: Pantheon Books. Repchek, Jack. 2003. The Man Who Found Time: James Hutton and the Discovery of the Earth’s Antiquity. Cambridge, MA: Perseus Books. Ridley, Matt. 2000. Genome: The Autobiography of a Species in 23 Chapters. London: Perennial. Rudwick, Martin J. S. 2014. Earth’s Deep History: How It Was Discovered and Why It Matters. Chicago: University of Chicago Press Rutherford, Adam. 2016. A Brief History of Everyone Who Ever Lived: The Human Story ­Retold Through Our Genes. New York: The Experiment. Ryan, Frank. 2009. Virolution. London: HarperCollins. Sagan, Carl. 1977. The Dragons of Eden: Speculations on the Evolution of Life. New York: Random House. Silk, Joseph. 1989. The Big Bang. Revised and enlarged edition. New York: W. H. Freeman & Co.

55

Barry Wood

Slon, Viviane, et al. 2018. “The Genome of the Offspring of a Neanderthal Mother and a Denisovan Mother.” Nature 561, 113–116. Spier, Fred. 1996. The Structure of Big History: From the Big Bang until Today. Amsterdam: Amsterdam University Press. Steno, Nicolaus. 1667. Reprint edition: The Earliest Geological Treatise by Nicolaus Steno. Trans. Axel Garboe. London: Macmillan, 1958. Steno, Nicolaus. 1669. Reprint edition: The Prodromus of Nicolaus Steno’s Dissertation Concerning a Solid Body Enclosed by Process of Nature within a Solid. Trans. John Garrett Winter. New York: Hafner Publishing, 1968. Sullivan, Walter. 1974. Continents in Motion: The New Earth Debate. New York: McGraw-Hill. Swimme, Brian Thomas, and Mary Evelyn Tucker. 2011. Journey of the Universe. New ­Haven: Yale University Press. Tully, R. Brent, et al. 2014. “The Laniakea Supercluster of Galaxies.” Nature 513 (4 S­ eptember), 71–73. Uexkull, Jakob von. 1957. “A Stroll through the Worlds of Animals and Men: A Picture Book of Invisible Worlds,” Instinctive Behavior. Trans. Claire H. Schiller. Madison, CT: International Universities Press, 5–80. Weinberg, Steven. 1977. The First Three Minutes. New York: Basic Books. Wilde, S. A., J. W. Valley, W. H. Peck, and C. M. Graham. 2001. “Evidence from Detrital Zircons for the Existence of Continental Crust and Oceans on the Earth 4.4 Gyr Ago”. Nature 409, 175–178. Winchester, Simon. 2001. The Map that Changed the World: William Smith and the Birth of Modern Geology. New York: HarperCollins. Wood, Barry. 2013. “Bridging the ‘Two Cultures’: The Humanities, the Sciences, and the Grand Narrative.” The International Journal of Humanities Education 10, 53–63. Wood, Barry. 2018. “Refueling the Magic Furnace: Kilonova 2017 Rewrites the Element Creation Story.” Journal of Big History 3 (3), 1–16. Wood, Barry. 2019. “Petrotemporality at Siccar Point: James Hutton’s Discovery of the Deep-Time Narrative.” The Study of Time, XVI, 157–178. Zimmer, Carl. 2011. A Planet of Viruses. Chicago: University of Chicago Press.

56

3 BIG HISTORY AND ASTRONOMY – SPACE IS BIG1 The Fermi paradox: its relevance to big history and the human race Jonathan Markley A typical big history course starts at the Big Bang and proceeds through the ­formation of galaxies and stars, gradually narrowing to a single star system, and then looking at the formation of a single planet, then narrowing still further to look at the creation and spread of the living organisms that inhabit that planet, and it ultimately culminates with the impact and possible future of a single species amongst those life-forms. In short, it starts with everything, but ends with human beings. In the early part of the course billions of years are covered per class, then later millions, then tens of thousands, and by the end the focus narrows to ­thousands or even mere centuries. To a purest, this is a violation of the very essence of big history, which seeks to understand the history of everything, and to put everything in its proper context. Imagine a course on the Twentieth Century that finished with the first sixty years in the first two weeks, three more decades in the  next two weeks, and by the second half of the course was entirely focused on the later parts of 1999, eventually narrowing to single hours and even minutes in the final day of the century. Walter Alvarez provides a standard justification for this big history approach in his new book A Most Improbable Journey. A Big History of Our Planet and Ourselves. The broader history of everything might seem irrelevant to someone interested in human history, but it is not. The human situation in which we find ourselves is the result of a history that has unfolded across enormous stretches of time and space, and almost everything that has taken place in human history has been strongly influenced by events deeper in the past… For me the astonishing realization that comes from the study of Big History is just how unlikely our world is. At innumerable points in its history, events could have led to totally different results – to a human situation completely different from what we know today or to a world with no humans at all.2

57

Jonathan Markley

A similar but more critical description was given by Eric Chaisson (in which he quoted David Christian and Fred Spier). Even big historians’ work is limited. Big History, as most often defined – ‘human history in its wider context’ (Christian…) or ‘an approach to history that places human history in its wider context’ (Spier 2010…) – pertains mostly to the meandering cosmic trek that led specifically to us on Earth. As such, it mainly concerns, in reverse order of appearance, changes that led to humankind, the Earth, the Sun, and the Milky Way Galaxy. Scant treatment is given, or need be given, to other galaxies, stars, or planets throughout the almost unimaginably vast Universe, for the goal of Big History is to place humanity itself into a larger cosmic perspective.3 In this chapter, I argue that big historians must indeed consider other galaxies and other planets, and that this is actually an essential part of investigating the final part of Walter Alvarez’s point: the improbability of the presence of human beings on a planet like this. There is more at stake than the particular forces that shaped a single star system, and the wider story does in fact help us understand the “trek” that led to us here. The Fermi paradox was first expressed by Enrico Fermi, probably around May 1950, “where is everybody?”4 A small group of scientists had been discussing the ­possibility of visits to Earth by flying saucers and led Fermi to wonder why this didn’t seem to have occurred and speculated about a number of reasons why. (It might be impossible, or not worth the effort, or it might take too long, etc.) Since this initial conversation, the question has grown more broadly to wonder why we have never been able to detect the presence of intelligent life anywhere in the Universe beyond our single planet. Building from this abstract question and the increasing possibility that radio telescopes might be able to detect the presence of intelligent life elsewhere, Frank Drake formulated the Drake Equation in 1961.5

N = R* • fp • ne • fl • fi • fc • L

N = the number of advanced aliens that emit electro-magnetic signals that we can detect. R* = the speed at which suitable stars form that can support intelligent life. fp = the proportion of those stars that have planets. ne = the number of planets, per star system, that have an environment that can sustain life. fl = the proportion of those planet on which you actually get life. fi = the proportion of those planets on which you get intelligent life. fc = the proportion of those intelligence species that develop technology so that they emit signals we can detect. L = the amount of time those signal emitting intelligent species have been doing so. The Drake Equation was not meant as a tool to actually calculate how many alien civilizations there are, but rather is a tool to consider what information is needed in 58

Big History and astronomy

order to answer that question. At this point we are completely unable to estimate some of these variables. For example, until we actually start finding planets with evidence of life, any attempt to assign a value to fl (the fraction of planets on which life appears) is pure conjecture. However, in 1961 fp (fraction of stars with planetary systems) was also a complete unknown. The first exoplanet (a planet circling a star other than our own) was not confirmed until 1995 (the planet was named 51 Pegasus b), but since that time the number has skyrocketed, particularly since the launch of the Kepler space mission in 2009. As of the end of May 2018, NASA listed 3,730 confirmed exoplanets, in 2,783 star systems, with another 4,496 candidates awaiting confirmation. 929 of those planets are defined as “terrestrial” (as opposed to gas giants, etc.)6 TESS (Transiting Exoplanet Survey Satellite) was launched aboard a SpaceX ­Falcon 9 on April 18, 2018, and (at the time of writing this chapter) was on track to reach its target orbit by mid-June 2018. Once results from TESS start coming in, the count is likely to increase rapidly. Assuming that the James Webb Space Telescope ( JWST) is launched successfully in 2020, the number of confirmed planets will ­reliably rise into the hundreds of thousands. By that point it should become relatively simple to make a close estimate of the value of fp but even without that, we can now say that planets are very common, and the value is likely to be relatively high. The quality of the data that will be produced by TESS and JWST will also make it possible to reasonably estimate the value of ne, the number of planets with an environment suitable for life. The last factor (L) changes over time, because it really relates to how long humans have themselves had the technology to detect signals from other planets. Heinrich Hertz was the first man to transmit and receive controlled radio waves in the 1880s, and the use of radio did not become widespread until the twentieth century.7 Serious attempts at SETI (Search for Extra-Terrestrial Intelligence) have only been made since the 1960s. That means that in 2018, there is only a 60-year window in which we could have detected intelligent life. If a civilization rose and fell before that time so that its signals didn’t reach us in the correct window, we would never know. If a technological civilization arose at about the same time as ours, then it would have to be within 100 light years for us to be able to detect its emissions, as otherwise those signals wouldn’t have arrived here yet. But year by year, the viable window grows longer as we continue to search. Since the first four numbers in the Drake Equation are now known to be of reasonable size, it makes Fermi’s question increasingly puzzling. Either one or more of the final three variables must be zero (or so close to zero as makes no difference), or we have missed something crucial. We come back to the question, “where are they?” Arthur C Clarke is reputed to have said, “two possibilities exist: Either we are alone in the Universe or we are not. Both are equally terrifying.” Why are they terrifying questions, and not merely an abstract intellectual exercise? Why do these questions matter for big historians, if the ultimate purpose is to tell the story of how humans came to be here and now? As discussed below, the answer to this problem is incredibly important for understanding both our past and our future. Possible explanations for the Fermi paradox fall into two broad categories. The first is that there simply are no aliens to be discovered. The second is that they exist, 59

Jonathan Markley

but we can’t find them for various reasons. They might not exist because something prevents life from coming into existence or prevents life developing to a point where we could detect them. Alternately, we may simply be the first, perhaps because conditions prior to this did not allow for the emergence of a technological species (and this possibility suggests there may be other civilizations emerging at the same time as us). We might not be able to find them because they come and go too rapidly, and we have not been looking for a long enough time, or perhaps they exist but are too far away for us to find them. Perhaps civilizations get to a point where they no longer have any interest in expanding beyond their own star system. More frighteningly, there might be an advanced civilization intent on preventing any rivals from succeeding, and they hide in wait ready to destroy any newly emerging species. Perhaps we are simply looking for the wrong things, and advanced civilizations simply do not broadcast electro-magnetic signals. (With fiber-optic cables etc., our own planetary civilization has become less “noisy” and therefore it is more difficult for an Extra-­ Terrestrial intelligence to spot us.) In something called the “Zoo hypothesis” it has been suggested that they are there, they are watching us, but they are deliberately staying hidden so as not to interfere with our development. Numerous articles have been written on the topic of the Fermi paradox, and a good summary is by David Brin. Brin is better known as a science fiction author but has published numerous non-fiction works. His 1983 article on the “Great Silence” extends the discussion of the Fermi paradox into the idea of the Great Filter.8 One of the key aspects of the first possibility (that technological aliens do not exist) is the concept of the Great Filter. This term was introduced by Robin H ­ anson in 1998 in his online essay “The Great Filter – Are We Almost Past It?”9 This attempt to explain the Fermi paradox (sometimes also called “The Great Silence”) posits the idea that there must be certain bottlenecks that are incredibly hard (or even ­impossible) to pass through. Hanson suggested nine possible candidates: (1) (2) (3) (4) (5) (6) (7) (8) (9)

The right star system (including organics) Reproductive something (e.g. RNA) Simple (prokaryotic) single-cell life Complex (archaeatic & eukaryotic) single-cell life Sexual reproduction Multi-cell life Tool-using animals with big brains Where we are now Colonization explosion

It now appears that step 1 is probably relatively easy to satisfy. Step 2, the first biogenesis where “not life” becomes “life” for the first time, is a strong candidate. It is a step which is still largely inexplicable, and despite many attempts to formulate theories, no single view has succeeded in gaining much traction. Each of these steps is included in a standard big history approach, and hopefully it now becomes apparent why these topics are more important than merely being a “back-story” to the rise of the human race. The likelihood (or extreme improbability) of each of the above steps addresses the Fermi paradox, but it also addresses the survival prospects of our own species. 60

Big History and astronomy

If it is nearly impossible to pass through steps 1 to 7, then the human race has good future prospects. Life has already accomplished the most crucial steps, and the future is assured. One day human colony ships will depart our Solar System, and we will become the interstellar species we have been hunting for all along. Perhaps we will be the first, and perhaps we will be the only one. Arguably proponents of big history are best positioned of all to appreciate that this is in the very nature of our species. In the last few years with the discoveries of Denisovans and homo florensis in Indonesia, we have come to appreciate that the instinct and tendency to migrate huge distances pre-dates our own species and was also present in our older genetic relatives. We have known that modern humans spread from Africa to almost every landmass on Earth with the exception of Antarctica. No matter how difficult, humans have found a way to populate Australia, the Americas, etc. In a later epoch, Polynesians overcame astonishing distances to migrate across the Pacific Ocean. Given the technology available to them, this accomplishment was no less challenging than today’s proposed colonies on Mars championed by the likes of Elon Musk. In 2013 the Mars One Foundation called for volunteers for a one-way mission to colonize Mars in 2023. They received 200,000 volunteers.10 Media tended to focus on the one-way aspect of the mission as if this is something unusual in human history. The first humans to cross the Bering Strait land-bridge (who would become the ancestors of modern Native Americans) were also on a one-way mission. Few colonists voyaging to the Americas on the Mayflower can have expected to return to England. Given the weight of evidence over hundreds of thousands of years, we can surely say that once it becomes possible, humans will attempt to colonize other planets in our Solar System, and eventually other stars in the sky, no matter the danger and cost. In 2013, Stuart Armstrong and Anders Sandberg went so far as to argue that any civilization capable of colonizing other star systems would not find it too much of a stretch to extend their reach even into other galaxies. “This result implies that the absence of aliens is more puzzling than it would be if we simply considered our own galaxy. This makes the Fermi paradox more puzzling and more relevant to the future fate of humanity.”11 Elsewhere Armstrong stated, “almost any answer to the Fermi paradox gives rise to something uncomfortable.”12 A disturbing possibility presents itself. What if Hanson’s steps 1–7 are relatively common? We do not yet possess the technology to be able to detect if multi-cellular life has emerged on any of the exoplanets we have discovered. Even if the possibility of life successfully overcoming the first steps is as low as one in a thousand, or even one in a billion, then that would make simple life very common considering that there must be billions of suitable star systems. If that is true, that suggests step 8 or step 9 is the critical time at which intelligent life fails. If the Great Filter is still ahead of us, then the “Great Silence” suddenly becomes an ominous and terrifying threat. Perhaps all sufficiently advanced species learn to split the atom and eventually destroy themselves in global nuclear war. Perhaps it is the biological scientific revolution which makes it possible to create super-viruses that can wipe out all life. Perhaps instead of annihilating themselves, intelligent life simply runs out of resources before it can break out of its star system, or perhaps instead they run out of spiritual energy and die of simple ennui. Consider that our own species homo sapiens sapiens, even with 61

Jonathan Markley

recent discoveries pushing the date back, is only thought to have been in existence for about 300,000 years,13 yet any interstellar civilization would probably have to be millions of years old. Big historians must face the possibility that there may be no future ahead for the human race. Answering the Fermi paradox is therefore not just important for understanding the course of life elsewhere in the Universe, but also for judging our own chances of survival. The idea that the Great Filter might be ahead of us has led Nick Bostrom of the Future of Humanity Institute at Oxford University to write of the search for evidence for life on Mars, and the SETI project in general: I hope that our Mars probes will discover nothing. It would be good news if we find Mars to be completely sterile. Dead rocks and lifeless sands would lift my spirit… If we discovered traces of some simple life form… it would be bad news. If we found fossils of something more advanced… it would be very bad news… Scientifically interesting, certainly, but a bad omen for the future of the human race…. I conclude that the silence of the night sky is golden, and why, in the search for extraterrestrial life, no news is good news. It promises a potentially great future for humanity.14 David Brin in his guise as a science-fiction author has explored these ideas extensively, including in his 2012 novel Existence15 in which he describes a Universe dominated by the Great Filter. No advanced species ever survives in the long term, due to the combined impossibility of overcoming hostile competitors, cultural decline, exhaustion of resources, and the sheer impossibility of maintaining an advanced civilization for many millennia. He also explores the idea that the Great Silence isn’t the silence of loneliness, but a dark and forbidding silence in which every intelligent species is smart enough to stay quiet if they want to survive. These ideas are not mere science fiction, and Brin has taken a leading role in opposing projects to broadcast signals into space (METI = Message to Extraterrestrial Intelligence) as opposed to the passive search for evidence (SETI = Search for Extraterrestrial Intelligence). Amongst other questions, he asks16: If broadcasting is such a good idea, why aren’t other civilizations doing it? If we attract hostile attention, what could “they” do to us, worst-case? If broadcasting is potentially a bad idea, how can it be delayed long enough for further discussion? Historians of all fields will relate to Brin’s warning that technologically less advanced cultures tend to suffer devastating effects when encountering groups with better weapons, etc. Even if there are more advanced creatures out there, and even assuming they are not actively hostile, the consequences for humanity could be dire. Brin, perhaps optimistically, concluded his 1983 article (cited above) with the hope that perhaps older alien civilizations might deliberately conceal themselves and “speak softly, lest they disturb the infant’s extravagant and colourful time of dreaming.”17

62

Big History and astronomy

A possible solution to the Fermi paradox I would like to present a possible solution to the Fermi paradox that I believe is partially new. It contains elements of existing theories, including the idea that perhaps we are the first (or near first) intelligent species to emerge in the Universe because prior conditions did not allow for this to happen. It also builds on the idea of Panspermia first proposed by Svante Arrhenius in 1908.18 The idea is that life could be “seeded” on planets from space, spreading from planet to planet and star system to star system. We have one positive data point in this great mystery; one planet with life; one technologically advanced species. At this point, we ourselves are the only one to have been discovered.We also have one negative data point: billions and billions of planets, and no evidence of the presence of life (intelligent or otherwise) on any of them. It appears that the Drake Equation, however constituted, must lead to a number that is so tiny that it is very close to zero, but which also allows for a tiny chance of success that permits the existence of life on Earth. (If the number was actually zero it would mean that life is not possible anywhere in the Universe, including here. If it was anything greater than an infinitesimally small number, then there would be life teeming across the sky.) What is needed is an explanation that allows for intelligent life to exist but which requires an incredibly high threshold before it occurs. With somewhere between 100 to 200 billion stars in the Milky Way, and possibly many trillion galaxies in the ­Universe, a number as small as one in a billion would still be far too high. Given ­Armstrong and Sandberg’s argument that intergalactic spread is also possible, this means that the likelihood of intelligent technological life should be astonishingly small. They recognized this, stating that “the likelihood of intelligent life must be reduced by many orders of magnitude compared to previous arguments.”19 One common estimate is that once interstellar colonization begins, it would take less than four million years to colonize the entire Milky Way. For a species that is only a few hundred thousand years old this seems like a long time, but our planet is about 4.5 billion years old, and the Universe itself has been in existence close to 13.8 billion years. An alien species with only a tiny head-start (in relative terms) would already have completed this process before the emergence of complex life in the Cambrian Explosion, let alone the emergence of humanity. What possible explanation could explain a long delay in the emergence of intelligent life, and its extreme rarity? We need to find a process that is so utterly unlikely that it can almost never happen, but which still has some extremely small chance of occurring, perhaps a one in a trillion chance or even less. A possible answer was suggested by the work of Alexei Sharov and Richard ­Gordon.20 Sharov and Gordon argued that based on a statistical analysis, the life we have on Earth would have come into existence about 9.5 billion years ago. This is of course a major problem, since this predates the age of the Earth by 5 billion years. Their argument was inspired by Moore’s Law, a concept from computer ­technology. Gordon Moore, an early computing industry executive, made observations about the rate at which computing power was increasing, and various speeds were suggested. The term has now entered the English language such that the Meriam Webster

63

Jonathan Markley

Dictionary defines it as “an axiom of microprocessor development usually holding that processing power doubles about every 18 months especially relative to cost or size.” Sharov and Gordon reasoned (correctly) that you could extrapolate backwards using Moore’s Law to estimate that the microchip had been invented in the 1960s. They reasoned that the same methodology could be applied to the complexity of life, and this was the basis for their conclusion that the specific life forms in existence on Earth today must have origins pre-dating the formation of the Solar System (and therefore requiring a Panspermia solution to the problem of the origins of life on Earth).21 Assuming that very simple life might require around 10 billion years to emerge would certainly help solve the problem of why intelligent life has taken so long to appear in the Universe. It would take some time for the right types of stars, planets, and elements to appear in sufficient quantities. Assuming also that stars similar to our own are necessary for life, a ten-billion-year time frame is difficult to achieve, since stars like ours only last for about 10 billion years. As our sun ages, its temperature will increase, and conditions on Earth will become inhospitable long before the 10 ­billion year mark, possible as early as only 1 or 2 billion years from now. Larger stars complete their life cycle faster, and though red dwarf stars live considerably longer, this would require any planets to orbit much closer to their star in order to be within the habitable Goldilocks Zone, and this may well make them inhospitable because of their exposure to hard X-rays.22 It is worth pointing out as an aside that the reason we have to make so many assumptions about what is necessary for life is that we lack any basis of comparison. Is it necessary to have a planet the same size and composition as the Earth? Must it have a large moon? Must it be tilted on its axis? Can life only emerge around stars like our own? What are the exact parameters of an acceptable Goldilocks Zone? Is it enough to just have liquid water? Would some other liquid do, such as methane at ultra-low temperatures? Is a strong magnetic field essential? etc. Once (or if) we succeed in finding life elsewhere, we will be able to start drawing conclusions about what are the vital ingredients for life. It is very hard to draw conclusions based on a sample of one (life on Earth). The first step is that simple life begin on a planet roughly equivalent to Earth, around a star roughly equivalent to our own sun, and that this process begin about ten billion years ago and lasted for around five billion years. That simple life must then somehow be transferred to our own Solar System right at the beginning of its formation 4.5 billion years ago, and that this simple life somehow be seeded onto our own planet where it could continue the second half of its existence. The possibility that life might have started elsewhere is suggested by another great mystery: abiogenesis. How did life first begin? Many scenarios for the first emergence of life have been suggested, but none has gained widespread acceptance. How did chemistry become biology? Most hypotheses work on the assumption that life must have emerged in the conditions that were present in the first billion years of the Earth, with only rare exceptions. In 2013 Steven Benner argued that highly oxidized molybdenum was probably vital to the emergence of life, but the early Earth did not have sufficient oxygen available in the critical period, but early Mars at the same time probably did.23 According to Benner’s theory, life may have originated on 64

Big History and astronomy

Mars, and then was seeded via an asteroid collision that would have brought it to our own planet. This possibility certainly makes complex life less likely (starting on one planet further out in the Goldilocks Zone, and them being transferred to a fairly similar planet closer towards the sun), but it does not seem sufficiently improbable to satisfy the Drake Equation and the Great Silence. A starting point for life outside our Solar System would satisfy both problems: that elements be present other than were available on the early Earth, and that it requires a much longer time period than has to be assumed if life first emerged on our own planet. For this scenario to be plausible, it must be possible for life to emerge on an ­exoplanet, and then for biological material to be transferred from that alien star system into our own. It must be very unlikely (to satisfy the Drake Equation) but not actually impossible. In October 2017, the Pan-STARRS1 telescope in Hawaii spotted asteroid 1l/2017, now called ‘Oumuamua (“messenger from afar arriving first”). It was thought to have been the first interstellar object ever detected in our Solar System, but it is only a brief visitor. After slingshotting around our sun it is now outbound, and is already further from the sun than Jupiter (as of June 2018).24 It is estimated that ‘Oumuamua must have been in interstellar space for hundreds of millions of years, and it will soon return to that void. In May 2018, it was announced that ‘Oumuamua was not the first after all. ­Asteroid 2015 BZ509 (which was first detected in 2015) was determined to be another object that originated outside of our Solar System. Unlike ‘Oumuamua, 2015 BZ509 is a permanent visitor. It orbits around the sun in the opposite direction to all the planets, known as a retrograde orbit. Even more important, based on the way its orbit is linked to Jupiter, astronomers consider that it has been present in our Solar System for about 4.5 billion years, the period in which our system was first formed.25 While only two such objects have been identified to date, astronomers consider it likely that there are many more, and it is only now that our instruments have become sufficient sensitive to allow us to begin the process of identifying them. We do not have any clues about the origin points of either objects. What their presence does prove, however, is that it is possible for material from one star system to be transferred to another system. On the other hand, it seems less likely that biological material could be successfully transferred if the process required hundreds of millions of years. The nearest star system to our own is Proxima Centauri, which is about 4.25 light years away, or over 40 trillion kilometers. Proxima b, a rocky planet about 1.3 times the size of the Earth was discovered orbiting Proxima Centauri in 2016. Its orbit is much closer to its star than the Earth, but the relative coolness of Proxima Centauri means that liquid water is at least physically possible on Proxima b.26 Four light years seems too far away, and even if it is not, the closer the star, the greater the chance that a transfer of biological material could be accomplished. What is often not appreciated is that stars regularly come closer to the Solar System than Proxima Centauri. Our own system orbits the center of the Milky Way Galaxy, as do many millions of other systems. Just as planets within the Solar System move nearer or further away from the Earth as they orbit the sun,27 other star systems move in and out in relation to us as they follow their galactic orbits. Proxima Centauri is actually drawing closer to us, and over the course of the next 30,000 years the range will shrink by about 1light year, before it draws away again. About 10,000 years 65

Jonathan Markley

after Proxima Centauri’s closest approach, Ross 248 will approach slightly closer still. About 45,000 years from now Gliese 445 will also approach, giving us three separate star systems all closer than 4.25 light years (our current nearest galactic neighbor). None of them will approach closer than around 3 light years, which still seems too far. The star system Gliese 710 is currently 63 light years away, but in 1.35 million years it will come stunningly close to our Solar System. Initial estimates put it at about 0.8 light years at closest approach, but thanks to much more accurate data from the Gaia Space Telescope launched in December 2013, it is now estimated that at the point of its closest approach, it will be only 0.21 light years distant. It will be about three times brighter than Sirius. “After the passage of Gliese 710, up to 0.1% of all Oort Cloud comets might be removed from the cloud, and up to 0.01% might be pushed into potentially observable orbits.”28 This may seem like a tiny number, but there are huge numbers of comets in the Oort Cloud. In real terms, the authors estimate that this will mean approximately 10 comets every year continuing for around three to four million years. While not close enough to interfere with the orbits of planets around our sun, it is not only possible but likely that material will be shared between the two systems as they pass by each other. Gliese 710 is hardly unique. Scholz’s Star and its companion brown dwarf is now 20 light years away, but 70,000 years ago it approached to within 0.8 light years, close enough to interact with the Oort Cloud, though the authors of the study conclude that it would have “likely caused negligible impact on the flux of long-period ­comets.”29 These examples are enough to demonstrate that star systems can and do come close enough to the Solar System to interact and share material. The next question is whether simple living organisms, such as those that first inhabited the Earth, could survive a journey in space to reach our planet. Two studies describe experiments in which cyanobacteria (which are therefore similar to bacteria from the early Earth) were collected from a limestone cliff in Devon, England, and placed in space. In the first, they were launched by a Russian Soyez rocket and the bacteria were exposed to space for ten days. In the second, additional samples were placed outside the International Space Station for 548 days.30 Remarkably, some of the bacteria were able to survive. Even assuming that life could survive on a rock in space for an extended period of time, it does not prove that it would be able to survive the forces required to propel it into space from a planetary surface (such as a large asteroid collision) and to survive re-entry into the Earth’s early atmosphere. An experiment with just such a scenario in mind attempted to answer this question.31 The experiment involved firing frozen samples of a single-celled algae into water using a two-stage light gas gun. The authors concluded, “life forms that can serve as the base of a food chain and transform an environment making it suitable to life as we know it, could survive the ejection and re-impact onto a planetary body (Mars, the Moon, or Europa, for example), thus giving a foothold to life on another world.” Thus, it is possible that simple life could emerge around another star system, ­taking billions of years to reach the level of complexity as the organisms that inhabited the early Earth. Those organisms could then have been flung into space after a major impact on their home planet, and the rocks containing the dormant life forms could have been propelled into the outer edges of that star system. That 66

Big History and astronomy

star system could then have made a close fly-by of our own Solar System in its first billion years, and our Solar System could have captured that asteroid, which could have been flung into the inner system where it impacted on the early Earth, seeding our planet with the life, from which complex organisms such as ourselves could emerge after billions more years of development. Each one of these steps is improbable, but just slightly possible. When multiplied together, it makes it almost but not quite impossible. This is exactly the level of probability that would explain the Great Silence. Any scenario that has a realistic chance of occurring would happen far too often. Only a scenario that is very rare (perhaps one in a trillion chance?) could allow for the emergence of one intelligent species, and perhaps no more than one, out of all the trillions of stars in the sky. It does not exclude the possibility that there could be other species yet to be discovered, but they would likely be very rare and far away, and hence it would be natural that we had not been able to detect them yet.

Concluding thoughts This chapter has sought to present the argument that big historians should look to the heavens with more than just the goal of finding out where humanity came from. We should also realize that other star systems, and the potential for both simple and intelligent life in those star systems, are of critical importance to understanding our own place in the Universe, and for estimating our chances of survival in the future. Is the Great Filter ahead of us, or behind us? We should realize that we can never hope to answer many of the fundamental questions about why life exists on our planet until we have some basis of comparison. Otherwise it is nearly impossible to know which factor is critical, and which factor is just random happenstance. Finally, a new explanation for the Fermi paradox has been offered. Like all other explanations it must be acknowledged that it is mere speculation, but it is speculation based on one inescapable fact: we have not found any evidence of life elsewhere, and there must be something that makes it very improbable indeed. I hope my explanation may be close to the truth, because it would mean that the Great Filter is behind us. While we remain confined to a single planet, we continue to be threatened with the same fate as the dinosaurs. If we can look forward to a future of reaching for the stars and to eventually becoming an interstellar species, then we will have eliminated all but the most dire threats of extinction.

Notes 1 “Space is Big” is a reference to a larger quote from Douglas Adams’ Hitchhikers’ Guide to the Galaxy: “Space is big. You just won’t believe how vastly, hugely, mind-bogglingly big it is. I mean, you may think it’s a long way down the road to the chemist’s, but that’s just peanuts to space.” 2 Walter Alvarez, A Most Improbable Journey. A Big History of Our Planet and Ourselves, W.W. Norton, New York, 2017, p. 4. 3 Eric J. Chaison, “Cosmic Evolution – More Than Big History by Another Name.” ­Evolution. A Big History Perspective, edited Leonid E. Grinin, Andrey V. Korotayev, Barry H. Rodrigue, Uchitel Publishing House, Volgograd, 2011, p. 38. 67

Jonathan Markley

4 Eric M. Jones, “‘Where is everybody?’ An Account of Fermi’s Question.” Los Alamos ­National Laboratory, 1985, pp. 1–12. The story is based entirely on oral testimony of several men who were present at the conversation, and none of them were certain of the exact date. 5 https://www.seti.org/drakeequation. 6 https://exoplanets.nasa.gov/. 7 My grandfather owned a radio shop for many years in Wellington, the capital of New ­Zealand. In Feb 1933, one of his newspaper advertisements asked potential customers, “Good Evening! Have you realised what pleasure you are missing when you don’t own a radio?” Evening Post column 1, p. 9, 4 Feb 1933. Even then many households did not yet own a radio. 8 Glen David Brin, “The ‘Great Silence’: The Controversy Concerning Extraterrestrial Intelligent Life” Journal of the Royal Astronomical Society, 24, 1983, pp. 283–309. See also Tim Urban’s blog Waitbutwhy. His essay length post about the Fermi paradox is written in a casual tone (including two uses of the phrase ‘we’re f**cked’ to describe some of the more frightening possible answers to the paradox) but it gives an accessible introduction to the main concepts and I have in fact assigned it as course readings in my Big History course. https://waitbutwhy.com/2014/05/fermi-paradox.html 9 https://mason.gmu.edu/~rhanson/greatfilter.html. 10 https://www.space.com/22758-mars-colony-volunteers-mars-one.html. 11 Stuart Armstrong & Anders Sandberg, “Eternity in six hours: Intergalactic spreading of intelligent life and sharpening the Fermi paradox.” Acta Astronautica 89, Aug–Sep 2013, pp. 1–13. 12 https://phys.org/news/2013-08-silence-skybut.html. 13 Jean-Jacques Hublin, Abdelouahed Ben-Ncer, Shara E. Bailey, Sarah E. Freidline, Simon Neubauer, Matthew M. Skinner, Inga Bergmann, Adeline Le Cabec, Stefano Benazzi, Katerina Harvati & Philipp Gunz, “New fossils from Jebel Irhoud, Morocco and the pan-African origin of Homo sapiens” Nature 546, June 2017, pp. 289–292. 14 Nick Bostrom, “Where are they? Why I hope the search for extraterrestrial life finds nothing.” MIT Technology Review May/June 2008, pp. 72–77. https://www.technology review.com/s/409936/where-are-they/. 15 David Brin, Existence, Tor, New York, 2012. 16 http://www.davidbrin.com/nonfiction/setisearch.html. 17 Brin “The Great Silence” p. 307. 18 Svante Arrhenius, Worlds in the Making: The Evolution of the Universe, Harper & Row, New York, 1908. 19 Armstrong and Sandberg, “Eternity in six hours”. 20 A. A. Sharov, “Genome increase as a clock for the origin and evolution of life” ­Biology Direct, 1, June 2006, pp. 17–26; A. A. Sharov & R. Gordon, “Life before Earth” which was issued as a “pre-publication article” in 2013 https://arxiv.org/ftp/arxiv/­ papers/1304/1304.3381.pdf. Their ideas have been subject to heavy criticism. See for example the reviewers’ comments for the earlier article, and for the second one, the rebuttal by Caren Marzban, Raju Viswanathan and Ulvi Yurtsever, “Earth before life”, B ­ iology Direct 9, no.1, Dec 2014, https://link.springer.com/article/10.1186/1745-6150-9-1# citeas. Interestingly, none of the critics reject the idea of Panspermia and some expressly state that it is a viable hypothesis despite the flaws within the articles under review. A modified version of the article “Life before Earth” appears in a chapter of the book ­Habitability of the Universe Before Earth: Exploring Life on Earth and Beyond, ed. R. ­Gordon & A. Sharov, Amsterdam, Elsevier, 2018. This book contains many chapters dealing with concepts relevant to this chapter, including chemical conditions for life, the concept of Panspermia, etc.

68

Big History and astronomy

21 I wish to stress that I do not necessarily accept the conclusions of Gordon and Sharov. Criticism of their work leads me to the conclusion that their claims are unproven. Rather, their argument was the inspiration that started me thinking about a possible solution for the Fermi paradox. 22 https://www.ewass.ras.ac.uk/10-x-rays-could-sterilise-alien-planets-in-otherwisehabitable-zones. 23 Steven Benner, paper presented at the Goldschmidt Geochemistry Conference, ­F lorence Italy, August 29th, 2013. See also his contribution (as well as other chapters) in Frontiers of Astrobiology, Cambridge, Cambridge University Press, 2012. 24 https://solarsystem.nasa.gov/small-bodies/asteroids/oumuamua/in-depth/. 25 F. Namouni & M. H. M. Morais, “An interstellar origin for Jupiter’s retrograde co-­ orbital asteroid.” Monthly Notices of the Royal Astronomical Society: Letters 477, no.1, 11 June 2018, pp. L117–L121, https://doi.org/10.1093/mnrasl/sly057. 26 https://www.space.com/33838-planet-proxima-b-full-coverage.html. 27 For example, the distance to Mars varies from 54.6 million km to 401 million km ­depending on where the Earth and Mars are as they orbit the sun. 28 F. Berski & P. A. Dybczynski, “Gliese 710 will pass the Sun even closer Close approach parameters recalculated based on the first Gaia data release.” Astronomy & Astrophysics, 595, Nov. 2016. https://doi.org/10.1051/0004-6361/201629835. See also, V. Bobylev, “Stars outside the Hipparcos list closely encountering the Solar system.” Astronomy ­L etters, 36, no.11, Nov 2010, pp. 816–822. 29 Eric E. Mamajek, Scott A. Barenfeld, Valentin D. Ivanov, Alexei Y. Kniazev, Petri Väisänen, Yuri Beletsky, Henri M. J. Boffin, “The Closest Known Flyby of a Star to the Solar System.” The Astrophysical Journal Letters 800, no.1, Feb 12, 2015. 30 Karen Olsson-Francis, Rosa de la Torre, Charles S. Cockell, “Isolation of Novel ­Extreme-Tolerant Cyanobacteria from a Rock-Dwelling Microbial Community by ­Using Exposure to Low Earth Orbit.” Applied and Environmental Microbiology, 76, ­no. 7, Apr. 2010, pp. 2115–2121; Charles S. Cockell, Petra Rettberg, Elke Rabbow, Karen Olsson-Francis, “Exposure of phototrophs to 548 days in low Earth orbit: microbial selection pressures in outer space and on early earth.” The ISME Journal 2011, 1–12. In a Freudian slip, the story at the science news website PhysOrg.com called the lead scientist “Charles Cockwell.” https://phys.org/news/2010-08-microbes-survive-year-space.html 31 D.L.S. Pasini, M.C. Price, M.J. Burchell, M J. Cole, “Survival of Nannochloropsis Phytoplankton in Hypervelocity Impact Events up to Velocities of 6.07 KM S.” EPSC Abstracts 8, European Planetary Science Congress 2013. https://meetingorganizer. copernicus.org/EPSC2013/EPSC2013-396.pdf.

Bibliography Alvarez, Walter A Most Improbable Journey. A Big History of Our Planet and Ourselves. New York, W.W. Norton, 2017. Armstrong, Stuart, & Sandberg, Anders “Eternity in Six Hours: Intergalactic Spreading of Intelligent Life and Sharpening the Fermi Paradox.” Acta Astronautica 89, Aug–Sep 2013, pp. 1–13. Arrhenius, Svante Worlds in the Making: The Evolution of the Universe. New York, Harper & Row, 1908. Berski, Filip, & Dybczynski, Piotr A. “Gliese 710 will Pass the Sun Even Closer Close Approach Parameters Recalculated based on the First Gaia Data Release.” Astronomy & Astrophysics 595, Nov. 2016. doi:10.1051/0004-6361/201629835

69

Jonathan Markley

Bobylev, B. “Stars Outside the Hipparcos List Closely Encountering the Solar System.” Astronomy Letters 36, no. 11, Nov 2010, pp. 816–822. Bostrom, Nick “Where Are They? Why I Hope the Search for Extraterrestrial Life Finds Nothing.” MIT Technology Review May/June 2008, pp. 72–77. www.technologyreview. com/s/409936/where-are-they/ Brin, David “The ‘Great Silence’: The Controversy Concerning Extraterrestrial Intelligent Life” Journal of the Royal Astronomical Society, 24, 1983, pp. 283–309. Brin, David Existence, New York, Tor, 2012. Chaison, Eric J. “Cosmic Evolution – More Than Big History by Another Name.” In ­Evolution. A Big History Perspective, edited Leonid E. Grinin, Andrey V. Korotayev, Barry H. Rodrigue, Volgograd, Uchitel Publishing House, 2011, pp. 37–48. Cockell, Charles S., Rettberg, Petra, Rabbow, Elke, Olsson-Francis, Karen “Exposure of Phototrophs to 548 Days in Low Earth Orbit: Microbial Selection Pressures in Outer Space and on Early Earth.” The ISME Journal 5, 2011, 1671–1682. Gordon, Richard, & Sharov, Alexei, eds., Habitability of the Universe Before Earth: Exploring Life on Earth and Beyond, Amsterdam, Elsevier, 2018. Hublin, Jean-Jacques, Ben-Ncer, Abdelouahed, Bailey, Shara E., Friedline, Sarah E., ­Neubauer, Simon, Skinner, Matthew M., Bergmann, Inga, Le Cabec, Adeline, Benazzi, Stefano, Harvati, Katerina, & Gunz, Philipp “New Fossils from Jebel Irhoud, Morocco and the pan-African Origin of Homo Sapiens” Nature 546, June 2017, pp. 289–292. Impey, Chris, Lunine, Jonathan, & Funes, José, eds, Frontiers of Astrobiology, New York, Cambridge University Press, 2012. Jones, Eric M. ‘Where Is Everybody?’ An Account of Fermi’s Question, Los Alamos, Los A ­ lamos National Laboratory, 1985, pp. 1–12. Mamajek, Eric E., Barenfeld, Scott A., Ivanov, Valentin D., Kniazev, Alexei Y., Väisänen, Petri, Beletsky, Yuri, & Boffin, Henry M.J. “The Closest Known Flyby of a Star to the Solar System.” The Astrophysical Journal Letters 800, no.1, 4 Feb 12, 2015. Marzban, Caren, Viswanathan, Raju, & Yurtsever, Ulvi “Earth before life”, Biology Direct 9, no. 1, Dec 2014, https://link.springer.com/article/10.1186/1745-6150-9-1#citeas Namouni, Fathi, & Morais, Maria Helena Moreira “An Interstellar Origin for Jupiter’s Retrograde Co-Orbital Asteroid.” Monthly Notices of the Royal Astronomical Society: ­L etters 477, no. 1, 11 June 2018, pp. L117–L121, doi:10.1093/mnrasl/sly057 Olsson-Francis, Karen, de la Torre, Rosa, & Cockell, Charles S. “Isolation of Novel ­Extreme-Tolerant Cyanobacteria from a Rock-Dwelling Microbial Community by Using Exposure to Low Earth Orbit.” Applied and Environmental Microbiology 76, no. 7, Apr. 2010, pp. 2115–2121. Pasini, D.L.S, Price, M.C., Burchell, M.J., & Cole, M.J. “Survial of Nannochloropsis ­Phytoplankton in Hypervelocity Impact Events up to Velocities of 6.07 KM S.” EPSC Abstracts 8, European Planetary Science Congress 2013. https://meetingorganizer.copernicus.org/EPSC2013/EPSC2013-396.pdf Sharov, Alexei A. “Genome Increase as a Clock for the Origin and Evolution of Life.” Biology Direct, 1, June 2006, pp. 17–26. Sharov, Alexei A., & Gordon, Richard “Life before Earth”: A “Pre-Publication Article.” 2013. https://arxiv.org/ftp/arxiv/papers/1304/1304.3381.pdf

Websites www.davidbrin.com/nonfiction/setisearch.html https://exoplanets.nasa.gov/ https://mason.gmu.edu/~rhanson/greatfilter.html 70

Big History and astronomy

https://phys.org/news/2010-08-microbes-survive-year-space.html https://phys.org/news/2013-08-silence-skybut.html https://solarsystem.nasa.gov/small-bodies/asteroids/oumuamua/in-depth/ https://waitbutwhy.com/2014/05/fermi-paradox.html www.ewass.ras.ac.uk/10-x-rays-could-sterilise-alien-planets-in-otherwise-habitablezones www.seti.org/drakeequation www.space.com/22758-mars-colony-volunteers-mars-one.html www.space.com/33838-planet-proxima-b-full-coverage.html

71

4 BIG HISTORY AND MACRO-EVOLUTION Evolutionary principles and mechanisms at biological and social phases of the big history Leonid E. Grinin, Andrey Korotayev and Alexander Markov Introductory remarks In the recent decades the research of the Universe as a whole has made a significant progress. This is especially clearly manifested in the development of big history (Christian, Brown and Benjamin 2013; Grinin et al. 2014; Spier 2015; Rodrigue, ­Grinin and Korotayev 2015, 2016, 2017; Christian 2017). Big History provides a unique opportunity to consider the development of the Universe as a single process. big history approach regards the process of evolution as a continuous and unified process from the origins of the Universe to the present state of humankind and further to the future. Cosmic, geological/biological and social phases of universal evolution have a genetic and structural continuity. The importance of this approach is doubtless. It aims to cover within a single theoretical framework all phases of the universal evolution from the Big Bang to the forecasts of the foreseeable future, to show that the present state of the humankind is the result of the process of self-organization of the matter. Within big history researchers distinguish some main evolutionary laws and principles (e.g., concerning energy and complexity). However, it is very important to recognize that there are many more such integrating principles, laws, mechanisms and patterns of evolution at all its levels than it is usually supposed. In our chapter, we will demonstrate that one can find common traits in development, functioning and interaction of apparently rather different processes and phenomena of Big History if we regard this in connection with the macro-evolutionary paradigm. On the one hand, big history is very similar in content to a coherent description of the universal macro-evolutionary process, because it shows the formation of new ­levels of complexity in the process of historical development of the Universe. At the same time, the general features of the development, operation and interaction are found in many seemingly dissimilar processes and phenomena, showing a particular aspect of universal evolution as real similarities that occur in a variety of

72

Big History and macro evolution

manifestations at all levels and in all lines. Therefore, the combining of the potential of the big history with evolutionary approaches can open new horizons in this direction. We plan to analyze the biological and social phases of the big history and ­macro-evolution to give ideas about the driving forces and patterns of transition from one phase to another.We will discuss in common, i.e., cross phasing, evolutionary and big history rules, principles, patterns and laws; at the same time, we will highlight the peculiarities of realization of those rules at each phase. We believe that it is possible to identify a considerable number of such similarities and to group them into large blocks. In particular, we will consider such features as: • • • • • •

Ability for self-preservation and self-organization. Law of life cycle of various objects. Rule of evolutionary block assemblage. Unevenness and catastrophes. Typical and unique objects. Recombination or circulation of matter in nature.

Comparison between different phases of big history is especially important, because such an analysis suggests new promising possibilities to deepen our understanding of the course, trends, mechanisms and peculiarities of different phases of big history. Each subsequent phase of big history is accompanied by the emergence of new evolutionary mechanisms. For example, certain prerequisites and respective preadaptations can be normally detected within the previous phase; emergence of new evolutionary mechanisms does not imply the invalidation of those active during the previous phases. One can observe the emergence of a complex system of interaction between ­various forces and mechanisms determining the evolution of new forms. For example, biological organisms operate in the framework of certain physical, chemical and geological laws; behavior of social systems and people has certain biological ­limitations and so on. Some models (similar in principle) may emerge not only in the breakthrough points, changing the main line of evolution, but also in some directions that may be regarded as dead ends (from the big history leading line perspective). For example, the emergence of social life forms took place in many phyla and classes – from bacteria to insects, birds and mammals. Additionally, among insects, we can find rather high forms of social organization. Thus, in our chapter, we plan to give overview of the perspective big history brings to bear on evolution on the macroscale. In our chapter, we will attempt at combining big history potential with the potential of Evolutionary Studies. Such an approach opens up some new perspectives for our understanding of evolution and big history, their driving forces, vectors and trends; it creates a consolidated field for interdisciplinary research. In this chapter, we analyze the similarities and differences between social and biological evolution. Since the comparison of biological and social evolution is an important but (unfortunately) understudied subject, we shall re-state a few of the salient points from our previous article.

73

Leonid E. Grinin et al.

We are still at the stage of a vigorous discussion about the applicability of  ­Darwinian evolutionary theory to social/cultural evolution. Unfortunately, we all are mostly dealing with a polarization of views. On the one hand, we confront a total rejection of Darwin’s theory of social evolution (see, e.g., Hallpike 1986). On the other, we deal with those who stress that cultural evolution demonstrates all the key Darwinian evolutionary characteristics (Mesoudi, Whiten and Laland 2006). We believe that, instead of following the outdated objectivist principle of   ‘either – or’, we should concentrate on the search for methods that could allow us to apply achievements of biological evolutionary science to social evolution and vice versa. In other words, we should concentrate on the search for productive generalizations and analogies for analysis of evolutionary mechanisms. The big history approach aims for inclusion of all mega-evolution within a single paradigm.1 Hence, this approach provides an effective means to address the above-mentioned task. As is known, not only systems evolve, but mechanisms of evolution evolve too (see more on this in section ‘Some preconditions of the transition from biological to social phase of the big history’). This concept also appears rather fruitful as regards the ­development of big history itself. Let us consider some of the parameters and examples that we might consider. Each sequential phase of big history is accompanied by the emergence of new evolutionary mechanisms; therefore, certain prerequisites and preadaptations can be detected within the previous phase. So, development of new mechanisms of evolution does not imply invalidation of evolutionary mechanisms that were active during previous phases. As a result, one can observe the emergence of a complex system of interaction of forces and mechanisms determining the evolution of new forms. Biological organisms operate in the framework of certain physical, chemical and geological laws (see Kutter 2015 on this topic and also on the comparison between physical and biological evolution). Likewise, the behaviors of social systems and people have certain biological limitations. New forms of evolution that determine big history transition into a new phase may result from activities going in different directions. Some evolutionary forms that are similar in principle may emerge not only at a breakthrough point, but may also result in a dead end – from the overall view of big history. For example, the emergence of social forms of life took place in many phyla and classes – bacteria, insects, birds and mammals. Additionally, among insects, we can find rather high forms of socialization (see, e.g., Reznikova 2011; Ryabko and Reznikova 2009; Robson and Traniello 2002). Despite the common trajectory and interrelation of social behaviors by these various life forms, there has been a large overall difference in the impact that each has had on the Earth. What is more, as regards information transmission mechanisms, it appears possible to speak about certain ‘evolutionary freaks’. Some of those mechanisms (in particular, the horizontal exchange of genetic information) were spread rather widely in the biological evolution of simple organisms but were later discarded (or transformed into highly specialized mechanisms, e.g., sexual reproduction) among more complex organisms. Today, they are mostly confined to the simplest forms of life. We mean the horizontal exchange of genetic information (genes) among microorganisms, which makes many useful genetic ‘inventions’ literally a sort of ‘commons’ of microbe 74

Big History and macro evolution

communities. Among the bacteria, the horizontal transmission of genes contributes to the fast development of antibiotic resistance (e.g., Markov and Naymark 2009). For the present chapter, the following turns out to be important: The horizontal exchange of genetic information (in its general function) is distantly similar to those forms of information exchange that became extremely important for social evolution – the direct borrowing of innovations and their introduction into social life. Hence, principles and mechanisms that appear as marginal relevance at a certain phase of big history may turn out to be extremely important in a later phase.2 These parallels suggest that analysis of similarities and differences between the mechanisms of evolution may help us to understand the general principles of megaevolution and big history in a much fuller way.3 They may also help us to better understand their driving forces and supra-phase mechanisms (mechanisms that ­operate in two or more phases of big history) (see also Grinin, Markov and K ­ orotayev 2011; also Grinin and Korotayev 2008, 2009a, 2009b; Grinin, Markov and Korotayev 2009a, 2009b; Korotayev 2018). Let us return now to a comparison of biological and social evolution. It is important to describe similarities and differences between these two types of macro-evolution – at various levels and in various aspects. This is necessary because such comparisons tend to be deformed by conceptual extremes and tend to be incomplete.4 These limitations are true even in respect of the above-mentioned paper by Mesoudi,Whiten and Laland (2006), as well as a rather thorough monograph by Christopher Hallpike, Principles of Social Evolution (1986). There, Hallpike offers a fairly complete analysis of similarities and differences between social and biological organisms, but does not provide a clear and systematic comparison between social and biological evolution.

Biological and social organisms: a comparison at various levels of evolution There are a few important and understandable differences between biological and social macro-evolution, nonetheless, it is possible to identify a number of fundamental similarities. One may single out at least three basic sets of shared factors: • •



First of all, there are similarities that stem from very complex, non-equilibrium, but stable systems whose principles of function and evolution are described by General Systems Theory, as well as by a number of cybernetic principles and laws. Secondly, we are not dealing with isolated systems but with a complex interaction between organisms and their external environment. As a result, the reaction of systems to external challenges can be described in terms of general principles that express themselves within a biological reality and a social reality. Thirdly, it is necessary to mention a direct ‘genetic’ link between the two types of macro-evolution and their mutual influence.

It is important to emphasize that similarity between the two types of macro-­ evolution does not imply commonality. Rather significant similarities are frequently accompanied by enormous differences. For example, the genomes of chimpanzees and the ­humans are 98% similar. However, there are enormous intellectual and social 75

Leonid E. Grinin et al.

differences between chimpanzees and humans that arise from the apparently ‘insignificant’ variations between the two genomes.5 It appears reasonable to continue the comparison between the two types of ­macro-evolution on the basis of the analysis used by Hallpike (1986: 33), who singles out the following similarities between social and biological organisms: (1) The institutions of societies are interrelated in a manner analogous to the organs of the body and preserve their continuity despite changes of individual membership, just as individual cells are renewed in organs. (2) There is a specialization of organic functions analogous to the social division of labor. (3) In both cases self-maintenance and feedback processes occur. (4) There are adaptive responses to the physical environment. (5) In organisms, we find the transmission of matter, energy and information analogous to trade, communication, etc., in societies. According to Hallpike (1986: 33–34), societies are unlike organisms in the ­following respects: (1) They are not physical entities at all, since their individual members are linked by information bonds, not by those of a purely physical nature. (2) Societies are not clearly bounded, for example, two societies may be distinct politically, but not culturally or religiously. (3) Societies do not reproduce, so that cultural transmission from generation to generation is indistinguishable from general processes of self-maintenance.6 (4) Societies are capable of metamorphosis to a degree only found in organic phylogeny. (5) The individual members of a society, unlike cells, are capable of acting with purpose and foresight, and of learning from experience. (6) Structure and function are far less closely related in societies than in organisms. Hallpike also comes to the sound conclusion that similarities between social and biological organisms are in general determined by similarities in organization and structure (we would say similarities between different types of systems). As a result, Hallpike believes that one can use certain analogies when institutions can be represented as similar to some organs. In this way, cells may be regarded as similar to individuals; central government similar to the brain and so on. Spencer (1898) and Durkheim (1893/1991) are important representatives of this tradition (see also ­Heylighen 2011). Hallpike also has sufficient grounds to add Alfred Radcliffe-Brown and Talcott Parsons. When comparing biological species and societies, Hallpike (1986: 34) singles out the following similarities: (1) ‘Species, like societies, do not reproduce.’ (2) ‘Both have phylogenies and metamorphosis.’ (3) ‘Both are composed of competing individuals.’ 76

Big History and macro evolution

He also singles out the following difference: ‘Unlike species, however, societies are organized systems, whereas species are simply collections of individual organisms’ (Hallpike 1986: 34). Further, Hallpike tries to demonstrate that, because of such differences between biological and social organisms, the very idea of natural selection does not appear to be very productive with respect to social evolution.7 We believe that his proofs are not very convincing, although they make some sense in certain respects. In addition, his analysis is concentrated mostly at the level of an individual organism and an i­ndividual society. He hardly moves at the supra-organism level (though he, of course, discusses the evolution of species). We believe that with this, Hallpike (­notwithstanding his desire to demonstrate the sterility of the application of Darwinian theory to social evolution) involuntarily amplifies the effect of similarity between biological and ­social evolution, because the analogy between the biological organism and society (as Hallpike admits himself) is rather salient indeed. On the other hand, Hallpike does not take into account the point in social evolution where a few substantially new supra-socium levels of development emerge. We contend that it is very important to consider not only evolution at the level of a society but also at the level above individual societies, as well as the point at which both levels are interconnected. The supra-organism level is very important, as regards biological evolution (but, perhaps, less so in respect to social evolution).Thus, it might be more productive to compare societies with ecosystems rather than with organisms or species. However, this would demand the development of special methods, as in this case it would be necessary to consider the society not as a social organism, but as a part of a wider system, which includes the natural and social environment. We identify the following differences between the social and biological evolution:

At the level of an individual society and an individual biological organism (1) As Hallpike (1986: 33) notes, societies are capable of such rapid evolutionary metamorphoses that they were not observed in the pre-human organic world. However, social systems are not only capable to change and transform, they are also capable to borrow innovations and new elements. (2) They may be also transformed consciously and with a certain purpose. Such characteristics are absent in natural biological evolution in any form. (3) In the process of social evolution the same social organism may experience radical transformation more than once. (4) Key information transmission differs significantly in biological and social evolution (we shall consider this point in more detail in section ‘As regards the results of social/natural selection’). (5) In biological evolution, the acquired characteristics are not inherited; thus, they do not influence the biological evolution that proceeds very slowly. This point will be also considered in more detail in section ‘As regards the results of social/ natural selection.’ (6) It appears very important to note that, though biological and social organisms are significantly (actually ‘systemically’) similar, they are radically different in their 77

Leonid E. Grinin et al.

capabilities to evolve.The biological organism does not evolve by itself; evolution may only take place at a higher level (population, species, etc.), whereas social evolution can often well be traced at the level of an individual social organism. What is more, it is frequently possible to trace the evolution of particular institutions and subsystems within a social organism.

As regards the results of social/natural selection (1) Biological evolution is more additive (cumulative) than substitutive; put in another way: ‘the new is added to the old’. In contrast, social evolution (especially during the two recent centuries) is more substitutive than additive: ‘the new replaces the old’ (Grinin, Markov and Korotayev 2008, 2011). (2) Since social evolution is different from biological evolution, in respect of mechanisms of emergence, fixation and diffusion of evolutionary breakthroughs (aromorphoses), this leads to long-term restructuring in size and complexity of social organisms. It is important to note that, in contrast to biological evolution, where some growth of complexity is also observed, such social reorganization becomes continuous. In recent decades, societies that do not experience a constant and significant evolution look inadequate and risk extinction. In addition, size of societies (and systems of societies) tends to grow constantly through more and more tightly integrative links (this trend has become especially salient in recent millennia), whereas the trend toward increase in the size of biological organisms in nature is rather limited and far from general. (3) Within social evolution, we observe the formation of special suprasocietal systems that also tend to grow in size. This can be regarded as one of the results of social evolution and serves as a method of aromorphosis fixation and diffusion.

At supra-organic (suprasocietal) level As a result of the above-mentioned differences, within the process of social evolution, we observe the formation of two types of special suprasocietal systems: (A)  ­amalgamations of societies with varieties of complexity that have analogies to biological evolution; (B) emergence of elements and systems that do not belong to any society, in particular that lack many analogies to biological evolution. Naturally, type B needs a special comment.The first type of supra-organic amalgamation is rather typical not only for social but also for biological evolution.8 However, within biological evolution, amalgamations of organisms with more than one level of organization are usually very unstable and are especially unstable among highly organized animals.9 Within social evolution, we observe the emergence of more and more levels: from groups of small sociums to humankind as a whole. Of course, it makes sense to recollect analogies with social animals: social insects, primates and so on. Neither should we forget to compare society with the individual biological organism but also with groups of organisms bound by cooperative relationships. Such groups are widely present among bacteria and even among viruses. It should be noted that modern biologists have developed well-respected theories that account for the emergence of intragroup cooperation and altruism, including 78

Big History and macro evolution

competition, kin selection, group selection and so on (see, e.g., Reeve and Hölldobler 2007). However, it is not clear if societies should be really compared with groups of organisms rather than individual organisms, whether we should not consider societies within the system of numerous intersocietal links? In any case, it is clear that the level of analysis is very important for comparison of biological and social evolution. Which systems should be compared? Such analogies are more frequent when society (the social organism) is compared to a biological organism or species. However, in many cases, it may turn out to be more productive to compare societies with other levels of the biota’s system organization: with populations, ecosystems and communities, with particular structural elements or blocks of communities (e.g., with particular fragments of trophic networks or with particular symbiotic complexes), with colonies (with respect to colonial organisms), or finally – and this is the closest analogy – with groups of highly organized animals (cetaceans, primates and other social mammals or termites, ants, bees and other social insects). Thus, here we are confronting a rather complex and hardly studied methodological problem: Which levels of biological and social processes are most congruent? What are the levels whose comparison could produce the most interesting results? In general, it seems clear that such an approach should not be a mechanical equation of ‘social organism = biological organism’ at all times and in every situation. The comparisons should be operational and instrumental. That means that we should choose the scale and level of social and biological phenomena, forms and processes that are adequate for their respective tasks. We would say again that sometimes it is more appropriate to compare an individual biological organism with a society, whereas in other cases, it could well be more appropriate to compare a society with a community (of, say, ants or bees), a colony, a population or a species. We believe that the issue of the ‘presence’ of the social ontogenesis (and its comparison with the biological ontogenesis) should be studied in this framework (see Grinin, Markov and Korotayev 2008: Chap. 6 for more details). However, there are some cases when it seems more appropriate to compare social ontogenesis with biological phylogenesis. Hence, different scales and types of scientific problems need special approaches. This subject will be discussed further in the subsequent section of the present article.

Similarities and differences at the level of evolutionary mechanisms Biological and social aromorphoses In certain respects, it appears reasonable to consider biological and social macro-­evolution as a single macro-evolutionary process. This implies the necessity to comprehend the general laws and regularities that describe this process, though their manifestations may display significant variations, depending on properties of a concrete, evolving entity (­biological or social).We believe that many similarities and differences in laws and driving forces in the biological and social phases of big history can be comprehended more effectively if we apply the concepts of biological and social aromorphosis. For a detailed consideration of aromorphoses and their regularities, see our other contribution to this volume (see also Grinin, Markov and Korotayev 2011). 79

Leonid E. Grinin et al.

The application of the notion of biological and social aromorphosis has helped us to detect a number of regularities and rules that are common for biological and social evolution – ‘payment for the arogenic progress’, ‘special conditions for the aromorphosis emergence’ and so on. Such rules and regularities are similar for both biological and social phases of big history. However, we shall not analyze them in the present article.

On the peculiarities of key information transmission at various phases of big history Replication on the basis of the matrix principle is a fundamental feature of all forms of life (see, e.g.,Timofeev-Ressovsky,Vorontsov and Yablokov 1969: 15–16). However, the process of such replication cannot be conducted with a 100% accuracy; hence, the replication of a complete genome without any errors is virtually impossible. That is why the emergence of practically any new biological organism is accompanied by random change in genes (i.e., mutations). However, a significant change of the genotype occurs extremely rarely. Yet, the role of mutations in biological evolution is extremely important and very well known, because the mutations are one of the main sources providing ‘raw materials’ for evolution (Timofeev-Ressovsky,Vorontsov and Yablokov 1969: 72). However, there is also an opinion that the importance of mutations for evolution has been exaggerated, whereas the main source of new genetic material for major morphobiological reorganizations was provided by the gene duplication (see, e.g., Shatalkin 2005: 30). The gene duplication may indeed be a source of new material; yet the studies that try to prove that the morphobiological reorganizations are, first of all, results of duplications that have been conducted just for 15 years, and at the moment we are rather dealing with accumulation of data in this field, that is why we still prefer to keep to the classical point of view on the role of mutation in the process of biological evolution. However, it is important to emphasize that the number of distortions by which transmission of information is accompanied from generation to generation within social evolution (especially in complex societies) is orders of magnitude higher than that observed within biological evolution. There are grounds to maintain that the role of such ‘distortions’ in social macro-evolution tends to increase (in addition to conscious and purposeful alteration of cultural information). In the meantime, it appears that we observe just the opposite within biological macro-evolution. For example, among viruses and some ­bacteria, mutational variability is constantly necessary for their mere survival; on the other hand, in complex biological organisms, it is necessary only up to a very limited extent. Within social evolution, some unconscious distortion of transmitted cultural information always takes place, which may be regarded to some extent as analogous to biological mutations.10 This, by itself, may lead to certain socio-evolutionary shifts (Korotayev 1997, 2003; Grinin and Korotayev 2007b, 2009b). However, the conscious directed alteration of the information by its carriers is significantly more important. Though many are still sure that ‘history never teaches anything to anybody’, already the elites of many complex Agrarian societies quite often tried to take into account the errors made by their predecessors and to modify the ‘socio-cultural genotype’ accordingly in order to avoid them in future. 80

Big History and macro evolution

One may recollect, for example, the conscious alteration of the social position of the military elite by the founders of the Sung dynasty in China (960–1279 CE), in order to prevent the military coups that jeopardized the political stability of their predecessors (Wright 2001). Similarly, there was the conscious and purposeful replacement of traditional military systems with the modernized military systems of  Western Europe by Peter the Great in Russia, Muhammad Ali in Egypt (see, e.g., Grinin 2006a; Grinin and Korotayev 2009c, 2009d) and so on. Thus, the major part of fixed socio-cultural alterations (supported by social selection) emerges not as a result of ‘random errors of copying’ (though, of course, such random errors do exist), but as a result of purposeful alteration of respective memes. Such ‘mutations’ are directional from the very beginning and do not seem to have any analogues in natural biological evolution.

On the inheritance of acquired characteristics The other (and perhaps even more important) difference is that, in the process of biological (but not social) evolution, the acquired characteristics are not inherited.11 That is why socio-evolutionary changes are accumulated much faster than biologically useful changes of phenotype determined by mutation processes. Thus, because the acquired characteristics do not influence biological evolution, biological evolutionary processes go extremely slowly (in comparison with social evolution). On the other hand, within social evolution, the acquired characteristics can be inherited, and, hence, social evolution goes ‘according to Lamarck’ rather than ‘according to Darwin’. This point has been noted many times by a number of evolutionists (see, e.g., Mesoudi, Whiten and Laland 2006: 345–346). Consequently, social evolution proceeds much faster. In addition, as social evolution tended to go more and more ‘according to Lamarck’, it became more and more Lamarckian rather than Darwinian, which was one of the main factors for the acceleration of social evolution. Still, it appears necessary to mention that in some rare cases, one can observe the inheritance of acquired characteristics in complex biological organisms (Zhivotovsky 2002a). For example, somatic mutations may well be inherited in plants both with vegetative and sexual reproduction. In animals, viruses can insert themselves into the genome of gametes – subsequently the offspring inherit quite an ‘acquired characteristic’, the virus infection. The ability to inherit acquired characteristics is found in many plant-eating insects, in which specialized symbiotic bacteria live. Biochemical and ecological characteristics of such symbiotic complexes are determined up to a very large extent by bacteria (see, e.g., Dunbar et al. 2007). The possibility of inheritance of acquired characteristics through special particles (pangenes) was proposed by Darwin himself (Darwin 1883). Within the genomes of complex biological organisms, one can find a very large number of retropseudogenes and even working copies of genes that emerged as a result of the ‘copying’ of genetic information from RNA molecules to the chromosome with special enzymes (such genes are characterized by the absence of introns). Thus, in biological evolution, one may observe the ‘copying’ into the genome of information on the structure of mature matrix RNA. Because the alternative splicing is quite a controlled process, regulated by the cell and subject of intermediate influence of external conditions 81

Leonid E. Grinin et al.

(see, e.g., Lareau et al. 2007), mature mRNA may actually carry some (albeit rather incomplete and fragmentary) information on ‘acquired phenotypic characteristics’, and this information may be transmitted to the genome of the germ line. The impossibility of genetic inheritance ‘according to Lamarck’ postulated by the Synthetic Theory of Evolution exists because the mechanism of reverse translation does not appear to have emerged. That is why there is no way for changes that occur in an organism during its lifetime, at the level of proteins, to be recorded back into the genome.12 On the other hand, at present, we know that the phenotype at the cellular level is determined not only by proteins but also by a great variety of functional RNAs, whereas intravital changes of those molecules may well be written into the genome because here the mechanism of reverse transcription exists and is rather widely spread in biological organisms (including complex organisms). Hence, the point is not that within the biological evolution the ‘Lamarckian’ inheritance is totally impossible; rather the point is that such an inheritance is rather disadvantageous in most cases (see also Steele et al. 2002; Zhivotovsky 2002b). Consequently, such an inheritance is not usually an important mechanism of evolution (and, especially, of arogenic evolution). For example, it is evident that the hereditary fixation of adaptive modifications (‘modification genocopying’) is disadvantageous in many cases. Note that this includes those very consequences of the organ exercise whose inheritance played such an important role in Lamarck’s theory. In order for an adaptive modification to appear, we should observe first a genetically determined capability for such a modification (e.g., the muscles’ ability to grow as a result of exercise or the lymphocytes’ ability to develop immunity against new pathogens). However, if such a genetically determined ability has appeared, the firm fixation in the genotype (the genocopying) of only one of many possible versions of the final state of the trait (e.g., a precise size of a muscle or an immunity toward a specific pathogene) will not be a progressive evolutionary change; it will be a degenerative evolutionary change, accompanied by a decrease of the organic complexity and a loss of one of the ontogenetic regulatory circuits. In biological evolution, such events take place rather frequently, but this is not the arogenic evolutionary pathway. Within social evolution, there is no significant difference in the inheritance mechanisms between those traits that have been inherited from ‘ancestral’ societies and the ones that have been acquired throughout the history of existence of a given society. There could be some insignificant difference as regards the firmness of the fixation of the respective alterations, the easiness of their acceptance and so on, but it is impossible to say that acquired social characteristics are transmitted to new generations with significantly more difficulties (especially in complex societies). A serious obstacle for the operation of the ‘Lamarckian’ mechanism can be seen in traditionalism, which holds negative attitudes toward innovation and glorifies everything inherited from ancestors. This was very typical for simple traditional societies. However, such attitudes have weakened in a significant way in modern complex social systems.13 This might be connected with the development of the means, ­methods and technologies of forecasting, which is the conscious evaluation of innovation. Forecasting makes those characteristics that might be considered dangerous or disadvantageous by traditionalists (in particular, a very low precision of the ‘memotype’ replication and ‘Lamarckian’ inheritance) to become more acceptable in a society. 82

Big History and macro evolution

On the nature of hereditary variation Hereditary variation is a key issue in the theory of evolution.This is the issue, around which the main discussions between representatives of various schools of evolutionary thought (classical Darwinism, Synthetic Theory of Evolution, Orthogenesis, ­Nomogenesis, NeoLamarckism and so on) are concentrated. Variation is the main material basis of evolution; its character, mechanisms, factor, and emergence rates determine to a very high extent the character of the evolutionary process.These mechanisms of variation are one of the most fundamental areas of difference between biological and social evolution.14 Starting with Darwin, biologists have based their evolutionary theories on the idea that hereditary variation is basically ‘indeterminate’ or undirected, that is, random. However, as we have noted, within biological evolution, one can still detect a trend toward a decrease of randomness, both in mutational and recombinational variation. In some sense, this trend continues into social evolution, where variation is even less random and more directed.15 As mentioned above, there are significant differences between biological and social evolution in regard to the accuracy of copying (reproduction of replicators), because in general the precision of copying of genes (and, correspondingly, periods of their existence in a recognizable form) exceeds by orders of magnitude values of analogous indicators for memes. That is why ‘memetics’ (in contrast with genetics) has to deal with a much lower precision of replication and with a much higher speed of mutagenesis, though some replicators (memes) may have rather long periods of life. For example, according to some recent estimates, roots of some most widely used words may be preserved in a recognizable form for about 10,000 (and even more) years of linguistic evolution (Pagel et al. 2007). Another example can be provided by ‘long-lived’ folklore-mythological motifs that can survive for dozens thousand years (see, e.g., Korotayev and Khaltourina 2011; Berezkin 2007; Korotayev 2006; Korotayev et al. 2006). The same can be said about a very long life of some technical methods, for example, the production of stone tools. However, it makes sense to distinguish between various types of information transmission, depending on the number of copies in which the information is stored and reproduced (as well as the forms of that reproduction). There could be situations in which there is just a single carrier of important information. An ancient engineer could take his secrets of construction to the grave so that nobody could continue his techniques any more. There are lots of historical facts known to us from just one source; and if, in the process of transmission, there was distortion of the initial text, this could affect our current knowledge of the past. Those unique ancient books that disappeared in fire did not let us know the important information contained in them, and so on. These are examples of distortion or loss of information by functioning social systems. It seems appropriate here to recollect the information irreplaceability principle (Lyapunov principle). According to this principle, information that has entirely disappeared cannot be reconstructed in its entirety – what can be replaced are portions of information coming from a common source (See Rautian 1988a, 1988b). We confront a different case when we deal with information that is used by numerous carriers (as in the case of the use of a mass language). In such cases, changes in a living 83

Leonid E. Grinin et al.

language should not be always regarded as information distortion; we should rather speak about some drift in the use of linguistic matrices and patterns (similar to gene drift in populations), because language carriers may well know older forms, but prefer new ones. One may even observe the coexistence of persons using different linguistic forms and lexemes (similarly within one population, there could be different phenotypes). However, with time, some forms win the competition and language changes. When we speak about the accuracy of transmission of biological information, it is necessary to take into account that biological evolution has worked out rather effective molecular mechanisms that allow for sharply reduced precision of DNA replication when necessary (for example, SOS-response among bacteria). For some primitive biological objects, such as viruses, too high a precision of replication can even be lethal; in order to successfully go through their life cycles, they need very low precision of replication or, in other words, a very high rate of mutation (mutagenesis). For such organisms, evolutionary changes turn out to be necessary components of their everyday life (Vignuzzi et al. 2005)! Generally, though, in biological evolution, replication accuracy increases rather than decreases with the growth of the organismal complexity. In this sense, the reduction of precision that is observed in the transition from biological to social phase of the big history looks as if this were a ‘step backward’. However, this observation is rather superficial, as it does not take into account the nature of those errors that emerge in the process of replication, notably the degree of their randomness/ directionality. Within biological systems, replication errors are basically random. Taking into consideration the decrease of randomness, this may be interpreted in the following way: Nature has not developed any biological mechanisms that allow us to forecast results of concrete genetic changes and to plan them.Though a cell (for example, a lymphocyte) may ‘know’ in advance that, in order to achieve a needed result, it should alter some particular part of the genome, it, however, lacks mechanisms that would allow it to forecast results of a concrete genetic alteration.16 That is why, in the framework of biological evolution, the acceleration of adaptation-genesis through a radical reduction of the precision of replication is a very expensive and risky strategy that can be afforded only by very primitive forms of life. The situation changes radically if the replication ‘errors’ become not random, but actually purposeful, based on forecast of the possible results of concrete changes introduced into the ‘memotype’ of a social system. The presence of ‘directed mutations’ (in addition to undirected ones) radically distinguishes the process of ‘mutational variation’ in the evolution of memes from what is observed within the evolution of genes, where ALL the mutations are basically undirected. That is why we believe that the difference between biological and social evolution in respect to randomness/directionality of hereditary variation is more fundamental than the differences in precision of replicator copying or mutation rate. In the process of ‘socio-cultural mutagenesis’, the element of randomness is significantly smaller, because people possess the ability (albeit limited) to foresee results of certain concrete ‘mutations’. That is why human creativity (say, in development of new judicial laws or new technologies) may differ qualitatively from the ‘creativity’ of biological ­evolution – especially, as regards the effectiveness and the speed with which the respective ­results are achieved. 84

Big History and macro evolution

On the other hand, one should not exaggerate the role of conscious planning in relation to social innovation. Random search, trial and error, remains very important in social evolution (Grinin 1997, 2006b, 2007a, 2011a; Korotayev 2003), although there has been a clear decreasing trend in randomness in recent centuries (see, e.g., Korotayev 1999, 2003, 2004; Korotayev, Malkov and Khaltourina 2006; Grinin 1997, 2007a, 2009a). Thus, it is not sufficient just to have respective challenges in order that serious transformations could take place. Most societies ‘respond’ to new problems in old, habitual, tested and familiar ways, as they choose – not from a set of hypothetical alternatives – but from a set of accessible alternatives (Van Parijs 1981: 51). In other words, they use actually known measures instead of potential ones (Claessen 1989). In these situations, their behavior is often quite similar to that of social animals. ­Naturally, not all such ‘responses’ are effective. As a result, many societies perish, disappear or lose their independence (Grinin 2011a). For example, after the Roman regiments were withdrawn from Britain in 410 CE, the Britons (Romanized British Celts) sought protection from the raids of their Irish and Scottish neighbors. They invited Saxons to defend them in return for plots of land in Britain. Actually, this was a variation of the very well-known Roman method to use barbarians to fight barbarians. However, the Saxons, after they had seen the Britons military weakness, stopped obeying local authorities and became masters of the country (together with Angles and Jutes). In this way, the Britons, notwithstanding their fierce and long resistance, were partly evicted, partly destroyed and partly enslaved. As a result, barbarian Anglo-Saxon states were found in place of the state of the Britons (Blair 1966: 149–168; Chadwick 1987: 71; Philippov 1990: 77). If we take into account general historical contexts, we see that an extremely small fraction of all responses to various challenges turned out to be capable of becoming sources for system aromorphoses. This implies that most societies turned out to be incapable to move to new qualitative levels: They did not have the necessary potential for change, their construction had certain ‘defects’, the system might have been too rigid to transform easily, or some necessary conditions were lacking and so on (­Grinin 2011a, 2011b; Grinin and Korotayev 2009e).

The ability to borrow and the horizontal exchange of genetic information These facts illustrate a rather strange situation. There are similarities in biological and social evolution, such as the transmission of information, variability, community complexity, etc. However, these similarities occur at the lower stages of biological evolution (involving simple biological organisms), whereas they are absent in higher stages of biological evolution (involving complex biological organisms).17 One of the main differences between social and biological evolution is the ability of social systems to not just change and transform, but also to borrow new elements. However, in this respect, social evolution resembles the biological processes that prevailed during the epoch of the ‘prokaryotic biosphere’ (and those processes continue up to the present among prokaryotes and monocellular eukaryotes). Among the prokaryotes, we find the ability to ‘transform naturally’ – to absorb DNA from the environment and to insert it into their genome, which leads to an immediate transformation of the phenotype. There is also, of course, a significant difference 85

Leonid E. Grinin et al.

between this biological and social analogue: In society the borrowings are usually made consciously. Horizontal gene transfer produces many useful genetic ‘inventions’, a sort of commons for microbe communities. For example, communities of marine planktonic microbes use the genes of proteorhodopsins – proteins that allow them to partly utilize sunlight. In contrast to the proteins that participate in real photosynthesis, proteorhodopsins do not need the help of many other specialized proteins. Thus, in order to acquire a useful function, it is sufficient for a microorganism to borrow a single gene (Frigaard et al. 2006). Complex borrowing of entire gene systems is observed much less frequently, but when they occur, they have more significant consequences. An original and widespread version of such ‘borrowing’ results in the emergence of symbiotic systems, which sometimes actually leads to the formation of a new organism out of several other organisms.The role of such systems is often underestimated, but all functioning of the modern biosphere is based on them. There are many examples. Terrestrial plants would not have been able to achieve evolutionary success without symbiosis with mycorrhizal fungi and nitrogen-fixing bacteria. Herbivorous animals, both insects and vertebrates, are unable to digest plant food without symbiosis with specialized microorganisms. Indeed, the principle ecological, biospheric role of animals is precisely to process plant food! In highly complex biological organisms, in contrast to social organisms and human societies, large-scale ‘borrowings’ in the form of symbiotic relations or alien genetic material rarely take place, but many of the most important aromorphoses are connected just with them.

Analogues of ‘suprasocietal institutions’ in biological evolution Let us come back to the question: Are there analogues of such structures in the evolution of the biosphere? The answer will depend on the level of the biosphere’s system organization. Society is frequently compared with biological organisms, but – in this case – we are comparing suprasocietal amalgamations with supra-organic systems: populations, species, ecosystems, groups of social animals and so on. However, this is probably not quite an appropriate scale of analysis, so we need to compare suprasocietal institutions of a global scale (like the United Nations) with biological objects of immeasurably smaller scale, for example, with particular ‘casts’ of the ant family.18 At any scale, it is difficult to find good analogies to the formation of suprasocietal institutions within biological evolution. This becomes even more evident if we compare societies, not with organisms, but with supra-organic biological systems (e.g., populations or species). Although those biological systems (like societies) can amalgamate into systems of a higher order (ecosystems or the biosphere), these higher-­ order systems are not centralized but are weakly integrated – nothing like suprasocietal institutions as the World Health Organization, UNESCO, or even a complex tribal confederation with its own supra-tribal regulation organs. For example, one can observe the formation of rather complex links between species in ecosystems; certain key species may produce a decisive influence on other species in the community, but this does not result in the formation of any ‘supra-species institutions’. 86

Big History and macro evolution

On the one hand, it is possible to see in this comparison one of the fundamental differences between social and biological macro-evolution. On the other hand, some biological analogues of ‘suprasocietal institutions’ did emerge. In the Holocene (the last 10,000 years, starting with the Agrarian Revolution), human societies developed suprasocietal institutions. In the course of the socio-biological evolution of the resulting ‘anthroposphere’, we observe a parallel growth in the integration of humankind and integration and coordination of evolutionary changes of biological populations, species and ecosystems. In other words, the development of the global human community (the World System) may be regarded as a factor of integration of biological evolution at its upper level. Thus, social and biological evolutions are related processes that supplement and maintain each other. Indeed, there is a tendency toward their fusion into a single complex process, one leading to the development of an ‘anthropo-biosphere’. In this respect, it appears to be possible to speak about the co-evolution of biological and social development.

On the role of selection in biological and social evolution The role of selection in social evolution differs significantly from the one in biological evolution. In the biological world, the main source of stable, heritable innovations (mutational and recombinational variation) is characterized by a high degree of randomness and unpredictability (although, of course, it is also necessary to take into consideration all the above-mentioned qualifications about the means of optimization). In this situation,‘post-factum selection’, the selection among the deviations that have already emerged and have found their realization in the phenotype, becomes the only way to give the process a certain directionality (in this case – to secure the additive character of changes). In the social world, the main sources of heritable innovations are not random errors of copying and reproduction but conscious and purposeful correction and alteration of memes. However, such purposefulness is unable to foresee not only all the consequences of its actions but even the near consequences. That is why intentional actions may appear random. Throughout human history, failures of some societies have been a sort of payment for the success of others (what we denote as a rule of payment for the arogenic progress), from which the role of selection in the search for successful aromorphic variants acquires an especially important meaning (Grinin, Markov and Korotayev 2011; Grinin 1997, 2007a; Grinin and Korotayev 2009b). Societies frequently confront such situations when an old system does not work. Those who do not change or look for more effective means perish. Selection at the gene/meme level plays a less important role in social evolution than it does in biological evolution. However, selection in social evolution takes place not so much at the level of memes but more at the level of organizations, institutions and social systems. At the level of intersocietal competition, until recently, social selection acted in an especially tough way: ‘the victor gets more or everything; the defeated may lose himself ’ (Grinin 2003, 2004, 2009a, 2009b, 2010, 2011a, 2011b). So, this is a selection mechanism that is rather different from the one found in biological evolution. 87

Leonid E. Grinin et al.

One more important aspect of social selection that is absent in biological evolution is the struggle for the selection of a certain model (model of reforms, model of unification, ideological model) at the level of individual societies, as well as at the intersocietal level. Everywhere, we can observe the selection of leaders, models, courses, central positions and so on. The decisive advantage could be rather different in different cases. In some cases, this could be a very capable and talented leader; in others this could be an advantageous geographic position; in still other situations this could be just a lucky contingency. Thus, although we are dealing with rather different mechanisms of selection in biological and social evolution, their roles are very important in both cases. Still, within biological evolution, selection process is more important, because there is no alternative, whereas such an alternative exists within social evolution.

Some preconditions of the transition from biological to social phase of the big history Social evolution as a logical result of the development of adaptiogenesis mechanisms In addition to what has been already said about the organic links between biological and social evolution, one should consider another aspect of adaptiogenesis. The process of adaptation that constitutes the principal contents of biological evolution may proceed at different levels: (1) the organism structure; (2) its behavior; (3) structure and behavior of a socium as a superorganic amalgamation. At all those levels, one may observe the transition from primary, primitive and slow methods of adaptiogenesis based on random mutations, recombination and selection to more progressive, effective and rapid ways of evolutionary change. Not only organisms, species and societies evolve, but mechanisms of evolution evolve too. The general direction of this evolutionary movement may be characterized as a trend to the reduction of the role of random processes and the growth of systematic controlled processes. The initial and primary evolutionary algorithm is the random search, the trial-and-error method. However, at all levels of adaptiogenesis, one may observe a gradual development of such mechanisms that decrease the role of randomness and, thus, optimize this algorithm; though it appears impossible to exclude entirely an element of randomness either from biological or from social evolution. (1) The organism structure level. Even at the basic level of biochemistry, physiology and morphology, many forms of life have developed ways of adaptiogenesis that are faster and more effective than the random search conducted according to the scheme of ‘random mutations + selection’. One of these mechanisms is regulation of the mutagenesis rate, depending on available conditions: Under favorable conditions, the mutagenesis rate is low; in unfavorable conditions, it increases (Grinin, Markov and Korotayev 2008: Chap. 6, §6.8). It is also appropriate to mention epigenetic changes of hereditary material that are transmitted to a number of generations, in particular parental genomic imprinting 88

Big History and macro evolution

that became especially developed in the most complex organisms, such as mammals and flowering plants ( Jablonka and Lamb 1999). Imprinting is actually a sort of purposeful manipulation of hereditary properties of offspring. With the maturation of male and female gametes, certain parts of the genome are marked in a special way, for example, through methylation. The methylation of  DNA is not a chaotic process but is regulated by complex molecular systems.What is especially important is that methylation of particular nucleotides increases the probability of their mutating. Thus, through the methylation (or non-methylation) of particular nucleotides, cells can in principle regulate the probability of their mutation (Vanyushin 2004). Another example of the purposeful change of hereditary information is provided by the development of adaptive (acquired) immunity through combining genetic blocks, subsequent somatic hyper-mutation and clonal selection. Both the combining of DNA fragments (V-(D)-J recombination) and hyper-mutation are processes that are only partly random. In other words, the limits of randomness in this case are rather accurately demarcated (Grinin, Markov and Korotayev 2008: Chap. 4, §4.2.4). The combination of DNA fragments is conducted from a precisely defined set, and the hyper-mutation takes place at a rather accurately demarcated part of a gene, while the selection of lymphocyte clones makes the whole process unequivocally ­directional. As a result, the final outcome of such a ‘sequence of random events’ turns out to be quite deterministic. Such a mechanism may be designated as ‘optimized random search’.19 Note that in the case of the acquired immunity, from a ‘technical’ point of view, the achieved result may well be transmitted to the offspring, for example, via the mechanism of reverse transcription and transmission of the genetic material from lymphocytes to gametes through endogenous retroviruses (Steele et al. 2002). However, this does not happen, because it is more advantageous to transmit not a concrete immunity to a particular pathogen to the offspring but a universal capability to develop immunity against any pathogen. In general, such mechanisms of purposeful genome alteration do not have a universal presence in biological organisms, and the overwhelming majority of mutations take place in a quite random way. Biologists rarely consider that assortative (selective) mating, mediated sometimes through extremely complex mechanisms of mate-choice, is nothing but an extremely effective mechanism for management of recombinational variation. However, in the real biological world, absolutely unselective, random mating is hardly ever observed. Indeed, random mating is a scientific abstraction, like an ‘ideal gas’, or an ‘absolutely dark body’. With growth in the level of organization of biological organisms, the complexity and effectiveness of mate-choice also grew, whereas the recombinational variation became less random as a result. (2) Level of individual behavior. One can trace the transition from predominantly hereditary and genetically determined behavioral patterns to more flexible learning-based ones. As we saw above, in the case of immunity, it was more advantageous to transmit to the offspring a universal capability to ‘learn’ instead of a rigidly determined means of resistance to a particular pathogen. In an analogous way, in the general course of evolution, it has turned out to be more 89

Leonid E. Grinin et al.

advantageous to transmit the ability to learn rather than to transmit rigidly fixed behavioral stereotypes.20 No doubt, the emergence of the capability to learn is a major aromorphosis, though it is very stretched over time. Even unicellular organisms have some inchoate abilities to learn (sensitization, habituation), let alone such highly organized animals as ants or bees. (3) Biological socium level (social adaptiogenesis). A wide variety of living ­organisms – from bacteria to mammals – lead a social way of life.The socium as a whole has certain system characteristics that can be more or less adaptive (Popov 2006). Here, we also observe the transition from rigidly genetically determined forms of social relationships to more flexible versions, within which a social system may adequately (adaptively) react to changes in its environment. For example, the size of subsidiary colonies of an anthill may change in a reasonable, that is, adaptive way, depending on resource availability (Zakharov 1978: 49). However, in general, for all the pre-human forms of life, such possibilities are limited. The human development of the ability to evolve socially, which implies the possibility of an almost limitless change in the structure of social systems, appears to be a natural (though qualitatively higher) continuation of this evolutionary trend.

One of the ‘preadaptations’ that facilitated the transition from biological to social evolution The issue of how biological evolution transformed into social evolution is among the most important questions of big history and evolutionary studies. What ‘preadaptations’ were needed for the transition from biological to social phase of the big history? This is a very complex subject. And here we shall restrict ourselves to consideration of just one of those preconditions. Social macro-evolution (and, hence, the start of the big history social phase) became  possible due to the emergence of a uniquely human ability denoted as ‘­ultra-sociality’ (Boyd and Richerson 1996). This is only found among humans and designates the ability to change their social organization radically and almost limitlessly in response to internal and external challenges. Only humans are capable of forming collectivities that could be entirely different as regards their structure, their traditions, their norms of behavior, their modes of subsistence, their systems of intragroup relationships, their family types, etc. Whatever the complexity of the collectivities of non-human primates, they do not have such flexibility. Each species usually has only one type of social organization; some cultural differences are observed, but they are incomparable with the ones observed in Homo sapiens sapiens.Yet, some animals possess a limited ability to adaptively change the structure of their socium. For example, in disadvantageous circumstances, one may observe growth in the rigidity of social hierarchy (the ‘power vertical’), whereas the relationships become more egalitarian under more favorable conditions. Sometimes the transition to a social way of life occurs during unfavorable conditions, whereas the same animals may return to solitary life with improvement of conditions (Popov 2006). Those adaptive modifications of social structure in animal communities are still significantly inferior in their scale to what is observed in human societies; in addition, among other animals, they are characterized by a much higher degree of predictability. 90

Big History and macro evolution

The emergence of ultra-sociality was a natural result of the preceding co-­ development of intellect and social relations among our ancestors. The progressive development of the brain and intellectual capacities in primates is inseparably linked with a social way of life – with the necessity to predict actions of other members of their group, to manipulate them, to learn from them, to achieve an optimum combination of altruism and egoism in their behavior. At present, this is the point of view of the majority of primatologists (e.g., Byrne and Whiten 1988; Byrne and Bates 2007). The idea that the primates intellect developed first of all for, say, effective search for fruit (the ecological intellect hypothesis) does not now have many supporters. It cannot explain why primates need such a large brain, if many other animals, such as squirrels, perfectly manage similar tasks, though their brain remains small. In contrast, the ‘social intellect hypothesis’ is supported by facts. Scientists have detected a significant positive correlation between brain size in primates and the size of their social groups (Dunbar 2003). It is necessary to note that primates (in contrast to the majority of other social animals) know all the members of their group ‘by sight’ and have particular relationships with each of them. There are grounds to maintain that individualized pair relationships are the most intellectually ‘resource-intensive’ (­Dunbar and Shultz 2007). A positive feedback appears to have existed between the development of a social intellect and the growth of complexity in social relationships of hominids.21 Those individuals that managed to achieve a higher status within a social hierarchy, due to a higher intellect or a better ability to foresee actions of others, left more numerous offspring, which in turn led to the general intellectual growth of the socium. As a result, in subsequent generations, in order to move up the social ladder, it was necessary for individuals to possess an even more developed social intellect, and so on. Interesting experimental facts have been recently obtained. They indicate that intellectual abilities of a ‘social’ character, which allow for resolution of social tasks, developed in our ancestors earlier in comparison with the intellectual capabilities of the other types (e.g., the ones that allow to solve ‘physical’ and instrumental tasks) (Herrmann et al. 2007). In order to function effectively in a complex, constantly changing socio-cultural environment, our pre-human ancestors had to develop intellectual abilities of a rather concrete type: abilities of effective communication, learning and – most importantly – of understanding not only actions, but also thoughts and desires of members of their groups (Vygotsky 1978). It is quite evident that abilities of this kind should become apparent in early childhood, in the period of active learning and social adaptation. There are two alternative hypotheses about possible mechanisms in the evolutionary development of these social abilities. The first hypothesis suggests that they emerged as a result of the uniform development of the intellect as a whole (general intelligence hypothesis).The second suggests that this was the directed development of specific socially oriented abilities, whereas all the other abilities (such as abilities to think logically, to detect cause-and-effect links in the physical world and so on) developed later, as something additional and secondary. This is called the cultural intelligence hypothesis (Barkow et al. 1992; ­Shettleworth 1998; Herrmann et al. 2007). 91

Leonid E. Grinin et al.

At first glance, the general intelligence hypothesis looks more plausible, but it is also possible to provide evidence in support of the cultural intelligence hypothesis. For example, it is known that specific intellectual abilities develop locally in many animals, but their overall intellectual level does not grow (or grows insignificantly). One can mention, for example, the birds’ unique orientation abilities (Shettleworth 1998). Special experiments have been conducted in order to test these hypotheses. The experiments were based on the following reasoning: If the cultural intelligence hypothesis is true, then there should be an age in the individual development of ­humans when we are not different in our ‘physical’ intellect from the apes, even though we are already far above them in our ‘cultural–social’ intellect. Experiments have confirmed the cultural intellect hypothesis: It turns out that 2.5-year-old children have the same level of development as adult chimpanzees and orangutans in respect to solving physical tasks (spatial, quantitative, detection of cause-and-effect relationships and so on), but they are significantly superior as regards the effectiveness with which they solve tasks of a social nature, such as those connected with the prediction of others’ actions, communication, learning and so on (Herrmann et al. 2007). In general, present-day anthropological data suggest the following: (1) The development of social relationships and intellectual abilities in the higher primates (in general) and the hominids (in particular) proceeded within a single evolutionary process that was accelerated by the above-mentioned positive feedback; (2) This process tended to lead to the growth of complexity and flexibility of social relationships. Thus, the development of ultra-sociality and the ability to evolve socially within one of the groups of primates was a natural and logical result of the development of a trend that started among the primates long before the emergence of Homo sapiens sapiens.

Afterword. The formation of social evolution’s own mechanisms The transition from the biological to social phase of big history was a very complex process that we do not quite understand even now. Within this transition it appears possible to speak about a phase change of a few subtypes of macro-evolution: the biological type of macro-evolution was first transformed into the biological-social type, then the biological-social type was transformed into the social-biological type; and, finally, the latter was transformed into the social type of macro-evolution already in the framework of the unequivocally human society (See Grinin and Korotayev 2009b: Chap. 1 for more details). In the course of anthropogenesis, biological macro-evolution was transformed into bio-social evolution.The discoveries of recent decades have moved the dating of the emergence of our species deep in the past to about 200,000 BP (see, e.g., Stringer 1990; Bar-Yosef and Vandermeersch 1993; Pääbo 1995; Gibbons 1997; Holden 1998; Culotta 1999; Kaufman 1999;White et al. 2003; Lambert 1991; Zhdanko 1999; Klima 2003: 206). However, the borderline around 50,000–40,000 BP still retains an immense importance. This is the point from which we can speak with a complete confidence about humans of a contemporary cultural type, in particular about the presence of full-fledged languages, as well as ‘really human’ culture (e.g., Bar-Yosef and 92

Big History and macro evolution

Vandermeersch 1993: 94). There is, of course, some hypothesis that human language appeared long before 50,000–40,000 BP. Although this is contested by other scientists, everybody agrees that by 40,000 BP language existed wherever humans lived (e.g., Holden 1998: 1455). Richard Klein, an anthropologist from Stanford University proposes the following hypothesis to explain the gap between the emergence of anatomically modern Homo sapiens sapiens and the emergence of language and cultural artifacts that took place much later. According to Klein, the modern brain is a result of rapid genetic changes. He hypothesizes that such changes took place around 50,000 BP, pointing out that the affluence of cultural artifacts starts just after that date, as well as the migration of anatomically modern humans out of Africa (see Zimmer 2003: 41ff). Thus, the emergence of Homo sapiens sapiens did not automatically result in social macro-evolution proper. We believe that the evolutionary driving forces were still mostly biological when modern humans first emerged, but that the social forces gradually increased their importance and prevailed over the biological ones at a certain point. Naturally, this was a rather prolonged process, within which the breakthrough point could hardly be identified. We contend that the social component became dominant after 50,000–40,000 BP. However, it did not become absolutely dominant, as biological adaptation and physical anthropological transformation continued in many important ways. The point is that they did not disappear, but their role significantly decreased.22 This transition to modern human society is sometimes denoted as the Upper Paleolithic Revolution. If we use the title of the book by Mellars and Stringer (Mellars and Stringer 1989), we may call this radical transformation: The Human Revolution.Thus, starting with the Upper Paleolithic Revolution, we may speak about the transition from socio-biological evolution to social evolution, a process that was finalized by the Agrarian Revolution. There were not many major aromorphoses in the hunter-gatherer epoch (Grinin 2006b, 2009a), which is why the overall rates of socio-evolutionary processes were slow and their directionality rather vague. Such a type of social macro-evolution may be denoted as socio-natural. As a result of a system of interrelated aromorphoses connected with the Agrarian Revolution, one could observe the transition to the socio-­ historical type of macro-evolution. As a result of this, social macro-evolution changed its algorithm in a rather significant way, resulting in modification of certain evolutionary laws. We shall consider below how the significance of laws of evolution and the process of social macro-evolution changed as a result of the Agrarian Revolution. Main factors of social change in foraging societies were the result of adaptation to new and various environments – from the deserts of Australia to the pack ice of the Arctic. This was only possible through the modification of socio-cultural systems. This made it possible for humans to people most of the world’s landmass, to create an enormous variety of tools and crafts, as well as social and other institutions. Effective adaptations let people not only survive, but sometimes also live relatively ‘comfortable’ lives that Sahlins (Sahlins 1972) called the original affluent society.The character of human relations with their environment varied significantly, but generally these were one of human adaptations to the natural world (see, e.g., Leonova and Nesmeyanov 1993; see also Grinin 2006b: 82–83). 93

Leonid E. Grinin et al.

In the Agrarian epoch, the character of those interrelations changed significantly through the transition to much more conscious and effective change of the environments at a rather wide scale (irrigation, clearing of forests, plowing of steppes, soil fertilization, construction of cities, roads and so on). Natural forces (animal, wind and water energy) started to be used on a much wider scale (earlier humans actively used only fire). Natural raw materials started to be transformed into entirely new products (metals, fabrics, ceramics and glass). Thus, within social evolution process a more and more significant role started to be played by peculiarly social factors that (in contrast with natural factors) are connected to conscious goal-setting and goal-achieving. Gradually, with economic-­ technological progress, the growth of surplus accumulation capacities, as well as general cultural complexity of social systems, their evolution became almost purely social. As a result, the ‘vector’ of evolutionary selection turned out to be directed toward societal capabilities to adapt to social (rather than natural) environments, which implies the capacity to compete with neighboring social systems in economic, military, commercial, cultural, ideological and other spheres. Finally, we would like to mention the following important changes in the ‘algorithm’ of social evolution: •

The start of the mechanism for resource accumulation.

In the tens of thousands of years of the human foraging epoch, long-term material resource accumulation was relatively insignificant when compared to subsequent epochs. There was, of course, a certain amount of accumulation, of knowledge, traditions and technologies, albeit at a limited scale. This accumulation took place not in every society, but was observed at the global scale and was due to the overall demographic growth, increase in numbers of social systems, emergence of new tools, products, etc. There was practically no special accumulation sector prior to the Agrarian Revolution.23 In many cases, people could produce more than they actually needed, and sometimes even so-called ‘original affluent societies’ could emerge (Sahlins 1972). For example, with respect to the gatherers of sago in New Guinea, people would spend a minor part of their time securing food for themselves, whereas they would spend the rest of the time at other activities and leisure (Shnirel’man 1983, 1989). The impossibility to accumulate and/or the absence of the desire to accumulate slowed down development, which contributed to the slow speed of social evolution (Grinin 2006b, 2009a). In simple social systems of agriculturalists and pastoralists, the emergence of the possibility (and, later, the desire) to accumulate led to numerous transformations in the spheres of functional differentiation, distribution, social stratification, exchange, trade, development of property relationships, increasing political complexity and so on. •

Strengthening of the ability of social systems to change.

Agrarian societies turned out to be more capable of serious social transformations than hunter-gatherers, while complex Agrarian societies turned out to be much more capable of such transformations than simple agriculturalist and pastoralist systems. 94

Big History and macro evolution

The growth of social systems’ ability to change provides a vivid demonstration of the main difference between social and biological evolution – that humans were capable of consciously transforming their social systems, with preconceived goals. •

Intersocietal contacts become the leading factor of social evolution.

The importance of various contacts increased sharply, and this contributed to a more active adaptation of social systems to their environments. The growth of the role of contacts dramatically raised the importance of external social driving forces (Grinin 1997–2001 [1997/2: 23]; 2007a: 177). Note that this had an enormous importance for the development of the World System and for the humankind as a whole. Military and other interactions stimulated improvements in administration, defense, culture, technology and so on. All this contributed to the development of a single global process involving numerous societies and peoples. It is also appropriate to note that the growth of societal size is not only due to natural demographic growth, but is more importantly due to the integration and unification of social systems. Thus, external contact factors become most important with respect to societal evolution.

Notes 1 This paradigm is discussed in Grinin et al. (2011). 2 Note that in the biological macroevolution the ‘borrowing’ is found mostly at lower levels of the biological evolution, whereas it is found much less frequently at higher levels. The opposite situation is observed in social macroevolution – in general, the older the society, the lower its borrowing rate (incidentally, this accounts to a considerable extent for a low rate of change in the majority of ancient societies). 3 We denote as megaevolution the process of evolution throughout the whole of big ­h istory, whereas we denote as macroevolution the process of evolution during one of its particular phases. 4 This is typical, for example, for a very interesting and controversial article by M ­ esoudi, Whiten and Laland (2006), where we clearly deal with an attempt to impose the Darwinian methodology on the study of social evolution. The importance of the above-mentioned differences (including such fundamental differences as the absence in social evolution of a clear distinction between genotype and phenotype) is downplayed by a statement that those differences are either illusory or unimportant (ibid.: 345). Such an approach also reduces the value of a rather interesting methodology that they propose. 5 It appears appropriate to mention that the genomes of the humans and the chimpanzees differ by ten major genome reorganizations. A few years ago, it turned out to be possible to sequence the genome of the rhesus macaque (a special issue of the Science was devoted to this subject; see in particular Rhesus Macaque…, 2007). This is the third primate genome that was sequenced (after the human and chimpanzee genomes). Up to that moment, when detecting differences between the genomes of the humans and the chimpanzees, specialist could not determine which of those differences emerged in the human evolutionary line, and which appeared in the evolutionary line of the chimpanzees. The reading of the rhesus macaque genome substantially facilitated this task. The comparison with the macaque genome allowed detecting that three of those 95

Leonid E. Grinin et al.

differences happened in the human evolutionary line, whereas the other seven occurred in the evolutionary line of the chimpanzees (see Markov and Naymark, 2009 for more details). 6 However, there are cases when societies create new societies rather similar (with basically the same ‘memotype’) to the ‘maternal’ ones, for example, with the establishment of settler colonies. See the next section for more information on the differences in ways of information transmission. 7 Because the systems of transmission of traits within biological and social systems are rather different; because of the higher degree of complexity of social systems, and so on. 8 There is, however, a major difference: any large enough society usually consists of a whole hierarchy of social systems (e.g., with respect to a typical agrarian empire these would be: nuclear family – extended family – clan community – village community – primary district – secondary district – province), so that it can hardly be compared with a single biological organism (though both systems can still be compared functionally, as is correctly noted by Hallpike [1986]). 9 We could mention various flocks and packs of animals as examples of such amalgamations with one level of organization. More complex superorganic amalgamations may be found in the biological evolution among less complex organisms. This trend seems to be opposite to what is observed in the social evolution, though, say, village communities in more complex societies tend to be less complex than in more simple ones (see, e.g., Korotayev 1995; 2003: 75–90; Korotayev et al. 2000, 2011). 10 Close results are arrived at by Dawkins (1993) in his theory of the ‘evolution of memes’. 11 As one of the differences between social and biological evolution is connected with the absence in the former of clear equivalents of genotype and phenotype (see, e.g., ­Mesoudi, Whiten and Laland, 2006: 344–345), it appears quite evident that the expressions ‘socio-cultural genotype’ and ‘socio-cultural phenotype’ should be regarded as metaphors rather than as exact scientific terms. 12 On the other hand, there is a hypothesis that such a mechanism may have existed at the earliest phases of biological evolution. What is more, scientists have experimentally obtained RNA molecules that can perform certain stages of reverse translation (Nashimoto 2001). 13 On the other hand, we observe another trend in connection with some sorts of regulation mechanisms. One should not think that the only evolutionary mechanism in social evolution is a conscious change of existing objects. There is also an opposite trend that may be denoted as institutionalization. In many cases, certain relationships are fixed by customs or laws in order to avoid excessive variation/equivocality that may often be harmful for a social system. For example, one could observe the development of rather rigid marriage institutions, various legal codes and constitutions that can be only altered with significant difficulties (that are usually consciously established by respective norms aimed at the provision of the stability of respective codes and constitutions). In this respect the trend toward the canalization of changes may be also traced in the social macro-evolution. 14 It appears that this is relevant not only for the biological and social phases of big history, but also to all its preceding phases. 15 When we make such comparisons, we compare genotype with that totality of socio-­ cultural information (it may be denoted as ‘memotype’), which is transmitted from generation to generation and determines main characteristics of social systems. In social systems, in addition to biological generations, parents and children, we find other types of continuity (that could be sometimes even more important) like institutional and legal 96

Big History and macro evolution

16

17

18 19

20

21 22

23

continuity whose role increases constantly. That is, we observe the growth of the importance of information transmission in the framework of institutions, corporations, organizations and so on, that is conducted not between biological generations (from parents to children), but, say, from an experienced worker to an inexperienced one, or from a teacher to a pupil. In addition the emergence of external information carriers (in form of books, electronic records and so on) allows conducting a distance transmission of information without any direct contact between respective people, which, incidentally, contributes to the growth of the socio-cultural evolution rate. Actually, as a result, in complex social systems the number of information transmission channels grows by orders of magnitude (especially with the emergence of external information carriers). In some sense, this growth already starts with the development of social life among the animals. Such a mechanism (in the form of scientific methods and genetic engineering) was finally developed in the course of socio-cultural evolution; this mechanism, however, could still hardly be called perfect. We do not have a full explanation of this phenomenon, but one may think about the application to the macro- and even megaevolution of the law of the negation of the negation, which in this case may be interpreted in the following way: From a free borrowing of information to its rigid isolation and canalization, and then again to its free (but now conscious) borrowing. From contraposition of biological (genetic) and social mechanisms of evolution (within the process of anthropogenesis and sociogenesis) to genetic evolution controlled by the humans. On the other hand, a large anthill or termitary may well be compared with a large village community. In this way, a more flexible reaction to unknown situations develops; this may be compared with multifunctional institutions in human societies that while remaining apparently the same institutions may allow social systems to behave differently in different situations, whereas respective institutions would experience certain changes with the change of situations. Thus, army may be relatively small during the time of peace, and then it would grow sharply in size as a result of mobilization, whereas its functions also substantially change. The same can be said about the flexibility of the family, the village community and many other social groups and institutions. It appears necessary to note that in both cases the ability to learn does not replace entirely the genetically determined concrete adaptations; the former is added to the latter. In the immunity system of higher organisms, the system of innate immunity is preserved in addition to a new system of adaptive (acquired throughout the life) immunity; similarly, in the behavior of higher animals, behavioral patterns developed throughout the life through the learning are combined with innate genetically determined behavioral traits. This social intellect is also called the ‘Machiavellian intellect’, e.g., Byrne and Whiten (1988). There are sufficient grounds to maintain that the biological evolution of the humans did not stop 200–150,000 BP; it did not stop either after the Upper Paleolithic Revolution (see, e.g., Alexeev, 1984: 345–346; 1986: 137–145; Yaryghin et al., 1999, vol. 2: 165; Borinskaya, 2005; Borinskaya and Korotayev, 2007). Thus, the above-mentioned factor must have played some role in the biosocio-cultural evolution of Homo sapiens sapiens. With a possible exception of some highly specialized hunters (usually of large aquatic animals), gatherers and fishers – for example, some social systems described ethnographically for the North-Western Coast of America (see, e.g., Averkieva 1978; ­Shnirel’man, 1986). 97

Leonid E. Grinin et al.

Bibliography Alexeev, Valery. Formation of the Mankind. Moscow: Politizdat, 1984. (Алексеев, В. П. Становление человечества. М.: Политиздат, 1984.) Alexeev, Valery P. Ethnogenesis. Moscow: Vysshaya shkola, 1986. (Алексеев, В. П. Этногенез. М.: Высшая школа, 1986.) Artzrouni, Marc; with John Komlos. “Population Growth through History and the Escape from the Malthusian Trap: A Homeostatic Simulation Model.” Genus 41(3–4), 1985: 21–39. Averkieva, Yulia. “The Indians of the Northwest Coast of the North America (the Tlingit).” The North American Indians. Yulia Averkieva (editor). Moscow: Progress, 1978: pp. 318–360. (Аверкиева, Ю. П. “Индейцы северо-западного побережья Северной Америки (Тлинкиты).” Североамериканские индейцы/Ред. Ю. П. Аверкиева. М.: Прогресс, 1978. C. 318–360.) Barkow, Jerome; with Leda Cosmides and John Tooby (editors). The Adapted Mind: Evolutionary Psychology and the Generation of Culture. New York, NY: Oxford University Press, 1992. Bar-Yosef, Ofer; with Bernard Vandermeersch. “Modern Humans in the Levant.” Scientific American April, 1993: 94–100. Berezkin, Yuri. Myths Populate America. Areal Allocation of Folk Motive and Early Migration to the New World. Moscow: OGI, 2007. (Березкин, Ю. Е. Мифы заселяют Америку. Ареальное распределение фольклорных мотивов и ранние миграции в Новый Свет. М.: ОГИ). Blair, Peter. Roman Britain and Еarly England 55 B.C. – A.D. 871. New York – London: W. W. Norton and Company, 1966. Borinskaya, Svetlana. “The Role of Genetic Factors in Social Evolution.” History and Complexity Studies: Methodology of the Study. Sergei Malkov and Andrey Korotayev (editors). Moscow: KomKniga/URSS, 2005: pp. 62–75. (Боринская, С. А. “Роль генетических факторов в социальной эволюции.” История и синергетика: Методология исследования/Ред. С. Ю. Малков, А. В. Коротаев. М.: КомКнига/URSS, 2005. C. 63–75.) Borinskaya, Svetlana; with Andrey Korotayev. “Genes, Peoples and Sociocultural Evolution.” History and Mathematics: The Analysis and Modeling of Sociohistorical Processes. Sergei Malkov, Leonid Grinin and Andrey Korotayev (editors). Moscow: KomKniga/ URSS, 2007: pp. 232–242. (Боринская, С. А., Коротаев, А. В. “Гены, народы и социокультурная эволюция.” История и Математика: анализ и моделирование социально-исторических процессов/Ред. С. Ю. Малков, Л. Е. Гринин, А. В. Коротаев. М.: КомКнига/URSS, 2007. C. 232–242.) Boyd, Rob; with Peter Richerson. “Why Culture Is Common, but Cultural Evolution Is Rare.” Proceedings of the British Academy 88, 1996: 77–93. Byrne, Richard; with Lucy Bates. “Sociality, Evolution, and Cognition.” Current Biology 17(16), 2007: R714−R723. Byrne, Richard; with Andrew Whiten (editors). Machiavellian Intelligence. Oxford: Oxford University Press, 1988. Chadwick, Nora. The Celts. London: Penguin, 1987. Christian, David. (2017). What Is Big History? Journal of Big History, 1(1): 4–19. Christian, David, with Cynthia Stokes Brown and Craig Benjamin. Big History: Between Nothing and Everything. (New ed.) New York, NY: McGraw Hill Education, 2013. Claessen, Henri. “Evolutionism in Development.” Vienne Contributions to Ethnology and Anthropology 5, 1989: 231–247. 98

Big History and macro evolution

Culotta, Elizabeth. “A New Human Ancestor?” Science 284, 1999: 572–573. Darwin, Charles. The Variation of Animals and Plants under Domestication. 2nd ed., rev. New York: D. Appleton, 1883. Dawkins, Richard. The Selfish Gene. Moscow: Mir, 1993. (Докинз, Р. Эгоистичный ген. М.: Мир, 1993.) Dunbar, Helen; with Alex Wilson, Nicole Ferguson and Nancy Moran. “Aphid Thermal Tolerance is Governed by a Point Mutation in Bacterial Symbionts.” PLoS Biology 5(5), 2007: e96. Dunbar, Robin. “The Social Brain: Mind, Language, and Society in Evolutionary Perspective.” Annual Review of Anthropology 32(1), 2003: 163–181. Dunbar, Robin; with Susanne Shultz. “Evolution in the Social Brain.” Science 317, 2007: 1344–1347. Durkheim, Emile. On Social Division of Labor. A Sociological Method. Moscow: Nauka, 1991 [1893]. (Дюркгейм Э. О разделении общественного труда. Метод социологии. М.: Наука, 1991 [1893].) Frigaard, Niels-Ulrik; with Asuncion Martinez, Tracy Mincer and Edward DeLong. “Proteorhodopsin Lateral Gene Transfer between Marine Planktonic Bacteria and Archaea.” Nature 439, 2006: 847–850. Gibbons, Ann. “Chromosome Shows that Adam was an African.” Science 278, 1997: 804–805. Grinin, Leonid. Formations and Civilizations. Ch. 2. Filosofia i obschestvo 2, 1997: 5–89. (Гринин, Л. Е. Формации и цивилизации. Гл. 2. Философия и общество 2, 1997: 5–89.) Grinin, Leonid. Formations and Civilizations. [The book was published in the journal ­Filosofia i obschestvo in 1997–2001]. (Гринин, Л. Е. Формации и цивилизации. [Книга печаталась в журнале Философия и общество с 1997 по 2001 г.]). Grinin, Leonid. “The Early State and Its Analogues.” Social Evolution & History 2(1), 2003: 131–176. Grinin, Leonid. “Early State and Democracy.” The Early State, Its Alternatives and Analogues. Leonid Grinin, Robert Carneiro, Dmitri Bondarenko, Nikolay Kradin and Andrey Korotayev (editors). Volgograd: Uchitel, 2004: pp. 419–463. Grinin, Leonid. “The Transformation of the Egypt State System in the 19th – Early 20th Centuries: From Developed State to the Mature one.” Egypt, Near East and Global World. Alexander Logunov (editor). Moscow: Krankes, 2006a: pp. 123–132. (Гринин, Л. Е. “Трансформация государственной системы Египта в XIX – начале XX вв.: от развитого государства к зрелому.” Египет, Ближний Восток и глобальный мир/ Ред. А. П. Логунов, М.: Кранкэс, 2006a. C. 123–132.) Grinin, Leonid. Productive Forces and Historical Process. 3rd ed. Moscow: KomKniga, 2006b. (Гринин, Л. Е. Производительные силы и исторический процесс. Изд. 3-е. М.: КомКнига, 2006b.) Grinin, Leonid. “The Problems of Analysis of Driving Forces of Historical Development, Social Progress and Social Evolution.” Philosophy of History: Problems and Prospects. Yuri Semyonov, Ivan Gobozov and Leonid Grinin (authors). Moscow: KomKniga/ URSS, 2007a: pp. 148–247. (Гринин, Л. Е. “Проблемы анализа движущих сил исторического развития, общественного прогресса и социальной эволюции.” Философия истории: проблемы и перспективы/Ю. И. Семенов, И. А. Гобозов, Л. Е. Гринин. М.: КомКнига/URSS, 2007a. C. 148–247.) Grinin, Leonid. “Production Revolutions as the Most Important Thresholds of History.” The Man and Nature: Confrontation and Harmony. Eduard Kulpin (editor). Series ‘Socionatural history. Genesis of Natural and Societal Crises in Russia’. Moscow: Publishing Analytical Centre ‘Energy’, 2007b: pp. 191–221. (Гринин, Л. Е. “Производственные революции как важнейшие рубежи истории.” Человек и природа: противостояние 99

Leonid E. Grinin et al.

и гармония/Ред. Э. С. Кульпин. Серия «Социоестественная история. Генезис кризисов природы и общества в России». М.: Издательско-аналитический центр «Энергия», 2007b. C. 191–221.) Grinin, Leonid. State and Historical Process. Political Cut of Historical Process. 2nd ed. ­Moscow: LIBROCOM, 2009a. (Гринин, Л. Е. Государство и исторический процесс: Политический срез исторического процесса. Изд-е 2-е. М.: ЛИБРОКОМ, 2009а.) Grinin, Leonid. “The Pathways of Politogenesis and Models of the Early State Formation.” Social Evolution & History 8(1), 2009b: 92–132. Grinin, Leonid. State and Historical Process. The Evolution of the Statehood: From an Early to a Mature State. 2nd ed. Moscow: LIBROCOM, 2010. (Гринин, Л. Е. Государство и исторический процесс: От раннего государства к зрелому. Изд. 2-е. М.: ЛИБРОКОМ, 2010.) Grinin, Leonid. State and Historical Process. The Epoch of the State formation. Moscow: ­Izdatel’stvo LKI, 2011a. (Гринин, Л. Е. Государство и исторический процесс: Эпоха формирования государства: Общий контекст социальной эволюции при образовании государства. Изд. 2-е. М.: Изд-во ЛКИ, 2011a.) Grinin, Leonid. The Evolution of Statehood: From Early States to Global Society. Saarbücken: Lambert Academic Publishing, 2011b. Grinin, Leonid E., with David Baker, Esther Quaedackers and Andrey V. Korotayev. 2014. (Eds.). Teaching & Researching Big History: Exploring a New Scholarly Field. Volgograd: Uchitel, 2014. Grinin, Leonid; with Robert Carneiro, Andrey Korotayev and Fred Spier (editors). Evolution: Cosmic, Biological, and Social. Volgograd: Uchitel, 2011. Grinin, Leonid; with Andrey Korotayev. “Political Development of the World System: A Formal Quantitative Analysis.” History & Mathematics: Historical Dynamics and Development of Complex Societies. Peter Turchin, Leonid Grinin, Victor de Munck and Andrey Korotayev (editors). Moscow: KomKniga, 2006: pp. 63–114. Grinin, Leonid; with Andrey Korotayev. “Political Development of the World System: A Formal Quantitative Analysis.” History and Mathematics. Macrohistorical Dynamics of the Society and State. Sergei Malkov, Leonid Grinin and Andrey Korotayev (editors). Moscow: KomKniga, 2007a: pp. 49–101. (Гринин, Л. Е., Коротаев, А. В. “Политическое развитие Мир-Системы: формальный и количественный анализ.” История и Математика: макроисторическая динамика общества и государства/Ред. С. Ю. Малков, Л. Е. Гринин, А. В. Коротаев. М.: КомКнига, 2007a. C. 49–101.) Grinin, Leonid; with Andrey Korotayev. “Social Macroevolution and Historical Process.” Filosofia i obschestvo 2, 2007b: 19–66; 3: 5–48; 4: 17–50. (Гринин, Л. Е., Коротаев, А. В. “Социальная макроэволюция и исторический процесс.” Философия и общество 2, 2007b: 19–66; 3: 5–76; 4: 17–50). Grinin, Leonid; with Andrey Korotayev. “Macroevolution and the World-System: New Facets of Conceptualization.” Istoriya i sovremennost 1, 2008: 3–31. (Гринин, Л. Е., Коротаев, А. В. “Макроэволюция и Мир-Система: новые грани концептуализации.” История и современность 1, 2008: 3–31.) Grinin, Leonid; with Andrey Korotayev. “Social Macroevolution: Growth of the World System Integrity and a System of Phase Transitions.” World Futures 65(7), 2009а: 477–506. Grinin, Leonid; with Andrey Korotayev. Social Macroevolution. Genesis and Transformations of the World System. Moscow: LIBROCOM, 2009b. (Гринин, Л. Е., Коротаев, А. В. Социальная макроэволюция: Генезис и трансформации Мир-Системы. М.: ЛИБРОКОМ, 2009b.)

100

Big History and macro evolution

Grinin, Leonid; with Andrey Korotayev. “On Some Peculiarities of Social-Political Development of Ottoman Egypt (the 16th – 18th Centuries).” Vostok (Oriens) 1, 2009c: 46–62. (Гринин, Л. Е., Коротаев, А. В. ”О некоторых особенностях социальнополитического развития османского Египта (XVI–XVIII вв.).” Восток 1, 2009c: 46–62.) Grinin, Leonid; with Andrey Korotayev. ”On the Typological Characteristics of Statehood of Ottoman Egypt (the 16th – 19th Centuries).” Vostok 3, 2009d: 35–51. (Гринин, Л. Е., Коротаев, А. В. “О типологических характеристиках государственности в османском Египте XVI–XIX вв. (К постановке проблемы).” Восток 3, 2009d: 35–51.) Grinin, Leonid; with Andrey Korotayev. “The Epoch of the Initial Politogenesis.” Social Evolution & History 8(1), 2009e: 52–91. Grinin, Leonid; with Alexander Markov and Andrey Korotayev. Macroevolution in Wildlife and Society. Мoscow: LKI, 2008. (Гринин, Л. Е., Марков, А. В., Коротаев, А. В. Макроэволюция в живой природе и обществе. М.: ЛКИ, 2008.) Grinin, Leonid; with Alexander Markov and Andrey Korotayev. “Aromorphoses in Biological аnd Social Evolution: Some General Rules for Biological and Social Forms of Macroevolution.” Social Evolution & History 8(2), 2009a: 6–50. Grinin, Leonid; with Alexander Markov and Andrey Korotayev. “Aromorphoses in Wildlife and Society: An Experience of Comparing Biological and Social Forms of Macroevolution.” Evolution: Cosmic, Biological, and Social. Leonid Grinin, Alexander Markov and Andrey Korotayev (editors). Moscow: LIBROCOM, 2009b: pp. 176–225. (Гринин, Л. Е., Марков, А. В., Коротаев, А. В. “Ароморфозы в живой природе и обществе: опыт сравнения биологической и социальной форм макроэволюции.” Эволюция: космическая, биологическая, социальная/Ред. Л. Е. Гринин, А. В. Марков, А. В. Коротаев. М.: ЛИБРОКОМ, 2009b. C. 176–225.) Grinin, Leonid; with Alexander Markov and Andrey Korotayev. “Biological and Social Aromorphoses: A Comparison between Two Forms of Macroevolution.” Evolution: Cosmic, Biological, and Social. Leonid Grinin, Robert Carneiro, Andrey Korotayev and Fred Spier (editors). Volgograd: Uchitel, 2011: pp. 162–211. Hallpike, Christopher. Principles of Social Evolution. Oxford: Clarendon, 1986. Herrmann, Esther; with Josep Call, Maria Hernàndez-Lloreda, Brian Hare and Michael Tomasello. “Humans have Evolved Specialized Skills of Social Cognition: The Cultural Intelligence Hypothesis.” Science 317, 2007: 360–366. Heylighen, Francis. “Conceptions of a Global Brain: An Historical Review.” Evolution: Cosmic, Biological, and Social. Leonid Grinin, Robert Carneiro, Andrey Korotayev and Fred Spier (editors). Volgograd: Uchitel, 2011: pp. 274–289. Holden, Constance. “No Last Word on Language Origins.” Science 282, 1998: 1455–1458. Jablonka, Eva; with Marion Lamb. Epigenetic Inheritance and Evolution: Lamarckian Dimension. Oxford: Oxford University Press, 1999. Kaufman, Daniel. Archeological Perspectives on the Origins of Modern Humans. A View from Levant. Westport, CT: Bergin & Garvey, 1999. Klima, B. “The Period of Homo Sapiens of Modern Type till the Beginning of Food Production (Producing Economy): Overall Review (Except the Art).” History of Humanity. Vol. 1. Prehistoric Period and the Origins of Civilization. Z. De Laat (editor). Moscow: ­U NESCO, 2003: pp. 198–207. (Клима, Б. “Период человека разумного современного вида до начала производства пищи (производящего хозяйства): общий обзор (за исключением искусства).” История Человечества. Т. 1. Доисторические времена и начала цивилизации/Ред. З. Я. Де Лаат. М.: ЮНЕСКО, 2003. C. 198–207.)

101

Leonid E. Grinin et al.

Korotayev, Andrey. “Mountains and Democracy: An Introduction.” Alternative Pathways to Early State. Nikolay Kradin and Valery Lynsha (editors). Vladivostok: Dal’nauka. 1995: pp. 60–74. Korotayev, Andrey. The Sabaean essays. Some General Tendencies and Factors of Evolution of the Sabaean Civilization. Moscow: Vostochnaya literatura, 1997. (Коротаев, А. В. Сабейские этюды. Некоторые общие тенденции и факторы эволюции сабейской цивилизации. М.: Вост. лит-ра, 1997.) Korotayev, Andrey. “Objective Sociological Laws and a Subjective Factor.” Vremya mira 1, 1999: 204–233. (Коротаев, А. В. “Объективные социологические законы и субъективный фактор.” Время мира 1, 1999: 204–233.) Korotayev, Andrey. Social Evolution: Factors, Laws, Tendencies. Moscow: Vostochnaya literatura, 2003. (Коротаев, А. В. Социальная эволюция: факторы, закономерности, тенденции. М.: Восточная литература, 2003.) Korotayev, Andrey. World Religions and Social Evolution of the Old World Oikumene Civilizations: A Cross-cultural Perspective. Lewiston, NY: The Edwin Mellen Press, 2004. Korotayev, Andrey. “‘Midwest-Amazonian’ Folklore-Mythological Parallels?” Acta ­Americana 14(1), 2006: 5–24. Korotayev, Andrey. The 21st Century Singularity and Its Big History Implications: A Re-Analysis. Journal of Big History, 2(3), 2018: 73–119. Korotayev, Andrey; with Yuri Berezkin, Artem Kozmin and Alexandra Arkhipova. “Return of the White Raven: Postdiluvial Reconnaissance Motif A2234.1.1 Reconsidered.” Journal of American Folklore 119, 2006: 472–520. Korotayev, Andrey; with Nikolay Kradin, Victor de Munck and Valery Lynsha. ”Alternatives of Social Evolution: Introductory Notes.” Alternatives of Social Evolution. Nikolay Kradin, Andrey Korotayev, Dmitri Bondarenko, Victor de Munck and Paul Wason (editors). Vladivostok: FEB RAS, 2000: pp. 12–51. Korotayev, Andrey; with Nikolay Kradin, Victor de Munck and Valery Lynsha. “Alternatives of Social Evolution: Introductory Notes.” Alternatives of Social Evolution. Nikolay Kradin, Andrey Korotayev and Dmitri Bondarenko (editors). 2nd ed. Saarbücken: ­Lambert Academic Publishing, 2011: pp. 12–51. Korotayev, Andrey; with Darja Khaltourina. Myths and Genes. Moscow: LENAND/URSS, 2011. (Коротаев, А. В., Халтурина, Д. А. Мифы и гены. М.: ЛЕНАНД/URSS, 2011.) Korotayev, Andrey; with Artemy Malkov and Darja Khaltourina. Introduction to Social Macrodynamics: Compact Macromodels of the World System Growth. Moscow: KomKniga/ URSS, 2006. Kutter, G. S. 2015. In: B. Rodrigue, L. Grinin, A. Korotayev. 2015. (Eds.). From Big Bang to Galactic Civilizations. Vol. 1: Our Place in the Universe. An Introduction to Big History. Delhi: Ratna Sagar, 2015. pp. 24–41. Lambert, David. The Prehistoric Man. The Guide-book of Cambridge. Leningrad: Nedra, 1991. (Ламберт, Д. Доисторический человек. Кембриджский путеводитель. Л.: Недра). Lareau, Liana; with Maki Inada, Richard Green, Jordan Wengrod and Steven Brenner. “Unproductive Splicing of SR Genes Associated with Highly Conserved and Ultraconserved DNA Elements.” Nature 446, 2007: 926–929. Lekevičius, Edmundas. “On Some Analogues between Ecosystems’ Evolution and Economical Development: From A. Smith and Ch. Darwin to the Newest Ideas.” ­Evolution: Cosmic, Biological, and Social. Leonid Grinin, Alexander Markov and Andrey ­Korotayev (editors). Moscow: LIBROCOM, 2009: pp. 226–259. (Лекявичюс, Э. “О некоторых аналогиях между эволюцией экосистем и развитием экономики: от А. Смита и Ч. Дарвина до новейших идей.” Эволюция: космическая, биологическая, социальная/Ред. Л. Е. Гринин, А. В. Марков, А. В. Коротаев. М.: ЛИБРОКОМ, 2009. C. 226–259.) 102

Big History and macro evolution

Lekevičius, Edmundas. “Ecological Darwinism or Preliminary Answers to Some ­Crucial though Seldom Asked Questions.” Evolution: Cosmic, Biological, and Social. Leonid ­Grinin, Robert Carneiro, Andrey Korotayev and Fred Spier (editors). Volgograd: ­Uchitel, 2011: pp. 101–121. Leonova, Natalia; with Sergey Nesmeyanov (editors). The Problems of the Ancient Societies’ Paleoecology. Moscow: Russian Open University, 1993. (Леонова, Н. Б., Несмеянов, С. А. (Ред.) Проблемы палеоэкологии древних обществ. М.: Российский открытый университет, 1993.) Markov, Alexander; with Elena Naymark. “On Some Newest Achievements of Evolutionary Biology.” Evolution: Cosmic, Biological, and Social. Leonid Grinin, Alexander Markov and Andrey Korotayev (editors). Moscow: LIBROCOM, 2009: pp. 306–363. (Марков, А. В., Наймарк, Е. Б. “О некоторых новейших достижениях эволюционной биологии.” Эволюция: космическая, биологическая, социальная/Ред. Л. Е. Гринин, А. В. Марков, А. В. Коротаев. М.: ЛИБРОКОМ, 2009. C. 306–363.) Mellars Paul; with Chris Stringer (editors). The Human Revolution: Behavioural and Biological Perspectives on the Origins of Modern Humans. Princeton, NJ: Princeton University Press, 1989. Mesoudi, Alex; with Andrew Whiten and Kevin Laland. “Towards a Unified Science of Cultural Evolution.” Behavioral and Brain Sciences 29, 2006: 329–383. Nashimoto, Masayuki. “The RNA/Protein Symmetry Hypothesis: Experimental Support for Reverse Translation of Primitive Proteins.” Journal of Theoretical Biology 209, 2001: 181−187. Pääbo, Svante. “The Y-Chromosome and the Origin of All of Us (Men).” Science 268, 1995: 1141–1142. Pagel, Mark; with Quentin Atkinson, and Andrew Meade. “Frequency of Word-Use Predicts Rates of Lexical Evolution throughout Indo-European History.” Nature 449, 2007: 717–720. Philippov, Igor. “The Emergence of Feudalism in Western Europe.” History of the Middle Ages. Zinaida Udaltsova, Sergei Karpov (editors). Moscow: Vysshaya skola, 1990: vol. 1, pp. 42–84. (Филиппов, И. С. “Возникновение феодального строя в Западной Европе.” История средних веков/Ред. З. В. Удальцова, С. П. Карпов. М.: Высшая школа, 1990. Т. 1. C. 42–84.) Popov, Sergei. “The Problem of Adaptation in Social Structure Studies.” Jurnal obschey biologii 67(5), 2006: 335–343. (Попов, С. В. Проблема адаптивности при исследовании социальных структур. Журнал общей биологии 67(5), 2006: 335–343. Rautian, Alexander. “Paleontology as a Source of the Data on Evolutionary Laws and Factors.” Modern Paleontology. Vladimir Menner and Vladimir Makridin (editors). Moscow: Nedra, 1988a: vol. 2, pp. 76–118. (Раутиан, А. С. “Палеонтология как источник сведений о закономерностях и факторах эволюции.” Современная палеонтология/ Ред. В. В. Меннер, В. П. Макридин. М.: Недра, 1988a. T. 2. C. 76–118.) Rautian, Alexander. “Dictionary of Terms and Subjects.” Modern Paleontology. Vladimir Menner and Vladimir Makridin (editors). Moscow: Nedra, 1988b: vol. 2, pp. 356–372. (Раутиан, А. С. “Словарь терминов и наименований.” Современная палеонтология/ Ред. В. В. Меннер, В. П. Макридин. М.: Недра, 1988b. T. 2. C. 356–372.) Reeve, H. Kern; with Bert Hölldobler. “The Emergence of a Superorganism through Intergroup Competition.” Proceedings of the National Academy of Sciences of the USA 104(23), 2007: 9736−9740. Reznikova, Zhanna. “Evolutionary and Behavioural Aspects of Altruism in Animal ­Communities: Is there Room for Intelligence?” Evolution: Cosmic, Biological, and Social. ­Leonid Grinin, Robert Carneiro, Andrey Korotayev and Fred Spier (editors). ­Volgograd: Uchitel. 2011: pp. 162–211. 103

Leonid E. Grinin et al.

Rhesus Macaque Genome Sequencing and Analysis Consortium. “Evolutionary and ­Biomedical Insights from the Rhesus Macaque Genome.” Science 316, 2007: 222–234. Robson, Simon; with James Traniello. “Transient Division of Labour and Behavioral Specialization in the Ant.” Formica schaufussi. Naturwissenschaften 89, 2002: 128–131. Rodrigue Barry, with Leonid Grinin and Andrey Korotayev (Eds.). From Big Bang to ­Galactic Civilizations. Vol. 1: Our Place in the Universe. An Introduction to Big History. Delhi: Ratna Sagar, 2015. Rodrigue Barry, with Leonid Grinin and Andrey Korotayev (Eds.). From Big Bang to ­Galactic Civilizations. Vol. 2: Education and Understanding Big History around the World. Delhi: Ratna Sagar, 2016. Rodrigue Barry, with Leonid Grinin and Andrey Korotayev (Eds.). From Big Bang to ­Galactic Civilizations. Vol. 3: The Ways that Big History Works: Cosmos, Life, Society and our Future. Delhi: Ratna Sagar, 2017. Ryabko, Boris; with Zhanna Reznikova. “The Use of Ideas of Information Theory for Studying ‘Language’ and Intelligence in Ants.” Entropy 11(4), 2009: 836–853. Sahlins, Marshall. Stone Age Economics. New York: Aldine de Gruyter, 1972. Shatalkin, Anatoly. “Molecular Phylogenies – Revolutionary Breakthrough in Systematics.” Evolutionary Factors of Forming Animal World. Emilia Vorobyeva and Bella Striganova (editors). Moscow: KMK, 2005: pp. 30–42. (Шаталкин, А. И. “Молекулярные филогении – революционный прорыв в систематике.” Эволюционные факторы формирования разнообразия животного мира/Ред. Э. И. Воробьева, Б. Р. Стриганова. М.: КМК, 2005. C. 30−42.) Shettleworth, Sara. Cognition, Evolution, and Behavior. New York, NY: Oxford University Press, 1998. Shnirel’man, Victor. “Gatherers of Sago.” Voprosy istorii 11, 1983: 182–187. (Шнирельман, В. А. “Собиратели саго.” Вопросы истории 11, 1983: 182–187.) Shnirel’man, Victor. “Late Primitive Community of Farmer-Breeders and Advanced Hunters, Fishers and Gatherers.” The History of Primitive Society. The Epoch of the Primitive Tribal Community. Yulian Bromley (editor). Moscow: Nauka, 1986: pp. 236–426. (Шнирельман, В. А. “Позднепервобытная община земледельцев-скотоводов и высших охотников, рыболовов и собирателей.” История первобытного общества. Эпоха первобытной родовой общины/Ред. Ю. В. Бромлей. М.: Наука, 1986. C. 236–426.) Shnirel’man, Victor. Emergence of Producing Economy. Moscow: Nauka, 1989. (Шнирельман, В. А. Возникновение производящего хозяйства. М.: Наука, 1989.) Spencer, Herbert. “Principles of Sociology.” Complete Works. Herbert Spencer (author). Vol. 1. St. Petersburg: T-vo I. D. Sytina, Otd. N. A. Rubakina, 1898. (Спенсер, Г. “Основания социологии.” Соч. Т. 1. СПб.: Т-во И. Д. Сытина, Отд. Н. А. Рубакина, 1898.) Spier, F. (2015). Big History and the Future of Humanity. Chichester: John Wiley & Sons. Steele, Edward; with Robin Lindley and Robert Blanden. What If Lamark Is Right? ­Immunoge-netics and Evolution. Moscow: Mir, 2002. (Стил, Э., Линдли, Р., Бланден, Р. Что, если Ламарк прав? Иммуногенетика и эволюция. М.: Мир, 2002) Stringer, Christopher. The Emergence of Modern Humans. Scientific American December, 1990: 68–74. Тimofeev-Ressovsky, Nikolay; with Nikolay Vorontsov and Alexei Yablokov. Brief Essay on Evolution Theory. Moscow: Nauka, 1969. (Тимофеев-Ресовский, Н. В., Воронцов, Н. Н., Яблоков, А. В. Краткий очерк теории эволюции. М.: Наука, 1969.). Van Parijs, Philippe. Evolutionary Explanation in the Social Sciences: An Emerging Paradigm. Totowa, NY: Rowman and Littlefield, 1981. Vanyushin, Boris. “Materialization of Epigenetics, or Little Changes and their Big Consequences.” Himiya i zhizn’ 2, 2004: 32–37. (Ванюшин, Б. Ф. “Материализация 104

Big History and macro evolution

эпигенетики, или Небольшие изменения с большими последствиями.” Химия и жизнь 2, 2004: 32–37.) Vignuzzi, Marco; with Jeffrey Stone, Jamie Arnold, Craig Cameron and Raul Andino. “Quasispecies Diversity Determines Pathogenesis through Cooperative Interactions in a Viral Population.” Nature 439, 2005: 344–348. Vygotsky, Lev. Mind in Society: The Development of Higher Psychological Processes. Cambridge, MA: Harvard University Press, 1978. White, Tim D.; with Berhane Asfaw, David DeGusta, Henry Gillbert, Gary D. ­R ichards, Gen Suwa and F. Clark Howell. “Pleistocene Homo sapiens from Middle Awash, ­Ethiopia.” Nature 423, 2003: 742–747. Wright, David. The History of China. Westport, CT: Greenwood, 2001. Yaryghin, Vladimir; with Veronica Vasilyeva, Igor Volkov and Valerija Sinelschikova. ­Biology. Moscow: Vysshaya shkola, 1999. (Ярыгин, В. Н., Васильева, В. И., Волков, И. Н., Синельщикова, В. В. Биология. М.: Высшая школа, 1999.) Zakharov, Anatoly. Ants, Family, Colony. Moscow: Nauka, 1978. (Захаров, А. А. Муравей, семья, колония. М.: Наука.1978.) Zhdanko, Alexei. “Letter to the Editors. Remarks on the Prehistory (Modern Data of ­Paleontology and Paleoarchaeology about Origins of Homo sapiens.” Filosophia i obschestvo 5, 1999: 175–177. (Жданко, А. В. “Письмо в редакцию. Заметки о первобытной истории (современные данные палеоантропологии и палеоархеологии о возникновении Homo sapiens).” Философия и общество 5, 1999: 175–177). Zhivotovsky, Lev. “On Inheritance of Acquired Characteristics.” Materials of Academic Genetic Conference. February 26–27, 2002. Moscow: Timiryazev Agricultural Academy, 2002a: pp. 110–119. (Животовский, Л. А. “О наследовании приобретенных признаков.” Материалы научной генетической конференции. 26−27 февраля 2002 г.. М.: Изд-во Московской сельскохозяйственной академии им. К. А. Тимирязева, 2002a. C. 110−119.) Zhivotovsky, Lev. “A Model of the Early Evolution of Soma-to-Germline Feedback.” Journal of Theoretical Biology 216, 2002b: 51–57. Zimmer, Carl. Great Mysteries of Human Evolution. Discover 24(9), 2003: 34–44.

105

PART II

Big history, social science and the humanities

5 BIG HISTORY AND ANTHROPOLOGY Our place in the multiverse: anthropology, civilization and big history Barry H. Rodrigue We humans engage in a constant process of enlarging our understanding of the world around us. As our ancestors spread throughout Africa and beyond, they developed innovative strategies for survival – from tools and clothing to languages and customs. When dispersed human groups came into contact with each other, they shared ideas and genes. Self-awareness is a result of such interactions. This merging led to intercultural thinking of humanity as a global community, which in turn led to the birth of what we call anthropology. It was a method of self-actualization – by better comprehending our place in the world, we adapted ourselves and our surroundings. We are again at the threshold of a new self-awareness, a product of the consolidation of scholarship and global contacts to form what has been called cosmic evolution, big history and universal studies. This expanded worldview is a product of taking a sense of a unified humanity to the next level – to the recognition that we are but one of many symbiotic life forms on Earth and but one entity within a much larger cosmos.1

Anthropography When did human self-consciousness expand beyond the level of other animals to become a focus of society? Did it happen for our Australopithecine ancestors three million years ago? Were the half-million-year-old shell-etchings of Homo erectus in Java an expression of that curiosity, or the Neanderthal stone rings in Occitania? Certainly, the cave and rock art of early humans in Africa and Eurasia and Australia demonstrates the studied engagement with the world (Figure 5.1). At some point, mystical conjecture fused with practical knowledge to form a continuum of human understanding about the world. Inside the Epic of Gilgamesh, for example, a story of the search for eternal life mingled with encounters between Palaeolithic and Neolithic peoples. This fusion of ideas certainly did not provide a seamless fabric of awareness, but it was a beginning, an incipient form of anthropology and big history. 109

Barry H. Rodrigue

Figure 5.1  R  ecently discovered cuneiform Tablet V of the Gilgamesh Epic, c 1800 bce, which provides fresh insights about ethno-geographic encounters in the epic’s Cedar Forest. Slemani Museum, Sulaymaniyah, Kurdistan, Iraq.2 Photograph by Osama Amin, 2015, Wikimedia Commons. Greyscale conversion of original image.

Something as simple as the repurposing of artefacts in Palaeolithic times indicates human connections with their past. This process increased during the Neolithic, as growing population densities and more extensive forms of farming, industry and habitation led to the recycling of artefacts to supplement new activities. Reuse of structural remains became so common in ancient Egypt that admonitions against such spoliation were proclaimed by dynastic officials in the second millennium bce.3 We can thus see the roots of anthropography, the study of cultural literacy, as having its origin in the deep shadows of our ancestors’ existence. Writing systems led to better preservation of human ideas about their engagement with the world. We see how Shang dynasty oracle-bone inscriptions from 3,200 years ago expressed concern about their votaries’ place in society, the landscape and the cosmos.4 Soon afterwards, a surge in human self-reflection appeared in works by Wenamun, Sappho, Laozi and Mahavira. Some refer to this as an axial age, but it was as much a result of the wider use of writing (Figure 5.2a and b).5 The rise of Neolithic class structure and leisure time for elites led to a more focused curiosity about the past. The Tisbury Hoard from Wiltshire (England) included items whose dates span a millennium, and so they are considered an artefact collection of the ninth century BCE. Neo-Babylonian King Nabonidus, in the sixth century BCE, established a museum in Ur (Iraq) curated by his daughter, Ennigaldi.7 In this fashion, collecting artefacts became a strategy to empower leadership, along with its artisanal panoply of grave-robbers, treasure hunters and traders. This self-reflection about heritage resulted from wider social interactions and deeper political structures, which manifested themselves in increased cross-cultural expressions, as when King Darius I of Persia commissioned a proclamation etched in three cuneiform languages onto a cliff face in the Zagros Mountains, along the road to Mesopotamia, 2,500 years ago. The Bisotun inscription is an imperial account of conquest, but, since the Persian Empire was a multicultural polity, it also illustrated intentional interaction between its different societies (Figure 5.3).8 110

(a)

(b)

Figure 5.2  P  etroglyphs and a Picenean inscription found in the vicinity of Mt. Conero in Italy. Photographs by Alessandro Montanari (a) and Roland Saekow (b).6 ­Greyscale conversion of original image.

Figure 5.3  B  isotun inscription. John Quackenbos, Illustrated History of Ancient Literature, Oriental and Classical, New York: Harper & Brothers, 1882, p. 65. Greyscale conversion of original image.

Barry H. Rodrigue

Anthropological studies are often dated to Herodotus, a scholar who had set out to document the struggle of the Persian Empire to annex the Greek states in the fifth century BCE. Since the invading Persian forces were multicultural, his resulting Histories expanded into a wide-ranging regional study of cultures, customs and languages.This contemplation of human societies was continued by later Mediterranean scholars, from Tacitus and Ptolemy to Pausanias. Although deep literacy was still not common, public reading by lettered retainers allowed narratives to reach a wider audience and encouraged interest in other peoples and places.9 Some, like philosopher-poet Lucretius in his verses On the Nature of Things in the first century bce, expressed a materialist view of the universe and a unitary sense of humanity, presaging anthropology and big history by hundreds of years. Literacy grew with the technology and infrastructure supporting it. Travellers’ accounts became popular, as with Faxian’s journey from China to India to collect Buddhist documents in the fifth century CE or Aḥmad ibn Faḍlān’s envoy from the Caliph of Baghdad to the Volga Bulgars in the tenth century.10 Their observations of foreign customs entered a variety of activities, from political policy and geographic reports to religious discourse and popular story-telling. Along with literacy, schools began for elite males and those with noted skills. Centres for learning sprang up in places like Nalanda (India), which drew students from around Asia, while Inca aristocracy along the Andes attended the yachay wasi (house of knowledge) for instruction in reading quipu, mathematics and public affairs. In this way, questioning of the wider world was debated in academic settings but then percolated through society as students returned to their home communities. Much cross-cultural material also existed in non-public works. Evidence of this came from the Cairo Geniza, a Jewish document repository of the last millennium that held rich detail of cultural interaction by merchants from the Mediterranean through the Arabian Peninsula to India and beyond. Indeed, Marco Polo’s thirteenth-century Travels, with its rich descriptions of Asia, was the narrative of a business venture that only came to be recorded by happenstance.11 In Europe, Renaissance discovery of Classical knowledge in scriptoriums and the libraries of Al-Andalus not only revived interest in ancient scholars like Herodotus and Lucretius but also required researchers to negotiate exotic cultural traditions so as to access the materials. A manifestation of this upsurge in knowledge was a demand for relics, so much so that Papal authorities interdicted the black-market export of Roman artefacts in 1461. The Vatican also began public exhibition of heirlooms on Capitoline Hill a decade later to foster cultural pride.12 Nor was this new intercultural awareness confined to Eurasia, as we see contact with the Americas going back prior to the Columbian exchange, as with Polynesian voyages to South America and transient European settlement in North America over a 1,000 years ago. What effect did these meetings have on the societies involved? It is unclear, but a market developed for imaginative stories, as in the aja’ib and mirabilia genres of ‘wonder’ tales (Figure 5.4).13 The slow pace of these contacts and regional political consolidations set the stage for a more intense period of globalization. Just as millennia of encounters slowly encouraged intergroup reflection, a fierce period of worldwide engagement began in the fifteenth century and forged an even more integrated understanding of 112

Big history and anthropology

Figure 5.4  A  ntler carving of faces interpreted as Dorset (below) and European (above), c. fourteenth century ce, Baffin Island, Nunavut (Canada).14 Photograph by David Coventry. Greyscale conversion of original image.

humanity’s place in the world.The process was not new, but it was more frenzied and is still going on today. It was the profound impact of this new globalization that led to the modern formulation of both anthropology and big history.

Globalization European colonial expansion in the fifteenth century led to profound changes in understandings about humanity, but there was no metaphysical quality of European society that unleashed their hegemony on the world. The process can be described in a ‘Global Algorithm’: Asian invention + Afro-American resource + European gestalt = Global civilization The formula’s categories are shorthand for physical and intellectual exchange. Of course, this global dynamism was far more complex than a mere algorithm can convey, but it seeks to illustrate that – far from being just a European-driven ­phenomenon – the new global engagement had grown from the vast silk-route network into a 113

Barry H. Rodrigue

Figure 5.5  S ultan Firoz Shah Tughlaq of the Delhi Sultanate had this Ashokan pillar (third century bce) removed from Topra Kalan (Haryana) to Firozabad as part of his antiquarian collections in 1356 ce. The use of monumental stoneworks fused with the collection of artefacts to create a form of antiquarianism that sought to legitimize rulers.16 The column now is within the bounds of Delhi, India. Photograph by Samuel Bourne, 1860, British Library Online Gallery. Greyscale conversion of original image.

planetary sphere of interaction that is more properly designated as ‘global civilization’ (Figure 5.5).15 This global process can be seen at work when Mughal emperor Akbar I (1542–1605) institutionalized the concept of Sulh-i-Kul (universal peace), which sought to harmonize cultural and spiritual traditions from around Eurasia and the world. Originally developed by Sufi scholar Ibn Arabi of Iberia in the twelfth century, his philosophical ideal drew courtiers to South Asia and led to wide communication, as in Akbar’s celebrated correspondence with Phillip II of Spain and others.17 In Europe, aggressive colonialism and Enlightenment efforts to understand their expanding worldview led to new thoughts of how things fit together. Giambattista Vico’s The New Science (1725) built a framework of universal history, while Carl ­Linnaeus’ Systema Naturae (1735) constructed a matrix of biological connectivity. ­Denis Diderot and Jean le Rond d’Alembert’s Encyclopédie (1751+) sought to amalgamate knowledge into useful forms for people’s use. Johann Friedrich Blumenbach is considered one of the first modern anthropologists. His dissertation at the University of Göttingen, On the Natural Variety of Humankind (1775), was an anatomical classification of human races, but it premised a single human species with all groups having equal potential and variations being the result of environmental factors.18 Human studies were not well demarcated at this time, as artefacts, languages and folklore were lumped together as ‘antiquities’ and came to be linked to notions of historical romanticism and nationalism.19 As the wealth generated by global trade trickled through society, new infrastructures for the study of humanity resulted. The Society of Antiquaries of London emerged from a Westminster tavern in 1717, when a group met to discuss how to preserve historic buildings. It was a domain of amateurs, where the banal coexisted with the erudite. In 1794, William Shakespeare’s 114

Big history and anthropology

head was looted from his tomb to satisfy an influential English antiquarian. Two decades later, folklorist and author Walter Scott wrote of these contrasting values in his best-selling novel, The Antiquary.20 Scholars struggled to make sense of the facts, minutiae and notions that were jumbled together by antiquarians. Neo-Confucian scholar Miura Baien (1723–1789) merged Japanese concepts with Chinese and European ideas to develop a new vision of the world, one that has been compared favourably with the later studies of Alexander von Humboldt.21 In Tibet, such intercultural views coalesced to produce new visions of interaction, as in The Detailed Description of the World (1830), a synthesis of   Tibetan worldviews with cosmopolitan ideas acquired in China. However, when Lobsang Palden Chopal, the Chief Minister in Shigatse, sought to expand on such new knowledge 50 years later, he was executed by a Tibetan government fearful of Russian, British and Chinese encroachment.22 Many Europeans viewed human society as a stage of global and progressive advancement, but one in which western European society formed the leading edge, a logic that often became justification for overseas expansion.23 In 1784, European colonials established the Asiatic Society of Bengal, which set up the Oriental Museum in Kolkata 30 years later, the oldest collection of its kind in India. Initially run by Europeans, local elites were finally allowed to join after 1829, including members of the celebrated Tagore family of artists and scholars.24 European colonialism exacerbated problems of overseas appropriation of cultural materials, such as the Parthenon marbles’ acquisition by the Earl of Elgin in 1801 (during the Ottoman Empire’s control of Greece) and their placement at the British Museum. The colonial traffic in antiquities led Egyptian ruler Mohamed Ali Pasha to ban artefact exports in 1835 and establish a storehouse for them. In 1858, his son created the Antiquities Service, which oversaw excavations in Egypt, and began the first museum in the Middle East five years later.25 The mid-century impact of evolution and materialist thought on society was profound. Although evolutionary theories and timeframes had earlier existed, the mechanisms were not so well enunciated as those by Charles Darwin (1859), Herbert Spencer (1862) and Karl Marx (1867). Spencer’s theories especially had an impact on anthropological thought, along with those of anthropologists Edward Burnett Tylor (1871) and Lewis Henry Morgan (1877) (Figure 5.6).26 People certainly knew of the many varieties of cultural expression at this time, but the overarching concept of culture was just beginning to be identified and elaborated. In 1871, Tylor provided one of the first definitions of culture in its social context. Culture or Civilization, taken in its wide ethnographic sense, is that complex whole which includes knowledge, belief, art, morals, law, custom, and any other capabilities and habits acquired by man as a member of society.27 Tylor saw anthropology to be in the lineage of universal history and interdisciplinarity.28 It thus came to be appreciated that anthropology was not just about ‘things’ that antiquarians collected but a matter of ‘process’ – in an interactive and evolutionary dynamic. Human studies continued to be hampered by pseudo-scientific theory, which included crude notions of biological and social evolution that set up hierarchies of superior and inferior species, races and societies.These populist views served to justify 115

Barry H. Rodrigue

Figure 5.6  H  athor shrine (fifteenth century bce), Deir el-Bahari, Egypt. The remains are preserved in the Museum of Egyptian Antiquities in Cairo. Photograph by Henri Édouard Naville, 1907, Wikimedia Commons. Greyscale conversion of original image.

institutions of class, caste and slavery, since those lower on a biased evolutionary tree could be considered non-human or un-civilized and deserving of diminished existence. Social contradictions came into a heated debate in Europe and its colonial outposts, as during the U.S. Civil War (1861–1865), a bloodbath that liberated African-American slaves but left them at the bottom of a racist social system. Scholarship began to more forcefully challenge such conceits, as research strengthened the precept that, despite differences, humans were a unified species. It was a long process that is still going on today.29 As the concept of a global humanity gained wider support, it led to the professionalization of human studies. Begun by amateurs, anthropology required little training or equipment. Folklore and linguistics were accessed by conversation and observation, while anyone with a shovel could begin archaeological excavation. Biological anthropology was seen as an extrapolation from everyday hybridizing of farm crops and livestock. In 1878, the secretary of the Smithsonian Institution noted that anthropology was the most popular branch of science.30 Anthropologists did not need certification: One could enter the field as a self-trained amateur. For example, as a young man, Edward Sylvester Morse (1838–1925) was known for his study of snails and his drafting abilities along the Gulf of Maine. He came to the attention of zoologist Louis Agassiz, who hired him to work at nearby Harvard University. Although lacking a degree, Morse became co-founder of the journal, The American Naturalist, as well as a lecturer at Harvard and a fellow of the National Academy of Sciences.31 In the 1870s and 1880s, Ned Morse went to Japan as an advisor on educational reform during the Meiji Restoration. His collection of brachiopods led him to discover the Ōmori shell mound and to identify Jōmon culture (16,500 BP), while his study of Japanese ceramics and architecture cemented his reputation as an anthropologist. What began as an amateur passion developed into professional work. 116

Big history and anthropology

Figure 5.7  E  xcavation of the Ōmori shell mound, Jōmon culture, Shinagawa (Tokyo), ­Japan, c 1877. Frontispiece, Edward Sylvester Morse, ‘Shell Mounds of Omori’, Memoirs of the Science Department, University of Tokio, Japan, vol. 1, part 1, Tokyo: University of Tokyo, 1879. Greyscale conversion of original image.

His international work and questions about the origins of Japanese society led to the founding the Anthropological Society of Japan in 1884 and the first academic journal of anthropology two years later. In 1892, a member of this group of scholars, Tsuboi Shogoro, became the first professor of anthropology at the University of ­Tokyo (­Figure 5.7).32 In order to take advantage of amateur endeavours in anthropology, the British ­Association for the Advancement of Science published Notes and Queries on Anthropology: For the Use of T   ravellers and Residents in Uncivilized Lands in 1874. Adopting ­Tylor’s vision of progressive social evolution, the volume proclaimed an inclusive view of humanity: ‘History has confined itself chiefly to the achievements of special races; but the anthropologist regards all races as equally worthy of a place in the records of human development’. It also provided less noble suggestions: ‘If after a battle, or other slaughter, the head of a native can be obtained with the soft parts in it, it might be preserved and transmitted carefully and perfectly closed up in a small keg filled up with spirit, or brine thoroughly saturated with salt’.33 Different cultural traditions had fascinated people for millennia, and, as more studious works came into being, they continued to mingle with populist representations of humanity. The Great Exhibition in London and the Smithsonian Institution in Washington D.C. featured cultural exhibits from around the world, but so did P.T. Barnum’s circus and Buffalo Bill’s Wild West Show. Such populist and professional tensions led scholars to further specialize cultural studies, which led to the establishment of more schools of anthropology, journals, societies, a standard vocabulary and regulations to protect antiquities (Figure 5.8).34 It is difficult to estimate the numbers of anthropologists engaged in work at this time. As university infrastructure divided into disciplines and departments in the late 117

Barry H. Rodrigue

Figure 5.8  F  rench postcard from Buffalo Bill’s Wild West Show, 1903. MS 327, James Wojtowicz Collection, McCracken Research Library, Buffalo Bill Center of the West, Cody, Wyoming, USA. Greyscale conversion of original image.

nineteenth century, anthropology likewise segmented. Folklorists were found in literature departments, physical anthropologists in medicine, linguists in languages, archaeologists in geology and ethnologists in sociology. In 1876, ethnographer Alfred Russel Wallace noted the ‘chaotic state of the infant science of anthropology’.35

Professionalization Anthropologists worked to document traditional culture at a time when indigenous societies were being rapidly transformed by industrial society and colonial contact. These inquiries led to more involvement by state agencies. When the United States became embroiled in wars with native peoples in its western territories, anthropologists like Alice Fletcher and James Mooney worked to establish the Bureau of Ethnology (1879) as a repository for tribal materials. In Britain, the Ordnance Survey had been charged with mapping as an adjunct to its military operations, but, by the mid-nineteenth century, its scope of work expanded to include archaeology, folklore and other landscape-related topics. Similar partnerships took place elsewhere, from South Asia to South America.36 In 1899, Franz Boas set up the first department of anthropology in the United States, at Columbia University, by bringing together the study of archaeology, ethnology, linguistics and physical anthropology into what became known as the four-field system. Other scholars were even more inclusive.Wilhelm Wundt began his academic career in Germany as a professor of anthropology and, although later celebrated as the founder of psychology, saw the ‘philosopher and historian, theologian and ethnologist 118

Big history and anthropology

in unified work’ to understand humanity.37 A key purpose of anthropology was to identify what it was to be human through interdisciplinary methods. Although a study of culture was central to anthropology, identifying it as a concept only came as a result of comparative work, which Ruth Benedict illustrated in her Patterns of Culture (1934). Culture is almost unconscious – we don’t think about how we walk, for example, we just do it as something learnt in our society. But by studying other traditions of walking, we come to appreciate our own tradition and are able to then formulate general theories about pedestrian behaviour. In this way, an examination of differences and commonalities between species, races, ethnicities and languages developed, which led to better theoretical understandings about how society functions.38 For example, ethnographers documented ceremonial exchange among remote peoples around the world. Franz Boas investigated the potlatch among the Kwakiutl in northern British Columbia (1921), while Bronislaw Malinowski studied the Kula Ring of the Trobriand Islanders in eastern New Guinea (1922). When Marcel Mauss conceptualized gift exchange as a form of social reciprocity (1925), these disparate phenomena were appreciated as a cultural artefact of humanity as a whole.39 The pendulum swing between data collection and concept synthesis lies at the heart of anthropology. Indigenous anthropology developed alongside Western models. Translator George Hunt was a Tlingit/English metis who assisted Franz Boas and others in British Colombia and Alaska, becoming a celebrated ethnologist in his own right. Some mediated cross-cultural issues, such as Sarat Chandra Roy, who served in the colonial judicial system for Bengal. He cultivated an appreciation for tribal society, published widely and, in 1921, established Man in India, the country’s first journal of anthropology. These resident anthropologists had the benefit of already being in the field and knowing the local languages and local societies. What they usually lacked were connections to bases of power, a common problem of centre/periphery scholarship.40 Culture came to be understood as a facilitator of natural selection, illustrating links between social and biological adaptation.41 Anthropology also developed an awareness of society’s potential for intentional transformation. This understanding resulted in contrary efforts, from the negative eugenics movement and the Holocaust to the positive accomplishments of tribal empowerment and post-conflict capacity-building.42 Anthropologists tend to adopt two general strategies, which can be identified as materialist and mentalist approaches. The materialist approach leans towards tangible, evolutionary models, like cultural ecology, as seen in the work of Lucy Mair, while the mentalist approach engages psychological and symbolic analysis, like structuralism, as employed by Claude Lévi-Strauss.43 Ironically, anthropologists themselves mirrored the social structures studied. The materialist and mentalist approaches can be thought of as anthropology’s two moieties, while their thematic studies can be thought of as clans, and, as in tribal structures, these groups interacted in complex ways. Anthropology thus became a society in its own right. As a result of such professional tribalism, many scholars had been professionally limited in the anthropological discourse. The depreciation of women, people of colour and those from non-elite social orders delayed the study of topics that anthropology would later come to address. This is seen in Zora Neale Hurston’s work on 119

Barry H. Rodrigue

Figure 5.9  A  nthropologist Irawati Karve conducting a field interview, Maharashtra, India c 1960. Photograph courtesy of the Karve Family, Pune, Maharashtra, India. Greyscale conversion of original image.

internal colonialism, Irawati Karve’s emic study of caste and kinship and Arnold Van Gennep’s reassessment of the individual in society (Figure 5.9).44 The flowering of diversity, technology and global contacts by the mid-twentieth century allowed anthropology to expand its repertoire. As Sally Slocum observed: ‘It is our task, as anthropologists, to create a ‘study of the human species’ in spite of, or perhaps because of, or maybe even by means of, our individual biases and unique perspectives’.45 Heritage came to be seen as a product of all humanity, as a ‘commons’. In this tradition, the first edition of the United Nations’ History of Humanity came out in 1966, followed by a network of World Heritage Sites a decade later.46 Curator and historian Neil MacGregor at the British Museum pioneered a new way to envision artefacts as a shared human experience, an expression of ‘global civilization’.47 A recent text on anthropology describes this agenda: In our rapidly changing and increasingly interconnected world, where longstanding cultural boundaries between societies are being erased, new social networks and cultural constructs have emerged, made possible by long-distance mass transportation and communication technologies. To better describe, explain, and understand these complex but fascinating dynamics in a globalizing world, anthropologists today are adjusting their theoretical frameworks and research methods and approaches.48 The older areas of concern to anthropology, such as kinship and tool use, expanded and came to have a continuum with other disciplines, from psychology and engineering to medicine and demographics. Advances in cognitive and neurological science 120

Big history and anthropology

Figure 5.10  G  reen Dragon Bridge, near Nankau Pass, Great Wall of China, photograph by Frances Bode, c. 1929. Bode was a pioneering woman who photographed social settings around the world for textbook use and whose work is held at the Museum of Modern Art in New York City. Courtesy of Penelope Markle. Greyscale conversion of original image.

led to the study of human brains and behaviour. Older dichotomies were elaborated by richer varieties, as when LGBT awareness amplified the study of sexuality and gender, which brought into question other binary categories. Applied anthropology began to resolve problems with new subfields and techniques, from cyberethnography and metagenomics to hyperspectral imaging (Figure 5.10).49 Among the important lessons to be derived from studying social systems is how many of our cultural traditions are of relatively recent origin. Only four sets of parents take us back to a century ago. By such calculations, we are but 100 generations removed from the Egyptian pyramids and the Great Wall of China, 500 from the start of agriculture and 5,000 from the human migration out of Africa. Those 10,000 people are the size of a small town’s population – the May Day Stadium in Pyongyang, North Korea, holds more than ten times that number! Nonetheless, in those few generations, humanity has developed extraordinary social abilities that far surpass other life forms on Earth and lead scholars to wonder about our potential abilities to effect larger changes in the multiverse (Figure 5.11).50 When we look at our existence through the lens of anthropology, our spotlight is by definition on ourselves, Homo sapiens. As in Paul Gauguin’s painting, D’où Venons Nous/Que Sommes Nous/Où Allons Nous – we want to know where we came from, 121

Barry H. Rodrigue

Figure 5.11  P  aul Gauguin, Where Do We Come From? What Are We? Where Are We Going? D’où Venons Nous/Que Sommes Nous/Où Allons Nous, Tahiti, 1897. Greyscale conversion of original image.

what is happening now and where we are going. This hominid vision has been critiqued, and change is being suggested to expand anthropology to even wider realms. Just as humans grew out of a focus on their own kin and tribe to global humanity, so are we now in the process of expanding into a much more generous vision of existence. Indeed, a number of anthropologists are also big historians and have been engaged in seeking wider understanding of this more universal outlook.51

Universal humanity and big history As anthropology had begun to develop in the early nineteenth century, a more rigorous form of universal history also had started to come together, as in Wilhelm von Humboldt’s model of integrated education and his brother Alexander von ­Humboldt’s five-volume study of existence, Kosmos (1845–1862). Anthropologist ­Johann Friedrich Blumenbach was among Alexander’s professors at the University of Göttingen, where their School of History sought to develop a modern and scientific universal history. Their goal was to unify knowledge and deploy it so the individual, society and nature could more harmoniously coexist, if not prosper.52 Ironically, this synthesis of knowledge took place just as a movement towards academic specialization emerged. Those subjects that had been united under the broad rubric of philosophy or arts split into physics, history, sociology and the proliferation of disciplines and university departments that we see today.53 This new academic infrastructure slowed efforts to generate a universal history, but the goal to holistically understand existence never died away. The ongoing work to assemble a meta-narrative of existence included works like geographer Alfred Russel Wallace’s Man’s Place in the Universe (1903), engineer Hiram Maxim’s Life’s Place in the Cosmos (1933) and ecologist Imanishi Kinji’s The World of Living Things (1941).54 A vast aggregation of new data then came out of the World War and Cold War eras, requiring larger frames of reference. This resulted in intensified cross-disciplinary studies, as seen in bio-chemistry, electrical engineering and the expansion, earlier described, for anthropology. By the 1970s, this beau ideal was beginning to be formulated as cosmic evolution and other rubrics. 122

Big history and anthropology

It was a global conjuncture that occurred in multiple disciplines, regions, languages and societies.55 Soviet astrophysicist Joseph Shklovsky wrote Universe, Life, Intelligence in 1962, which was expanded with U.S. astrophysicist Carl Sagan four years later. Other works followed, including U.S. bio-geologist Preston Cloud’s Cosmos, Earth and Man (1978), ­Austrian physicist Erich Jantsch’s The Self-Organizing Universe (1980), Colombian mathematician Antonio Vélez’ Humanity: Inheritance and Conduct (1986), U.S. biologist Lynn ­Margulis’ Microcosmos (1986) and Chinese rocket-scientist Qian Xuesen’s complexity studies (1991) on what his team called 开放的复杂巨系统 [Open Complex Giant System].56 Social scientists also joined these new directions, as when economic historian ­Andre Gunder Frank and sociologist Immanuel Wallerstein described global networks outside of Cold War models as a one-world system. Geographer Georges Nicolas saw a need for humanity to bridge the widening chasm between meaning and science, drawing inspiration from the French traditions of geo-anthropology, such as those expressed by Claude Levi-Straus, Paul Vidal de la Blache and Antoine Bailly. These efforts then merged with even larger paradigms, as when economist Graeme Snooks amplified his Theory of Global Dynamic Systems to encompass all of Earth’s history.57 This scholarship began to enter classrooms. In 1974, astrophysicists George Field and Eric Chaisson gave a course on what they called cosmic evolution. Other scientists moved in this direction, as when astrophysicist G. Siegfried Kutter produced Universe and Life: Origins and Evolution (1987), based on two decades of research and teaching.58 In 1985, John Mears advocated for a general-education curriculum based on macro-history and, four years later, began teaching such a course, as did Australian historian David Christian, Dutch anthropologist Fred Spier and Russian psychologist Akop Nazaretyan. It was a dynamic process, as academics began to expand beyond the confines of their disciplinary boundaries.59 These scholarly activities reflected a ferment in holistic thinking that also had been taking place in popular culture. Humanity’s search for meaning stretched beyond traditional confines to embrace wider horizons, finding expression in both faith traditions and secular engagement. Philosopher Jiddu Krishnamurti sought to generate an all-encompassing expression of understanding that embraced humanity, nature and the cosmos, as in his Beginnings of Learning (1975). In the tradition of archaeologist and philosopher Teilhard de Chardin, cultural historian Thomas Berry developed a ‘new story’ that integrated a global narrative of humanity and nature in works like The Dream of the Earth (1988). Both Krishnamurti and Berry left behind active organizations that developed educational programmes, multimedia productions and a legacy that converged with the new science and scholarship in a global articulation of holistic thinking.60 Such overviews entered community life and social organizing. The modern environmental movement began with an understanding of the interconnectedness of life with the world around it. This ‘deep ecology’ appeared in the work of biologist ­Rachel Carson, whose study, Silent Spring (1962), led to a revisioning of humanity’s interaction with nature. Her book is acknowledged as second only to Charles ­Darwin’s The Origin of Species for having changed scientific thinking. Likewise, in 1983, when the United States announced its Strategic Defense Initiative and an orbiting weapons system – the ‘Star Wars’ Program – Osamu Nakanishi, an international relations professor and dean at Soka University in Tokyo, founded the Institute for Global and Cosmic Peace to advocate for cooperation in space.Their work led to 123

Barry H. Rodrigue

publications of what they called universal studies, which incorporates big history, and then to the first university courses on big history in Japan.61 In the early twenty-first century, a variety of organizations came together around this general trend, including the International Big History Association (2010), the ­Eurasian Centre for Megahistory and Systems Forecasting (2011), the Big History ­Project (2011), the Deep Time Journey Network (2014), the Asian Big History Association (2014), the Indian Association for Big History (2016) and the African Big History Association (2017). Each of these groups adopted particular themes and strategies of interest. For example, the Eurasian Centre has a special focus on cliodynamics, while the Asian Association has an interest in social action and the search for meaning. It is not my intent to recapitulate the big history/universal studies movement here, as it already has been well documented elsewhere.62 However, it is very important to emphasize three central points: (1) Universal thinking and disciplinary thinking worked together to produce the big history paradigm that we see today. Universal models provided a framework, while the disciplines provided depth. The interplay of these two tendencies led to the development of cosmic evolution, big history and universal studies. (2) This transdisciplinary study of existence materialized as an independent human invention all around the world, occurring to many people from many backgrounds at the same time. Such a global conjuncture serves as an important reminder of the common humanity of our world, a central theme of modern anthropology. (3) As a consolidation of knowledge, big history does not replace but networks disciplines into larger, more holistic understandings of how things work together, as a form of super-interdisciplinarity. Nor is this process restricted to just big history. It can be seen at work in more commonplace areas, as in the recent formation of the International Science Council (2017), when over 40 social and physical science organizations merged, including the U.S. ­National Academy of Sciences and the National Research Foundation of South ­Africa.63 In this way, a continuum of studies evolves into a holistic network of knowledge. In a big history model, disciplines continue their research and teaching in a usual manner, but they do so in a broader, self-conscious context.This is seen in global historian Craig Benjamin’s ‘little big history’ of Jericho, one of the world’s oldest cities. As he summarizes it: ‘The history of Jericho is a 14,000-year-long reminder that the big story of humanity can only really be understood if it is embedded deeply into the natural context in which it has played out, for the environment is truly the great physical stage upon which our human drama continues to unfold’.64 This transdisciplinary approach also engages with essential community outreach, as when palaeobiologist Nigel Hughes paused in his study of trilobite fossils in the Himalayas to compose a story about a village girl and her quest to find a natural explanation for gatchpathor (petrified wood) that is common throughout much of the region. Monisha and the Stone Forest introduces Earth history to children and was produced by the Geological Society of India in Bangla and English.65 It not only fuses issues of science and society but helps expand this new continuum of big history in a form of applied anthropology that helps lay people understand this quest for wider understandings (Figure 5.12). 124

Figure 5.12  C  over of the Bangla edition of the book by Nigel Hughes and Rati Basu, ­Monishar Pathorer Bon [Monisha and the Stone Forest], Kolkata: Monfakira Press, 2012. Greyscale conversion of original image.

Figure 5.13  A  nthropology and big history students from Symbiosis International University doing fieldwork at a megalith in Lohegaon, Maharashtra, India, on 10 May 2018. Left to right: Tanvi Shah, Anupoma Bandyopadhyay, Arshiya Dutt, Najiba Yasmin, Abhiman Paul, Sakshi Saldanha and Barry Rodrigue. Photograph by Tanvi Shah. Greyscale conversion of original image.

Barry H. Rodrigue

An academic example of this new continuum can be seen at my own institution of Symbiosis International University. Our Symbiosis School for Liberal Arts (SSLA) was founded in Pune, Maharashtra in 2011, the first liberal arts programme in India, one in which interdisciplinarity was central to its mission. Six years later, SSLA adopted big history as a form of super-interdisciplinarity, since it was seen to be a natural extension of the university’s philosophy of वसुधवै कुटमु ब् कम् (vasudhaiva kutumbakam, ‘the world is one family’).The Indian Association for Big History is based at SSLA, while the first course of study and the first academic conference on big history in South Asia were held there in 2018. In this way, Sarat Chandra Roy’s 1938 call for a wider view of anthropology, one that would lead to a sense of ‘universal humanity’ was fulfilled (Figure 5.13).66

Anthropology and big history Anthropology’s understanding of the flexibility of society leads to an awareness that humans have the ability to not only adapt to a wide range of conditions but to intentionally modify their surroundings and themselves. A challenge however lies in our ‘natural’ limits and our abilities to overcome them. These limitations include brain function and sensory range, from wavelength perception by the eye to frequency reception by the ear. Despite the constraints, many insights and strategies to enhance the human potential for connectivity and problem-solving occur, such as intelligence amplification (IA) and artificial intelligence (AI).67 Traditional societies and global civilization coexist and evolve. Because civilization is an expression of modernity, it has less ‘baggage’ in the form of heritage to slow its transformation (except when heritage is invoked by demagogues). Individuals possess multiple identities, including those bounded by language, religion, gender, tribe and many other traditional forms of cultural selfhood. The challenge we face today is how we can empower the least developed of these identities – our global and cosmic identities – as a new form of civilization.68 In order to keep a positive trajectory of human self-awareness moving forward, we first have to survive. Historical psychologist Akop Nazaretyan has documented how humans have managed to reduce violence over the last million and more years, despite the development of ever more lethal technologies. He codified this phenomenon as the Law of   Techno-Humanitarian Balance, in which human populations – those that managed to survive – advanced strategies to constrain the use of harmful acts. He does not limit his study to just intentional weaponry but also includes destructive mechanisms like chemical contamination, reduction in biodiversity or other behaviours that negatively impact humanity.69 In this way, pollution and racism pose as much of an ultimate threat to human survival as nuclear proliferation. A positive example of this axiom in action is the Spacewatch Program, which monitors near-Earth asteroids. In 1991, when international tensions were high, the program alerted the world of a small asteroid coming towards Earth. The concern was not for an impact, which would have been negligible, but for the fear that, if it burst in Earth’s atmosphere, it could be mistaken for a nuclear explosion and provoke missile strikes during the First Gulf War and the collapse of the Soviet Union.70 ­Nazaretyan therefore sees such efforts as important ongoing components of our social evolution and ability to survive. 126

Big history and anthropology

Anthropologists and big historians see humanity, life and Earth’s stability as related issues. This awareness by itself can help mitigate disputes, reduce tensions and find alternatives. International relations scholar Osamu Nakanishi provides an example of the positive effects of the simple transition from small thinking to big thinking. During Japan’s Age of Warring States, many battles were fought by warlords over the fertile river plain of Kawanakajima in Honshu in the sixteenth century. Conflict ceased with establishment of broad central control during the Edo Period. As ­Nakanishi summarizes the transition: ‘No one fights over Kawanakajima today’.71 With similar global awareness, we could also hope to reduce conflict worldwide. Anthropology informs the larger views of big history, and big history impacts anthropology. Since the progression of time converts the present and future into the past, anthropology can be seen as a form of big history, an incipient reflection of our on-going beginnings. This expansion of the conceptual framework of anthropology makes the resulting macro-anthropology all but synonymous with big history, in a trend we might describe as the study of change, of how all things evolve and are networked in the universe. An example of the cross-fertilization of ideas appears in the use of energy. In 1943, anthropologist Leslie White identified energy as central to studies of culture. His Law of Cultural Evolution explains how culture advances when energy is harnessed and its use increases.72 Everything in the universe may be described in terms of energy. Galaxies, stars, molecules, and atoms may be regarded as organizations of energy. Living organisms may be looked upon as engines which operate by means of energy derived directly or indirectly from the sun. The civilizations, or cultures of mankind, also, may be regarded as a form or organisation of energy. … Culture is a kind of behavior. And behavior, whether of man, mule, plant, comet or molecule, may be treated as a manifestation of energy. Thus we see, on all levels of reality, that phenomena lend themselves to description and interpretation in terms of energy.73 One of the strategies suggested by the field of cosmic evolution, and adopted by some big historians, is a measure of complexity called energy-rate density.The metric formula for this complexity ratio is: Φm = energy / time / mass It measures the amount of energy in a unit-of-time passing through a given mass, such as the calories active in a gram of carbon per second. Every object can be assigned a number based on this algorithm, while the number can be totalled for collective objects. This number is used as a measure of complexity (Φm) – the higher the number, the more complex an object is considered. In this way, a butterfly is more complex than a galaxy, as the energy passing through its small mass yields a much higher number than the energy passing through the huge mass of a galaxy.The most complex thing in the known universe (as derived from this ratio) is collective human society.74 Besides a general awareness, the complexity ratio might also be of practical use for more basic comparisons. 127

Barry H. Rodrigue

Table 5.1  R  endition of Eric Chaisson’s table of average energy rate densities Average energy rate densities System • • • • • •

Human society Animals Plants Earth Sun Milky way

Age (Gya)

Φm[erg/s/g]

0 0.5 3 4 5 12

5,00,000 40,000 900 75 2 0.5

Cosmic Evolution: Rise of Complexity in Nature, Cambridge: Harvard University Press, 2001, p. 139; idem, ‘Energy Rate Density as a Complexity Metric and Evolutionary Driver’, Complexity, vol. 16, no. 3, January 2011, p. 28.

Anthropology, like many studies, has the need to discern and assess patterns.75 How do we compare, say, the excavated remains from a medieval farming village in Tunisia with a present-day hunting ground in the Amazon Basin? The usual approach is an assessment of factors that vary with the interests of the scholars engaged in the work. This can be a very subjective. The complexity ratio could provide a metric with which to gauge objects and assemblages of objects from very different cultures, times and locations.76 It certainly would not be a formulation of inherent cultural value, but it could be of use in comparative studies, as a correlation that might lead to further studies (Table 5.1). Other innovative strategies are being similarly adopted for studies of human culture. Isaac Newton’s Law of Universal Gravitation (1686) calculated the attraction between objects (based on mass and distance). It was reformulated to assess economic networks in the late nineteenth century, but it only gained traction as a more pervasive theory – the ‘structural gravity model’ – a century later. Historians and archaeologists recently applied this analysis to set up a predictive model for discovering lost settlements and reconstructing ancient trade networks from cuneiform records of 4,000 years ago.77 This kind of interdisciplinary thinking has the potential to greatly enrich our understanding of our past. Moreover, as our understanding of the universe increases, so does our sense of humanity’s engagement with the cosmos. For example, expansion into outer space has led to experiments on the human ability to travel long distances in low-gravity, and in isolated, confined settings, from space ships to orbital stations. It also raises questions about non-Earth colonies and terraforming of habitats.78 These are questions that can be beneficially addressed by anthropological insights.

The future It used to be thought that self-reflection was just an ability of modern humans, one dating to a threshold of about 50,000 years ago. Recent research however shows that a capacity for problem-solving is not only available to others of our ancestral lineage 128

Big history and anthropology

and our primate cousins but to other vertebrates and invertebrates as well. Nonetheless, the cognitive ability with which Homo sapiens network with each other and with their environment is of monumental distinction from other intelligences on Earth.79 This ability opens the door on the potential for activities like extra-terrestrial communication. Dale Russell is a palaeontologist who worked with NASA on questions of intelligence. He speculated about how dinosaurs might have evolved into a species comparable to modern humans and how such a thought-experiment might assist in communication with similar extra-terrestrial intelligences.80 This kind of reasoning is important because humanity has engaged in scientific efforts to reach into the ­meta-galaxy for decades. In 1974, Cornell University’s National Astronomy and Ionosphere Center and the U.S. National Science Foundation broadcast a message from the Arecibo radio telescope in Puerto Rico that provided information about Earth and its solar system, elements of DNA and their configuration and the human shape and society. The message was transmitted in the direction of the Great Globular Star Cluster in the constellation of Hercules, 22,000 light years from Earth. Likewise, between 1972 and 1977, NASA’s Pioneer and Voyager space probes carried data about humanity. Pioneer 10 and 11 conveyed plaques with figures of humans and symbols for Earth and the solar system embellished on them. On Voyager 1 and 2, analog recordings on a 12-inch (30-centimeter) gold-plated copper disk contained sounds and images of the diversity of life and culture on Earth. These ‘Golden Records’ included greetings in 55 languages, music and images of humanity in a variety of its cultural expressions. Voyager 1 has now passed outside our solar system and is travelling through interstellar space towards Gliese 445, a star in the constellation of Camelopardalis, 18 light years from Earth, which it will reach in 40,000 years (Figure 5.14).81

Figure 5.14  N  ASA’s ‘Plaque of Humanity’, etching made for the Pioneer 10 space probe in 1972. Designed by Carl Sagan and Frank Drake; drawn by Linda Salzman ­Sagan. Courtesy of Eric Chaisson and NASA. Greyscale conversion of original image. 129

Barry H. Rodrigue

A variety of languages have been designed to facilitate potential communication between humans and otherworldly intelligences. One is Lincos (Lingua Cosmica), a mathematically based programme from which intercultural communication between Earth and extra-terrestrial societies could be developed. Produced in 1960, Lincos was configured by two Canadian astrophysicists to transmit a message to nearby stars from the Yevpatoria RT-70 radio telescope in Ukraine in 1999 and 2003, the first since the Arecibo broadcast. One of its targets was 16 Cyg A, part of a triple-star ­system in the constellation Cygnus, 69 light years from Earth.82 Besides these efforts in the field of astrolinguistics, other initiatives to better understand our place in the cosmos are underway. The Kepler space observatory, launched by NASA in 2009, has a dedicated mission to seek exoplanets with similarities to Earth and a potential for similar life forms. Ellen Stofan, NASA’s chief scientist, anticipates evidence of extra-terrestrial life will be found by 2045. And, in 2016, the Hubble space telescope discovered that the universe has ten times the number of galaxies than previously thought, which vastly amplifies the opportunities to find life beyond Earth.83 The Search for Extra-Terrestrial Intelligence (SETI) is essentially an anthropological study. Such contact will likely not be with surviving civilizations in our galaxy, let alone in the universe/multiverse. Our existence as a species encompasses 300,000 out of 13,800,000,000 years – only 0.002% of all known existence. So it is likely that the communications and remains that we discover will be of a civilization that has vanished, through extinction or through evolution.84 This opens the door on another potential use of anthropology in the future, as astro-archaeology. If contact with extra-terrestrial life should occur, anthropology’s experience with a wide range of topics would inform these encounters.This would include not just linguistics, cognitive behaviour and cultural analysis but also the discourse surrounding colonialism and issues of inclusion/exclusion. A parallel question is that if all sentient species eventually die out, or transcend to a different level, how do we Homo sapiens leave evidence of our knowledge for other sentient creatures in the meta-galaxy to discover and perhaps benefit from its use? This is essentially a question of exo-­ humanitarian values.85 Whatever the result of such efforts, big historians Alexander Panov and Joseph Voros note that in either event – finding extra-terrestrial intelligence or not – the result is significant: We are one of many or we are unique. Panov further speculates that natural science has possibly begun to run its course and that a new direction, strategy and inspiration for humanity are needed. He does not identify what that new configuration might be, but he sees it as important for humanity to consider.86 From the experiences of anthropology, it could be that a new strategy might be not one thing but a constellation of them. Indeed, Alexander Panov and Akop Nazaretyan highlight the role of superfluous diversity in human societies as an important way to overcome crisis. Nazaretyan posits this as the Rule of Redundant Variety, in which the myriad ways that societies around the world accomplish the same purpose allows later societies, including global civilization, to choose from divers and bountiful options.87 The members of the Eurasian Centre for Big History and System Forecasting in the Russian Academy of Sciences consider the scale-invariant sequence that results in

130

Big history and anthropology

the Snooks-Panov Vertical to be a curious anomaly and contemplate its significance. A version of this formula is: tn = t* – T / αn. The coefficient α > 1 is a compression ratio of duration of every subsequent phase of evolution in comparison to the previous one;T is the duration of the entire described period of time; n is the number of a phase transition; t* is the limit of the sequence of phase-transition moments {tn}. Independently developed by nuclear physicist Alexander Panov in Russia and systems theorist Graeme Snooks in Australia at the turn of the twenty-first century, the calculations highlight how the span of time between major events in Earth and human history has become increasingly compressed. For example, the time between the Industrial Revolution and the Information Revolution is shorter than the span between the Upper Palaeolithic and the Neolithic. This formulation then becomes ‘vertical’ – reaches 0 – at about the year 2026.The question is: Does this predictive ‘singularity’ have significance?88 Regardless of the formula, one just has to look out a window to see that the world is in crisis. In the last decades, we have become more aware that entire species of life are rapidly vanishing, along with fresh water supplies. Pollution makes many parts of the land and seas uninhabitable. Non-renewable resources are being exhausted. Global warming is impacting the entire planet, from the melting of the world’s ice sheets and permafrost to the related rise in sea levels and changing storm patterns. Local agriculture and business are destroyed by competition from multinational industry, resulting in the concentration of people in urban areas, as more and more residents are dropped to the lowest rungs of society. This human degradation of Earth and its life is now referred to as the Anthropocene epoch. It is estimated that the original foraging lifestyle of our ancestors who lived prior to 10,000 years ago had the capacity to provide sustenance for 15 million people. The adoption of simple agriculture then allowed the carrying capacity to rise to 750 million.The industrial production that began 300 years ago permitted twice that number to coexist. In the last century, our population has risen to 7 billion and is expected to reach more than 10 billion by the end of the century. This extraordinary growth has been possible because of a variety of intersecting factors, including adoption of petroleum as an energy source; the use of hybrid crops, pesticides and artificial fertilizers; medical advances and urbanization.These numbers are far beyond the natural carrying capacity of the Earth and are possible only because of a greatly stratified society that leaves a majority with hunger and few of life’s amenities, a situation that is worsening.89 Many of the social problems we see around us are not an erosion of values, they are the result of our global system not being able to cope with the vast numbers of people and the scarce resources.There is no going back to an archaic, primeval stage of society for the majority of people, indeed such quaint notions are part of the problem. If such an attempt were made by urban humanity to return to a simpler, primeval stage of production, billions of people would die.That is not to say that we cannot learn and adapt concepts from our past or that traditionalist societies cannot be encouraged, but we as a whole can only go forward with the configuration of new strategies for survival.90

131

Barry H. Rodrigue

And so, we are where we began, sitting around fires in our Palaeolithic caves millennia ago, asking questions of our existence. We have sketched out the pattern of our origins and more clearly see how we fit in the world, filling in the details and debating where to go with this new information. But our context has shifted. This time, we are seen as members of a global tribe contemplating a vast cosmos with a larger tool kit than that of our Acheulean ancestors. In this spirit, author Vandana Singh encourages us to ‘step out of the claustrophobia of the exclusively human and discover joy, terror, wonder, and meaning, in the greater universe’.91

Acknowledgements I would like to express my appreciation for insights from Penelope Markle, Nobuo Tsujimura, Anita Patankar, Afshan Majid, Priyadarshini Karve, Sakshi Saldanha, Shweta Sinha Deshpande, Harald Prins and Enid Still.They bear no responsibility for my extrapolations but did provide important ideas and inspiration.

Notes 1 The multiverse is commonly seen as a larger entity of which our universe is a component. Although there are many uses of the term, some quite divergent, a big history perspective by astronomer Tom Gehrels is the one that is in mind for this essay. The concept of a multiverse also serves as a metaphor for the variety of ways of seeing and interpreting existence. Tom Gehrels, ‘The Chandra Multiverse’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 45–70. 2 Amin, Osama, ‘The Newly Discovered Tablet V of the Epic of Gilgamesh’, Ancient History Et Cetera, 24 September 2015, , accessed 25 N ­ ovember 2017. 3 Manuel Vaquero et al., ‘Temporal Nature and Recycling of Upper Paleolithic Artifacts: The Burned Tools from the Molí del Salt Site (Vimbodí i Poblet, northeastern Spain)’, Journal of Archaeological Science, vol. 39, no. 8, August 2012, pp. 2785–2796. Quaternary International, vol. 361, The Origins of Recycling: A Paleolithic Perspective, March 2015. ­Barbara Gilli, ‘The Past in the Present: The Reuse of Ancient Material in the 12th Dynasty’, Aegyptus, vol. 89, nos. 1–2, 2009, pp. 89–110. Wilhelm Nicolaisen, ‘“Distorted Function” in Material Aspects of Culture’, Folklore Forum, vol. 12, nos. 2–3, pp. 223–235. 4 Zhenoao Xu, David Pankenier and Yautias Jiang, East-Asian Archaeoastronomy: Historical Records of Astronomical Observations of China, Japan and Korea, Boca Raton: CRC Press, 2000, pp. 13–24. The development of writing led to partial preservation of earlier understandings of the world. Tao Qi, ‘Echoes of the Palaeolithic: A Research Note on the Great Floods and the Origins of Chinese Civilization’, in From Big Bang to G ­ alactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 84–94. 5 For big history considerations of the Axial Age, see the following. Dmitri Bondarenko and Ken Baskin, ‘Big History, Complexity Theory, and Life in a Non-Linear World,’ op. cit., pp. 183–196. Lazar Puhalo, ‘The Rise of Personhood: Development of Social Justice and Natural Rights in the Axial Eras’, op. cit., Volume II, Education and

132

Big history and anthropology

Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 118–123. 6 Many representations of early peoples’ understanding of themselves and their surrounding world are still undecipherable, such as the petroglyphs and Picenean inscriptions found around Mt. Conero in Italy. Giuseppe Barbone et al., ‘La roccia con incisioni del Monte Cònero: relazione preliminare’, Preistoria Alpina, vol. 46, no. 2, 2012, pp. 93–98. Rodolfo Coccioni et al., Carta Geologica con Itinerari Escursionistici: Firenze, Italia, Parco Regionale del Conero, (1:20,000) Florence: SELCA, 1993. Gaia Pignocchi et al., I Petroglifi Preistorici del Monte Cònero: Progetto di Studio e Valorizzazione, Ancona: submitted to the administration of Conero Park, 8 May 2007. 7 United Kingdom, Department for Culture, Media and Sport, Portable Antiquities Scheme, Tisbury Hoard, WILT-E8DA70, , accessed 23 April 2017. Richard Gartner, Metadata: Shaping Knowledge from Antiquity to the Semantic Web, Cham: Springer, 2016, pp. 15–16. 8 The Bisotun inscription is a UNESCO World Heritage Site. United Nations Educational, Scientific and Cultural Organization, World Heritage Centre, List: Bisotun, , accessed 23 April 2017. 9 Stewart Flory, ‘Who Read Herodotus’ Histories?’ The American Journal of Philology, vol. 101, no. 1, Spring 1980, pp. 12–28. 10 Although known in the ancient Mediterranean world, Lucretius’ writing was lost and only rediscovered in an archive in 1417, providing an impetus to Renaissance thought. His ages of humankind was re-demarcated by antiquarian Christian Thomsen in 1834 as Stone Age, Bronze Age and Iron Age. Jacques Lezra and Liza Blake, eds., Lucretius and Modernity: Epicurean Encounters Across Time and Disciplines, New York: Palgrave ­Macmillan, 2016. The tradition of travellers’ narratives intensified during the Renaissance, as with those of Benjamin of Tudela in the twelfth century, Giovanni da Pian del Carpine in the thirteenth century and Muhammad Ibn Baṭūṭah in the fourteenth century. 11 Adina Hoffman, Sacred Trash: The Lost and Found World of the Cairo Geniza, New York: Schocken, 2011. Amitav Ghosh, In an Antique Land, New Delhi: Ravi Dayal Publishers, 1992. Stephen Haw, Marco Polo’s China: A Venetian in the Realm of Khubilai Khan, ­L ondon: Routledge, 2006. 12 John Rowe, ‘The Renaissance Foundations of Anthropology’, American Anthropologist, vol. 67, no. 1, February 1965, pp. 1–20. Tim Murray, Milestones in Archaeology: A Chronological Encyclopedia, Santa Barbara: ABC-Clio, 2007, pp. 44–45. 13 Caroline Roullier et al., ‘Historical Collections Reveal Patterns of Diffusion of Sweet Potato in Oceania Obscured by Modern Plant Movements and Recombination’, Proceedings of the National Academy of Sciences of the United States of America, vol. 110, no. 6, 5 February 2013, pp. 2205–2210. Kirsten Seaver, The Frozen Echo: Greenland and the Exploration of North America, ca. A.D. 1000–1500, Stanford: Stanford University Press, 1996. Christopher Fee, Mythology in the Middle Ages: Heroic Tales of Monsters, Magic, and Might, Santa Barbara: Praeger, 2011. 14 Hans Georg Bandi, Eskimo Prehistory, Fairbanks: University of Alaska Press, 1964, 1972, pp. 164–166. Heather Pringle. ‘Vikings and Native Americans’, National Geographic, November 2012, pp. 80–89. 15 This reassessment of global civilization parallels that of later interpretations in anthropology (see the text linked to endnotes 46 and 47 on professionalization in respect to UNESCO and the British Museum). 16 Firuz Shah Tughlaq’s collection of antiquities was in a tradition that had existed earlier in the Mamluk dynasty and would continue in the Mughal empire. He was also

133

Barry H. Rodrigue

17

18

19

20

21

22

23

a celebrated collector of other antiquities, having ancient documents translated into Arabic. Sayed Ali Nadeem Rezavi, ‘Antiquarian Interests in Medieval India: The Relocation of Ashokan Pillars by Firuzshah Tughluq’, in Proceedings of the Indian History Congress 70th Session, 2010, pp. 994–1010. I would like to thank historian Afshan Majid for sharing this example of South Asian antiquarianism. Afshan Majid, Badauni and Mughal Society and Culture as Reflected in his Works, Ph.D. Thesis, Department of History, Centre of Advanced Study, Aligarh Muslim University, 2015, pp. 84–103. The first edition of Blumenbach’s book was a privately produced version of his dissertation that he published in 1776. It went through many editions and appeared in many languages. For example, Johann Friedrich Blumenbach, On the Natural Variety of Humankind, Göttingen: Vandenhoek & Ruprecht, 1781 [Latin publication]. Peter Miller and François Louis, Antiquarianism and Intellectual Life in Europe and China, 1500–1800: A Comparative Consideration of the Fascination with Antiquity in European and Chinese Intellectual History, Ann Arbor: University of Michigan Press, 2012. James Urry, Before Social Anthropology: Essays on the History of British Anthropology, ­London: Routledge, 1993. Christopher Shea, ‘Alas, Poor William Shakespeare: Where Does His Skull Rest? New York Times, 24 March 2016. Walter Scott, The Antiquary, Edinburgh: James Ballantyne & Company, 1816. Miura’s work was hampered by foreign and domestic policies of the Tokugawa shogunate and so became lesser known than those of other scholars. Gino Piovesana, ‘Miura Baien, 1723–1789, and His Dialectic and Political Ideas’, Monumenta Nipponica, vol. 20, nos. 3–4, 1965, pp. 389–421. My appreciation goes to Nobuo Tsujimura for bringing Miura to my attention. Btsan po no mon han (1789–1839) was a lama of the Drepung monastery in Lhasa. In about 1814, he travelled to Beijing, where he spent the rest of his life. His Detailed Description of the World was first published in Mongolia in 1830, but it had been a work in progress for more than a decade. Lobsang Yongdan, ‘Tibet Charts the World: The Btsan po No mon han’s Detailed Description of the World, An Early Major Scientific Work in Tibet’, Mapping the Modern in Tibet, ed. Gary Tuttle, 2011, pp 73–134. University of Cambridge, Research, ‘The Tibetan Lama who Wrote a World Geography’, 14 June 2014, , accessed 12 June 2018. Turrell Wylie, ‘Dating the Tibetan Geography “Dzam Gling Rgyas Bshad” through its Description of the Western Hemisphere, Central Asiatic Journal, vol. 4, no. 4, 1959, pp. 300–311. Lobsang Palden Chophel was the Sengchen Lama and Chief Minister in Shigatse, Tibet. A modernizer and internationalist who took an active interest in science, languages and technology, his execution in 1887 was the result of his unwitting friendship with British spy Sarat Chandra Das, from whom he learned of advances in Western innovation. Sarat Chandra Das, A Journey to Lhasa and Central Tibet, London: John Murray, 1902, pp. 62, 68, 108–110. Alex McKay, ‘The Drowning of Lama Sengchen Kyabying: A Preliminary Enquiry from British Sources’, Tibet, Past and Present, eds. Henk Blezer and Abel Zadoks, Leiden: Brill, 2002, pp. 263–279. There had been a long tradition of documenting cultures to aid political and military goals. Programs dealing with intercultural affairs began in order to facilitate discovery and conquest of territories, then to acquire information for governing colonies and managing trade. China especially had a long tradition of such efforts. Laura Hostetler, ‘Qing Connections to the Early Modern World: Ethnography and Cartography in ­Eighteenth-Century China’, Modern Asian Studies, vol. 34, no. 3, July 2000, pp. 623–662. 134

Big history and anthropology

24 Lalita Prasad Vidyarthi, Rise of Anthropology in India: A Social Science Orientation, Volume 1, Delhi: Concept Publishing Company, 1976, pp. 214–215. Shakunt Pandey, ‘The Indian Museum Completes 200 Years’, Science Reporter, vol. 51, no. 10, October 2014, pp. 38–41. 25 John Merryman, ‘Thinking about the Elgin Marbles’, Michigan Law Review, vol. 83, no. 8, August 1985, pp. 1880–1923. Kate Fitz Gibbon, Who Owns the Past?: Cultural Policy, Cultural Property, and the Law, New Brunswick: Rutgers University Press, 2005, pp. 109–121. George Stocking Jr., Victorian Anthropology, New York: The Free Press, 1987. 26 Charles Darwin, On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life, London: John Murray, 1859. Herbert ­Spencer, First Principles of a New System of Philosophy, London: Williams and Norgate, 1862. Karl Marx, Capital: Critique of Political Economy, vol. 1, Hamburg: Otto Meissner, 1867 [­German publication]. Edward Tylor, Primitive Culture: Research into the Development of Mythology, Philosophy, Religion, Art, and Custom, vols. 1–2, London: John Murray, 1871. Lewis Morgan, Ancient Society; Or, Researches in the Lines of Human Progress from Savagery, Through Barbarism to Civilization, New York: Henry Holt and Company, 1877. Daniel Smail, On Deep History and the Brain, Berkeley: University of California Press, 2008. 27 Edward Tylor, op. cit., Primitive Culture, vol. 1, p. 1. 28 Edward Tylor only circumstantially references Alexander von Humboldt’s series K ­ osmos, ­ lexander but his intent is obvious from the context. Edward Tylor, op. cit., pp. 1–2, 17. A von Humboldt, Kosmos, vols. 1–5, Stuttgart: J.G. Cotta Publishing, 1845–1862. R. Jon McGee and Richard Warms, Anthropology Theory: An Introductory History, ­Lanham: Rowman & Littlefield, 2017, pp. 197–198. 29 An example of such transition is how physician Robert Knox saw a common base in not just human anatomy but between species, which led him to develop an early version of evolution that he called zoological history. He expanded his observations from anatomy to ethnology, which mirrored populist views of the time in privileging European society. Robert Knox, The Races of Men: A Philosophical Enquiry into the Influence of Race over the Destinies of Nations, London: Henry Renshaw, 1850, 1862. Efram Sera-Shriar, The Making of British Anthropology, 1813–1871, London: Routledge, 2013, 2015. Biologist Stephen Gould noted that science is embedded in subjective human culture and so has often failed to consciously address many social problems and its own biases in an objective way. Stephen Jay Gould, The Mismeasure of Man, New York: W.W. Norton, 1981, 1996. 30 Elisabeth Tooker, ‘A Note on Undergraduate Courses in the Latter Part of the ­Nineteenth Century’, Man in the Northeast, vol. 39, Spring 1990, pp. 45–51. As an example of the ease of entry into human studies and its pervasive interest to the general public, it has been and is still common for farmers and other residents of North America to collect and display indigenous arrowheads and other artefacts that they have found while tending their fields or visiting historic sites. This is something I have frequently encountered in my own fieldwork. 31 Arthur Spiess, ‘Wild Maine and the Rusticating Scientist: A History of Anthropological Archaeology in Maine’, pp. 101–129, Man in the Northeast, vol. 30, Fall 1985, pp. 104–106. Christopher Benfey, The Great Wave: Gilded Age Misfits, Japanese Eccentrics, and the Opening of Old Japan, New York: Random House, 2004, pp. 45–73. 32 Edward Morse, ‘Shell Mounds of Omori’, Memoirs of the Science Department, University of Tokio, Japan, Volume I, Part I, Tokyo: University of Tokyo, 1879. Benfy, op cit. Shinji Yamashita, Joseph Bosco and Jeremy Eades, The Making of Anthropology in East and Southeast Asia, New York: Berghahn Books, 2004, pp. 91–92. 33 British Association for the Advancement of Science, Notes and Queries on Anthropology: For the Use of Travellers and Residents in Uncivilized Lands, London: Edward Stanford, 1874, pp. iv, 142. 135

Barry H. Rodrigue

34 Susan Brownell, The 1904 Anthropology Days and Olympic Games: Sport, Race, and ­American Imperialism, Lincoln: University of Nebraska Press, 2008. Rosemarie Bank, ‘“Show Indians”/Showing Indians: Buffalo Bill’s Wild West, the Bureau of Indian Affairs, and American Anthropology’, Journal of Dramatic Theory and Criticism, vol. 26, no. 1, Fall 2011, pp. 149–158. Bruce Trigger, A History of Archaeological Thought, Cambridge: Cambridge University Press, 1990, viz. ‘Archaeology and its Social ­Context’, pp.  ­370–411. Mingxin Liu, ‘A Historical Overview on Anthropology in China’, ­Anthropologist, vol. 5, no. 4, 2003, pp. 217–223, , accessed 13 April 2017. 35 In 1898, it was estimated that only 50 archaeologists were engaged in scientific research in the United States but that almost 5000 amateurs were also at work. Andrew ­Christenson, ‘Who were the Professional North American Archaeologists of 1900? Clues from the Work of Warren K. Moorehead,’ Bulletin of the History of Archaeology, vol. 21, no. 1, May 2011, pp. 6, 13, 17. Jay Bernstein, ‘First Recipients of Anthropological Doctorates in the United States, 1891–1930, American Anthropologist, vol. 104, no. 2, 2002, pp. 551–564. Reba Soffer, ‘The Development of Disciplines in the Modern English University’, The Historical Journal, vol. 31, no. 4, December 1988, pp. 933–946. Alfred Russel Wallace, ‘Peschel’s “Races of Man”’, Nature, vol. 15, 28 December 1876, p. 174. 36 Curtis Hinsley, The Smithsonian and the American Indian: Making a Moral Anthropology in Victorian America, Washington, DC: Smithsonian Institution Press, 1981, 1994. Stiofán Ó Cadhla, Civilizing Ireland: Ordnance Survey 1824–1842: Ethnography, Cartography, Translation, Dublin: Irish Academic Press, 2007. Indra Sengupta, ‘A Conservation Code for the Colony: John Marshall’s Conservation Manual and Monument Preservation ­Between India and Europe’, in ‘Archaeologizing’ Heritage?: Transcultural Entanglements between Local Social Practices and Global Virtual Realities, eds. Michael Falser and Monica Juneja, Berlin: Springer-Verlag, pp. 21–37. 37 Franz Boas’ Anthropology Department institutionalized the four-field system to regain some of the synthesis lost when universities adopted departments and disciplines. ­Elsewhere, anthropology focused on social-cultural studies, with archaeology, linguistics and physical anthropology remaining in other departments. Variations resulted around the world. In my own Department of Anthropology in the School for Liberal Arts at Symbiosis International University, we use the four-field system, while neighbouring universities use the social-cultural approach. As anthropologist ­Elisabeth Tooker has pointed out, much of Boas’ framework for anthropology at Columbia followed the existing trajectory established by amateur anthropologists earlier in the century, including the four-field system. Tooker, ‘A Note on Undergraduate Courses in the Latter Part of the Nineteenth Century’. Wilhelm Wundt, Völkerpsychologie: Eine Untersuchung der Entwicklungsgesetze von Sprache, Mythus und Sitte; Mythus und Religion, vol. 2, part 1, Leipzig: Wilhelm Engelmann, 1905, p. iii [German publication]. 38 Ruth Benedict, Patterns of Culture, Boston: Houghton Mifflin Company, 1934. An early student in the ethnography of body movement, including walking, was sociologist Marcel Mauss. Marcel Mauss, ‘Les techniques du corps’, Journal de Psychologie, vol. 32, nos. 3–4, 15 March 1936, pp. 271–293. 39 Franz Boas, ‘Ethnology of the Kwakiutl: Based on Data Collected by George Hunt’, 35th Annual Report of the Bureau of American Ethnology, Parts 1 and 2, Washington, DC: Bureau of American Ethnology, 1921. Bronislaw Malinowski, Argonauts of the Western Pacific: An Account of Native Enterprise and Adventure in the Archipelagoes of Melanesian New Guinea, London: G. Routledge & Sons, 1922. Marcel Mauss, ‘Essai sur le don: Forme et 136

Big history and anthropology

40

41

42

43

4 4

45

raison de l’échange dans les sociétés archaïques’, L’Année Sociologique, vol. 2, no. 1, 1925, pp. 30–126. Theodore Timreck, dir., Franz Boas, 1858–1942, Public Broadcasting Associates, 1980. Sangeeta Dasgupta, ‘Recasting the Oraons and the “Tribe”’, in Anthropology in the East: Founders of Indian Sociology and Anthropology, eds. Patricia Uberoi, Nadini Sundar and Satish Deshpande, Ranikhet: Permanent Black, 2007, pp. 132–171. Sarat Chandra Roy, ‘An Indian Outlook on Anthropology’, Man ( Journal of the Royal Anthropological Institute), vol. 38, no. 9, September 1938, pp. 146–150. Mondher Kilani, ‘Is a Peripheral Anthropology Possible? The Issue of Universalism’, Kroeber Anthropological Society Papers, vol. 101, no. 1, 2012, pp. 98–105. In 1896, psychologist James Mark Baldwin postulated connections between biological and cultural evolution. Others continued in this investigation. James Mark Baldwin, ‘A New Factor in Evolution’, The American Naturalist, vol. 30, no. 354, June 1896, pp. 441–451. Bruce Weber and David Depew, Evolution and Learning: The Baldwin Effect Reconsidered, Cambridge: MIT Press, 2003. Leonid Grinin, Andrey Korotayev and Alexander ­Markov, ‘Biological and Social Phases of Big History: Evolutionary Principals and Mechanisms’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry R ­ odrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 141–174. A side effect of anthropology’s participation in projects of social engineering has been a temporary decline of the field in some areas, as in post-Nazi Germany and post-British India. In other locales, partisan visions of the field have struggled, as in the United States and Russia during and after the Cold War. David Price, ‘Lessons from Second World War Anthropology: Peripheral, Persuasive and Ignored Contributions’, Anthropology Today, vol. 18, no. 3, June 2002, pp. 14–20. Shweta Sinha Deshpande, ‘India: Evolving a Big History’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, Barry Rodrigue, Leonid Grinin, Andrey Korotayev (editors), Delhi: Primus Publishing, 2016 pp. 255–264. JeanYves Durand, ‘“Traditional Culture” and “Folk Knowledge”: Wither the Dialogue between Western and Post-Soviet Anthropology?’, Current Anthropology, vol. 16, no. 2, April 1995, pp. 326–330. Roy D’Andrade, ‘Moral Models in Anthropology’, Current Anthropology, vol. 16, no. 3, June 1995, pp. 399–408. William Haviland and his colleagues elegantly defined this dichotomy in anthropology as materialist and mentalist approaches. William Haviland et al., Cultural Anthropology: The Human Challenge, Boston: Cengage Learning, 2017, pp. 67–68. Lucy Mair, Primtive Government, Harmondsworth: Penguin Books, 1962. Claude Lévi-Strauss, The Savage Mind, Paris: Librairie Plon, 1962 [French publication]. Irma McClaurin, ‘Zora Neale Hurston: Enigma, Heterodox, and Progenitor of Black Studies’, Fire!!!: The Multimedia Journal of Black Studies, vol. 1, no. 1, March 2012, pp.  49–67. Nandini Sundar, ‘In the Cause of Anthropology: The Life and Work of Irawati Karve’, in Anthropology in the East: Founders of Indian Sociology and Anthropology, eds. Patricia Uberoi, Nadini Sundar and Satish Deshpande, Ranikhet: Permanent Black, 2007, pp. 360–416. Nicole Belmont, Arnold Van Gennep: Le Créateur de l’ethnographie française, Paris: Payot, 1974. Marilyn Strathern, ‘An Awkward Relationship: The Case of Feminism and Anthropology’, Signs: Journal of Women in Culture and Society, vol. 12, no. 2, Winter 1987, pp. 276–292. Susan Seymour, Cora DuBois: Anthropologist, Diplomat, Agent, Lincoln: University of Nebraska Press, 2015. Sally Slocum, ‘Woman the Gatherer: Male Bias in Anthropology’, in Toward an Anthropology of Women, ed. Rayna Reiter, pp. 36–50, New York: Monthly Review, 1975, p. 50. Clifford Geertz noted an additional problem of anthropology in that practitioners 137

Barry H. Rodrigue

46

47

48 49

50

51

had largely restricted their readings to those of other anthropologists up until the 1960s. Clifford Geertz, in ‘An Interview with Clifford Geertz’, by Richard Handler, Current Anthropology, vol. 32, no. 5, December 1991, p. 611. United Nations, History of Humanity, Paris: United Nations Educational, Scientific and Cultural Organization, 1966, 2009; International Commission for a History of the Scientific and Cultural Development of Mankind, , accessed 4 December 2014. Poul Duedahl, ‘Selling Mankind: UNESCO and the Invention of Global History, 1945-1976’, Journal of World History, vol. 22, no. 1, March 2011, pp. 101–133. United Nations Educational, Scientific and Cultural Organization, ‘World Heritage List’, , accessed 13 April 2018. This diversification of anthropology was similar to the re-envisioning of Western Civilization as Global Civilization, a crediting of the reality of the field in its full context. Enrico Bertacchini et al., eds., Cultural Commons: A New Perspective on the Production and Evolution of Cultures, Cheltenham: Edward Elgar Publishing, 2012. Neil MacGregor, presentation, and conversation with Barry Rodrigue, Symbiosis School for Liberal Arts, 31 January 2018. MacGregor’s discourse was part of the exhibit, India and the World: A History in Nine Stories, mounted by the Chhatrapati Shivaji Maharaj Vastu Sangrahalaya (Mumbai), the British Museum (London) and the National Museum (New Delhi). For reference to MacGregor’s theses, see the following. Neil MacGregor, A History of the World in 100 Objects, London: Allen Lane, 2010. Haviland et al., op. cit., p. 47. An example of this broadening of anthropology across the boundaries of disciplines is a recent edition of Science News Magazine that featured the theme of ‘protein archaeology’, which presented advances in understanding protein synthesis and how it informs us about more expansive aspects of evolution. The magazine’s featured article was­Jennifer ­M ichalowski, ‘Proteins of the Past: Reconstructing Tiny Pieces of History Deepens Understanding of Evolution’, Science News Magazine, vol. 189, no. 12, 11 June 2016, pp. 16–20. These calculations are, of course, rough estimates, as there were many variables, including how previous societies encouraged having children much earlier in life than does modern industrial society. Among those who contemplate humanity’s potential to bring about change in the multiverse are the following. Akop Nazaretyan, ‘Life’s Meaning as a Global Problem of Modernity: A View from a Big History and ­Complexity-Studies Perspective’, op. cit., pp. 317–338. Alexander Panov, ‘Singularity of Evolution and Post­Singular Development’, op. cit., pp. 370–402. Joseph Voros, ‘Big Futures: M ­ acroHistorical Perspectives on the Future of Humankind’, op. cit., pp. 403–436. Professor Zhu Weibin of Sun Yat-sen University noted the connection of Gauguin’s work to human studies and big history. Weibin Zhu, ‘Big History and World History in China’s Colleges and Universities’, in From the Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin, Andrey Korotayev, Delhi: Primus Books, 2016, p. 323n5. Among the big historians who are anthropologists are Andrey Korotayev, Dmitri Bondarenko, Magomedkhan Magomedkhanov and Nikolay Kradin. This preponderance of Russian anthropologists highlights the macro-perspective in Soviet and post-Soviet scholarship. Big historian Fred Spier has written of his development into big history by way of biochemistry, sociology and anthropology. Fred Spier, ‘Roads towards Big History’, op. cit., Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, 2017, pp. 253–260. In addition, there are many other scholars from other disciplines who are engaged in cultural studies. 138

Big history and anthropology

52 The Göttingen School of History was not all progressive, but the stream represented by Blumenbach and the von Humboldts were more enlightened. Gerd Hohendorf, ‘Wilhelm von Humboldt’, Prospects: The Quarterly Review of Comparative Education, Paris: UNESCO, International Bureau of Education, vol. 23, no. 3/4, 1993, pp. 665–676. Alexander von Humboldt, op. cit. Gerard Helferich, Humboldt’s Cosmos, New York: Gotham Books, 2004. Fred Spier has linked big history thought to Alexander von Humboldt. Fred Spier, Big History and the Future of Humanity, Oxford: Wiley-Blackwell, 2010, p. 10. 53 The classical structure of the medieval university included the faculties of theology, medicine, law and arts. Certain specializations had developed in other training programs, such as in mining schools. The new movement towards university specialization in the last half of the nineteenth and early twentieth centuries occurred by the development of departments from the philosophy and arts faculty. For example, ‘science’ originally meant ‘knowledge’ and had been considered ‘natural philosophy/natural history’ in the arts before it broke into its present departments. For this reason, it is common to have a joined university faculty of ‘arts and sciences’ even today. Reba Soffer, op. cit. Michel Foucault, Discipline and Punish: The Birth of the Prison, Paris: Gallimard, 1975, Part 3, Sections 2–3 [French publication]. Immanuel Wallerstein, Unthinking Social Science: The Limits of Nineteenth-Century Paradigms, Cambridge: Polity Press, 1991. 54 Alfred Russel Wallace, Man’s Place in the Universe: A Study of the Results of Scientific Research in Relation to the Unity or Plurality of Worlds, New York: McClure, Phillips & Company, 1903. Hiram Percy Maxim, Life’s Place in the Cosmos, New York: D. ­Appleton, 1933. Kinji Imanishi, ‘A Proposal for Shizengaku: The Conclusion to my Study of Evolutionary Theory’, Anthropology Quarterly, vol. 14, no. 3, 1983, pp. 3–18 [ Japanese publication]; idem, ‘A Proposal for Shizengaku: The Conclusion to my Study of ­Evolutionary Theory’, trans. Rick Davis, Journal for Social and Biological Structures, vol. 7, no. 4, October 1984, pp. 351–368. An important part of Imanishi’s integrated studies included his concept of 自然学 (deep nature thought). I would like to thank Nobuo Tsujimura for sharing his insights about Imanishi. Nobuo Tsujimura, personal communications (e-mail), to Barry Rodrigue, 4 June 2017. I am also grateful to Alex ­Holowicki for his presentation on ‘Big History and Big Anxieties in the Interwar Period: Rethinking ­H iram Percy Maxim’s Life’s Place in the Cosmos’ at the third conference of the International Big History Association in Amsterdam in 2016. 55 In 2009, when I began assembling an international directory and bibliography of big history, I noted a global conjuncture of macro-studies occurred in the second half of the twentieth century, when scholars from many disciplines independently began to generate holistic studies under the names of cosmic evolution, the open complex giant system, universal studies, big history, etc. So I began an open-ended essay, ‘Big ­H istory – A Study of All Existence’, that I update with newly discovered insights as I learn about them. Barry Rodrigue, ‘A Big History Directory’, 2011; idem, ‘Big ­H istory: A Working Bibliography of References, Films and Internet Sites’, 2011; idem, ‘Big ­H istory – A Study of All Existence’, 2010+; . 56 Joseph Shklovsky, Universe, Life, Intelligence, Moscow: USSR Academy of Sciences, 1962 [Russian publication]. Joseph Shklovsky and Carl Sagan, Intelligent Life in the Universe, New York: Random House, 1966. Preston Cloud, Cosmos, Earth and Man: A Short History of the Universe, New Haven: Yale University Press, 1978. Erich Jantsch, The Self-­ Organizing Universe: Scientific and Human Implications of the Emerging Paradigm of Evolution, Oxford: Pergamon Press, 1980. Antonio Vélez, Man: Inheritance and Conduct, Medellín: Editorial Universidad de Antioquia, 1986 [Spanish publication]. Lynn Margulis and Dorion Sagan, Microcosmos: Four Billion Years of Evolution from Our Microbial Ancestors, 139

Barry H. Rodrigue

57

58

59

60

61

62

63

New York; Summit Books, 1986. Xuesen Qian, Jingyuan Yu and Ruwei Dai, ‘A New Discipline of Science – The Study of Open Complex Giant System and Its Methodology’, Chinese Journal of Nature, vol. 1, 1990, pp. 3–10 [Chinese publication]. Andre Gunder Frank, World Accumulation, 1492–1789, London: Macmillan Press, 1978. Immanuel Wallerstein, The Politics of the World Economy: The States, the Movements and the Civilizations, Cambridge: Cambridge University Press, 1984. Georges Nicolas, ‘­Humanisme – Cosmisme’, Cahiers de Géographie du Québec, vol. 33, no. 90, December 1989, pp. 379–385. Graeme Snooks, The Dynamic Society: Exploring the Sources of Global Change, London: Routledge, 1996. Eric Chaisson, Syllabus, Astronomy 8, Cosmic Evolution, Harvard University, ­Cambridge, Massachusetts, Fall 1975; idem, ‘The Scenario of Cosmic Evolution’. ­Harvard Magazine, November–December 1977, pp. 21–33; idem., ‘The Broadest View of the Biggest Picture’. Harvard Magazine, January–February 1982, pp. 21–25. George Field, Interview by Richard Hirsh, Smithsonian Astrophysical Observatory, Cambridge, Massachusetts, 14 July 1980. Center for the History of Physics, American Institute of Physics, , accessed 4 December 2014. George Field, Gerrit Verschuur and Cyril Ponnamperuma, Cosmic Evolution: An Introduction to Astronomy, Boston: Houghton Mifflin, 1978. Eric Chaisson, Cosmic Dawn: The Origins of Matter and Life, Boston: Little Brown, 1981. G. Siegfried Kutter, Universe and Life: Origins and Evolution, Boston: Jones & Bartlett, 1987; idem., ‘Big History: A Personal Perspective’, in Evolution: A Big History Perspective, eds. Leonid Grinin, Andrey Korotayev and Barry Rodrigue, Volgograd: Uchitel Publishing, 2011, pp. 101–120. John Mears 1986; personal communication, to Barry Rodrigue, Western History Association, Conference, Incline Village, Nevada, 14 October 2010. David Christian, ‘The Return of Universal History’, History and Theory, vol. 49, no. 4, December 2010, pp. 6–27. Jiddu Krishnamurti, Beginnings of Learning, Worthing: Littlehampton Book Services, 1975. Thomas Berry, Dream of the Earth: The Universe Story, San Francisco: Sierra Club Books, 1988. On some of these on-going initiatives, see Orla Hazra, ‘Tarumitra: Friends of Trees, Understanding and Practicing an Integrated Cosmology’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 191–202. Arne Næss, ‘The Shallow and the Deep, Long-Range Ecology Movements: A Summary’, Inquiry, vol. 16, nos. 1–4, 1973, pp. 95–100. David Attenborough, Ask Me Anything, Reddit, 8 January 2014, , accessed 18 May 2018. Osamu Nakanishi, ed., Applying Big History: Nature, War and Peace, Yokohama: Institute for Global and Cosmic Peace, 2016; idem., Universal Studies and the Modern World: Becoming Global and Cosmic Humanity, Yokohama: Institute for Global and Cosmic Peace, 2017. Osamu Nakanishi and Nobuo Tsujimura, ‘Universal Studies and Big History in Japan’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin, Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 289–294. Barry Rodrigue, ‘The Study of all Existence: Big History, Universal Studies and the Global Conjuncture’, International Journal for the Transformation of Consciousness, vol. 3, no. 1, June 2017, pp. 15–34. Gordon McBean and Alberto Martinelli, ‘Blurring Disciplinary Boundaries’, Science, vol. 358, no. 6366, 24 November 2017, p. 975. 140

Big history and anthropology

64 Craig Benjamin, ‘The Little Big History of Jericho’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume 1, Our Place in the Universe: An Introduction to Big History, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2015, p. 261. 65 Nigel Hughes and Rati Basu, Monisha and the Stone Forest, Bengaluru: Geological Society of India, 2012. Nigel Hughes and Rati Basu, Monishar Pathorer Bon, Kolkata: Monfakira Press, 2012. It has also been produced online. Nigel Hughes, Rati Basu, Bipattaran and Ensemble, Monisha and the Stone Forest, , accessed 10 December 2017. Nigel Hughes, Payel Ghosh and Dipen Bhattacharya, ‘The Monishar Pathorer Bon (Monisha and the Stone Forest) Book Project: Novel Educational Outreach to Children in Rural Communities, Eastern Indian Subcontinent’, Journal of Geoscience Education, vol. 63, no. 1, February 2015, pp. 18–28. 66 The Indian Association for Big History is based at the Symbiosis School for Liberal Arts, with its founding president being anthropologist Shweta Sinha Deshpande. India’s first big history course was built from an anthropology core in 2018, while the SSLA Collaborative for Asian Anthropology has an interdisciplinary mission of universal studies. Sarat Chandra Roy, op. cit., p. 150. 67 Robin Dunbar, Human Evolution: Our Brains and Behavior, New York: Oxford University Press, 2016. Bruce Bower, ‘Evolution’s Ear’, Science News, vol. 174, no. 5, 15 August 2008, pp. 22–25. Barry Rodrigue, ‘A New Design for Living’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume I, Our Place in the Universe: An Introduction to Big History, New Delhi: Primus Books, 2015, pp. 183–187. Francis Heylighen, ‘Conceptions of a Global Brain: An Historical Review’, op. cit., Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 341–356. Vernor Vinge, ‘The Coming Technological Singularity: How to Survive in the Post-Human Era’, op. cit., pp. 357–369. 68 In an attempt to more intentionally cultivate global identity and awareness, students at J.F. Oberlin University in Japan and Symbiosis International University in India worked together to promote a ‘Guideline for Global Humanity’. Barry Rodrigue had proposed developing this guideline in a Japanese collection of essays about big history and universal studies in 2017, so volume co-authors Nobuo Tsujimura and Hirofumi Katayama worked to assemble this inter-university project. Their students collaborated on an assignment called, Applying Big History to Problems of Peace and Symbiosis, via the Edmodo online educational network , which was administered by Rachael Guarnaccia in autumn 2017. The assignment was for the students to sort out what cultural artefacts in their heritage would be conducive to promoting global identity and cooperation, versus those traditions that would hinder cooperation or be neutral. This effort is on-going. Barry Rodrigue, ‘An Emergent Future: Evolving A Global Revolution’, in Universal Studies and the Modern World, Yokohama: Institute of Global and Cosmic Peace, 2017, pp. 196–218. One of the resulting essays was published by Symbiosis student Anaga Krishna in a big history publication. Anaga Krishna, ‘Suggestions for a Global Guideline for Humanity’, Origins, vol. 8, no. 3, pp. 7–10. We plan to continue this effort and invite participation by others; contact for this can be made through the Oberlin Big History Project at . 69 Barry Rodrigue op cit. Akop Nazaretyan presented his concept of techno-­humanitarian balance in the Russian journal, Social Sciences Today, in 1993. Initially, he called it the Law of Evolutionary Correlations, but, as a result of further work, he refined it into its present form. It has appeared in many publications, two of which follow. Akop ­Nazaretyan, ‘Technology and Psychology: On the Concept of Evolutionary Crises’, 141

Barry H. Rodrigue

70

71

72

7 3 74

75

76

Social Sciences Today, no. 3, 1993; Evolution of Non-Violence: Studies in Big History, Self-­ Organization and Historical Psychology, Saarbrucken: Lambert, 2010. Psychologist Steven Pinker later documented a similar trend. Steven Pinker, The Better Angels of Our Nature: Why Violence Has Declined, New York: Viking Books, 2011. The Spacewatch Program was begun by astronomer Tom Gehrels at the Kitt Peak National Observatory in Arizona (USA) in 1980. University of Arizona, Lunar and Planetary Laboratory, Spacewatch, , accessed 19 May 2018. Tom Gehrels, On the Glassy Sea: In Search of a Worldview, Charleston: BookSurge, 2007, pp. 183–202; idem., personal communication, to Barry Rodrigue, 24 March 2010. In this way, anthropologists Sada Mire in Somaliland and Shweta Sinha Deshpande in India contemplate how a macro-views help resolve issues in a diffuse range of social settings, from individual and family discord to scholarly and national identities. Sada Mire, ‘Somalia: Studying the Past to Create a Future’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 279–288. Shweta Sinha Deshpande, op. cit., pp. 255–264. Osamu Nakanishi and Nobuo Tsujimura, op. cit., p. 290. Leslie White, ‘Energy and the Evolution of Culture’, American Anthropologist, vol. 45, no. 3, Part 1, July 1943, p. 338. This study of society through the lens of energy and its products continued among neo-materialists and cultural materialists like Marvin Harris and Robert Carneiro. Big historians have similarly engaged in this study. Frank Neile, Energy: Engine of Evolution, Amsterdam: Elsevier, Shell Global Solutions, 2005; idem., ‘The Next Energy Revolution: Evolutionary Energetics, Models and Scenarios’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 197–250. Leslie White, op. cit., p. 335. Eric Chaisson, Cosmic Evolution: Rise of Complexity in Nature, Cambridge: Harvard University Press, 2001. Some advocates of cosmic evolution and big history seek to more firmly establish their field as a more concrete field of study. One of their goals is to set out research agendas, with one of these agendas being a comparative study of complexity. Fred Spier, ‘Big History Research: A First Outline’, Evolution: A Big History Perspective, Volgograd: Uchitel Publishing, 2011, pp. 26–36. Michael Smith and Peter Peregrine, ‘Approaches to Comparative Analysis in Archaeology,’ The Comparative Archaeology of Complex Societies, ed. Michael Smith, Cambridge: Cambridge University Press, 2012, pp. 4–20. A way that the complexity ratio might work in anthropology is that one could tally the energy-rate density of individual artefacts and then add them together to arrive at the complexity ratio for a site, which would then yield a number to compare disparate sites. A table of assigned values of complexity would need to be developed so that one could calculate such numbers. A researcher could then go to a table and look up the complexity value of an iron sickle, a chicken or a chert projectile point, and then tally them with other artefact values to arrive at a complexity number for a given site. It would not be a final word in analysis, but it could provide common numbers for comparative purposes. In 2012, I proposed the creation of a table of complexity values to proponents of energy-rate density, but no one – understandably – wanted to go through the labour-­intensive process of assembling this kind of reference. There are other models that might be of similar use, such as Graeme Snooks’ dynamic model of society. The synthesis of such new strategies could result in new conceptualizations. Graeme Snooks, op. cit. 142

Big history and anthropology

77 Yoto Yotov et al., An Advanced Guide to Trade Policy Analysis: The Structural Gravity Model, Geneva: United Nations and World Trade Organization, 2016. Gojko Barjamovic, Thomas Chaney, Kerem Coşar and Ali Hortaçsu, ‘Trade, Merchants, and the Lost Cities of the Bronze Age’, National Bureau of Economic Research, 2017, , accessed 24 April 2018. 78 Gilles Clément and Millard Reschke, Neuroscience in Space, Berlin: Springer, 2010. ­Matthew Ralphs et al., ‘Water Extraction on Mars for an Expanding Human Colony’, Life Sciences in Space Research, vol. 7, November 2015, pp. 57–60. Asian Scientist Newsroom, ‘Chinese Taikonauts To Grow Vegetables On Moon’, Asian Scientist, 10 December 2012, , accessed 20 May 2018. 79 Francesco D’Errico and Chris Stringer, ‘Evolution, Revolution or Saltation Scenario for the Emergence of Modern Cultures?’, Philosophical Transactions of the Royal Society B, vol. 366, no. 1567, 12 April 2011, pp. 1060–1069. Steven Mithen, Problem-Solving and the Evolution of Human Culture, London: Institute for Cultural Research, 1999. Graziano Fiorito and Pietro Scotto, ‘Observational Learning in Octopus vulgaris’, Science, vol. 256, no. 5056, 24 April 1992, pp. 545–547. Paul Patton, ‘One World, Many Minds’, Scientific American Mind, December 2008, pp. 72–79. Robin Dunbar, op. cit. 80 Dale Russell, An Odyssey in Time: The Dinosaurs of North America, Toronto: University of Toronto Press and the Canadian National Museum of Natural Sciences, 1993, pp. 208–218; idem, ‘Speculations on the Evolution of Intelligence in Multicellular Organisms’, in Life In The Universe, ed. John Billingham, Proceedings of a Conference Held at the National Ames Research Center, Moffet Field, California, June 19–20, 1979 (CP–2156), National Aeronautics and Space Administration, Scientific and Technical Information Branch, 1981, pp. 259–275. 81 United States, National Aeronautics and Space Administration, Jet Propulsion Laboratory, ‘Voyager’, , accessed 20 May 2018; idem, ‘The Pioneer Missions’, 26 March 2007, , accessed 20 May 2018. 82 Alexander Ollongren, Astrolinguistics: Design of a Linguistic System for Interstellar Communication Based on Logic, Berlin: Springer, 2012. Lincos was designed by Hans Freudenthal of the University of Utrecht in the Netherlands in 1960. Hans Freudenthal, Lincos: Design of a Language for Cosmic Intercourse, Part 1, Amsterdam: North-Holland Publishing Company, 1960. Astrophysicists Yvan Dutil and Stéphane Dumas from Québec (Canada) created a noise-resistant coding system and used Lincos in their broadcasts of messages from Ukraine. Yvan Dutil and Adrian Hon, ‘Lincos with Dr. Yvan Dutil’, Astrobiology, 4 July 2009, , accessed 6 May 2018. 83 Mike Wall, ‘Signs of Alien Life will be Found by 2025, NASA’s Chief Scientist Predicts’, Space.com, 7 April 2015, , accessed 19 April 2015. United States, National Aeronautics and Space Administration, ‘Hubble Reveals Observable Universe Contains Ten Times More Galaxies than Previously Thought’, 13 October 2016, , accessed 25 November 2017. 84 Douglas Vakoch, Archaeology, Anthropology and Interstellar Communication, Washington, DC: National Aeronautics and Space Administration, 2014. Nobuo Tsujimura and Hirofumi Katayama, ‘Think Cosmically, Act Globally: Emerging Clues for the Big History Movements’, International Journal for the Transformation of Consciousness, vol. 3, no. 1, June 2017, p. 51. In this spirit of detecting remnants of extra-terrestrial civilizations, computer scientist Vernor Vinge opens his novel, A Fire Upon the Deep, with archaeologists 143

Barry H. Rodrigue

85

86

87

88

89 90 91

‘excavating’ a five-billion-year-old data archive in outer space and inadvertently releasing a malevolent super-intelligence into the galaxy. Vernor Vinge, A Fire Upon the Deep, New York: Tor Books, 1992. Vinge’s story acts as a link between cultural vectors of change and new paradigms of expression. For example, the common image of archaeologists releasing cursed entities into the living world harkens back to a primordial human fear of the dead. Akop Nazaretyan, ‘Fear of the Dead as a Factor in Social Self-Organization’, Journal for the Theory of Social Behaviour, vol. 35, no. 2, June 2005, pp. 155–69. Alexander Panov, op. cit., pp. 381–384. Joseph Voros, op. cit. pp. 419–425. Since we are made of elements generated in first- and second-generation stars billions of years ago, this vastly expanded vision allows us to contemplate how it is possible that we are made of elements once part of older and now extinct extra-terrestrial civilizations, just as some of the tiles in many homes are made of limestone from quarries that held Homo erectus bones. Speculative fiction has engaged with these kinds of syntheses far longer than academia, as poets, artists and creative visionaries have seen humanity in the wider realms of nature and the cosmos. While Panov sees the need for a new paradigm to at least invigorate science, he discards religion as an antiquated structure that already served as a social framework. Alexander Panov, op. cit. Joseph Voros, op. cit. Akop Nazaretyan, Non-Linear Future: Mega-History, Complexity Theory, Anthropology and Psychology for Global Forecasting, ­Moscow: Agarmak Media, 2017 [Russian publication]. As an example of Panov’s critique, there has been trend away from depth of knowledge in recent decades, at least in some venues. It is related to the expansion of digitized technology in the 1990s, as new teachers and workers who were technologically proficient but lacked disciplinary depth were hired and promoted. The goal was to make institutions modern, relevant and competitive, but a side effect was the shallowing of disciplinary knowledge. An obvious solution would have been the encouragement of partnerships between the two sets of professionals, but this has not happened because of poor understanding of the process and therefore of the problem. This concept disciplinary shallowing was developed by digital historian W ­ illiam J. Turkel. William J. Turkel, personal communication (conversation), with Barry Rodrigue, Oakville, Ontario (Canada), 1 August 2011. It is possible that big history and universal studies could be one of the tools to mitigate problems such as these. Alexander Panov, op. cit.. Akop Nazaretyan, Civilization Crises in a Universal History Context: Self-Organization, Psychology and Forecasting, Moscow: Mir, 2004 [Russian publication]; idem., ‘Life’s Meaning as a Global Problem of Modernity’, op. cit. In a practical application of this concept to my own work in anthropology, the need to purposively review and select useful strategies from the worldwide cultural commons for present and future use is a process that I call ‘retrofitting the future’. Barry Rodrigue, ‘Retrofitting the Future’, in Teaching and Researching Big History: Exploring a New Scholarly Field, Leonid Grinin et al., eds., Volgograd: Uchitel Publishing, 2014, pp. 276–282. The synthesis in this paragraph is the result of discussions with members of the ­Eurasian Center for Megahistory and System Forecasting. Panov’s description of the scale-­ invariant sequence is found at Alexander Panov, op. cit., pp. 371–375. These numbers are derived from the calculations of physicist and ecologist Priyadarshini Karve at Symbiosis International University. Some of my thoughts about this needed transformation of society are discussed in more detail in Barry Rodrigue, ‘A New Design for Living’, op. cit. Vandana Singh, ‘A Speculative Manifesto’, in The Woman Who Thought She Was a Planet and Other Stories, New Delhi: Zubaan, 2013, p. 201.

144

Big history and anthropology

Bibliography Amin, Osama, ‘The Newly Discovered Tablet V of the Epic of Gilgamesh’, Ancient History Et Cetera, 24 September 2015, http://etc.ancient.eu/exhibitions/giglamesh-enkidu-humbabacedar-forest-newest-discovered-tablet-v-epic/, accessed 25 November 2017. Asian Scientist Newsroom, ‘Chinese Taikonauts to Grow Vegetables on Moon’, Asian Scientist, 10 December 2012, www.asianscientist.com/2012/12/topnews/chinese-taikonautsto-grow-vegetables-on-moon-2012/, accessed 20 May 2018. Attenborough, David, Ask Me Anything, Reddit, 8 January 2014, www.reddit.com/r/ IAmA/comments/1uqlmf/i_am_sir_david_attenborough_im_on_reddit_ama/, accessed 18 May 2018. Baldwin, James Mark, ‘A New Factor in Evolution’, The American Naturalist, vol. 30, no. 354, June 1896, pp. 441–451. Bandi, Hans-Georg, Eskimo Prehistory, Fairbanks: University of Alaska Press, 1964, 1972. Bank, Rosemarie, ‘“Show Indians”/Showing Indians: Buffalo Bill’s Wild West, the Bureau of Indian Affairs, and American Anthropology’, Journal of Dramatic Theory and Criticism, vol. 26, no. 1, Fall 2011, pp. 149–158. Barbone, Giuseppe et al., ‘La roccia con incisioni del Monte Cònero: relazione preliminare’, Preistoria Alpina, vol. 46, no. 2, 2012, pp. 93–98. Barjamovic, Gojko et al., ‘Trade, Merchants, and the Lost Cities of the Bronze Age’, National Bureau of Economic Research, 2017, www.nber.org/people/gojko_barjamovic, accessed 24 April 2018. Belmont, Nicole, Arnold Van Gennep: Le Créateur de l’ethnographie française, Paris: Payot, 1974. Benedict, Ruth, Patterns of Culture, Boston: Houghton Mifflin Company, 1934. Benfey, Christopher, The Great Wave: Gilded Age Misfits, Japanese Eccentrics, and the Opening of Old Japan, New York: Random House, 2004. Benjamin, Craig, ‘The Little Big History of Jericho’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume 1, Our Place in the Universe: An Introduction to Big History, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2015, pp. 247–262. Bernstein, Jay, ‘First Recipients of Anthropological Doctorates in the United States, ­1891–1930’, American Anthropologist, vol. 104, no. 2, 2002, pp. 551–564. Berry, Thomas, Dream of the Earth: The Universe Story, San Francisco: Sierra Club Books, 1988. Bertacchini, Enrico et al., eds., Cultural Commons: A New Perspective on the Production and Evolution of Cultures, Cheltenham: Edward Elgar Publishing, 2012. Blumenbach, Johann Friedrich, On the Natural Variety of Humankind, Göttingen: ­Vandenhoek & Ruprecht, 1781 (Latin publication). Boas, Franz, ‘Ethnology of the Kwakiutl: Based on Data Collected by George Hunt’, 35th Annual Report of the Bureau of American Ethnology, Parts 1 and 2, Washington, DC: ­Bureau of American Ethnology, 1921. Bondarenko, Dmitri and Ken Baskin, ‘Big History, Complexity Theory, and Life in a Non-Linear World,’ in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 183–196. Bower, Bruce, ‘Evolution’s Ear’, Science News, vol. 174, no. 5, 15 August 2008, pp. 22–25. British Association for the Advancement of Science, Notes and Queries on Anthropology: For the Use of Travellers and Residents in Uncivilized Lands, London: Edward Stanford, 1874.

145

Barry H. Rodrigue

Brownell, Susan, The 1904 Anthropology Days and Olympic Games: Sport, Race, and American Imperialism, Lincoln: University of Nebraska Press, 2008. Chaisson, Eric, Syllabus, Astronomy 8, Cosmic Evolution, Cambridge, MA: Harvard ­University, Fall 1975. ———, ‘The Scenario of Cosmic Evolution’, Harvard Magazine, November–December 1977, pp. 21–33. ———, ‘The Broadest View of the Biggest Picture’, Harvard Magazine, January–February 1982, pp. 21–25. ———, Cosmic Dawn: The Origins of Matter and Life, Boston: Little Brown, 1981. ———, Cosmic Evolution: The Rise of Complexity in Nature, Cambridge: Harvard University Press, 2001. ———, ‘Energy Rate Density as a Complexity Metric and Evolutionary Driver’, Complexity, vol. 16, no. 3, January 2011, pp. 27–40. Christian, David, ‘The Return of Universal History’, History and Theory, vol. 49, no. 4, December 2010, pp. 6–27. Christenson, Andrew, ‘Who were the Professional North American Archaeologists of 1900? Clues from the Work of Warren K. Moorehead’, Bulletin of the History of Archaeology, vol. 21, no. 1, May 2011, pp. 4–23. Clément, Gilles and Millard Reschke, Neuroscience in Space, Berlin: Springer, 2010. Cloud, Preston, Cosmos, Earth and Man: A Short History of the Universe, New Haven: Yale University Press, 1978. Coccioni, Rodolfo, et al., Carta Geologica con Itinerari Escursionistici: Firenze, Italia, Parco Regionale del Conero, (1:20,000) Florence: SELCA, 1993. D’Andrade, Roy, ‘Moral Models in Anthropology’, Current Anthropology, vol. 16, no. 3, June 1995, pp. 399–408. Darwin, Charles, On the Origin of Species by Means of Natural Selection, or the Preservation of Favoured Races in the Struggle for Life, London: John Murray, 1859. Das, Sarat Chandra, A Journey to Lhasa and Central Tibet, London: John Murray, 1902, pp. 62, 68, 108–110. Dasgupta, Sangeeta, ‘Recasting the Oraons and the “Tribe”’, in Anthropology in the East: Founders of Indian Sociology and Anthropology, eds. Patricia Uberoi, Nadini Sundar and Satish Deshpande, Ranikhet: Permanent Black, 2007, pp. 132–171. D’Errico, Francesco and Chris Stringer, ‘Evolution, Revolution or Saltation Scenario for the Emergence of Modern Cultures?’, Philosophical Transactions of the Royal Society B, vol. 366, no. 1567, 12 April 2011, pp. 1060–1069. Deshpande, Shweta Sinha, ‘India: Evolving a Big History’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 255–264. Duedahl, Poul, ‘Selling Mankind: UNESCO and the Invention of Global History, ­1945–1976’, Journal of World History, vol. 22, no. 1, March 2011, pp. 101–133. Dunbar, Robin, Human Evolution: Our Brains and Behavior, New York: Oxford University Press, 2016. Durand, Jean-Yves, ‘“Traditional Culture” and “Folk Knowledge”: Wither the Dialogue between Western and Post-Soviet Anthropology?’, Current Anthropology, vol. 16, no. 2, April 1995, pp. 326–330. Dutil, Yvan and Adrian Hon, ‘Lincos with Dr. Yvan Dutil’, Astrobiology, 4 July 2009, http://library.thinkquest.org/C003763/pdf/interview01.pdf, accessed 6 May 2018. Fee, Christopher, Mythology in the Middle Ages: Heroic Tales of Monsters, Magic, and Might, Santa Barbara: Praeger, 2011. 146

Big history and anthropology

Field, George, Interview by Richard Hirsh, Smithsonian Astrophysical Observatory, Cambridge, MA, 14 July 1980; Interview and transcript at the Center for the History of Physics, American Institute of Physics www.aip.org/history/ohilist/4602_1.html, accessed 4 December 2014. Field, George, Gerrit Verschuur, and Cyril Ponnamperuma, Cosmic Evolution: An Introduction to Astronomy. Boston: Houghton Mifflin, 1978. Fiorito, Graziano and Pietro Scotto, ‘Observational Learning in Octopus vulgaris’, Science, Vol. 256, no. 5056, 24 Apr 1992, pp. 545–547. Fitz Gibbon, Kate, Who Owns the Past?: Cultural Policy, Cultural Property, and the Law, New Brunswick: Rutgers University Press, 2005. Flory, Stewart, ‘Who Read Herodotus’, Histories?’ The American Journal of Philology, vol. 101, no. 1, Spring 1980, pp. 12–28. Foucault, Michel, Discipline and Punish: The Birth of the Prison, Paris: Gallimard, 1975 [French publication]. Frank, Andre Gunder, World Accumulation, 1492–1789, London: Macmillan Press, 1978. Freudenthal, Hans, Lincos: Design of a Language for Cosmic Intercourse, Part 1, Amsterdam: North-Holland Publishing Company, 1960. Gartner, Richard, Metadata: Shaping Knowledge from Antiquity to the Semantic Web, Cham: Springer International Publishing, 2016. Geertz, Clifford, in ‘An Interview with Clifford Geertz’, by Richard Handler, Current Anthropology, vol. 32, no. 5, December 1991, p. 611. Gehrels, Tom, On the Glassy Sea: In Search of a Worldview, Charleston: BookSurge, 2007. ———, ‘The Chandra Multiverse’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 45–70. Gilli, Barbara, ‘The Past in the Present: The Reuse of Ancient Material in the 12th ­D ynasty’, Aegyptus, vol. 89, nos. 1–2, 2009, 89–110. Ghosh, Amitav, In an Antique Land, New Delhi: Ravi Dayal Publishers, 1992. Gould, Stephen Jay, The Mismeasure of Man, New York: W.W. Norton, 1981, 1996. Grinin, Leonid, Andrey Korotayev and Alexander Markov, ‘Biological and Social Phases of Big History: Evolutionary Principals and Mechanisms’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey ­Korotayev, Delhi: Primus Books, 2017, pp. 141–174. Grinin, Leonid, Andrey Korotayev and Barry Rodrigue eds., Evolution: A Big History Perspective, Volgograd: Uchitel Publishing, 2011. Haviland, William et al., Cultural Anthropology: The Human Challenge, Boston: Cengage Learning, 2017. Haw, Stephen, Marco Polo’s China: A Venetian in the Realm of Khubilai Khan, London: Routledge, 2006. Hazra, Orla, ‘Tarumitra: Friends of Trees, Understanding and Practicing an Integrated Cosmology’, in From Big Bang to Galactic Civilizations: A Big History Anthology, ­Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, ­Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 191–202. Helferich, Gerard. Humboldt’s Cosmos, New York: Gotham Books, 2004. Heylighen, Francis, ‘Conceptions of a Global Brain: An Historical Review’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 341–356. 147

Barry H. Rodrigue

Hinsley, Curtis, The Smithsonian and the American Indian: Making a Moral Anthropology in Victorian America, Washington, DC: Smithsonian Institution Press, 1981, 1994. Hoffman, Adina, Sacred Trash: The Lost and Found World of the Cairo Geniza, New York: Schocken, 2011. Hohendorf, Gerd, ‘Wilhelm von Humboldt’, Prospects: The Quarterly Review of Comparative Education, vol. 23, no. 3/4, 1993, Paris: UNESCO, International Bureau of Education, pp. 665–676. Hostetler, Laura, ‘Qing Connections to the Early Modern World: Ethnography and ­Cartography in Eighteenth-Century China’, Modern Asian Studies, vol. 34, no. 3, July 2000, pp. 623–662. Hughes, Nigel and Rati Basu, Monisha and the Stone Forest, Bengaluru: Geological Society of India, 2012. ———, Monishar Pathorer Bon, Kolkata: Monfakira Press, 2012. Hughes, Nigel, Rati Basu, Bipattaran and Ensemble, Monisha and the Stone Forest, www. youtube.com/watch?v=UDrRBwp9vds, accessed 10 December 2017. Hughes, Nigel, Payel Ghosh and Dipen Bhattacharya, ‘The Monishar Pathorer Bon (­Monisha and the Stone Forest) Book Project: Novel Educational Outreach to Children in Rural Communities, Eastern Indian Subcontinent’, Journal of Geoscience Education, vol. 63, no. 1, February 2015, pp. 18–28. Humboldt, Alexander von, Kosmos, vols. 1–5, Stuttgart: J.G. Cotta Publishing, 1845–1862. Imanishi, Kinji, The World of Living Things, Tokyo: Kobundo shobo, 1941 [ Japanese publication]. ———, ‘A Proposal for Shizengaku: The Conclusion to my Study of Evolutionary T ­ heory’, Anthropology Quarterly, vol. 14, no. 3, 1983, pp. 3–18 [ Japanese publication]. ———, ‘A Proposal for Shizengaku: The Conclusion to my Study of Evolutionary Theory’, trans. Rick Davis, Journal for Social and Biological Structures, vol. 7, no. 4, October 1984, pp. 351–368. Jantsch, Erich, The Self-Organizing Universe: Scientific and Human Implications of the Emerging Paradigm of Evolution, Oxford: Pergamon Press, 1980. Kilani, Mondher, ‘Is a Peripheral Anthropology Possible? The Issue of Universalism’, ­Kroeber Anthropological Society Papers, vol. 101, no. 1, 2012, pp. 98–105. Knox, Robert, The Races of Men: A Philosophical Enquiry into the Influence of Race over the Destinies of Nations, London: Henry Renshaw, 1850, 1862. Krishna, Anaga, ‘Suggestions for a Global Guideline for Humanity’, Origins, vol. 8, no. 3, pp. 7–10. Krishnamurti, Jiddu, Beginnings of Learning, Worthing: Littlehampton Book Services, 1975. Kutter, G. Siegfried, Universe and Life: Origins and Evolution, Boston: Jones & Bartlett, 1987. ———, ‘Big History: A Personal Perspective’, in Evolution: A Big History Perspective, eds. Leonid Grinin, Andrey Korotayev and Barry Rodrigue, Volgograd: Uchitel Publishing, 2011, pp. 101–120. Lezra, Jacques and Liza Blake, eds., Lucretius and Modernity: Epicurean Encounters Across Time and Disciplines, New York: Palgrave Macmillan, 2016. Lévi-Strauss, Claude, The Savage Mind, Paris: Librairie Plon, 1962 [French publication]. Linnaeus, Carl, Systema Naturae, Leiden: Lugduni Batavorum, 1735 [Latin Publication]. Liu, Mingxin, ‘A Historical Overview on Anthropology in China’, Anthropologist, vol. 5, no. 4, 2003, pp. 217–223, http://krepublishers.com/02-Journals/T-Anth/Anth-050-000-000-2003-Web/Anth-05-4-217-303-2003-Abst-PDF/Anth-05-4-217-2232003-143-Liu-M/Anth-05-4-217-223-2003-143-Liu-M-Text.pdf, accessed 13 April 2017.

148

Big history and anthropology

Lucretius, Titus (Carus), De Rerum Natura (On the Nature of Things), c. 55 bce [Latin Publication]. MacGregor, Niel, A History of the World in 100 Objects, London: Allen Lane, 2010. Mair, Lucy, Primtive Government, Harmondsworth: Penguin Books, 1962. Majid, Afshan, Badauni and Mughal Society and Culture as Reflected in his Works, Ph.D. Thesis, Department of History, Centre of Advanced Study, Aligarh Muslim University, 2015, pp. 84–103. Malinowski, Bronislaw, Argonauts of the Western Pacific: An Account of Native Enterprise and Adventure in the Archipelagoes of Melanesian New Guinea, London: G. Routledge & Sons, 1922. Margulis, Lynn and Dorion Sagan, Microcosmos: Four Billion Years of Evolution from our Microbial Ancestors, New York: Summit Books, 1986. Marx, Karl, Capital: Critique of Political Economy, vol. 1, Hamburg: Otto Meissner, 1867 [German publication]. Mauss, Marcel, ‘Essai sur le don: Forme et raison de l’échange dans les sociétés archaïques’, L’Année Sociologique, vol. 2, no. 1, 1925, pp. 30–126. ———, ‘Les techniques du corps’, Journal de Psychologie, vol. 32, nos. 3–4, 15 March 1936, pp. 271–293. Maxim, Hiram Percy, Life’s Place in the Cosmos, New York: D. Appleton, 1933. McBean, Gordon and Alberto Martinelli, ‘Blurring Disciplinary Boundaries’, Science, vol. 358, no. 6366, 24 November 2017, p. 975. McClaurin, Irma, ‘Zora Neale Hurston: Enigma, Heterodox, and Progenitor of Black Studies’, Fire!!!: The Multimedia Journal of Black Studies, vol. 1, no. 1, March 2012, pp. 49–67. McGee, R. Jon and Richard Warms, Anthropology Theory: An Introductory History, Lanham, MD: Rowman & Littlefield, 2017. McKay, Alex, ‘The Drowning of Lama Sengchen Kyabying: A Preliminary Enquiry from British Sources’, Tibet, Past and Present, eds. Henk Blezer and Abel Zadoks, Leiden: Brill, 2002, pp. 263–279. Merryman, John, ‘Thinking about the Elgin Marbles’, Michigan Law Review, vol. 83, no. 8, August 1985, pp. 1880–1923. Michalowski, Jennifer, ‘Proteins of the Past: Reconstructing Tiny Pieces of History Deepens Understanding of Evolution’, Science News Magazine, vol. 189, no. 12, 11 June 2016, pp. 16–20. Miller, Peter and François Louis, Antiquarianism and Intellectual Life in Europe and China, 1500–1800: A Comparative Consideration of the Fascination with Antiquity in European and Chinese Intellectual History, Ann Arbor: University of Michigan Press, 2012. Mire, Sada, ‘Somalia: Studying the Past to Create a Future’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 279–288. Mithen, Steven, Problem-Solving and the Evolution of Human Culture, London: Institute for Cultural Research, 1999. Morgan, Lewis Henry, Ancient Society; Or, Researches in the Lines of Human Progress from Savagery, Through Barbarism to Civilization, New York: Henry Holt and Company, 1877. Morse, Edward, ‘Shell Mounds of Omori’, Memoirs of the Science Department, University of Tokio, Japan, Volume I, Part I, Tokyo: University of Tokyo, 1879. Murray, Tim, Milestones in Archaeology: A Chronological Encyclopedia, Santa Barbara: ABCClio, 2007.

149

Barry H. Rodrigue

Næss, Arne, ‘The Shallow and the Deep, Long-Range Ecology Movements: A Summary’, Inquiry, vol. 16, nos. 1–4, 1973, pp. 95–100. Nakanishi, Osamu, ed., Applying Big History: Nature, War and Peace, Yokohama: Institute for Global and Cosmic Peace, 2016 [ Japanese, Chinese and English publication]. ———, Universal Studies and the Modern World: Becoming Global and Cosmic Humanity, ­Yokohama: Institute for Global and Cosmic Peace, 2017 [ Japanese, Chinese, Russian, and English publication]. Nakanishi, Osamu and Nobuo Tsujimura, ‘Universal Studies and Big History in Japan’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2016, pp. 289–294. Nazaretyan, Akop and Akop Nazaretyan, ‘Technology and Psychology: On the Concept of Evolutionary Crises’, Social Sciences Today, no. 3, 1993 [Russian publication]. ———, ‘Fear of the Dead as a Factor in Social Self-Organization’, Journal for the Theory of Social Behaviour, vol. 35, no. 2, June 2005, pp. 155–169. ———, Civilization Crises in a Universal History Context: Self-Organization, Psychology and Forecasting, Moscow: Mir, 2004 [Russian publication]. ———, The Anthropology of Violence and Culture in Self-Organization: Essays on Evolutionary ­Historical Psychology, Moscow: LKI, 2008 [Russian publication]. ———, Evolution of Non-Violence: Studies in Big History, Self-Organization and Historical Psychology, Saarbrucken: Lambert Academic Publishing, 2010. ———, ‘Life’s Meaning as a Global Problem of Modernity: A View from a Big History and Complexity-Studies Perspective’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 317–338. ———, Non-Linear Future: Mega-History, Complexity Theory, Anthropology and Psychology for Global Forecasting, Moscow: Agarmak Media, 2017 [Russian publication]. Nicolaisen, Wilhelm, ‘“Distorted Function” in Material Aspects of Culture’, Folklore ­Forum, vol. 12, nos. 2–3, 1979, pp. 223–235. Nicolas, Georges, ‘Humanisme – Cosmisme’, Cahiers de Géographie du Québec, vol. 33, no. 90, December 1989, pp. 379–385. Neile, Frank, Energy: Engine of Evolution, Amsterdam: Elsevier, Shell Global Solutions, 2005. ———, ‘The Next Energy Revolution: Evolutionary Energetics, Models and Scenarios’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 197–250. Ó Cadhla, Stiofán, Civilizing Ireland: Ordnance Survey 1824–1842: Ethnography, Cartography, Translation, Dublin: Irish Academic Press, 2007. Ollongren, Alexander, Astrolinguistics: Design of a Linguistic System for Interstellar Communication Based on Logic, Berlin: Springer, 2012. Pandey, Shakunt, ‘The Indian Museum Completes 200 Years’, Science Reporter, vol. 51, no. 10, October 2014, pp. 38–41. Panov, Alexander, ‘Singularity of Evolution and Post-Singular Development’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, New Delhi: Primus Books, 2017, pp. 370–402. Patton, Paul, ‘One World, Many Minds’, Scientific American Mind, December 2008, pp. 72–79.

150

Big history and anthropology

Pignocchi, Gaia et al., I Petroglifi Preistorici del Monte Cònero: Progetto di Studio e Valorizzazione, Ancona: Submitted to the administration of Conero Park, 8 May 2007. Pinker, Steven, The Better Angels of Our Nature: Why Violence Has Declined, New York: Viking Books, 2011. Piovesana, Gino, ‘Miura Baien, 1723–1789, and His Dialectic and Political Ideas’, ­Monumenta Nipponica, vol. 20, nos. 3–4, 1965, pp. 389–421. Price, David, ‘Lessons from Second World War Anthropology: Peripheral, Persuasive, and Ignored Contributions’, Anthropology Today, vol. 18, no. 3, June 2002, pp. 14–20. Pringle, Heather, ‘Vikings and Native Americans’, National Geographic, November 2012, pp. 80–89. Puhalo, Lazar, ‘The Rise of Personhood: Development of Social Justice and Natural Rights in the Axial Eras’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2016, pp. 118–123. Qi Tao, ‘Echoes of the Palaeolithic: A Research Note on the Great Floods and the Origins of Chinese Civilization,’ in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 84–94. Qian, Xuesen, Jingyuan Yu and Ruwei Dai, ‘A New Discipline of Science – The Study of Open Complex Giant System and Its Methodology’, Chinese Journal of Nature, vol. 1, 1990, pp. 3–10 [Chinese publication]. Quaternary International, vol. 361, The Origins of Recycling: A Paleolithic Perspective, March 2015. Ralphs, Matthew et al., ‘Water Extraction on Mars for an Expanding Human Colony’, Life Sciences in Space Research, vol. 7, November 2015, pp. 57–60. Reining, Conrad, ‘A Lost Period of Applied Anthropology’, American Anthropologist, vol. 64, no. 3, Part 1, June 1962, pp. 593–600. Rezavi, Sayed Ali Nadeem, ‘Anitiquarian Interests in Medieval India: The Relocation of Ashokan Pillars by Firuzshah Tughluq’, Proceedings of the Indian History Congress 70th Session, 2010, pp. 994–1010. Rodrigue, Barry, ‘A Big History Directory’, 2011; www.rodrigue-global.org/universalstudies--archives.html, accessed 15 April 2017. ———, ‘Big History: A Working Bibliography of References, Films and Internet Sites’, 2011, www.rodrigue-global.org/universal-studies--archives.html, accessed 15 April 2017. ———, ‘Big History – A Study of All Existence’, www.rodrigue-global.org/universal-­ studies--archives.html, accessed 15 April 2017. ———, ‘Retrofitting the Future’, in Teaching and Researching Big History: Exploring a New Scholarly Field, eds. Leonid Grinin et al. Volgograd: Uchitel Publishing, 2014, pp. 276–282. ———, ‘A New Design for Living’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume I, Our Place in the Universe: An Introduction to Big History, eds. Barry ­Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2015, pp. 183–187. ———, ‘An Emergent Future: Evolving A Global Revolution’, in Universal Studies and the Modern World, Yokohama: Institute of Global and Cosmic Peace, 2017, pp. 196–218. ———, ‘The Study of all Existence: Big History, Universal Studies and the Global Conjuncture’, International Journal for the Transformation of Consciousness, vol. 3, no. 1, June 2017, pp. 15–34.

151

Barry H. Rodrigue

Rodrigue, Barry, Leonid Grinin and Andrey Korotayev, From the Big Bang to Galactic Civilizations: A Big History Anthology, Volume I, Our Place in the Universe: An Introduction to Big History, Delhi: Primus Books, 2015. Roullier, Caroline et al., ‘Historical Collections Reveal Patterns of Diffusion of Sweet Potato in Oceania Obscured by Modern Plant Movements and Recombination’, Proceedings of the National Academy of Sciences of the United States of America, vol. 110, no. 6, 5 February 2013, pp. 2205–2210. Rowe, John, ‘The Renaissance Foundations of Anthropology’, American Anthropologist, vol. 67, no. 1, February 1965, pp. 1–20. Roy, Sarat Chandra, ‘An Indian Outlook on Anthropology’, Man ( Journal of the Royal Anthropological Institute), vol. 38, no. 9, September 1938, pp. 146–150. Russell, Dale, An Odyssey in Time: The Dinosaurs of North America, Toronto: University of Toronto Press and the Canadian National Museum of Natural Sciences, 1993. ———, ‘Speculations on the Evolution of Intelligence in Multicellular Organisms’, in Life In The Universe, ed. John Billingham, Proceedings of a Conference Held at the National Ames Research Center, Moffet Field, California, June 19–20, 1979 (CP-2156), National Aeronautics and Space Administration, Scientific and Technical Information Branch, 1981, pp. 259–275. Scott, Walter, The Antiquary, Edinburgh: James Ballantyne & Company, 1816. Seaver, Kirsten, The Frozen Echo: Greenland and the Exploration of North America, ca. A.D. 1000–1500, Stanford: Stanford University Press, 1996. Sengupta, Indra, ‘A Conservation Code for the Colony: John Marshall’s Conservation Manual and Monument Preservation Between India and Europe’, in ‘Archaeologizing’ Heritage?: Transcultural Entanglements between Local Social Practices and Global Virtual Realities, eds. Michael Falser and Monica Juneja, Berlin: Springer-Verlag, pp. 21–37. Sera-Shriar, Efram, The Making of British Anthropology, 1813–1871, London: Routledge, 2013, 2015. Seymour, Susan, Cora DuBois: Anthropologist, Diplomat, Agent, Lincoln: University of ­Nebraska Press, 2015. Shea, Christopher, ‘Alas, Poor William Shakespeare: Where Does His Skull Rest?, New York Times, 24 March 2016. Shklovsky, Joseph, Universe, Life, Intelligence, Moscow: USSR Academy of Sciences, 1962 [Russian publication]. Shklovsky, Joseph and Carl Sagan, Intelligent Life in the Universe, New York: Random House, 1966. Singh, Vandana, ‘A Speculative Manifesto’, in The Woman Who Thought She Was a Planet and Other Stories, New Delhi: Zubaan, 2013, p. 200–203. Slocum, Sally, ‘Woman the Gatherer: Male Bias in Anthropology’, in Toward an Anthropology of Women, ed. Rayna Reiter, New York: Monthly Review, 1975, pp. 36–50. Smail, Daniel, On Deep History and the Brain, Berkeley: University of California Press, 2008. Smith, Michael and Peter Peregrine, ‘Approaches to Comparative Analysis in Archaeology’, The Comparative Archaeology of Complex Societies, ed. Michael Smith, Cambridge: Cambridge University Press, 2012. Snooks, Graeme, The Dynamic Society: Exploring the Sources of Global Change, London: Routledge, 1996. Spencer, Herbert, First Principles of a New System of Philosophy, London: Williams and ­Norgate, 1862. Soffer, Reba, ‘The Development of Disciplines in the Modern English University’, The Historical Journal, vol. 31, no. 4, December 1988, pp. 933–946. 152

Big history and anthropology

Spencer, Frank, ed., History of Physical Anthropology, Volume 1, New York: Garland, 1997. Spier, Fred, ‘Big History Research: A First Outline’, in Evolution: A Big History Perspective, Volgograd: Uchitel Publishing, 2011, pp. 26–36. ———, Big History and the Future of Humanity, Oxford: Wiley-Blackwell, 2010; new edition, 2015. ———, ‘Roads towards Big History’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Ways that Big History Works: Cosmos, Life, Society and Our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Publishing, 2017, pp. 253–260. Spiess, Arthur, ‘Wild Maine and the Rusticating Scientist: A History of Anthropological Archaeology in Maine’, Man in the Northeast, vol. 30, Fall 1985, pp. 101–129 (pp. 104–106). Stocking Jr., George, Victorian Anthropology, New York: The Free Press, 1987. Strathern, Marilyn, ‘An Awkward Relationship: The Case of Feminism and Anthropology’, Signs: Journal of Women in Culture and Society, vol. 12, no. 2, Winter 1987, pp. 276–292. Sundar, Nandini, ‘In the Cause of Anthropology: The Life and Work of Irawati Karve’, in Anthropology in the East: Founders of Indian Sociology and Anthropology, eds. Patricia Uberoi, Nadini Sundar and Satish Deshpande, Ranikhet: Permanent Black, 2007. pp. 360–416. Thompson, William Irwin, The Imagination of an Insurrection, Dublin, Easter, 1916: A Study of an Ideological Movement, Oxford: Oxford University Press, 1967. Timreck, Theodore, dir., Franz Boas, 1858–1942, Public Broadcasting Associates, 1980. Tooker, Elisabeth, ‘A Note on Undergraduate Courses in the Latter Part of the Nineteenth Century’, Man in the Northeast, vol. 39, Spring 1990, pp. 45–51. Trigger, Bruce, A History of Archaeological Thought, Cambridge: Cambridge University Press, 1990. Tsujimura, Nobuo and Hirofumi Katayama, ‘Think Cosmically, Act Globally: Emerging Clues for the Big History Movements’, International Journal for the Transformation of Consciousness, vol. 3, no. 1 June 2017, pp. 45–71. Tylor, Edward, Primitive Culture: Researches into the Development of Mythology, Philosophy, Religion, Art, and Custom, Volume I, London: John Murray, 1871. Uberoi, Patricia, Nadini Sundar and Satish Deshpande, Anthropology in the East: Founders of Indian Sociology and Anthropology, Ranikhet: Permanent Black, 2007. United Kingdom, Department for Culture, Media and Sport, Portable Antiquities Scheme, Tisbury Hoard, WILT-E8DA70, https://finds.org.uk/database/artefacts/record/id/467433, accessed 23 April 2017. United Nations Educational, Scientific and Cultural Organization, World Heritage Centre, List: http://whc.unesco.org/en/list/, accessed 23 April 2017. ———, History of Humanity, seven volumes, Paris: United Nations Educational, Scientific and Cultural Organization, 1966, 2009. ———, International Commission for a History of the Scientific and Cultural Development of Mankind, http://atom.archives.unesco.org/international-commission-for-history-ofscientific-and-cultural-development-of-mankind-2;isad, accessed 4 December 2014. University of Arizona, Lunar and Planetary Laboratory, Spacewatch, http://spacewatch.lpl. arizona.edu/, accessed 19 May 2018. United States, National Aeronautics and Space Administration, ‘The Pioneer Missions’, 26 March 2007, www.nasa.gov/centers/ames/missions/archive/pioneer.html, accessed 20 May 2018. ———, ‘Hubble Reveals Observable Universe Contains 10 Times More Galaxies than Previously Thought’, 13 October 2016, http://hubblesite.org/news_release/news/2016-39. 153

Barry H. Rodrigue

———, ‘Voyager’, https://voyager.jpl.nasa.gov/, accessed 20 May 2018. University of Cambridge, Research, ‘The Tibetan Lama who Wrote a World Geography’, 14 June 2014, www.cam.ac.uk/research/features/the-tibetan-lama-who-wrote-aworld-geography, accessed 12 June 2018. Urry, James, Before Social Anthropology: Essays on the History of British Anthropology, London: Routledge, 1993. Vaquero, Manuel et al., ‘Temporal Nature and Recycling of Upper Paleolithic Artifacts: The Burned Tools from the Molí del Salt Site (Vimbodí i Poblet, northeastern Spain)’, Journal of Archaeological Science, vol. 39, no. 8, August 2012, pp. 2785–2796. Vakoch, Douglas, Archaeology, Anthropology and Interstellar Communication, Washington, DC: National Aeronautics and Space Administration, 2014. Vélez, Antonio, Man: Inheritance and Conduct, Medellín: Editorial Universidad de Antioquia, 1986 [Spanish publication]. Vico, Giambattista, Principi di Scienza Nuova d’intorno alla Comune Natura delle Nazioni, (Principles/Origins of New/Renewed Science About/Surrounding the Common Nature of Nations), Naples: Giambattista Vico, 1725. Vidyarthi, Lalita Prasad, Rise of Anthropology in India: A Social Science Orientation, Volume 1, Delhi: Concept Publishing Company, 1976. Vinge, Vernor, A Fire Upon the Deep, New York: Tor Books, 1992. ———, ‘The Coming Technological Singularity: How to Survive in the Post-Human Era’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2017, pp. 357–369. Voros, Joseph, ‘Big Futures: Macro-Historical Perspectives on the Future of Humankind’, in From Big Bang to Galactic Civilizations: A Big History Anthology, Volume III, The Way that Big History Works: Cosmos, Life, Society and our Future, New Delhi: Primus Books, 2017, pp. 403–436. Wallace, Alfred Russel, ‘Peschel’s “Races of Man”’, Nature, vol. 15, 28 December 1876, pp. 174–176. ———, Man’s Place in the Universe: A Study of the Results of Scientific Research in Relation to the Unity or Plurality of Worlds, New York: McClure, Phillips & Company, 1903. Wall, Mike, ‘Signs of Alien Life will be Found by 2025, NASA’s Chief Scientist Predicts’, Space.com, 7 April 2015, www.space.com/29041alien-life-evidence-by-2025-nasa.html, accessed 19 April 2015. Wallerstein, Immanuel, The Politics of the World Economy: The States, the Movements and the Civilizations, Cambridge: Cambridge University Press, 1984. ———, Unthinking Social Science: The Limits of Nineteenth-Century Paradigms, Cambridge: Polity Press, 1991. Weber, Bruce and David Depew, Evolution and Learning: The Baldwin Effect Reconsidered, Cambridge: MIT Press, 2003. White, Leslie, ‘Energy and the Evolution of Culture’, American Anthropologist, vol. 45, no. 3, Part 1, July 1943, pp. 335–356. Wundt, Wilhelm, Völkerpsychologie: Eine Untersuchung der Entwicklungsgesetze von Sprache, Mythus und Sitte; Mythus und Religion, vol. 2, part 1, Leipzig: Wilhelm Engelmann, 1905. Wylie, Turrell, ‘Dating the Tibetan Geography “Dzam Gling Rgyas Bshad” through its Description of the Western Hemisphere, Central Asiatic Journal, vol. 4, no. 4, 1959, pp. 300–311. Xu, Zhenoao, David Pankenier and Yaotiao Jiang, East-Asian Archaeoastronomy: Historical Records of Astronomical Observations of China, Japan and Korea, Boca Raton: CRC Press, 2000, pp. 13–24. 154

Big history and anthropology

Yamashita, Shinji, Joseph Bosco and Jeremy Seymour Eades, The Making of Anthropology in East and Southeast Asia, New York: Berghahn Books, 2004. Yongdan, Lobsang, ‘Tibet Charts the World: The Btsan po No mon han’s Detailed Description of the World, An Early Major Scientific Work in Tibet’, Mapping the Modern in Tibet, ed. Gary Tuttle, 2011, pp 73–134. Yotov, Yoto, Roberta Piermartini, José-Antonio Monteiro and Mario Larch, An Advanced Guide to Trade Policy Analysis: The Structural Gravity Model, Geneva: United Nations and World Trade Organization, 2016. Zhenoao Xu, W. Pankenier, Yaotiao Jiang, East-Asian Archaeoastronomy: Historical Records of Astronomical Observations of China, Japan and Korea, Boca Raton: CRC Press, 2000. Zhu Weibin, ‘Big History and World History in China’s Colleges and Universities’, in From the Big Bang to Galactic Civilizations: A Big History Anthology, Volume II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev, Delhi: Primus Books, 2016, p. 319.

155

6 BIG HISTORY AND ARCHAEOLOGY Archaeology is big history Brian Fagan Stuart Piggott, an eminent British prehistorian of the 1950s, once described archaeology as the science of garbage. He was, of course, perfectly correct—up to a point. We archaeologists do indeed spend much of our time dissecting the discards of ancient human behavior, sometimes almost to the point of trivial obsession. We’re masters of arcane scientific methods that can date a single wheat grain, trace the life histories of Bronze Age archers, and conjure up images of the long-vanished hinterland of Angkor Wat in Cambodia using technology from space. All this seemingly miraculous archaeological detective work may dazzle the casual onlooker and has indeed revolutionized our understanding of early humanity. But archaeology is far more than trash heaps and buried cities. It is the only historical discipline that is completely multidisciplinary and that encompasses the humanities, the sciences, and social sciences. Even more important, it is unique in that it studies changing human behavior over immensely long periods of time, through three million years—the human past in its entirety (Kelly, 2016). Archaeology is big history and always has been, concerned with long- and short-term biological and cultural evolution, with emerging human diversity, and with human history on both a global and local scale. Archaeology has an anachronistic popular image. It suffers from Hollywood and lingering images of absent-minded professors digging in arid, remote lands, in the shadow of gold laden pyramids. It is still often considered a romantic pastime, despite generations of serious television documentaries and National Geographic Magazine articles. This misleading legacy stems in part from the spectacular discoveries of ancient civilizations by nineteenth century archaeologists, many of them little more than adventurers. Those days, there were not many excavators around. A century ago, archaeology was largely an amateur pursuit, often the realm of wealthy amateurs. Europe and the Mediterranean world were the major archaeological playgrounds. ­Archaeology in the Americas was also mostly an amateur pastime. The achievements of the pioneers were remarkable, culminating in the spectacular discovery of the tomb ­ arter in of the Egyptian pharaoh Tutankhamun by Lord Carnarvon and Howard C 1922, and with Leonard Woolley’s dissection of the Royal Burials at Ur in Iraq a decade later. These discoveries and the unearthing of the virtually unknown Sumerian, 156

Big History and archaeology

Hittite, and Minoan civilization were the work of archaeological giants, who worked with tiny budgets and often small armies of local workers. Small wonder a mosaic of these men formed the prototype for Hollywood’s Indiana Jones. Today’s archaeology owes much to the pioneers, but today’s researchers spend as much time in the laboratory as they do in the field.They often converse with one another in arcane scientific dialects that are unintelligible to non-specialists. Archaeologists also have an enduring popular image as time detectives, who conjure miracles from the tiniest of artifacts. This is something we have achieved with early hominins, some of whom we now know were left-handed, because of the direction of the flake scars on their artifacts. State-of-the-art medical sciences and meticulous analysis of seed remains have tracked down the home valley of Otzi, the Bronze Age Ice Man from the Alps, who died high in the mountains in 2300 BC, soon after the Pyramids rose by the Nile. The British humorist P.G. Wodehouse (1919: 38) once remarked of spectators at a London building site that “a mere hole in the ground is enough to grip their attention for hours at a time.” One can say the same of archaeologists, many of whom spend inordinate amounts of time peering into, and excavating, small trenches. Over the past half century, archaeology has become a highly specialized field of research that has developed its own forms of ardent academic provincialism. Herein lies the fundamental quality of archaeology. By its very nature, it is concerned with the minutiae of human behavior recovered by meticulous excavation from tiny hunting camps, from individual rooms of an ancient city, or from farming villages, temples, or palaces. Such focused research encourages narrow perspectives on the past, reinforced by the proliferation of high technology in archaeology since the 1950s. Today’s academic culture also fosters narrow perspectives by its insistence on peer-reviewed publication and a proliferation of academic journals. What is vernacularly known as “publish or perish,” combined with increasingly specialized training and research, militates powerfully against archaeologists becoming involved in big history in any numbers. This is a tragedy, but fortunately change is afoot. Histories of archaeology conventionally refer to the postwar years as the time when archaeology came of age (Fagan and Durrani, 2016).The major scientific catalyst was ­ ramatic fossil radiocarbon dating, which became commonplace during the 1960s. D discoveries, notably by Louis and Mary Leakey in East Africa, coincided with the development of potassium argon methods that provided the first relatively reliably chronology for the early chapters of human biological and cultural evolution. For the first time, we thought of human evolution in terms not of hundreds of thousands of years, but in millions. A huge chasm of human history appeared before us, which was big history on a hitherto unimaginable scale. For those of us engaged in archaeology in the early 1960s, the realization that there was big history out there was a truly liberating thought, especially for those who were excavating outside the comfortable frontiers of Europe and the Middle East and the more thoroughly trodden parts of the Americas. There were, of course, archaeologists who thought in broad terms long before radiocarbon dating. The European prehistorian Vere Gordon Childe was one of them. He wrote widely read syntheses of Near Eastern and European prehistory, which were pioneering studies of major issues like the origins of agriculture and the 157

Brian Fagan

beginnings and spread of the earliest civilizations (Childe, 1942). His Neolithic and Urban Revolutions became part of the historical canon and, or all their simplicity, are still widely used labels today. Childe was, ultimately, a Europeanist, whose perspective tended to end in the Nile Valley and Mesopotamia. Childe’s later career coincided with a significant expansion of archaeological research in far-flung places, among them Australia, New Zealand, Southeast Asia, the Pacific, and tropical Africa. Within a generation, archaeology changed from being a village where most people knew one another to a truly international enterprise conducted in every corner of the world. The diaspora of young archaeologists came about thanks to increased research funding in the United States, and from the untiring efforts of the Cambridge archaeologist Grahame Clark, who was one of the pioneers of the study of adaptations to environmental change by prehistoric societies (Fagan, 2001). He also encouraged his students (this author was one of them) to take up archaeological posts abroad, far from the familiar stamping grounds of Europe, Central America, and the Mediterranean. An explosion in the number of archaeologists world-wide has continued until today, triggered both by the expansion of academic archaeology, and by the legal requirements of cultural resource management. Just to give one example, there were about a dozen professional archaeologists in tropical ­Africa between the Sudan and the Cape of Good Hope in the early 1960s. Today, there are over a hundred in South Africa alone. The growing number of researchers, many of them based in Europe and North America, has led to an explosion of research in hitherto virtually unknown areas like Southeast Asia and the Amazon Basin. World prehistory was first taught at Cambridge University soon after World War II, when Dorothy Garrod was the Disney Professor of Archaeology. She had excavated Mt. Carmel in what is now Israel, where she found Neanderthals. She was one of the first, like Louis Leakey, to think of prehistory on global terms. Grahame Clark built on these foundations. He was a magisterial scholar, who firmly sent his student aboard. He was also an inveterate traveler, whose peregrinations in search of the past extended as far afield as Australia, India, and China. These travels and his broad perspective led him to write World Prehistory, the first such work that was a truly global, if incomplete, synthesis (Clark, 1961). But Clark also pushed archaeology squarely onto the stage of big history. It has been there ever since, largely because it is the only way of understanding human history without regard to national or civilizational boundaries, and without chronological limitations. A historian ponders centuries, even years, months, weeks, and years, even days. Archaeologists deal in millennia as well as centuries and shorter periods of time. World Prehistory set the stage at a time when archeologists were widening the scope of their research in dramatic days and adopting more sophisticated theoretical perspectives that relied significantly on other disciplines ranging from anthropology, biology, and ecology, to climatology, geology, statistical analysis, and zoology—to mention only a few. A tumultuous theoretical ferment unfolded between the 1960s and 1990s and continues somewhat abated to this day ( Johnson, 2010;Trigger, 2006).The debates unfolded as mainly younger archaeologists rebelled against mindless artifact classification and rigid chronological sequences. Their interests turned from description to analysis, and, above all, explanation.Why did early hominins come down from the trees? What were the motives behind Late Ice Age art? Why did people in the 158

Big History and archaeology

Middle East (and elsewhere) take up agriculture and animal husbandry immediately after the Ice Age? And why and how did the world’s pre-industrial civilizations come into being? By the late 1970s, world prehistory was taught in numerous colleges and universities. Archaeology, for all its ardent and highly technical specialization, had developed the methodology and theoretical sophistication to engage seriously, and, when appropriate, in meticulous detail, with the basic issues of big history.

Big questions for archaeologists Archaeology has morphed over recent generations into a team approach to studying history, albeit with artifacts and sites, animal bones, seeds, and environmental data as the archives, the tools if you will, for examining fundamental issues of early history. We contribute, of course, to more nuanced understanding of historical societies, but it is with the fundamental issues of the remote past that we are the major players. New, broad-based approaches to the study of very ancient societies now move far beyond stone hand axes or painted pottery. How, for example, do archaeologists study ancient biological and cultural diversity and reconstruct ancient environments? What do molecular biology, especially DNA and isotope analysis, bring to big history? Some of the most spectacular finds of recent years are little known, except to specialists. For instance, thanks to bone isotope analysis, we know that a Bronze Age archer buried near Stonehenge in southern Britain in about 2300 BC spent his childhood in Central Europe (Fitzpatrick, 2011). Underwater excavations on a shipwreck of the fourteenth century BC at Uluburun in southern Turkey recovered a rich cargo that included enough copper and tin to equip an entire regiment, and artifacts from nine different regions of the eastern Mediterranean world (Bass et al., 1989). These, and other cutting edge projects using scientific technology and astute excavation and observation, are gradually providing answers to the big questions that confront archaeologists on the global stage.

How, when, and where did we originate? Large numbers of people, especially in the United States, still believe that God created humanity as described in Genesis, Chapter 1. They also believe that Biblical chronologies date this event to 4000 BC. Such beliefs are a matter of religious faith, for science moved beyond such teachings in the mid-nineteenth century.  There is no question that Charles Darwin was correct when he identified tropical Africa as the cradle of humankind back in 1871. Until the 1960s, the chronology was measured in hundreds of thousands of years. Today, multidisciplinary paleoanthropology has traced early Homo back to at least two million years. If one defines a human as a toolmaking hominin, then recent research goes back over three million (Stringer and Andrews, 2002). Back in the 1960s, human evolution was almost ladder-like in its simplicity, from the Australopithecines and Homo habilis, “handy person” through Homo erectus, then the Neanderthals and modern humans. Today, we know that a considerable diversity of early hominins flourished in eastern Africa, and probably elsewhere, before two million years ago. Paleoanthropologists now tend to talk of a generic “early Homo” as 159

Brian Fagan

our remotest ancestor, probably a wise strategy at this stage in research. Quite apart from debates over chronology, hominin anatomy, and toolmaking, there is a proliferating literature on what humanness entailed. For generations, it has been defined by the uniquely human possession of culture, upright posture, and toolmaking—the classic doctrine of “coming down from the trees.” The first, and one of the most fundamental, questions facing archaeologists is that of human origins. We are certain that we originated in tropical Africa, that this is where the first toolmaking, upright standing hominins evolved. But the how, when, and where of human beginnings are still tantalizingly unresolved questions. There is, however, one certainty. From the beginning, there was a far greater diversity of early Homo than we once suspected. Biological and cultural diversity have been part of being human from the beginning.

When and why did we move out of Africa and what happened next? Two of the catalytic developments of early history coincided with what were inconspicuous moves of tiny human groups out of tropical Africa. Thanks to potassium argon dating, we now know that our archaic forebears left tropical Africa and settled in East and Southeast Asia, also in Europe sometime after two million years ago (Klein, 2009). These movements, which remain largely a mystery, were the result of the ­natural dynamics of hunter-gatherer life—the endless pursuit of different prey, migrations of herd animals, the seasons of scattered plant foods, and so on. Perhaps the greatest innovations were technological—the taming of fire, the development of stone-tipped spears and of multipurpose tools, but behind everything lay the imperatives of survival, of finding sustenance. The archaeological signature, the archive if you will left by archaic humans on the move is a matter of shreds and patches—scatters of stone tools, occasional, and rare, kill sites, like the 400,000-year-old camps at Schoningen in northern Germany with their long wooden spears. Hundreds of thousands of years of history between two million years ago and 200,000 thousand years before present are an enormous blank in our past. Generations of archaeologists have filled in some of the gaps, but these are mere snapshots of a sparsely inhabited archaic world that extended from western Europe into East and Southeast Asia. Australia and the Americas were uninhabited, as, most likely, was most of northern Eurasia. Generically, we call these scattered hunter-gatherers Homo erectus, but this broad term disguises complex, and almost unknown, processes of biological and cultural evolution, the slow modernization of humanity. But even as early as two million years ago, we can be confident that humanity was diversifying significantly. Future archaeologists will fill in many of the gaps from what is, at best, a highly fragmented and inconspicuous signature of human activities in a world inhabited by only a few million people. With the emergence of the Neanderthals over 300,000 years ago, the early chapters of the human past come into closer focus. Neanderthals cast a powerful spell over us (Papagianni and Morse, 2015).They have been portrayed as vicious, primitive brutes, as true primitives, but we now know that they were skilled, tough hunters and foragers, who mastered a remarkable range of natural environments, even open 160

Big History and archaeology

Ice Age steppe. Squat and powerfully built European Neanderthals adapted brilliantly to the extreme cold of the last Ice Age glaciation. There was considerable diversity among Neanderthal populations by 70,000 years ago, still poorly documented, especially the mysterious Denisovans of the Altai in Eurasia. Genetics are rapidly transforming our perceptions of the remote past, including convincing evidence that Neanderthals interbred with fully modern humans in several locations. But, in the final analysis, they were an evolutionary dead-end and were certainly not the direct ancestors of Homo sapiens sapiens, as earlier scholars believed.

Where did we moderns originate and what happened next? We are Homo sapiens, the clever and wise people, animals capable of subtlety, of manipulation, of self-understanding. We have fully articulate speech, design, and fabricate an astonishing variety of tools. We communicate, we make plans and innovate, pass on knowledge and ideas using an armory of skills that are unique to anatomically modern people. Everyone in today’s complex, biologically and cultural diverse world shares the same advanced cognitive skills that developed among Homo sapiens populations tens of thousands of years ago. One of the questions in big history revolves around the origins of modern humans: Where and when did our remotest direct ancestors evolve (Klein, 2009)? Intense controversy surrounds the origins of modern humans. Almost all biological anthropologists and archaeologists agree that modern humans developed in sub-Saharan Africa. The evidence comes predominantly from the rapidly developing science of molecular biology, which has long studied mitochondrial DNA (mtDNA), inherited through the female line. The research is now extending to other chromosomes, in attempts to unravel complex biological processes. There has long been talk of a theoretical “African Eve,” a primordial modern human woman who was ancestor of us all. This is, of course, a grossly simplistic scenario for what is obviously an intricate tapestry of human biological and cultural evolution that unfolded among isolated African hunting bands. The earliest known modern human fossils come from Herto in Ethiopia and date to about 150,000 years ago, which, in broad terms, confirms the genetic chronology for the appearance of modern humans around 150,000 to 200,000 years ago. Some of these very early humans may have spread, at least tentatively, out of Africa, perhaps into the Middle East and South Asia, perhaps as early as 100,000 years ago but we know almost nothing of these tiny population movements. The real diaspora out of African happened much later, perhaps around 70,000 years ago or even later, by which time Homo sapiens had developed the full cognitive abilities enjoyed by modern humans. Tracking down the appearance of these intangible skills is painstakingly difficult, reflected as it is in minor changes in stone toolkits detected in South African rock shelters. The great diaspora of anatomically modern humans that led ultimately to the settlement of Asia, Europe, the Pacific, and the Americas was one of the momentous developments of history, chronicled by genetics and archaeology. Only the bare outlines of these usually small scale population movements are known. But 161

Brian Fagan

there was growing biological and cultural diversity from the beginning, some of it adaptations to local environments, the others responses to challenges like extreme cold or heat (Fagan and Durrani, 2018). Quite when moderns settled, or even perhaps evolved, in South and East Asia remains uncertain, but they were well established in the southeastern mainland by at least 50,000 years ago. By 45,000 BP, Homo sapiens had settled New Guinea and Australia—the dates could be somewhat earlier, the chronology is not yet well established. There were hunter and fishers on the islands of the Bismarck Archipelago of the southwestern Pacific by 30,000. The greatest controversy of all surrounds the first settlement of the Americas, but this is now widely accepted to have taken place across the Bering Land Bridge between Siberia and Alaska during warming immediately after the Ice Age, perhaps by 15,000 BP. Almost all Americanists agree that the first settlers moved southward along the Pacific coast, into an uninhabited continent. Some of them were as far south as the Peruvian coast and northern Chile by 14,000 (Adovasio and Pedler, 2016; Meltzer, 2008). The late Ice Age, especially after 45,000 years ago, witnessed dramatic changes in human life and society. Hunting and foraging became more efficient, as more specialized toolkits created such artifacts as fine-bladed chisels for grooving and boring antler, ivory, and bone.The humble needle appeared, making it possible to make tailored, layered clothing that enabled people to survive nine-month sub-zero winters and to hunt in extremely cold environments (Fagan, 2010). These were the millennia when art became commonplace, not only body painting and decorated artifacts, but cave paintings and engravings that depicted the Ice Age bestiary, often elaborate symbols, even imprints of human hands. Much of this art must have been on leather, wood, clothing, and other perishable materials, but it is now certain that art was commonplace over large areas of the world by 25,000 years ago (Bahn, 2016). Archaeologists and art historians have argued over the meaning of Ice Age art for over a century, with little general agreement, except for a realization that human ­societies had acquired much greater symbolic and supernatural complexity.There can be no question that fluent speech and related cognitive skills placed a premium on transmitting knowledge from one generation to the next by word of mouth, through song, chant, dance, and ritual. With the end of the Ice Age some 15,000 years ago, natural warming, the retreat of ice sheets, rising sea levels, and other major environmental changes forced Homo sapiens to adapt to a radically different world (Mithen, 2006).

What were the consequences of agriculture and animal domestication? Everyone agrees that one of the defining, if not most defining turning point in history, was the shift from hunting and gathering to food production, to agriculture and animal domestication.Theories to account for the switchover date back to the 1870s, when the anthropologist Edward Tylor pointed out that every hunter-­gatherer knew that plants germinated, grew, and could be harvested. Today’s hypotheses revolve around complex processes of transition in which a wide variety of factors played a role, among them chronic drought in the Middle East, a trend in some regions toward 162

Big History and archaeology

more permanent settlement, and an increasing intimacy with herd animals such as goat and sheep. No genius “invented” agriculture or stock raising. They developed from the ­human penchant for innovation and opportunism, for grasping opportunities. Food production developed in several regions of the world—in the Middle East by 12,000 years ago, in South Asia by at least 8,000 years before present, and in China as early as the tenth millennium BC. Ancient Native Americans developed a remarkable expertise with native plants such as goosefoot, squashes, tomatoes, potatoes, maize, and beans, starting at least 6,000 years before present (Barker, 2006). The consequences of food production were far more important than its actual development. Hunting bands became anchored to their gardens and fields; herders rotated their flock and herds through their pasturelands. Permanent villages replaced temporary camps; exchange of raw materials, also exotic objects of symbolic value, expanded dramatically. Above all, there were fundamental, long-term changes in ­human societies everywhere, brought about by families and kin living cheek-byjowl in communities where factionalism and violence could erupt in short order. New social mechanisms came into being to deal with such problems, as kin ties and inheritance developed great importance in even small village societies. The changeover to food production was rapid, its consequences were dramatic. Steady, then more rapid, population growth followed. Within a few thousand years, agriculture and stock herding were near universal. Inevitably, there was a long-term trend toward greater economic, political, and social complexity, toward new forms of leadership vested in kin leaders and chieftains, who became adept at attracting and maintaining the loyalty of their followers. Over four to five millennia, societies in Southwest Asia became increasingly complex, leading, ultimately to the development of the first cities and literate pre-industrial civilizations in Mesopotamia and along the Nile. Similarly, pre-industrial states developed in South and East Asia, and in the Americas.

Why and how did pre-industrial civilizations develop? More than 70 years ago, Gordon Childe (1942) described the origins of civilization as an “Urban Revolution.” How great a revolution it was, if it indeed was, has been the subject of debate ever since. There was no one overriding cause for the emergence of state-organized societies, rather a complex set of interacting processes that fostered greater social complexity, literacy, long-distance trade, and all the panoply of pre-industrial civilization (Scarre and Fagan, 2016). Was the changeover the result of charismatic individuals who became powerful leaders? Ecological change, irrigation agriculture, warfare, and the development of social stratification have all been evoked as “prime movers” of state origins. The debate continues, the causes are unresolved. Childe was concerned almost entirely with Southwest Asia, with Egypt and ­Mesopotamia, with a doctrine of ex oriente lux, “out of the East came enlightenment” (Childe, 1952). Sixty years after his death, archaeologists are still arguing about the origins of state-organized societies, their despotic rulers, social stratification, and economic and social complexity. Egypt and Mesopotamia were far from unique. State-­ organized societies emerged in South Asia, East and Southeast Asia, also in Central 163

Brian Fagan

and South America. All these states share many general features, but they also have marked differences. Egyptian pharaohs ruled by precedent; early Sumerian leaders in ­Mesopotamia were secular and religious rulers. Early Chinese dynasties were governed by little more than war lords, until Emperor Shihuangdi unified a patchwork of small polities into a single kingdom by force and draconian governance in 221 BC. Maya lords in Central America were thought of as divine rulers who were intermediaries to the deities of the supernatural world. Centuries later, the supreme leader of the Aztec of the Mesoamerican highlands enjoyed absolute secular and supernatural authority, as did the Inca, as supreme ruler of Tawantinsuyu, the “Land of the Four Quarters” that radiated across the Andean highlands and lowlands from his capital at Cuzco. In studying the world’s early civilizations, archaeologists are researching societies with much compressed timescales compared with those of earlier millennia, their chronologies often defined by historical records, like Egypt’s dynastic time frame. Such researches have long been preoccupied with royal burials, with cities and palaces on a grand scale, with temple plazas, and the rise and fall of pre-industrial civilizations, as irregularly cyclical as those of later states and empires. In recent decades, the emphasis has tilted in two notable directions. Much of the research has shifted from the center to the periphery, from cities to their hinterlands. Such research began in Mesopotamia and Peru as early as the 1950s and 1960s, but has intensified dramatically since then, especially with the development of remote sensing methods such as LIDAR. For the first time, we can look at pre-industrial civilizations in the context of their wide landscapes. This is revolutionizing our understanding of changing settlement patterns and environmental challenges, as well as our knowledge of the ways in which cities and towns fed themselves. ­ onarchs The second shift has been subtler, but is, if anything, even more important. M and the elite of any state-organized society lived at the pinnacle of societies where most things flowed to the center—precious metals, valuable ornaments and symbols of religious, or supernatural power, the list is seemingly endless. With our, quite natural, preoccupation with the pinnacles of society, and with spectacular finds, we have tended to ignore the commoners, the anonymous farmers, fishers, herders, and artisans who fed everyone while remaining in the shadows. One of archaeology’s greatest contributions to big history is its concern with human behavior, not only the decisions and deeds of great people, of generals and their armies, but with humble folk. We study artifacts and food remains, source raw materials, deal with minutiae of climate change and caravan trade, markets, and the behaviors of different members of society—men and women, merchants from afar living in special quarters of great cities like Teotihuacan in the Basin of Mexico, priests and warriors, even hermits living in isolation. One of our most important, and indeed exciting, contributions to big history come from the inconspicuous, and often frankly rather dull, finds in garbage heaps and middens. For example, it is only in the past few years that we have learned about the great importance of bread and dried fish in the rations of those anonymous folk who built the Pyramids of Giza.

Archaeology in the realm of written records The Greeks called them ichthyophagi, “Fish Eaters.” They were barbaroi, savages who camped in sheltered bays of the Red Sea, subsisting off fish that they bartered with 164

Big History and archaeology

passing merchant ships for many centuries. Anonymous folk, these, known only from sparse Greek accounts, but excavations at the port of Myos Hormos on the western shore provide a window into their fishing practices. Myos Hormos was a major entrepot for the Indian trade at the time of Christ, a cosmopolitan port, where fisher folk camped by a shallow lagoon and mended their nets. Thanks to the waterlogged deposits, we know that they used bast fiber nets and basket traps, also stone barriers across shallow inlets. The Myos Hormos Fish Eaters were people on the margins of much wider societies and trade routes. Just for a moment, excavations bring them into a fleeting historical spotlight (Peacock and Blue, 2006). With the appearance of pre-industrial civilizations, written records and a much wider diversity of sources contribute to big history. Once historical documents, be they on clay, papyrus, or parchment, come into play, the role of the archaeologist changes significantly, just as we see at Myos Hormos, where excavations flesh out a superficial historical reference. This does not mean that archaeology becomes of lesser importance, for its great strengths are in documenting cultural change and continuity over long periods of time. One must not forget that many of the fundamental building blocks of pre-­ industrial, and indeed later, civilizations come from much earlier cultural developments. They stem from the basic continuity of human life that flourished under the ever-changing panoply of divine rulers, military campaigns, and new centers of economic and political power. Egypt provides a classic example (Kemp, 2017). A brilliant tapestry of pharaohs and carefully orchestrated deities presided over the Nile Valley. Like all pre-industrial civilizations, the rulers’ authority and power depended on the anonymous labors of thousands of commoners and artisans, regimented and orchestrated by a complex ideology. Pharaohs came and went, diplomacy and wars of conquest waxed and waned, but one thing remained constant. Each summer, the Nile flooded, crops were planted and harvested, and fishers cast seine nets into shallow lagoons.The life of most Egyptians continued as it had always done, ruled by the endless passage of the seasons. In Egypt and elsewhere, the continuity of pre-industrial civilizations depended on agriculture, on carefully stored food surpluses, and on huge numbers of anonymous folk, who provided the continuity that underlay, and still underlies, civilization. Archaeology breaks down their anonymity and brings them to the historical table, especially over the past five millennia. Written records go back about 5,000 years into the past, but are of limited use for the first thousand years or so. Then they proliferate, but most of the world was still preliterate. In the Americas, the Pacific, and tropical Africa, conventional, ­document-based history arrived with Europeans, especially during the Age of Discovery after AD 1500. In many parts of sub-Saharan Africa, written records did not come into use until the establishment of colonial rule in the 1890s—AD. It follows that, in many parts of the world, the long span of history relies almost totally on archaeology until very recent times indeed, a demanding intellectual issue for nations that are concerned with their heritage and with developing historical curricula. Africa presents a fascinating challenge for both archaeologists and historians. Any forms of historical record are little more than two thousand years old and often confined to the coast and caravan-accessible areas of the far interior. But humankind originated in sub-Saharan Africa some three million years ago, giving the continent the longest historical record on earth. Agriculture and herding societies arrived south 165

Brian Fagan

of the Sahara between two and three thousand years ago, long before written history. One possible source for earlier tribal histories are oral traditions, which are difficult to assess critically and extend back at the most a couple of centuries. Half a century ago, a small group of archaeologists, anthropologists, historians, and linguists realized that multidisciplinary research could throw light on such important issues as the Bantu expansion of 2,000 years ago. Decades of research have produced the first authoritative syntheses of African history as well as outlines of national histories that have stimulated further research (Connah, 2018). Some African scholars are actively involved in writing national histories. In some cases, research from an excavation has ended up in school and university textbooks within a year or less, sometimes even before its academic publication. At a time when the inexorable forces of globalization are blurring cultural diversity, archaeology reminds us that such identities have deep historical roots and are a critical part of being human. They may sometimes be used to fashion nationalist histories or to advance racist agendas, which is despicable. But the artifacts of the past are ultimately a neutral record of the rich and vitally important cultural diversity of humankind, which will always be with us. As generations of archaeologists have pointed out, archaeology is about people, not just artifacts and ruins. It’s about the strategy of a bison hunt on the North American Plains 6,000 years ago, where the excavators of the butchered carcasses could identify the direction of the wind on the day of the hunt. It’s discovering that Egyptian pharaoh Ramesses II, who died at 92, suffered from arthritis, painful dental abscesses, and poor circulation. Modern science knows more about the king’s health than he did. It’s teasing out fascinating historical detail. A classic example comes from the Colonial village at Martin’s Hundred,Virginia, where historical archaeologist Ivor Noel-Hume identified the owner of a house exposed in his excavation by combining written sources with a cannon ball and a fragment of gold thread found in his trenches. His name was William Harwood, the head of Martin’s Hundred, one of the few individuals allowed to wear gold on their clothing and to possess cannons ­Noel-Hume, 1982). His former colleague William Kelso has excavated the first colonists’ houses at Jamestown on Chesapeake Bay and identified the burials of some of the settlers (Kelso, 2017). In South Carolina, blacks outnumbered whites by almost 2 to 1, and one half of that majority were African born in 1740. By studying the containers used by women for cooking Leland Ferguson (1972) studied unconscious resistance to slavery, in an environment where African-Americans maintained their distinctive culture in the face of repressive enslavement. Archaeology, with its rich potential for studying the mundane and the trivial, the minutest details of daily life, is an unrivaled tool for the dispassionate study of social inequality and ethnicity, of trade and exchange, and contacts between widely separated communities and groups in historic times. What is loosely known as “historical archaeology” covers an enormous range of sites and societies (Orser, 2016). Excavations in Greenland have done much to clarify the early Norse settlement of this ice-mantled land and the reasons for its abandonment. Slow-moving analysis of cod bones from middens in northern Iceland and English towns have pinned down the period in the late tenth century when dried and salted cod became a European staple (Barrett and Orton, 2016). Excavations on slave plantations in the southern 166

Big History and archaeology

United States have shown how African-Americans retained much of their own culture during servitude. Malay quarters in Cape Town, South Africa, convict prisons in Tasmania, factories and bridges from the Industrial Revolution are but a few of the topics that are grist to the historical archaeologist’s mill. The archaeologist’s approach to big history encompasses all of humanity, and ­humankind’s universal cultural heritage. This does not necessarily give archaeologists unique authority over the past. In many societies, the past is a valued cultural commodity that preserves a group’s identity from one generation to the next. The past is vested not in science but in household, community, kin groups, and territory. Western science and its perspectives on the past are fundamentally incompatible with those of other societies. We think of the passage of the human past in terms of linear, albeit branching, time. The nineteenth-century German statesman Otto von ­Bismarck called this the “stream of time” upon which all human societies ride for a time. The analogy is apt for the archaeologist’s linear perspective, but there are numerous paradigms of the past in traditional societies. Many involve mythic creators, usually primordial ancestors, deities, or animals that have established the landscape and the cosmos. Such perspectives are often a matter of faith, of ritual belief, which one must respect.Those who espouse such ancient beliefs are as much stakeholders in the past, the universal cultural heritage of humankind, as archaeologists. So are other stakeholders, who range from the tourists who visit Stonehenge to the Navajo people who control the pueblos at Canyon de Chelly in the American Southwest. Nowhere do we see this more clearly than in today’s densely occupied cities, which are complex palimpsests of centuries, and sometimes millennia, of history, and a dense mosaic of stakeholders. Much urban archaeology involves complex excavations that explore the past with small trenches excavated ahead of the bulldozer. There are even instances of excavations in London that have been conducted in the basements of high rise buildings, Snapshots emerge from these excavations, of people going about their daily business, building and rebuilding houses and tenements, of shopkeepers and their wares, of cemeteries and individual burials. The huge Crossrail project that is bisecting London with a subway has yielded an astounding range of sites, among them plague burial pits, even the abandoned detritus of a factory that manufactured ketchup and other Victorian relishes. One can argue that archaeology is often the study of the trivial, but that misses the point. Even the humblest of artifacts and food remains add once hidden dimensions to the history of cities like New York, Paris, or Sydney with their many stakeholders. Our archives are finite, and, at best, fragmentary. Imagine writing history from a billiard cue, three glass beads, a shattered cup, and a spark plug. That is, ultimately, what archaeologists do and how they contribute to the later millennia of  big history, providing critical detail, subtle nuances of the past, and, above all, a perspective on the diversity of human behavior and society.We fill in gaps in the broad canvas of history, especially when it comes to the anonymous players, whose lives unfolded far from the historical radar. We study those who have been referred to as “people without history” (Wolfe, 2010). They certainly have a history, but it is inaccessible without archaeology. Archaeology makes unique, and important, contributions to big history, but, unfortunately, it is a discipline under threat from industrial society—from agriculture, 167

Brian Fagan

mining, deep plowing, and urban development, from looters, catastrophic damage caused by war, such as the destruction at Nimrud and Palmyra in Iraq and Syria, and even from the unintentional feet of thousands of tourists who are devastating sites like Angkor Wat and Mycenae with their devotion. A great deal of this work involves cultural resource and heritage management, which is now the dominant form of archaeological research in many countries, among them Britain, Canada, Denmark, Japan, Mexico, and the United States. Archaeology’s archives are finite; once disturbed by excavation or modern development, they have vanished into the proverbial mists of time.We are a discipline in long-term crisis, with a doubtful future.This means that huge segments of the world’s big history are under threat. The menace is real, compounded by the breadth of the archaeological enterprise, which embraces all manner of stakeholders in the past, from city dwellers to Amazon Indians still living by hunting and foraging. We may dig holes, study satellite images, and walk the countryside or urban landscapes, but, in the final analysis, we live and breathe big history. And our work also has the potential to provide insights into seemingly insurmountable problems of the future, such as racism and climate change. Robert Kelly (1916) calls this “The Fifth Beginning,” a future where cooperation is all-important and where, for the first time, human evolution could be, and must be up to us.

Further reading Adovasio, James, and David Pedler. 2016. Strangers in a New Land: What Archaeology Reveals About the First Americans. New York: Firefly Books. Bahn, Paul G. 2016. Images of the Ice Age. Oxford: Oxford University Press. Barker, Graeme. 2006. The Agricultural Revolution in Prehistory. Oxford: Oxford University Press. Barrett, James H., and David C. Orton. 2016. Cod & Herring. Oxford: Oxbow Books. Bass, G.F., et al. 1989. “The Bronze Age Shipwreck at Ulu Burun: The 1986 Campaign,” International Journal of Nautical Archaeology 93: 1–29. Childe, Vere Gordon. 1942. What Happened in History. Baltimore, MD: Pelican. ———. 1952. New Light on the Most Ancient East. London: Routledge and Kegan Paul. Clark, Grahame. 1961. World Prehistory: An Outline. Cambridge: Cambridge University Press. Connah, Graham. 2018. African Civilizations: An Archaeological Perspective. 2nd ed. ­Cambridge: Cambridge University Press. Fagan, Brian. 2001. Grahame Clark: An Intellectual Biography. Boulder, CO: Westview Press. ———. 2010. Cro-Magnon: How the Ice Age Gave Birth to the First Modern Humans. New York: Bloomsbury Press. ———. 2018. People of the Earth. 14th ed. Abingdon: Routledge. Fagan, Brian, and Nadia Durrani. 2016. A Brief History of Archaeology. 2nd ed. Abingdon: Routledge. Fitzpatrick, A.P. 2011. Amesbury Archer and Boscombe Bowmen: Early Bronze Age Burials at Boscombe Down, Avebury, Wiltshire, Great Britain. Salisbury: Wessex Archaeology Monograph. Johnson, Matthew. 2010. Archaeological Theory: An Introduction. 2nd ed. New York: Willey/Blackwell. Kelso, William M. 2017. Jamestown: The Truth Revealed. Charlottesville, VA: University of Virginia Press. 168

Big History and archaeology

Kelly, Robert L. 2016. The Fifth Beginning: What Six Million Years of Human History Can Tell Us About Our Future. Berkeley: University of California Press. Kemp, Barry. 2017. Ancient Egypt: The Anatomy of a Civilization. Abingdon: Routledge. Klein, Richard. 2009. The Human Career. 3rd ed. Chicago: University of Chicago Press. Meltzer, David. 2008. First People in a New World: Colonizing Ice Age America. Berkeley: University of California Press. Mithen, Steven. 2008. After the Ice: A Global History. Cambridge, MA: Harvard University Press. Noel-Hume, Ivor. 1982. Martin’s Hundred. New York: Knopf. Orser, Charles E. 2016. Historical Archaeology. 3rd ed. Abingdon: Routledge. Papagianni, Dimitra, and Michael A. Morse. 2015. The Neanderthals Rediscovered: How Modern Science Is Rewriting Their Story. Rev. ed. London and New York: Thames and Hudson. Peacock, David, and Lucy Blue, eds. 2006. Myos Hormos-Queir Al-Qadim: Roman and Islamic Ports on the Red Sea. Oxford: Oxbow Books. Scarre, Chris, and Brian Fagan. 2016. Ancient Civilizations. 2nd ed. Abingdon: Routledge. Stringer, Chris, and Peter Andrews. 2012. The Complete World of Human Evolution. 2nd ed. London and New York: Thames and Hudson. Trigger, Bruce G. 2006. A History of Archaeological Thought. 2nd ed. Cambridge: Cambridge University Press. Wodehouse, P.G. 1919. A Damsel in Distress. London: Herbert Jenkins. Wolfe, Eric. 2010. Europe and the People without History. 2nd ed. Berkeley: University of California Press.

169

7 BIG HISTORY AND PHILOSOPHY Philosophical foundations of big history: why big history makes sense Armando Menéndez Viso1 In the last decades, following the proposal by David Christian, many scholars have seen the need of a new kind of history, the big history, which completes the traditional approach by incorporating the scientific narrative, thus linking the most distant past to our present.Why is big history necessary, how (if at all) can it be distinguished from history tout court and what makes it a meaningful enterprise are the questions these pages aim to answer. It can be argued that the specificity of big history comes from the insertion of human history into a cosmic framework – or, in other words, from supplying a universal context for human deeds. In David Christian’s words, “just as we need world history to help us understand the significance of particular local histories, so, too, we need an even larger map to help us see the place of human history in the history of the Earth and the Universe”.2 In this sense, the big history would be a sort of an absolute reference, in which any deed can be placed, or the most general background in which any other histories take place.This is the idea underlying the ChronoZoom project, probably the best example of this facet of big history, defined by its authors as “a timeline for all of history: from the Big Bang, to the time of the dinosaurs, to the present”, with which “you can browse history, rather than digging it out piece by piece”.3 A tool like this is in itself a great achievement of the utmost usefulness, particularly for education, but does not suffice to take us beyond the limits of traditional history: historians, geographers, cosmographers, philosophers and, in general, sages of all times have tried to build something similar to a universal timeline or a universal map, summarising, ordering and contextualising the historical and scientific knowledge of the time. It is true that our world is now considered bigger than any time before, but this alone does not justify a new way of looking at it. Indeed, anything temporal can be situated somewhere between the Big Bang and now, and anything spatial can be placed within the known universe but an expanded framework does not necessarily entail a new vista, nor a new insight. Our relation to the sublime (the overwhelming, the universe at large) is not new either: the cosmic and the sempiternal have always been there, and the conscience of it is as old as humans. 170

Big history and philosophy

Just like us, or perhaps even more than us, our ancestors have perceived (with greater or lesser perspicacity) their dependence on the world around them, from the stars to the soil under their feet. Even time itself has been a matter of study for many cultures and individuals – ­historians, philosophers and physicists in particular.4 The peculiar character of big history cannot stem from the size of its object (the entire universe, also studied by astronomy and physics at large) or the dimensions of the product (the length of the timeline, the character of which is not different from any historical timeline). For big history to be something else than a mere quantitative extension of history tout court, it should bring in something new, something absent from traditional historical approaches, and this cannot be the cosmos. However, it is in the idea of cosmos (or, alternatively, in the idea of physis or nature) where the key can be found to explain the specific character of big history, and why it appeared only recently.

A natural history The Greek cosmos was mainly a set of orderly bodies, which moved regularly, following cyclic patterns. Thus, for the Pythagoreans, harmonious (and eternal) geometrical proportions resounded in nature; for Parmenides, there was no true change in cosmos; for Leucippus and Democritus, atoms were everlasting and inalterable; for Plato, the physical world was a pale counterpart of eternal, perfect ideas; for Aristotle, heavenly bodies belonged to immutable spheres; etc. Only Heraclitus can be considered an exception to this static vision of the universe, with his idea that everything flows. The cyclic vision of the universe remained dominant for centuries, and several turns were needed to alter it.The first one was a dramatic increase of size, both of the sublunary and the superlunary worlds, which began with the modern era. On the one hand, the exploration voyages of the Spaniards and the Portuguese in the fifteenth and sixteenth centuries started to complete the picture of the globe, thus giving an idea of its true proportions; on the other hand, the use of telescopes and microscopes made our sight reach the incredibly small and the biggest, which turned to be much bigger, and much further away, than anyone could have thought before that time.The observations of Tycho Brahe, Copernicus, Galileo, and others challenged the existence of perfect, immutable heavens and began to reveal the phenomenal distances between celestial bodies.The patient improvement and use of microscopes by Janssen, van Leeuwenhoek, or Hooke (to name a few) discovered the infinite in the small and opened a whole new world of complexity to scientific scrutiny. This expansion of the limits of the observable, with the consciousness of insignificance it entails, was nevertheless not enough to remove the concept of a stable universe. In fact, the mechanical image of the universe became triumphant, since it could be extended to every known realm: it was, properly speaking, universal.   The clock was still the preferred metaphor of the world: a cyclic, automatic machine which keeps its functioning unaltered throughout time. The underlying image of the universe was systematic but ahistorical. The book of nature, the mind of God, or the designs of the blind watchmaker were written in mathematical characters and could be deciphered once and for all. There was a mathesis universalis, and the most brilliant names of the time (Descartes, Leibniz, Newton, Wallis, etc.) strived to find it. 171

Armando Menéndez Viso

During the nineteenth century, something happened more decisive for the collapse of the idea of a constant universe – and it was revealed not by looking at the sky, but to the ground. Fossils were known since the Antiquity, and in general, they were correctly interpreted. But people like Gideon Mantell, the Murchisons, Adam Sedgwick and Charles Lyell revealed the connection between them, thus placing the origins of life in a past much more distant than anyone had ever imagined. Modern geologists found that the Earth has been continually changing for millions of years. Rivers, mountains, valleys, oceans, deserts, lakes, etc. were not parts of a permanent scenery but had their own history and a role in the history of life.The unprecedented extension of time brought about by geology made modern biology possible, creating room for the steady forces of evolution to operate. Thence Darwin’s words: “He who can read Sir Charles Lyell’s grand work on the Principles of Geology, which the future historian will recognise as having produced a revolution in natural science, yet does not admit how incomprehensibly vast have been the past periods of time, may at once close this volume.”5 The full extension of the changing nature of our planet began to be understood only from the 1960s, when plate tectonics was formulated, following the theory of the continental drift proposed by Alfred Wegener almost 50 years earlier. The idea proposed in the eighteenth century by James Hutton, that the Earth evolved together with life on it,6 could now be considered scientifically. Also in the 1960s, another key advancement was going to take place, extending the idea of change to the whole universe. It happened somewhat by chance in 1964, when Arno Penzias and Robert W. Wilson, using the Holmdel Horn Antenna, ­detected a stable hiss which seemed to come from everywhere7 – what was later called the cosmic microwave background radiation. This confirmed the hypothesis of an expanding universe, proposed in the 1920s by Edwin Hubble, and supported the Big Bang theory formulated by George Gamow and Georges Lemaître. In a few decades, humanity was moved from a world just a couple of millennia old to a universe of several billion years. But not just that: the new (or renewed) sciences showed that everything around us, from stars to atoms, is in perpetual change. This means a radical departure from the previous Western tradition: the classic (mainly ­Aristotelian) idea that only the sublunary world was subject to generation and corruption, carried into the middle ages, among others by Avicenna and Thomas Aquinas, was a resilient one. Only very recently we came to think that the cosmos at large is not a static system, but an evolving complex, thus producing one of the most shocking theoretical shifts in the history of knowledge. The universe as we conceive of it nowadays is continually changing, has an origin and an evolution: a history (in the sense of res gestae) – a natural history. “Natural history” is an old expression. Traditionally, it designated the description (historia) of natural things (animals, plants, rivers, mountains, etc.), as opposed to natural philosophy, which tried to explain the things described by natural historians, i.e., the natural world. The occupation of natural philosophers was what we now call science. Natural historians were often travellers, expeditioners, field workers who explored the world and gathered information from direct observation. Of course, many great scientists were natural historians and philosophers at the same time, like Mutis, Darwin, Lyell, Alexander von Humboldt, etc. Natural history was seriously 172

Big history and philosophy

undermined by the mathematical triumph of modern scientists, particularly after the dissemination of Newton’s Principia. The so many times commented success of a purely mathematical theory to predict unforeseen phenomena was interpreted as a definite proof of the mathematical structure of the universe, as well as a powerful motive to place analysis above data collection on the totem pole (despite the empiricist motto of the Royal Society, nullius in verba). Thus, natural history, which was a cherished activity in the classic world (Aristotle, Pliny the Elder, etc.) and more recently for Spanish cosmographers (Alonso de Santa Cruz, Diego Ribeiro, Alonso de Chaves, Pedro de Medina, etc.) and chroniclers (Gonzalo Fernández de Oviedo, Diego de Landa, Juan López de Velasco, José de Acosta, etc.) of the sixteenth and seventeenth centuries, ended somehow overshadowed by natural philosophy. Even if it continued to be practised, particularly in the field of what we now call biological sciences, it ceased to be at the core of scientific literature. A conspicuous exception can be found in the case of Alexander von Humboldt, who laid the foundations (and constructed a good deal of the building) of modern Earth sciences, to the point he can be considered one of the inventors of our concept of nature.8 And precisely von Humboldt’s Cosmos is a great model for big history – as is, in a different sense, Carl Sagan’s work of the same title, which constitutes an amplified, twentieth century natural history,9 and its twenty-first century continuation by Neil deGrasse Tyson. The key to understand von Humboldt’s contribution is that there is a natural history in a true sense. This makes it impossible to consider the natural world as a sheer stage on which proper history – i.e., human history – takes place. As von Humboldt saw very well, describing nature is not just an ornament in telling human affairs, or an expression of a taste for details: nature plays a role in our deeds, just as we play a role in the natural world. Historians do know this, of course, and have used climate, geography, plagues, etc. to give historical accounts. But normally they do not deal with the history of the natural phenomena in their work, taking the natural for granted. Big history, in turn, considers the natural as a true part of history. This way it resumes the tradition of natural history but with an important philosophical load: the assumption of the existence of a proper natural history. In sum, big history could have appeared earlier, but not much earlier, because it is the consequence of important shifts in the history of science, the most important of which is the awareness of time and change throughout the whole universe. As Nigel Calder put it back in 1983, in contrast to the knowledge of the Earth’s surface, which “approaches saturation”, the exploration of time on Earth is “beginning in earnest only now”.10 This is what big history is exploring. Some might dismiss it as another academic fashion but there are objective reasons to justify its existence. Big history brings together history and natural history without being reduced to any of them. It differs from world history in bringing natural history to the front: from the big history perspective, the natural world is not just a background or a source of external causes but a playing actor. The big history is more than the set of human deeds that took place all over the globe, including the history of the world at large, with all its natural components. On the other hand, big history departs from natural history in that it is proper history, interested not only in describing nature, but also in the causal explanations that help to understand the natural, the human, and the complex connections between them. 173

Armando Menéndez Viso

A difference in perspective The existence of a cosmic history makes big history possible and sensible. However, big history does not constitute a methodological or disciplinary departure from history tout court. Big history entails a new perspective, but not a new method, nor a disciplinary intention. That is why Fred Spier can define it as an approach, “the approach to history that places human history within the context of cosmic history, from the beginning of the universe up until life on Earth today. In a radical departure from established academic ways of looking at human history, in big history the past of our species is viewed from within the whole of natural history ever since the Big Bang. In doing so, big history offers modern scientific answers to the question of how everything has become the way it is now. As a consequence, big history offers a fundamentally new understanding of the human past, which allows us to orient ourselves in time and space in a way no other form of academic history has done so far. Moreover, the big history approach helps us to create a novel theoretical framework, within which all scientific knowledge can be integrated in principle”.11 Within big history, human and natural history are on a pair: there is a level of generality in which human and natural deeds can be explained by the same kind of causes, presenting comparable dynamics. Understanding the human past through big history requires that this past can actually be integrated into the history of cosmos not as an exceptional epiphenomenon, but rather like just another one in the series of (cosmic) events. It is true that from this perspective, many details of social history are missing, but the broader perspective allows us to perceive connections and continuities that otherwise would remain out of the picture. Trying to understand the human past is something that big history shares with every historic discipline. In fact, it should perhaps be said instead that every history deals with the present (the human present) through things that are always present (documents, testimonies, remains, ruins, traces, vestiges, etc.) but trying to figure out how those things have come to being in time (past time). Some historians could easily interpret the programme of big history as a criticism, as a veiled accusation of leaving out a significant deal of the past. Is big history not just trying to invade the territory traditionally occupied by academic historians? Not really: big history does not oppose “traditional” history in any sense, but rather opens a new field. “Defined as history from the Big Bang to now, [big history] is still a tiny subfield of the history subspecialty of world history”,12 it has been said. And this is true, if we understand world history as the history of the world, rather than the res gestae in the world. The latter is as interesting as it can be, but big history aims at a different kind of account, complementing, rather than rivalling, traditional history. It is not the object of big history what makes it different, but the look addressed to that object. The lenses of big history are wide-angle ones, allowing a substantial proportion of the surroundings to be portrayed together with the central motif. This does not imply, however, thicker or bigger lenses (i.e., a heavy theoretical apparatus): quite the opposite. Macro and telephoto lenses tend to be much bigger than wide-angle ones, since they are built to capture as many details as possible. Focusing on resolution restrains field of vision. Big history aims at widening the field, not at developing a method of its own. This implies taking some distance from the object, 174

Big history and philosophy

and therefore overlooking some of its details. In return, it includes significant parts of the surroundings, otherwise left out of the picture. In any case, what makes a ­photograph relevant is not the lens used to take it, but rather the quality of the resulting image. Even if traditional history is a discipline whereas big history is an approach, not a new theory or science, the latter needs philosophical foundations of its own and a distinctive style to set itself apart.The novel theoretical framework proposed in Spier’s definition is precisely the natural history, as described by modern sciences. But this framework is not given: it has to be built with the materials produced by individual disciplines. The world history of which big history can be a subspecialty is not the one traditionally contained in a world history handbook but rather the one that can be found, in different forms, in handbooks of astrophysics, geology, biology, meteorology, oceanography, etc. It does not compete with “traditional” history – among other reasons because the history in big history here is not historiography, Historie, but a set of deeds and facts, the Geschichte of the world around us. Big history is not necessarily about us humans. In this, big history is also different from traditional history: the deeds of the Phoenicians fit both in history and in big history (from a peculiar perspective in the case of the latter), but the physical history of the Gibraltar Strait fits only in the latter. Big history is not necessarily focused in human history – even if, of course, its stories are interesting from the human point of view. In this sense, the relation between big history and history can be considered analogous to the relation between biology and ecology. When biology is enlarged to encompass the links between living organisms of different species, ecology appears; when history is stretched to incorporate the links to and within the natural world, big history appears. Just as ecology does not preclude but enlightens biology, big history does not rival but enriches history. On the other hand, big history is certainly a story, not the story. Big history is not the history of the universe, let alone the history of everything. Furthermore, a history of the whole is by definition impossible and useless, like a 1:1 map. For big history to be useful and understandable it has to leave things out. And this means that there can be multiple big histories.

Matters of size When applied to history, big seems both significant and ambiguous: on the one hand, it generates something different from “ordinary” history, but, at the same time, it leaves the field completely open. It is easy to interpret “big” in big history in a similar way to “big” in Big Bang, and indeed big historians tend to depart from that original explosion, commencing their stories 13 or 14 billion years back into the past. There is, though, an important difference between both “big”: when added to bang, big immediately refers us to a particular event (the Big Bang, which can be contested or extensively unknown, but not confusing), whereas big history refers us to an account of events. In a way, big history stands for a particular history, the one told by the Big ­History Project. But this project itself has it differently: “Big History examines our past, 175

Armando Menéndez Viso

explains our present, and imagines our future. It’s a story about us. An idea that arose from a desire to go beyond specialized and self-contained fields of study to grasp history as a whole. This growing, multi-disciplinary approach is focused on high school students, yet designed for anyone seeking answers to the big questions about the history of our Universe. The Big History Project is a joint effort between teachers, scholars, scientists, and their supporters to bring a multi-disciplinary approach to knowledge to lifelong learners around the world”.13 But not every big history has to be related to the very origins of the universe: a galaxy, a geological period, an empire, or a technique can have equally big histories. Big history does not need to be the history of the big. Just like the dimensions of a map do not have to be proportionate to the territory represented in it (a map of the whole universe can be as big as one of a planet, a continent, a region, or a village), the bigness of history does not depend on the size of its player. There is no real reason to equate big history with the history of the cosmos: were that the case, what would be the difference between big history and cosmology? The universe is enormous, but its history does not have to be necessarily bigger than the history of bees, the history of Toledo, or the history of bread. Big history makes it possible to elaborate histories of things that could not fit within traditional historiographic frameworks. Individuals and social entities are seen on a pair with objects, both human-made and natural. A pebble, a bacterium or a valley are equally interesting for the big historian because all of them result from complex interactions which are in turn relevant for human knowledge. “Big” is always relative: the interesting thing for us is the relationships it unveils when used together with “history”. A picture of a planet (like the iconic Earthrise taken from the Apollo 8 on Christmas Eve, 1968) can be as big as a picture of a ­virus. Tiny things, like bacteria, their cilia and their molecules, can have their own big history. And, alas, cannot the life of cell be as big – fascinating, complex, daunting, etc. – as the life of a star? Humboldt wrote a physical description of the world, trying to depict nature as a whole, and called it Cosmos – the same title that Carl Sagan will later use for his famous TV series. Hooke, in turn, opened a whole universe devoting his Micrographia to the observation of the smallest. Savants and scientists have always perceived the relative character of bigness. According to Aristotle: Every realm of nature is marvellous: and as Heraclitus, when the strangers who came to visit him found him warming himself at the furnace in the kitchen and hesitated to go in, reported to have bidden them not to be afraid to enter, as even in that kitchen divinities were present, so we should venture on the study of every kind of animal without distaste; for each and all will reveal to us something natural and something beautiful. Absence of haphazard and conduciveness of everything to an end are to be found in Nature’s works in the highest degree, and the resultant end of her generations and combinations is a form of the beautiful.14 Recalling Kant,15 it can be said that the starry heavens above and the moral law within are equally big. Bigness is not in things, but in our relationships with them. Kantian philosophy brilliantly defends the internal character of knowledge by defining space 176

Big history and philosophy

and time as forms of our intuition. Later on, marginalist economists complete the same transition in the “moral” realm by placing the subject as the only source of value.16 Only we, with our look, can determine – variably, even volubly – what is small and what is enormous. Big history is as big as the perspective of those who elaborate it. Little big histories – to use the concept developed by Esther Quaedackers17 – are not only possible, but the most promising development for big history, since they can enrich the general big history ad infinitum.

Bringing sciences together Big history, and all the little big histories, can be identified by its pluridisciplinary character. Big history requires bringing together findings and specialists from many different fields. Anyone can do big history, but no one can do good big history without some historical, biological, physical, chemical, artistic, technical, astronomical, geological, mathematical, etc. knowledge. Big history, then, requires generosity and modesty on the part of the specialists, to learn from each other and elaborate with each other. Big history is transdisciplinary, rather than interdisciplinary, because it does not transit from one discipline to another, nor occupy gaps between them: it tries to develop through them. It is, therefore, a truly collective enterprise, which cannot be tackled by any single person, not even partially. Big history grows explicitly on cooperative work of many people devoted to different sciences and techniques, ranging from those who observe the positions of the stars or document the migration routes of a given species, to those who bake bread or make kites. There is no reason to consider big history closer to geology than to metallurgy or to leatherwork, even if, for various reasons, it has been more popular among astronomers and geologists than among cabinetmakers and graphic designers. Big history widens multi-causal processes, or, to be more precise, brings complexity to the front line of historical explanation, revealing the cooperation of individual, political and economic forces with natural ones in unprecedented ways. Historians, like anthropologists or economists, have always looked for non-human factors to explain human phenomena: climate, diseases, scarcity, crops, water, technology, etc. However, those factors can now be explained as well, and therefore not taken for granted; the natural world, traditionally seen as a set of independent variables from the historiographic perspective, becomes dependent. The explanans of some step was an explanandum in a previous one, and this is new to big history. For instance, in paleoanthropology the Great Rift Valley is just a given phenomenon, which explains the different paths of speciation on both sides of it; the explanation of the geological process remains outside the paleoanthropological account. On the contrary, from the big history perspective, the formation of the rift enters the story just as the evolution of certain species of primates, the climate of the area or the plants growing there. If some causal processes are left out of the picture is because they are conventionally considered less relevant, not because they do not require an explanation. A standard history brings together all the “external” factors necessary to account for the target phenomenon. Big history brings in more factors and is interested in the connection between those factors per se, not just in order to explain the target phenomenon. 177

Armando Menéndez Viso

Education and research Big history has been driven by an educational impulse since its inception. In physics, chemistry, paleography, logic, or geology, it is perfectly possible to publish for just a bunch of experts. This is not the case for big history: by its very nature, the practitioners of this field have to address the widest possible audience and write thinking in a whole array of readers, from undergraduate students to retired people. The educational facet of big history could be enough to justify it, but there are no reasons to circumscribe it to this task. However important, the educational face of big history is not the only one. It is perfectly possible to conceive a certain kind of research specific for big historians. Furthermore, without research the practitioners of big history would have a very limited field to work. It is necessary for big history (if it is to add something to human knowledge) to offer something beyond particular disciplines, and this can be the big perspective, the connections between their products that otherwise would be overlooked. Above certain size, sheer collection ceases to be sheer collection. Experts in big data know this very well: when a sufficiently big amount of data is put together, regularities and connections start to appear that would remain completely unnoticed without the collection. Gathering is a form or research, a way of expanding the borders of knowledge and gaining understanding of the world. Patterns emerge when we look at the whole picture – but that does not tantamount to the picture of the whole. Little big histories are necessary to discover new links and fuel the big history enterprise at large. Big history can never be completed: it is a living approach, evolving at the pace of scientific knowledge. Herodotus opened his Histories declaring that he wrote “so that things done by man not be forgotten in time, and that great and marvellous deeds, some displayed by the Hellenes, some by the barbarians, not lose their glory”.18 Our view has radically changed in the 2,500 years gone since then. The world is now incredibly bigger than Herodotus could have ever imagined, and history has reached a degree of complexity, richness and precision that his founding father could not even dream. Modern science showed us that nature too has a history. There are good reasons to make it definitely big, not to let their glory lose.

Notes 1 This chapter would not have been possible without the generous support of a Banco Santander/Universidad de Oviedo grant, the Institute of Interdisciplinary Studies of the University of Amsterdam, and the research project MINECO-18-FFI201782217-C2-1-1 of the Spanish Government. Special thanks should be given to Esther Quaedackers and Fred Spier for their stimulating discussion of the ideas presented here. 2 David Christian (2008), This Fleeting World: A Short History of Humanity. Great ­Barrington, MA: Berkshire. 3 http://eps.berkeley.edu/~saekow/chronozoom/. 4 Penelope J. Corfield (2007), Time and the Shape of History. New Haven & London: Yale University Press. 5 Charles Darwin [1859] (2008), On the Origin of Species (ed. by David Quammen). ­London & New York: Sterling; chapter IX, p. 281. 178

Big history and philosophy

6 James Hutton (1785), Abstract of a Dissertation read in the Royal Society of Edinburgh, upon the seventh of March, and fourth of April, MDCCLXXXXV, concerning the System of the Earth, its Duration, and Stability. Edinburgh: Royal Society of Edinburgh. www2.odl. ox.ac.uk/gsdl/cgi-bin/library?e=d-000-00---0munahi10--00-0-0-0prompt-10---4-----0-1l--1-en-50---20-about---00001-001-1-1utfZz-8-0&a=d&cl=CL3.3.9&d= munahi010-aep 7 A.A. Penzias & R.W. Wilson (1965), A Measurement of Excess Antenna Temperature at 4080 Mc/s, Astrophysical Journal, 142: 419–421. 8 Andrea Wulf (2015), The Invention of Nature: The Adventures of Alexander von Humboldt, The Lost Hero of Science. London: John Murray. 9 Carl Sagan (1980), Cosmos. New York: Random House. 10 Nigel Calder (1983), Timescale: An Atlas of the Fourth Dimension. New York: Viking Press; p. 11. 11 Fred Spier (2015), Big History and the Future of Humanity, 2nd ed. Oxford: John Wiley & Sons; p. 1. 12 Cynthia Stokes Brown (2007), Big History from the Big Bang to the Present. New York: The New Press; p. xiii. 13 www.bighistoryproject.com, as of February 2018. 14 Aristotle, De partibus animalium, I, 5 (645a). 15 KpV, Conclusion. 16 W. S. Jevons (1911) [1871], The Theory of Political Economy, 4th ed., London: Macmillan and Co.; C. Menger (1871), Grundsätze der Volkswirtschaftslehre. Vienna: Wilhelm Braumüller (English translation by James Dingwall and Bert F. Hoselitz, New York: The Free Press, 1950); C. Ehrenfels (1896), ‘The Ethical Theory of Value’, International Journal of Ethics, VI (3): 371–384. 17 https://blog.bighistoryproject.com/2018/03/01/a-short-history-of-little-big-histories/. 18 Ἡροδότου Ἁλικαρνησσέος ἱστορίης ἀπόδεξις ἥδε, ὡς μήτε τὰγενόμενα ἐξ ἀνθρώπων τῷ χρόνῳ ἐξίτηλα γένηται, μήτε ἔργαμεγάλα τε καὶ θωμαστά, τὰ μὲν Ἕλλησι τὰ δὲ βαρβάροισιἀποδεχθέντα, ἀκλεᾶ γένηται, τά τε ἄλλα καὶ δι᾽ ἣν αἰτίηνἐπολέμησαν ἀλλήλοισι. Translation by A. D. Godley (1920), Cambridge, MA: Harvard University Press.

179

8 BIG HISTORY AND POLITICAL SCIENCE Science, the deep past, and the political Lowell Gustafson Introduction The natural sciences place the development of politics within the development of an evidence-based account of the entire known past that has moved through stages from the Big Bang until today. While human politics do represent a new degree of complexity with new properties compared to previous ones, human nature – and human politics – emerged from earlier natural forms. Human political nature evolved from non-human nature. As a result, political science is placed within the natural sciences. The study of politics fits within big history in two ways.The first is in how it recasts a number of currently pressing political topics by placing them in a much longer time frame than is more usually done.This recasts how to use science to analyze major political issues. Evidence from the natural sciences substantiates an account of the deep past in ways that influence how to think about the politics of identity, nation, race, ethnicity, sex, gender, security, and globalization. Secondly, this approach demonstrates the increasingly complex ways that units are bound together, developing ever more complex structures, and leading to how humans are political now. Politics derives from polity, or the sustained, structured relationships among members. Humans’ political nature adds new properties to its natural components, but it remains rooted in its component parts which emerged before writing and even well before humans. In the section on Political Issues, contemporary political issues will be reframed by placing them in the context of the deep past. In From Polity to Human Politics, political science will be placed within the natural sciences.

Political issues Political identity The first political topic that is reframed by the deep past is political identity – or how groups of people develop a sense of belonging. The teaching of history has often had a political objective of identity formation through a national origin story. ­American political identity is bound up with being aware of the history of the American 180

Big history and political science

experience. When the American Historical Association (AHA) was founded in 1884, history had only recently emerged as a distinct academic discipline. “The first few professors in the field of history had only been appointed at major universities in the 1870s.” (AHA) The country had survived its Civil War and the last spike of the transcontinental railroad had been driven in 1869. The nation had achieved its Manifest Destiny of integrating territory from sea to shining sea. It was ready to tell its story. And the state was ready to sponsor it in public schools in part to foster nationalism and good citizenship. This just barely begins the topic of how nationalism is instilled through the teaching of history (Anderson, 1991; Ferro, 2003; Gellner, 1983; Hartnett et al., 2017; Hastings, 2017; Herb and Kaplan, 2018; Hutchinson, 2017; Kohl, 1996; Suny, 2001). There are many histories of many nations whose purpose is to encourage national political identity. Heroic figures, great battles, and epic events form the origin stories of many nations – and the political identities of many citizens. Similarly, the American Political Science Association was founded a few decades after the AHA, in 1903.The study of political science, like history, was associated with being American and even participating in American public life. Courses on the three branches of government were eventually supplemented by work on ethnic and gender politics, along with many other sub-fields. Knowing about and understanding the events leading to – and the text of – the Declaration of Independence, Constitution, Gettysburg Address, the Letter from a Birmingham Jail, and much else became part of being a good American citizen. Just over a century after the formation of the AHA, in 1982, some historians banded together in a World History Association to tell a story of globalization. Those with non-national agendas may teach more inclusive social or world histories. Still, most traditional history has limited itself to the period of the written record of the human past. Historians comb archives filled with primary documents, perhaps going back even as far as ancient Sumer. By the time writing had developed in Sumer by 2700 BCE, there were different groups of people living on all continents and regions in the world except Antarctica. Peoples were speaking different languages and had developed distinct cultures; the now familiar physical differences among peoples were visible. Civilizational, regional, national, ethnic, and other differences were already well developed. Beginning the study of humans within this period of time leaves out a very long prelude. Historians whose purpose is to foster a global identity face the task of trying to build bridges among various pre-existing cultural identities. By starting the teaching of history within the past few thousand years, the story starts with well-established differences that have often led to conflicts. This approach begins with difference and often with distrust and hostility. What political identity would be formed if the starting points of political stories are pushed back before the origins of nations? If the human political story is shown to begin in Africa at least 200,000 years ago, then new human identities may be a result of courses on Human Politics in addition to those on American, British, Chinese, and other nations.

Race and ethnicity Race or ethnic relations have been a long-standing issue in many nations, but certainly in the United States. A traditional account might begin with capturing slaves 181

Lowell Gustafson

in Africa early in the American colonial era, the slave trade, eventual emancipation, Jim Crow, the civil rights movement, accounts of recent abuses of African-Americans by police, and seemingly coded language for white supremacy in current American politics. Without ignoring or denying the tragedies of past centuries and the current manipulations of racial divisions, would a longer time frame for investigating race offer a possible way to improve current ethnic relations? It is possible now to tell a story of race and ethnicity that stretches back not just a few centuries, but much longer. Evidence accumulated and analyzed by physical anthropologists, archaeologists, and geneticists substantiates a narrative about humans over the past 200,000 years. These scientists have looked beyond archival texts to find evidence for the human past well before there was writing. They have found evidence to substantiate a story of humans evolving in Africa. A series of earlier hominin ancestors reach back to the australopithecines of some two million years ago, Homo erectus, Homo habilis, and others (Sarmiento et al. 2007). Humanity’s common ancestor with its closest living relative, chimpanzees, lived in Africa about seven million years ago. A thicket of hominin species evolved between then and about 200,000 years ago. Evolution has had relatively recent effect on humans as well. For example, before the domestication of animals some 10,000 years ago, no human drank cows’ milk and needed to be able to digest it. Once that milk was available and some people started to drink it, those who developed lactose tolerance had a new source of nutrition. The majority of humans are still lactose intolerant and often still find digesting dairy products to be uncomfortable. One possible importance of this for contemporary politics is how it reframes the questions of race and ethnicity. Physical evidence demonstrates that all living humans descended from a group of fewer than a couple thousand individuals in Africa about 200,000 years ago (Oppenheimer, 2003). The oldest existing human language seems to be the African Click languages. The oldest human art and ritual artifacts are in Africa. Humans are all Africans; it is just that some left that continent a little more recently than others. Humanity’s racial differences developed only in the relatively brief time since humans left Africa and migrated across the globe. The genetic, biological differences between human groups are minor, although they do seem to have effect on the distribution of certain diseases, such as cystic fibrosis, sickle-cell anemia, Thalassemia, Tay-Sachs disease, hereditary hemochromatosis, and lactose intolerance. It does seem as though natural selection has continued to play a role in human evolution since humans left Africa. For example, melatonin, produced by the pineal gland and accounting for darker skin pigmentation, provides protection from radiation and its effects of causing cancers (Reiter et al. 2012). As some populations migrated to regions that had less intense sunlight for less time of the year, less protection from the sun proved advantageous. Sunlight on the skin helps the body produce vitamin D, which helps in the use of calcium and helps prevent soft bones or rickets. People with dark skin in Africa suffer from less cancer; people with white skin in Europe have stronger bones. Varying skin color provides distinct benefits depending on the environment. Shorter, stockier people in cold climates conserve heat better. Tall, lanky people in warm climates dissipate body heat better. Is it better to be black or white; short or tall? Depends on your environment. And if 182

Big history and political science

you have access to furnaces, air conditioners, and artificial light, then the environment outside has less impact. All humans have a common origin; the scientifically substantiated story is of one human family, however dysfunctionally its members often behave. The biological traits among all humans are almost identical and none that significantly distinguish ethnic groups. None of this denies the more recent tragic history of racial relations in the United States and elsewhere. But by looking back further than the past few thousand years, it may be more possible to chart a way forward.

Sex and gender One of the great political issues of the last century has been the changing role of women and gender in politics. From the suffragettes’ struggle for the right to vote; women’s rights; women’s running for office; and the changes in public opinion and law about lesbians, gays, bi-sexual, and transgender people, the role of sex and gender in politics has been important in the United States and elsewhere in recent centuries. As with race, it is instructive to see the issues in a much longer time frame. It is especially important for political scientists who wish to understand not only the origins of the state, but of human politics. Gender relations were a principal part of the origin of human politics. The story of sexual differences is different from the one about racial difference. Different races did not develop until relatively recently and only after human all began with the same general genetic makeup. Racial differences are recent and not biologically significant. Sexual differences began much earlier, long before even humans existed. Racial differences among humans are some tens of thousands of years old. Sexual differences between males and females began with eukaryote cells about 1.2 billion years ago. Since life began about 3.8 billion years ago, that means that life had propagated itself asexually much longer (as many forms of life still do) than other forms of it have used sex for reproduction. Often reproduction could incorporate horizontal gene transfer in which prokaryote or eukaryote cells of the same or even different types could exchange certain genes. Hermaphroditic reproduction is practiced by some species in which an individual possesses the reproductive organs of both the male and female. Parthenogenesis, or asexual reproduction by a female, produces an exact replica or clone of itself. This is done by a number of invertebrate species, such as aphids, nematodes, some scorpion species, some crayfish species, water fleas, and even some of the vertebrate such as certain types of gecko and some hammerhead sharks. The benefits of sexual reproduction over the asexual methods used for much longer in the history of life is not obvious. But having more than one individual being involved in reproduction seems to permit better the ability to mutate quickly enough to keep pace with pathogens’ evolution (Ridley, 2003). There may be other sexier reasons for sex, but dealing with pathogens is at least one plausible reason for it. However it is done, a successful reproductive strategy is essential to a species’ survival. If sex is to be used, one element of that strategy is associated with the relative size and other characteristics of males and females. Reproductive strategies vary widely by species. Comparative studies of species relatively close to humans may 183

Lowell Gustafson

suggest behaviors of humanity’s and chimpanzees’ extinct common ancestor. The reproductive strategy of the great apes, with whom humans share a common ancestor about 10 million years ago, for example, includes pronounced sexual dimorphism. Among humans’ close relatives, the apes, males are larger than females. A dominant male will maintain a harem and have to fight other males to retain it, giving size and strength an advantage. Among common chimpanzees too, males are larger than females. Frans de Waal draws on his study of chimpanzees to suggest some possible comparisons between humans and pre-human ancestors (Waal, 1989, 2005, 2007). Political intrigue and strategies to become or remain an alpha male with reproductive rights is the stuff of chimpanzee politics. There are many reproductive strategies and many different demands for caring – or not caring – for offspring. Survival and reproductive strategies of species may be at the heart of  humans’ style of politics. Beginning the story of politics with the issue of reproduction is not new. It follows from the approach of Aristotle, St. Augustine, and others in ancient and medieval political philosophy who began their discussions about politics with the household (Aristotle; Doody et al., 2005). Evolving in Africa meant that humanity’s ancestors faced competitors that enjoyed many advantages. These competitors had sharp teeth, fearsome claws, shells, wings, and running speed – none of which humans’ forbearers did. Put someone down today alone on the plains of the Serengeti with only a rock and a stick, and the person would not be looking forward to a pleasant evening. Human ancestors’ unchosen strategy was to develop larger, more complex brains that permitted them to develop more complex ways of analyzing problems (such as competitors’ behavior and how to respond to them) and to relate in more complex political communities. Political, or social, organization was humanity’s weapon. Larger and more complex brains, with a larger cerebral cortex made more complex social or political relationships possible; which in turn improved humans’ chances of survival (Gustafson, 2009). Brains, not brawn, eventually won the day; and even more importantly, the organization among people that brains permitted won the day. Brain size increased well after human ancestors became bipedal and saw their jaws and canines decrease in size. The average size brain of Australopithecus (3.9–3 mya) was from 375 to 550 cm3, that of Homo habilis (1.9–1.6 mya) 500–800, Homo erectus (1.8M to 200K years ago) 750–1225, and Homo sapiens 200K to present 900–1880. A highly developed brain permitted increased social sophistication. There are, nevertheless, a number of disadvantages with a big brain. First, brains require a lot of energy. Each of a person’s three pounds of brain, 2% of total body mass, consume about 20% of a person’s energy. That means they must be fed a lot, and for most of human history, finding or growing enough food has not been easy. It also means that childbirth is more risky and painful. Mortality rates for mothers and children have often been high. Additionally, the baby has to be born before its brain and skull are fully developed. It takes a very long time for that baby’s brain to develop sufficiently for the child to become independent, much less sexually mature. It takes a lot of work and energy for its caretakers to get the baby not only from conception to birth, but from birth to maturity. Life spans were normally shorter early in hominin history than they are now. The normal life span was often about 30 years once a person survived to 5 years old. 184

Big history and political science

Child mortality rates were relatively high. To have a few children survive to maturity often meant having a number more. Unlike now, early polities needed more people if they were to survive. Women often would be involved in child bearing and raising, along with many other tasks from gathering, to farming, food preparation, weaving, and more throughout their child-bearing years. Additional resources were usually needed from males, who needed reason to be invested in making such provisions. Longer periods of childhood dependency, even if childhood was much briefer than it is in human society, required long-term cooperation in childcare if the species was to survive. Kinship organization was humans’ first political structure. In humans, size differences between genders are less pronounced. By and large, humanity has evolved to permit far more males to reproduce than the ape strategy permits. Human harems are not unknown, but it is more common to have many, if not most, males involved in reproduction. Males need not fight among themselves as much for the right to mate, and relative size and strength between males are less important than they are for strictly harem reproducers. More access to a female means there is less need for size and strength in fights with other males. It also encourages males to be more willing to provide goods if they feel invested in the long-term care of children. Reduced mortality rates, longer life spans, better health care, and population increases have transformed gender relations in very recent times. However, the help of the polity in raising children is no less an issue. Mothers get help in raising children in a variety of ways, but it is difficult for human mothers to follow the polar bear mother’s strategy of raising cubs entirely on their own.The long-term care of children, and long-term relationships among children’s caretakers, is made necessary by humanity’s large brains. It is those brains that make possible complex social and political relationships. These relationships, which start with the mother–child relationship and go on to include caretakers’ relationships, are humans’ principal source of power. Politics does not begin with electoral strategy and opinion polls either in the deep past or now. Politics does not begin even with the state, or full-time specialists in leading political communities such as cities. Politics has its origin in human reproduction. In both race and gender, pushing the origins of the human story back into deep time reframes the discussion about the origin of human polities and clarifies what is of fundamental importance today: the community’s long-term interest in child rearing.

Globalization Ethnicity and gender are recast by examining them in time frames that reach back hundreds of thousands or even billions of years. Another key contemporary issue – globalization – is as well. Globalization can be a shorthand way of talking about closer and denser relationships among people throughout the world through increased economic and cultural interactions. International trade and investments, tourism, and cultural exchanges are all part of this. Increased interactions may be leading to changes in political identity. Meeting, trading, visiting, and otherwise interacting with people from other nations and cultures may have an effect on the development of a more global identity. There is little reason for overconfidence about the development of global citizenship – but the phrase does resonate for many. 185

Lowell Gustafson

A narrative about globalization often begins with a choice of origin dates. Did it begin with the 1989 or so with the collapse of the Soviet Union and the end of dividing the world into three: the First (US, North America, and Japan), Second (­Communist nations), and Third (Africa, Asia, Latin America) Worlds? Did it begin with the British Empire on which the sun never set in the nineteenth century and the rule of the Pound Sterling? Did it begin with the European Imperial period beginning with the Age of Discovery in the fifteenth century? Did it begin with the Silk Road integrating political economies from Asia to the Middle East a couple thousand years ago? Again, by pushing its dates back, there are at least two major changes in the narrative about globalization. The first is about the role migration stories in political identity. The second reaches much further back to the origins and evolution of life.

Migration and political identity The importance of migration for political identity has long been part of the ­American story.The movement of explorers, settlers, trappers, and others across the Atlantic and then the continent is woven into the American mythos. The Lewis and Clarke expedition from 1804–1806, the Oregon Trail in the mid nineteenth century (and the video game about it that sold over 65 million copies since 1974), Ellis Island, and Pioneer Courage Park that sprawls across six city blocks in Omaha, Nebraska, are but a few examples of migration in traditional American identity. John Gast’s 1872 painting “American Progress” shows a personified America, schoolbook in hand, leading settlers, trains, and wagons across the land with fleeing Indians and wildlife in front of them. More recently, immigration has been a hotly contested political issue, with all of its many angles influencing contemporary American political identity. The story of global migration that has been developing in recent decades is of no less importance to human identity and politics. Total global migration has been steadily increasing in recent decades (United Nations; Wittgenstein). The movement of labor, no less than the trade of goods, and flow of capital across borders are transforming the political and economic world. Although the total numbers of people now migrating dwarf figures from the deep past, the story of early migration and globalization is being traced by archaeologists and geneticists (Schurr, 2015). It is generally accepted that Homo erectus migrated as far as Asia, but eventually went extinct. Modern humans left Africa about 70,000 years ago. From the time human ancestors left Africa, perhaps due to climate changes that were making available food sources more limited, people gradually spread out to inhabit the entire world, except for Antarctica, over the next 50,000 years. Humans reached the Americas via Beringia by 15,000–20,000 years ago. The story of peoples migrating across completely previously unknown territory is one of heroism, courage, ingenuity, and identity formation.

Humans and the globe As impressive as the global migration of humans from Africa across the globe is, the story of globalization of humans begins much further back than 70,000 years ago. It begins with the stuff that makes humans up right now. Humans – like all living 186

Big history and political science

beings on Earth – are made of the stuff that is commonly available on the surface of the globe. Each person is made of about 7 × 1027 atoms. About two-thirds of them are hydrogen, one-quarter oxygen, and one-tenth carbon.With nitrogen, they add up to about 87% of your body. Calcium accounts for 1.5% of human body weight, phosphorus just under 1.5%, potassium about 0.3%, sulfur 0.2%, sodium about 0.1%, and magnesium 0.05%. About 57% of an adult human body is water: H2O. All of these are relatively common atoms or chemicals found on the surface of Earth. Humans are made out of the same stuff of which Earth is made. In the well-known Biblical Genesis story, God made humans from the dust of the ground. The name Adam came from the Hebrew word adama, meaning ground or earth. So Adam might best be translated as Earthling. Current scientific views of the origins of life find evidence for the basic building blocks of life being made out of elements and chemicals that are common on the face of the earth (Hazen, 2005). Of course, a big difference between the Biblical and current scientific origin of humans is that in the former, humans were made directly from the earth. In the latter, the first life did rise from the earth, although probably in the seas rather than where it is dusty. And it took a very long time before the first prokaryote cells evolved in many stages to become humans, along with all other life forms. In either case, there are narratives about humans as having been made out of the stuff that makes up the surface of Earth. In both stories, the stuff of the globe is what humans are made of (Shubin, 2008, 2013). Humans are made from humus. The stuff of the earth, how it led to humans – and sustains humanity – now is part of the story of globalization. No less a part is humanity’s response to the Earth. Is it humanity’s goal to dominate, protect, and/or sustain the planet?

Globalization: protecting the homeland How groups seek to protect particularly prized portions of Earth has long been part of human politics. Even in the scavenging/gathering/and then hunting periods, groups may well have sought to protect particularly productive areas. During the agricultural/village era, some land that was rich and close to water sources was often considered to be more valuable than other land, and worthy of protection from use by other humans. An attachment to a portion of land often became part of what being a nation meant. In the international period since Westphalia, the most common cause for which wars have been fought is control of particular territory. There are seemingly endless cultural expressions about the motherland and fatherland. One might point to classics like Rig Veda, part of Hindu sacred writings, which says that “One should respect his motherland, his culture and his mother tongue because they are givers of happiness. A person who is respectful towards his land, civilization and language, attains greatness and he acquires all the happiness of life. His deeds should be such that makes the motherland, the culture and language proud” (Rigveda). One might also point, in a very different cultural setting, to the evocative painting by Jacek Malczewski (1854–1929) who expressed the martyrdom of the motherland. There is Má vlas, a set of six symphonic poems composed in the nineteenth century by the Czech composer Bedřich Smetana. The second poem is Vltava, Mein 187

Lowell Gustafson

Vaterland (My fatherland). There is the moving Finlandia by Jean Sibelius. These are expressions of the great significance attached to land as an ancestor from which humans have been born and that deserves protection or veneration. Many famous expressions of American attachment to the land easily come to mind. Irving Berlin’s “God Bless America,” Woodie Guthrie’s “This land is your land,” and “America the Beautiful” by Katharine Lee Bates are iconic American songs that celebrate the land. Not to be outdone, the Brazilian national anthem praises the “beloved, idolized homeland.” The “Lied der Deutschen,” written by Hoffmann von Fallersleben, from which the German national anthem was taken, praises the “­German fatherland.” A famous English poem by William Blake, whose words are still sung at some English sporting events, celebrates “England’s mountains green.” One might also recall Elton John’s tribute to princess Diana at her funeral, which closely follows Blake’s line with “England’s greenest hills.” These are but a few of the many expressions of reverence for the motherland or fatherland, the land which is an ancestor, the hills where the ancestors still walk. Nations have a powerful relationship with defined portions of land. Nationalists often seek to protect their nest, mourn the loss of their nation and the losses it has suffered. Sometimes they call for preemptive aggression against imminent or possibly future attacks. In World War II and the Cold War, nations were interested in how to deliver bombs by missile to enemies’ homelands. Following the Soviet Union’s successes with Sputnik in the late 1950s, the US was worried about what this might mean for future ICBMs. This concern is largely what motivated JFK on May 25, 1961, to tell Congress that he wanted America to send an American to the moon by the end of the decade. The experience of many astronauts transformed how they understood their home planet. Alexei Arkhipovich Leonov, a former Soviet cosmonaut, became in 1965 the first person to step out of a spacecraft and walk in space. His personal response was to observe how “the Earth was small, light blue, and so touchingly alone, our home that must be defended like a holy relic.” He later reflected that space exploration had shown “all of humanity that we are different… but can work together.” And he remembered “that time {the Cold War} – the insane mistrust, not just for people but between countries” (New Mexico). US astronaut William Anders was part of the Apollo 8 mission in 1968. While orbiting the moon, he took a picture of the earth over the horizon. Earthrise, the name given to the photograph, has had great influence on the environmental movement. No one could look at the picture and feel entirely secure that Earth could always protect humanity from dark, cold, foreboding space. Many felt that they had to protect it. The moon dust brought back helped tell a story about the formation of Earth and the moon. As Earth was being formed by gravity drawing space dust together, just over 4.5 billion years ago, it started to become a respectably sized planetoid when another one about the size of Mars hit it.This Big Thwack knocked all kinds of dust and debris into the space above Earth, where some of it accreted into the moon, which was at first much closer to Earth than it is today. The collision also knocked Earth off to the side a bit, resulting in the seasons Earth has today. The Moon has gradually added to the length of the day (it used to be six hours long), affects Earth’s tides, and

188

Big history and political science

much else. Astronauts went to the Moon and learned more about the history of the human homeland. An equally striking picture was taken in 1990 by the exploration of the solar system by Voyager 1, which has been traveling away from Earth at the rate of 40,000 miles an hour since 1977. The formation of the whole solar system was part of a single process, with the Sun grabbing over 99% of the matter in the area. The leftovers were put to good use, with planets from Mercury to Neptune (and the now deposed former planet of Pluto).Voyager 1 made it as far as Saturn by 1990. (It is now beyond Pluto; it would take another 165,000 years to get to Alpha Centauri, the next nearest star to us.) In 1990, astronomer Carl Sagan asked that camera of Voyager 1 be pointed at Earth to take a picture of humanity’s home. It is just barely possible to notice the pale blue dot, 3.7 billion miles away, on which humans live.1 The exploration of deep space made Sagan reflect on the human condition. “From this distant vantage point, the Earth might not seem of any particular interest. But for us, it’s different. Consider again that dot. That’s here. That’s home. That’s us. On it everyone you love, everyone you know, everyone you ever heard of, every human being who ever was, lived out their lives. The aggregate of our joy and suffering, thousands of confident religions, ideologies, and economic doctrines, every hunter and forager, every hero and coward, every creator and destroyer of civilization, every king and peasant, every young couple in love, every mother and father, hopeful child, inventor and explorer, every teacher of morals, every corrupt politician, every “superstar,” every “supreme leader,” every saint and sinner in the history of our species lived there – on a mote of dust suspended in a sunbeam.” The Earth is a very small stage in a vast cosmic arena. Think of the rivers of blood spilled by all those generals and emperors so that in glory and triumph they could become the momentary masters of a fraction of a dot. Think of the endless cruelties visited by the inhabitants of one corner of this pixel on the scarcely distinguishable inhabitants of some other corner. How frequent their misunderstandings, how eager they are to kill one another, how fervent their hatreds. Our posturings, our imagined self-importance, and the delusion that we have some privileged position in the universe are challenged by this point of pale light. Our planet is a lonely speck in the great enveloping cosmic dark. In our obscurity – in all this vastness – there is no hint that help will come from elsewhere to save us from ourselves. “The Earth is the only world known, so far, to harbor life. There is nowhere else, at least in the near future, to which our species could migrate.Visit, yes. Settle, not yet. Like it or not, for the moment, the Earth is where we make our stand. It has been said that astronomy is a humbling and character-building experience. There is perhaps no better demonstration of the folly of human conceits than this distant image of our tiny world. To me, it underscores our responsibility to deal more kindly with one another and to preserve and cherish the pale blue dot, the only home we’ve ever known.” (Sagan, 1977). For these people, and for those in the environmental movement, Earth is humanity’s homeland. It is a one-of-a-kind planet in the solar system. The current ability to get to another inhabitable planet is, at least not now, within humanity’s reach. It is Earth that keeps humans alive. Humans are made of the stuff that makes the Earth.

189

Lowell Gustafson

Its history is part of the solar system’s history, the Milky Way’s history, human history. Jonathan Yavelow notes that it was after the Earthrise photo that a series of environmental actions were taken: 1969 – National Environmental Policy Act 1970 – First Earth Day – 22 April 1970 – Clean Air Act 1970 – Environmental Protection Agency Formed Some people respond to the story of life emerging from Earth and sustained by it by supporting policies that will sustain it (Yavelow, 2013). As the editors of a recent book on Thomas Berry write, “He is particularly well-known for articulating a ‘universe story’ that explores the world-changing implications of contemporary science. Berry pointed the way to an ecological spirituality attuned to humanity’s place in nature and giving rise to an ethic of responsibility and care for the Earth.” In changing the focus from national to human identity, then the homeland includes not only sea to shining sea in North America, but the entire globe. A concern for sustainability is one way to protect the homeland and one meaning of globalization that may emerge from the deep past.

From polity to human politics The topics of national identity, race, ethnicity, sex gender, humanity, environmentalism, and globalization are influenced by the evidence that the natural sciences have provided about a deep past that is embedded in humans today. The deep past affects how a number of key political topics can be considered. But there is even a more fundamental way in which science and the deep past help to improve an understanding of the political. The highly complex politics of human societies emerged from earlier structures that can be called polity. Human politics has new properties and is of a different order than earlier polities, but includes many of the older, often simpler structures. Because of this, political science is necessarily distinct from other natural sciences, but still needs to be placed within them. What follows below begins with human politics and then works its way back to its relationships with earlier forms of polity. The numbers of transitions that had to go exactly as they did shows how improbable the current outcome was (Alvarez, 2017). Still, the development of complexity in universal history – including human history – is presented in the major works of big history (Brown, 2007; Chaisson, 2006; Christian, 2004; Christian et al., 2014; Spier, 2015).

Relationships within relationships: increasing complexity of polity The deep past is in part the story of what builds over time periods the relationship upon relationship that binds humans together as they are now. Historians might focus on the time periods – when did matter, stars, terrestrial planets, life, etc. first appear? The focus here is on the increasing complexity in the relationships between units, 190

Big history and political science

with earlier combinations often being incorporated within newer and more complex ones. A polity is a sustained community that has structured, sustained relations among its units; over time ever more complex relationships incorporate some earlier ones. With consciousness, or at least with the self-consciousness of social animals, there is a threshold from polity to the political. Human politics has new properties, but it is profoundly rooted in past levels of polity. Human nature evolved from nature. Although usually considered a human phenomenon, politics can be seen in a much wider context and as having emerged from earlier, simpler sustained, structured polities. It is as rooted in pre-human forms as spines are rooted in the evolution of the notochord of the Cambrian era.These relationships became more complex at certain thresholds, with each new level of complexity exhibiting new properties. Human politics is best understood when it is seen within such a holistic narrative and, as a result, human politics can be seen as the study of how polity has become more and more complex over time. The issue here is the origins of politics.

Political anthropology Political science is often about political relationships among people in relatively recent times. Some focus on the most recent election or the upcoming one. Some go back to the founding of the nation and its constitution. Some go back to the origins of the international system or of written political theory. Going backwards in time, there is much evidence for human polity before the great texts of political theory were written, before there was even writing. There is a considerable literature on the origins of the state and chiefdoms. Irrigation systems in ancient Egypt motivated political unification and increased that ancient nation’s organization. Long before that, beads, shell necklaces, tools, precious stones, and post holes hint at how humans lived in sustained, patterned polities. There is evidence of social organization in large ceremonial buildings, soldiers fighting battles, exchange between merchants, farmers, hunters and gatherers, cities with populations of 100,000 and bands of 50. The 2.5 million year old Oldowan tool industry, often associated with Australopithecines in East Africa, exhibited similar styles over large areas and long periods of time that suggest sustained polities in regular contact. Narratives about prehistoric polities are not told from careful readings of texts in archives or from opinion surveys, but from the interpretation of physical evidence. Political anthropology analyzes, among other topics, the development of more complex forms of human relationship over time. Human politics began with relatively small bands or kinship groups. The long period of childhood among hominids required extended child-rearing that required stable, long-term relationships. Given the great importance of fertility, many of these early polities may have been matriarchal. Extended families or bands of 50 or so people joined into villages in which some family relations became relatively distant.With cities, non-kinship groups began living together. New symbols fostered sustained relations among people in these larger political units, and patriarchal systems became the norm. As greater amounts of surrounding land became governed by city leaders and as cities were unified, humanity experienced the development of nations and empires. Each of these new and increasingly 191

Lowell Gustafson

complex polities sustained and incorporated earlier units. Present human societies are fumbling towards a unity among nations in a process of globalization. Many types of kinship groups continue to develop, as well as cities of various sizes, and different national structures in this globalizing period. Indeed, the most complex combinations of the greatest numbers of people in history are developing today. Physical anthropology and linguistics provide evidence that reaches back before written history. They substantiate the story of how hominins became bipedal, due to climate change, when humanity’s early ancestors lived in Africa, before venturing out to populate the globe. They tell how tongues and larynxes developed to permit versatile speech long before writing. But speech could never have developed without a long series of transformations. If grooming helps baboons create and maintain polities, a drawback is that a baboon can only groom one other at a time. Speaking around a campfire permits one to “groom” many individuals. Thus speech between foragers from different patches and thickets permitted them to develop multiple relationships during such gatherings. Speech permits the growth of the human type of polity (Bickerton, 2009). Syntactical, vocabulary-rich speech altered human polity and made it more complex. The increased complexity of organization made possible through speech provided a most powerful weapon for a species that lacked shells, talons, fangs, wings, or relative speed. Relationship, social organization, and polity have become the powerful force that supports the idea that humanity has reached a new age: the Anthropocene (Wilson, 2012). In this, humans have become so powerful a force of nature and have so reworked the natural environment that the Anthropocene has become a new age like the Jurassic or Cambrian.

Biology: from polity to politics Our having heads with sensory organs and four limbs goes back into the ancient oceans. Humans’ ability to hear is related to the gills of fish. Wrists are derived from an early amphibian now called Tiktaalik. Biological politics widens the story by looking at polity beyond humanity. As many profound insights as Aristotle had, he did not get it all right. In Politics, he continued his famous quote about humans being social by adding: “Anyone who either cannot lead the common life or is so self-sufficient as not to need to, and therefore does not partake of society, is either a beast or a god.” (Aristotle). It is a common study to investigate what makes humans different. Humans are said to be rational, speak in a syntactical, vocabulary-rich language, and write books.While it is true that humans and human society have unique abilities, these abilities are actually just more complex versions of earlier forms. The question of what is new about human polity makes sense only when it is seen as emerging from earlier forms, forms with which humanity still shares a great deal. An individual fish in a school, an antelope in a herd, a wolf in a pack, or a bird in a flock may be a beast, but each is social. The matriarch in an elephant herd decides when the group should migrate, the route to take, and the destination. Clearly her legitimacy depends on the group arriving at where the water is, but her leadership   he alpha male in a makes sense only in the context of other elephants following her. T 192

Big history and political science

chimpanzee troop seeks to organize mating, with himself as the only male who mates; females escaping to trysts with other males are notable, but there is a high price to pay if it is discovered (Waal, 2007). Bees communicate the location of newly discovered pollen to their hive members through aerial dances. Ants live in highly organized colonies whose members carry out specialized roles. If humans are social animals, they are by no means the only ones. Those who evolved long before humans lived highly complex social lives (Wilson, 2012). Is human polity more complex? To be sure. Is it a unique phenomenon without precedent? By no means. Sociable life may well have originated with microbes (Hird, 2009). Single cells had to go through an intricate process in order to live together in multicellular units. The earliest single-celled life forms included prokaryote cells, complex combinations of biochemicals that formed membranes defining territorial boundaries between themselves and the outside world.  They formed DNA to reproduce and were able to carry out metabolism.They were “prokaryote” since they did not have a cell nucleus where the DNA was kept separate from the rest of the cell. Prokaryote cells did quite well, surviving for two billion years before they changed. Most could not process food by using oxygen. Tiny organisms like cyanobacteria carried out anaerobic respiration in which they exhaled oxygen as a waste product. After two billion years, oxygen levels in the Earth’s atmosphere became so high that it was toxic to some prokaryote cells. In essence, they polluted the atmosphere with oxygen, not a desirable condition for anaerobic cells. However, at least one kind of cell developed an efficient way to get energy in this new atmosphere – by converting chemical energy from food into a stored form of fuel, adenosine triphosphate (ATP). This mitochondrial cell inhaled oxygen and carried out aerobic respiration, burning ATP with oxygen and exhaling CO2. A prokaryote cell then absorbed a nearby mitochondrial cell, but somehow managing to form a coalition with it rather than digesting it. It permitted the mitochondrial cell to maintain its own structure and co-opted its energy-producing abilities. In return, the host cell went out to find nutrients for both of them. The mitochondrial cell maintained its own DNA and lived in the cytoplasm of the host cell. The host cell’s DNA retreated to a protected kernel or nucleus. The new, more complex eukaryote cell had a more efficient internal structure that set it on a path to considerable evolution. In addition to a nucleus and mitochondria, these cells also developed centrioles, cilia, and other components, each contributing a particular function. If the mitochondrial organelle had been its own cell in the past, the eukaryote cell was itself something of a bicellular unit. Multicelled organisms and animals could never have existed without this greater complexity or without the mitochondrial organelle producing larger amounts of energy than had been available before aerobic respiration developed. How else can this single cell be considered other than as a polity? It represents an impressive increase in complexity over what had existed before, as well as being the ancestor of every redwood, whale, ant, person, and society that exists today. It had to establish ways to decide when and how to eat, move, and accomplish specified tasks. The history of any great modern nation or city is no more magnificent than the story of the origins of the first single cells. Eukaryotes seem to have retained their basic structure for a billion years before they began to experiment with cooperating with each other in larger units of multicellular life, around a billion years ago. 193

Lowell Gustafson

After single cells evolved, colonies of the same type of cells formed a multicellular unit. Stromatolites, accretions of single cyanobacteria, go back to 3.5 billion years ago. These very simple types of multicellular life have advantages relative to each cell trying to survive independently. The increase in cell numbers protect some of them from predators, facilitate reproduction, and provide for greater adaptability to their environment. The transition from the accumulations of the same type of cell to clusters of cells that control different functions happened about a billion years ago. They are clearly in evidence by 580 million years ago, when sponges were widespread. A sponge is an animal that lives in water and is made up of a single type of cell. All the cells cooperate with each other by siphoning nutrient-rich water through its cavities. The nutrients are absorbed and wastes excreted, which get pumped out of the cavities. It is a living irrigation system with no brain, nervous system, legs, eyes, or ears. But its cells somehow “know” how to work together. This one type of cell can adjust to any role within this animal. If you take a sponge, force it through a sieve so all its cells get separated and float to the bottom, they will scoot back together to reform a new sponge. No rugged individualists here. Sponge cells have long since decided that they have no interest in living separate lives. Each sponge is a polity of cells without central direction. Another example is Dictyostelium, a slime mold from a billion years ago, which is “like a society of amoebas that come together for a common cause, for which some will sacrifice themselves” (Zimmer, 2011). Or, there is the case of quorum sensing (QS), a process by which bacteria communicate with each other through chemical signal molecules, in order to synchronize the activities of large groups of cells (Waters and Bassler, 2005). This communication enables bacteria to “mount a co-operative response,” in order to gain access to resources or defend against external threats. As a culture grows, signal molecules are released that attract other bacteria, until a specific population density is reached. Once a threshold has been passed, a coordinated change in bacterial behavior is initiated (Wien, 2017). Although it is necessary for bacterial cells to act in concert for the greater good of the population, there is a second form of QS regulation that increases the ability of an individual cell to survive. Gram-negative bacterial cells have “neural regulatory networks that enable single bacterial cells to integrate environmental signals in order to ‘decide’ whether or not to join a quorum. Individual bacteria can adapt to a changing environment by ‘integrating multiple external signals’” (Withers et al., 2001). This is “social networking in the microbial world,” or a microbial polity (Atkinson et al., 2006). Life also discovered that there was an advantage in being able to move. Some bacteria have a flagellum, a propeller that allows them to move towards the sun, towards food, or away from predators. If a creature can move consistently in the same direction, it is a good idea to have some sense organs in the front and its excretion in the back. Front and back, right and left, up and down, all start to make a difference in animal structure. In order to be able to see what is coming and do something about it, a nervous system and the ability to analyze information and direct action is required. This is a very long way towards developing legs, brains, opposable thumbs, and all of those other components that define humans today. All of these complex body parts require cells to become specialized in function and organization, such as between liver and muscle cells. The requirements to harmonize all of these types of cells and organs became extraordinarily complex. Each 194

Big history and political science

organism became a highly sophisticated polity in its own right, interacting with what lies beyond its membrane or skin. It must be porous. It must take in nutrients and excrete waste. It must distinguish with reasonable accuracy between what “out there” is to its advantage and what is not. Each organism, in turn, becomes part of a larger system that includes so much else with which it must also become familiar. Actions within the body must be coordinated, along with decisions about how to do it, when and where to move in response to information, memories about what these decisions have led to in the past, so that improvements in decision-making can occur. All this is part of a heritage that predates humans or mammals. Collecting information, policy analysis, decision-making, communicating decisions, executing decisions, and much else does not begin with human political systems. Sustained, ordered relations among the members of each organism indicate how deeply rooted the practice of polity is. However, polity goes even more deeply than this.

Chemical and physical polity Polity reaches back to ordered, sustained relations among the biochemicals that made membranes, reproduction, and metabolism possible. The first “life” was not self-­sustaining, undifferentiated, or isolated. It required a way to protect itself from external threats, a way to renew and then reproduce itself, and a way to gain access to nutrients and energy. It needed borders, as well as defensive and economic policies. No one knows exactly how biochemicals came together to form components of the first prokaryote cells. Abiogenesis or biopoesis is the study of how amino acids can form via natural chemical reactions unrelated to life. It may be that “black smokers” deep underwater provided the geothermal energy and the sulfur that became food for these first organisms (Hazen, 2005). It may be that RNA and then DNA developed from self-replicating chemical cycles, such as the Krebs cycle, in which carbon atoms join in ever larger numbers until the process begins again. It may be that lipids were available for use as membranes (Deamer, 2011; Krauss, 2017; Smith and ­Morowitz, 2016).The first living cell was simple in comparison to later life forms, but much more complex than what had preceded them.The ordered relationships among chemicals that led to and sustained the life of an organism had to also be able to be renewed in a new, similar cell through reproduction. Chemical evolution indicates that molecules often can organize themselves into increasingly complex, sustained, and ordered relationships (Alvarez, 2014; Carroll, 2017). Molecular evolution refers to the development of DNA, RNA, and proteins, as well as to molecular development before biological evolution began. Proteins, carbohydrates, amino acids, nucleic acids, lipids, and other building blocks of life did not emerge full-blown in a cell after the formation of Earth. Before these could be combined in an ordered relationship called a cell, each had to be organized from simpler atomic components. Various types of atoms are able to share electrons, thus binding them together into a bi-atomic or multi-atomic unit called a molecule. This ability to share electrons permitted the development of many combinations of carbon, hydrogen, oxygen, nitrogen, and other elements long before there were cells. Once combined, many of these molecules went on to form more complex relationships with other molecules. 195

Lowell Gustafson

When two hydrogen atoms combine with an oxygen atom, a molecule is formed that has new properties that neither atom had in isolation – wetness etc. at least within certain temperatures, water is wet. Can it be meaningfully said that molecules are polities? They have sustained, ordered relationships among their members (atoms). The electromagnetic explanations for the relationships between atoms, within molecules, are part of the story about how humans’ brains and decision-making work. But what really is the chemistry of politics? How does the field of chemistry help to better understand polity? Is polity rooted in the most basic forms of matter, derived from the origins of the universe? Is what cannot be divided further a sustained, patterned relationship? Or is it a matter of isolation, separation, and non-responsiveness? Ever since the time of the philosopher Democritus in ancient Greece, many have sought the ultimate building block, the unit that cannot be divided further, the ἄτομος (atomos).What is an atomistic polity? At the simplest level, is there division, separation, unconnected fragments, or even antagonistic bits? Is the ultimate state of nature atomistic in this sense? At the base of nature, are there just the acts and interests of individual units? Is human nature consistent with such a nature at its most fundamental level? Are any attempts at community and cooperation merely a veneer that must somehow be pasted over a far more fundamental reality of separation and isolation? What politically can be learned from an atom? It is clear that some atoms are very sustainable. While C15 may last no more than a couple seconds, a C12 atom will probably stay intact for longer than the universe has so far existed. Either way, the atom is not the uncuttable, simplest element. Each atom is itself a polity – a sustained, patterned relationship – due to the electromagnetism between electrons, and between protons and neutrons. But what exists inside the atom’s nucleus, among its protons and neutrons? Is there at this level separation or polity? The simplest atom, hydrogen, is composed of an electromagnetic relationship between one proton and one electron. That single proton is composed of a relationship between two up quarks and one down quark. The relationship among these three quarks is structured by the strong force. The two different types of quarks do not unite to form one blob; they each maintain their relatively long distance from each other in constant movement. Reality is not ultimately at rest; it is spinning all the time. It incorporates differences between units. And it is defined by sustained, structured relationships. The strong force that structures the relationships among quarks is aptly named, as it is the most powerful of the four fundamental forces. As a result of it, the quarks in protons are tightly bound together, which is why nuclear fission and fusion involve extraordinary amounts of energy.The strong force also holds a neutron together. Each proton, each neutron, and each atom may thus be seen as a sustained, patterned relationship of members – a very simple polity. Are there units even more fundamental than quarks? Some theoreticians suggest that there may be vibrations in many more than four dimensions in the shape of strings or loops. So far, there is no way to measure or observe such vibrations, and the relationship between them and baryonic or normal matter will still need to be better defined. 196

Big history and political science

Transition to even more complex politics It is worthwhile to note that polity does not mean a lack of conflict and emergent complexity is not uniform or steady. The development of increasingly complex politics entails struggle as well as cooperation within sustained and patterned relationships. Astrophysicists discuss how the annihilation of matter and anti-matter took place soon after the Big Bang, when one particle out of a billion and one particles of matter survived.Why was one particle left over? No one yet knows, but the survivors of the mass destruction went on to form everything that can be seen today. Similarly, biologists talk about how prokaryote cells regularly consumed each other in a life and death struggle to eat or be eaten. That is until a prokaryote ate a mitochondrial cell, only to have it remain alive and form a cooperative relationship within it, perhaps giving birth to the first eukaryote cell. After that, there exists plenty of conflict between and within species, as well as five major mass-extinctions. After each disaster resulted in the death of a great many species, new opportunities opened up for other species. While there has been no consistently steady development of cooperative complexity, there are more complex polities that have evolved over great periods of time. Very often in the study of politics, students are most interested in conflict. ­International political history has often been the record of war. National politics are often strikingly conflictual. Even elections have warlike terminology, although it is better that electoral campaigns replace military ones. Still, more complex polities have been made possible after the destructive periods of wars. A unified nation replaced the separate colonies after the American Revolution. The World War II, which cost ­humanity 70 million lives, also left a legacy of rocket technology. This, and the Cold War’s motivation to send a rocket to the moon, provided the picture of the rising earth over the moon’s horizon that has so captured the imagination of humanists. After the Cold War and the fall of the Berlin Wall in 1989, many saw a period of optimistic globalism and the “End of History,” in which national borders were sometimes thought to be passé. Economic rationality and trade were considered harbingers of global polity. Then, after 9/11 and the “War on Terrorism,” many feared that politics was best characterized by a “Clash of Civilizations.” Are Christendom, the Islamic world, the Confucian and Chinese world, and others locked in conflict? Are states battling non-state actors for dominance? Is there any reason to hope for a way to create a transition to a more complex politics polity beyond the schisms of nationality, race, class, gender, and religion? Does the study of the emergence of politics from polity provide a scientifically based narrative that can help establish a more complex politics? Does it describe a process that is playing itself out? Can humans develop the most complex sustained, patterned relations that they have ever had? Will humans deepen their sense and practice of connectedness with each other and with the nature from which they have emerged and still depend? In the long run, many find little reason for optimism. Over 99% of all species that have ever existed are now extinct, and the rate of extinction due to human activities has quickened over the past century. Humans in their current form have existed for only 200,000 years or so. It is not hard to imagine that humanity will become extinct 197

Lowell Gustafson

in less time than that. If conditions become wildly fortunate, maybe humanity will survive for a few million years. But it is a virtual certainty that humans will be gone long before the earth is consumed by a red giant of a sun some five billion years from now.With earth and the rest of the solar system gone then, the universe will continue to expand until it all dissipates into a Big Chill. In the long run, entropy will overtake increasing complexity. Maybe there will be other universes, but the one for which there is evidence now will die. In the short run, over the next decades and centuries, the question remains if entropy, conflict, a lack of imagination, or other problems will thwart the emergence of even more complex, sustainable political structures.

Conclusions Drawing on the evidence that has been investigated by the natural sciences and the deep past permits the reframing of a number of political topics, such as nationality, race, gender, and globalization. An analysis of emergent complexity demonstrates how relationships have become increasingly sophisticated, reaching their most complex physical form in the human brain and producing increasingly sophisticated polities. From kinship to settled villages to nation and empires and now to global relationships, the current emergent complexity of polity is rooted in pre-human nature. This is a valuable way of using science to discuss the political. The transition from industrial to digital society has led to more complex relationships between more people than have ever existed. The electrical communication between billions of humans rivals the electrical communication between the 100 billion cells in each human brain. Humanity is fumbling towards a new and more complex polity without central direction but with patterns that reach back 13.82 billion years. Humans’ current polity is made up of components that are billions or millions of years old. The current transition towards the most complex set of political relationships that have ever existed are best understood – and fostered by – by a study of the context offered by the natural sciences and the deep past.

Note 1 To view the image go to www.nasa.gov/sites/default/files/images/540616ma in_pia 00452-43_full.jpg

References Alvarez, W. (2014). We Are Stardust Concentrated by Earth! Expositions: Interdisciplinary Studies in the Humanities, 8(1). Retrieved from http://expositions.journals.villanova.edu/ issue/view/130 Alvarez, W. (2017). A Most Improbable Journey: A Big History of Our Planet and Ourselves. New York: W. W. Norton & Company. American Historical Association. ‘Brief History of the AHA.’ www.historians.org/about-ahaand-membership/aha-history-and-archives/brief-history-of-the-aha, accessed September 22, 2017. 198

Big history and political science

Anderson, B. Imagined Communities: Reflections on the Origin and Spread of Nationalism, 2nd ed. London: Verso, 1991. Aristotle. ‘Politics.’ http://classics.mit.edu/Aristotle/politics.1.one.html, accessed ­September 21, 2017. Atkinson, S., Chang, C.-Y., Elizabeth Sockett, R., Cámara, M., & Williams, P. ‘Quorum Sensing in Yersinia Enterocolitica Controls Swimming and Swarming Motility’. Journal of Bacteriology vol. 188, no. 4, February 2006, pp. 1451–1461. Bickerton, D. (2009). Adam’s Tongue: How Humans Made Language, How Language Made Humans. New York: Hill and Wang. Brown, C. S. (2007). Big History: From the Big Bang to the Present. New York: New Press: Distributed by W. W. Norton. Carroll, S. (2017). The Big Picture: On the Origins of Life, Meaning, and the Universe Itself. New York: Dutton. Chaisson, E. (2006). Epic of Evolution: Seven ages of the Cosmos. New York: Columbia University Press. Christian, D. (2004). Maps of Time: An Introduction to Big History. The California World History Library. Berkeley: University of California Press. Christian, D., Brown, C. S., & Benjamin, C. (2014). Big History: Between Nothing and ­Everything. New York, NY: McGraw Hill Education. Deamer, D. W. (2011). First Life: Discovering the Connections between Stars, Cells, and How Life Began. Berkeley: University of California Press. Doody, J., Hughes, K. L., & Paffenroth, K. eds., (2005) Augustine and Politics. Lanham, MD: Lexington Books, p. 146. Ferro, M. The Use and Abuse of History: Or How the Past Is Taught to Children. London: Routledge, 2003. Gellner, E. (1983). Nations and Nationalism. Ithaca, NY: Cornell University Press, 1983. Gustafson, L. (2009). ‘Speaking Up: The Origins of Language.’ Villanova University. www.youtube.com/watch?v=H2Ma7dxu0O0, Falvey Library Lecture Series. Hartnett, S. J., Keranen, L. B., & Conley, D. eds. (2017) Imagining China: Rhetorics of ­Nationalism in an Age of Globalization. Rhetoric and Public Affairs Series. East Lansing, MI: Michigan State University Press. Hastings, D. (2017) Nationalism in Modern Europe: Politics, Identity and Belonging since the French Revolution. London: Bloomsbury Academic. Hazen, R. M. (2005). Genesis: The Scientific Quest for Life’s Origin. Washington, DC: Joseph Henry Press. Hazen, R. M. (2012). The Story of Earth: The First 4.5 Billion Years, from Stardust to Living Planet. New York: Viking. Herb, G. H., & Kaplan, D. H. eds. (2018) Scaling Identities: Nationalism and Territoriality. Lanham, MD: Rowman & Littlefield. Hird, M. (2009). The Origins of Sociable Life: Evolution after Science Studies. Basingstoke: Palgrave Macmillan, p. 35. Hutchinson, J. (2017) Nationalism and War. New York, NY: Oxford University Press. Kohl, F., eds. (1996). Nationalism, Politics and the Practice of Archaeology. Cambridge: ­Cambridge University Press. Krauss, L. M. (2017). The Greatest Story Ever Told–So Far. New York: Atria Books. Lederman, L. M., & Teresi, D. (2006). The God Particle: If the Universe Is the Answer, What Is the Question? Boston: Houghton Mifflin. New Mexico Museum of Space History, a Division of the New Mexico Department of Cultural Affairs. Alexei A. Leonov, The First Man to Walk in Space, www.nmspace museum.org/halloffame/detail.php?id=17, accessed September 21, 2017. 199

Lowell Gustafson

Oppenheimer, S. (2003). The Real Eve: Modern Man’s Journey Out of Africa. New York, NY: Carroll & Graf. Reiter, Russel J., Korkmaz, A., Ma, S., Rosales-Corral, S., & Tan, D.-X. ‘Melatonin ­Protection from Chronic, Low-Level Ionizing Radiation.’ Mutation Research/Reviews in Mutation Research, Vol. 751, no. 1, July–September 2012, pp. 7–14. Ridley, Matt. (2003). The Red Queen: Sex and the Evolution of Human Nature. New York: Harper Perennial. Rigveda, First Mandal, 13/9. Quoted at www.neelkanthdhaam.org/rigveda1.html, accessed September 21, 2017. Sagan, Carl. (1977). Pale Blue Dot: A Vision of the Human Future in Space. Reprint Edition. New York: Ballantine Books. Sarmiento, E., Sawyer, G. J., & Milner, R. (2007). The Last Human: A Guide to Twenty-Two Species of Extinct Humans. New Haven: Yale University Press. Schurr, T. G. (2014). Human Genetic Diversity in a Global Context. In: Spooner, B. L. (ed.) Globalization: The Crucial Phase. Philadelphia: University of Pennsylvania Press, pp. 71–114. Schurr, T. G. (2015). Tracing Human Movements from Siberia to the Americas: Insights from Genetic Studies. In: Frachetti, M., and Spengler, III R. (eds) Mobility and Ancient Society in Asia and the Americas. Cham: Springer. Shubin, N. (2008). Your Inner Fish: A Journey into the 3.5-Billion-Year History of the Human Body. New York, NY: Vintage. Shubin, N. (2013). The Universe within: Discovering the Common History of Rocks, Planets, and People. New York, NY: Pantheon Books. Singh, S. (2004). Big Bang: The Origin of the Universe. New York: Fourth Estate. Smith, E., & Morowitz, H. J. (2016). The Origin and Nature of Life on Earth: The Emergence of the Fourth Geosphere. Cambridge: Cambridge University Press. Spier, F. (2015). Big History and the Future of Humanity. Chichester: John Wiley & Sons Inc. United Nations, Department of Economic and Social Affairs, Population Division. ‘World Migration in Figures: A Joint Contribution by UN-DESA and the OECD to the United Nations High-Level Dialogue on Migration and Development, 3–4 October 2013.’ www.oecd.org/els/mig/World-Migration-in-Figures.pdf, accessed September 21, 2017. Waal, F. B. M. de. (1989). Peacemaking among Primates. Cambridge, MA: Harvard University Press. Waal, F. B. M. de. (2005). Our Inner Ape: A Leading Primatologist Explains Why We Are Who We Are. New York, NY: Riverhead Books. Waal, F. B. M. de. (2007). Chimpanzee Politics: Power and Sex among Apes. 25th anniversary ed. Baltimore, MD: Johns Hopkins University Press. Waters, C., & Bassler, B. ‘Quorum Sensing: Cell-to-Cell Communication in ­B acteria.’ Annual Review of Cell and Developmental Biology vol. 21, November 2005, pp. 319–346. Wien, Peter. (2017) Arab Nationalism: The Politics of History and Culture in the Modern Middle East. London: Routledge, Taylor & Francis Group. Wilson, E. O. (2012). The Social Conquest of Earth. New York: Liveright Pub. Corporation. Withers, H., Swift, S., & Williams, P. ‘Quorum Sensing as an Integral Component of Gene Regulatory Networks in Gram-Negative Bacteria.’ Current Opinion in Microbiology vol. 4, no. 2, 1 April 2001, pp. 186–193.

200

Big history and political science

Wittgenstein Centre for Demography and Global Human Capital. ‘Global Migration Data Sheet 2005–2010.’ www.global-migration.info/VID_Global_Migration_Datasheet_ web.pdf, accessed September 21, 2017. Yavelow, J. (2013). Star Gazing to Sustainability: Appreciating the Scientific Process. Dubuque, IA: Kendall-Hunt. Zimmer, Carl, ‘Can Answers to Evolution be Found in Slime?’ New York Times, 3 October 2011. www.nytimes.com/2011/10/04/science/04slime.html?pagewanted=all, accessed September 21, 2017.

201

9 BIG HISTORY AND HISTORIOGRAPHY Deep tides and swirling foam: the influence of macro-historical trends on micro-historical events David Baker Big history holds the potential to revolutionise how every historian thinks about micro-historical events and to reawaken one of the oldest debates on the relevance of meta-theory in historical scholarship, a debate which has long sat dormant, stagnant, and unresolved. Big history is well known for exploring broad trends that stretch across 13.8 billion years, rising complexity and collective learning being foremost amongst them. But those same broad trends trickle into every famine, every beheading, every palace coup, and every civil war in the past 5,000 years of conventional history. To revive Braudel’s metaphor, those events are surface disturbances, swirling foam atop the deep tides of big history.1 The broad trends of big history are integral to solving the puzzle of meta-theory in conventional historical scholarship. Not only could the old scholarly debates of the twentieth century be revived, but they could be dramatically enhanced with a fresh perspective on how the very large shapes the very small. In essence, a complexity generating mechanism like collective learning not only perpetuates the overarching trend that big history identifies as increasing complexity over 13.8 billion years, collective learning also determines the carrying capacity of the human population at any one time, which in turn determines whether a civilisation flourishes or struggles, which in turn drives the otherwise chaotic-seeming procession of micro-historical events.

The Annales school seeks meta-theory In a paper dating back to 1958, one of the foremost members of the Annales school of history, Fernand Braudel, coined the term ‘longue durée’ and decried how academic disciplines were busy defining their aims, methods, and superiorities, by drawing boundaries between each other. Braudel recognised that ‘each source encroaches on its neighbours, all the while believing it is staying in its own domain.’2 He deemed history the most flexible of all academic disciplines and said it could make use of all 202

Big history and historiography

that the other disciplines convey and can reflect them back again. Braudel encouraged the utmost interdisciplinary work in the examination of the cyclical movements of the short durée and the even longer term, the longue durée, which was the culmination of those cycles. ‘Science, technology, political institutions, conceptual changes, civilisations,’ Braudel said, ‘all have their own rhythms of life and growth, and the new history of conjunctures will be complete only when it has made up a whole orchestra of them all.’3 This is something that big history in the past few decades, by employing experts from cosmology, geology, biology, and the social sciences, has endeavoured to do. Around the same time, in 1955, a young graduate student, Le Roy Ladurie began conducting his investigation of the compoix of Languedoc (registered land surveys dating back to the fourteenth century) in southern France. Originally, he found that the model of sixteenth and seventeenth century accumulation of rural land by the wealthy elite reinforced the classic argument that the early modern period gave rise to the first capitalists in the transition from the medieval period to the modern one. But later, Ladurie went deeper into the countryside and mountain districts and found that this trend only held near the cities. A much more interesting set of trends manifested themselves across all of rural Languedoc.4 These trends could be divided into phases. Some phases saw the number of landowners grow in number as the average size of landholdings shrank, denoting overpopulation and rampant subdivision. Other phases saw average landholdings increase in size and the number of landholders decline, as prosperity reigned in a period of underpopulation. Ladurie identified overcrowding prior to the Black Death, sparse populations after the pandemic, and subdivision beginning again after 1500 AD. His work, The Peasants of Languedoc, focused on this demographic cycle between 1450 and 1730 AD. Ladurie found that this period maintained a certain ‘continuum’ of successive phases of growth and decline and that ‘these phases, taken together in chronological sequence (lift-off, rise, maturity, and decline) imply a unity and serve to describe a major, organised, secular rural fluctuation spanning eight generations.’5 Ladurie had successfully identified an ecological population cycle (or s-curve) in an agrarian civilisation. Ladurie recognised that it was only a microcosm of a set of cycles that happened elsewhere in what big history calls the agrarian era (c.10,000–250 years ago). He also recognised that the fluctuations of rise and decline occurred even though in the long term the population remained fairly stable in France as a whole, at around 20 million people, between 1300 and 1700 AD, hence his often misunderstood phrase histoire immobile, despite the numerous violent upswings and downturns.6 1300–1700 was a period where the agricultural carrying capacity remained fairly stable in France, before the population took off due to the agricultural innovations in the eighteenth century. Ladurie called early modern France a society ‘without a motor’ that developed slowly in comparison to the industrial world.7 Where he got it slightly wrong is that the motor was indeed there and quietly running, and that his snapshot of 1300–1700 was in fact bordered on either side by periods of innovation where the carrying capacity was indeed raised. In fact, the carrying capacity had been rising throughout the agrarian era – just not fast enough to cope with the explosive growth of populations. Ladurie also encouraged the notion of interdisciplinary work in history, saying: ‘As historians we are the rear guard of 203

David Baker

the avant-garde. We leave it to researchers in more sophisticated disciplines to embark upon the really dangerous missions. They are the pioneers… we historians draw very largely on the wealth created by established branches of quantitative science such as demography… [W]e have shamelessly pillaged – though we do try to give as good a return as possible – the resources of demography, to which we have given a historical dimension.’8 He also recognised that France served as a window to the world, and an example of how his case study was replicated in many other regions in many other time periods.9 Ladurie was correct, and needed only to push back the scope of the chronology to see just how connected all these trends really were. Periods of population growth and decline happened at intervals of centuries in many agrarian civilisations between 10,000 and 250 years ago. This trend itself was governed by an even broader rise of the carrying capacity and population numbers that resulted from our species’ unique ability to accumulate and pass on a vast amount of knowledge from generation to generation: collective learning. The groundswell of human population growth and innovation had been proceeding since the Palaeolithic. This generation of variations of ideas gradually raised the carrying capacity and the ability of human beings to harness the energy of their environments.Thus collective learning is tied to rising complexity in the Universe more generally. Collective learning itself is a generator of rising complexity. Just like natural selection is a generator of rising complexity. And all this connects right down to the population cycles of human societies and the micro-historical events they prompt.

The hunt for historical meta-theory Early works on population thinking In order to place Ladurie and his work in a larger historiographical context, this section will deal with the evolution of population thinking in scholarship over the course of several centuries, with a focus on the population debates among historians in the latter half of the twentieth century. The source of the debate rests in the relationship of population to economics and historical events and the apparent conflict between ecological and sociological variables in the longue durée of human history. The idea that population dynamics play a role in historical events, the strength of a civilisation, and the average standard of living for its people is not new. Many pre-­ modern thinkers drew a very general connection between overpopulation and the rise of famine, disease, and warfare. Plato and Aristotle both identified a connection between overpopulation and the rise of sociopolitical instability and this shaped their views on immigration and birth control.10 Confucius also saw that overpopulation could lead to civil strife and a drop in the standard of living, and modern Chinese scholars have endeavoured to find out how closely changes in dynasties are tied to population dynamics and sociopolitical instability.11 The concern of these great thinkers should not come as a surprise because in the era of agrarian civilisations, the carrying capacity did not rise quickly and population strain was felt every few centuries. Tertullian, an early Christian author who was particularly anxious about the question, lived in a period of overpopulation 204

Big history and historiography

and subsequent population decline in the Roman Empire (c.160–220 AD).12 While some have pointed out the absurdity of  Tertullian’s claims that the Earth did not have enough resources to provide for everyone in an age when the population was only a fraction of what it is now (c.250 million), the carrying capacity for the human population in his time was much lower than it is today.13 The high watermark of the Roman population (c.160 AD) was reduced in Tertullian’s lifetime by famine, pandemic, and frequent civil warfare, making the harshness of overpopulation very real at the time.14 Early modern scholars also drew the general connection between population strain in a region and the rise of disasters that reduced the population, the most frequently cited example being Machiavelli, who linked overpopulation with plagues, famines, and (less presciently) floods.15 The most thorough examination of these trends in the pre-modern era, however, was done by Ibn Khaldun.The late medieval Islamic philosopher not only recognised a connection between population and human history, but systematised it into a series of patterns or waves.16 He also drew a connection between population dynamics and state collapse. Khaldun lived in the Maghreb in Northwest Africa, on a thin strip of land between the Mediterranean and the desert. When a Maghrebian civilisation grew powerful, the elites (who practiced polygyny) would multiply rapidly causing an increase in competition for resources, an increase in faction and infighting, and a decline in social cohesion (which Khaldun referred to as ‘asabiyyah’). The agrarian civilisations that lined the coast faced nomadic pastoralists who inhabited the Sahara. When elites were few and social cohesion was high in the agrarian world, a united state could keep the nomads at bay. The moment that elites grew too numerous and the state degenerated into faction and civil war, they were easily conquered by the nomads leading to a change in dynasty or state collapse. Thereafter the pattern began anew. Due to the polygyny causing higher birthrates and the rapid proliferation of the elite, Khaldun assigned a total duration of 80–120 years for such cycles. His work was the first systematic application of population dynamics to shifts in social structure, state collapse, and historical processes.

Malthus identifies agrarian population cycles The modern application of long demographic cycles began with Thomas Robert Malthus, a well-to-do country vicar from the downs of Surrey in southern England. In 1798, he first published An Essay on the Principle of Population, which he gradually expanded into a full-length book.17 The core of his thesis was that human population growth has a consistent tendency to outstrip the resources of the land. Indeed, this is the founding principle of modern ecology for all animal species. It also applies very readily to humanity, both in the nomadic hunter-gatherer lifestyle of the Palaeolithic and in the agrarian era around 10,000 years ago to c.1800 AD. Malthus noted that the population grew much faster than the rate of agricultural production. He pointed out that during periods of overpopulation, food prices increased, real wages dropped due to an oversupply of labour, and shrinking incomes reduced the standard of living of the middling and lower classes to intolerable levels. The strain on the carrying capacity provoked recurrent famines, malnutrition created greater susceptibility to pathogens, and general discontent spilt forth into riot, faction, and war. The outcome 205

David Baker

was a population ‘crash’ until numbers became low enough for food prices to drop, wages to rise, and the standard of living to resuscitate itself. The result of the constant tension between the shortage of resources and the tendency of the human population to increase rapidly was an oscillation between periods of prosperity and disaster. Malthus had successfully identified the long cycles, or s-curves, that prevailed in the agrarian era. Malthus also established one of the founding principles of ecology and evolutionary biology. His Essay on Population was the direct stimulation for Charles Darwin’s theory of natural selection. In October 1838, that is, fifteen months after I had begun my systematic enquiry, I happened to read for amusement ‘Malthus on Population,’ and being well prepared to appreciate the struggle for existence which everywhere goes on from long-continued observation of the habits of animals and plants, it at once struck me that under these circumstances favourable variations would tend to be preserved, and unfavourable ones to be destroyed. The result of this would be the formation of new species. Here then I had at last got a theory by which to work.18 Malthus inspired Darwin’s population-based theory that explained what governs the selection of traits in animals and thus drives biological change, adaptation to niches, and speciation over time.The concepts of both agrarian s-curves and natural selection lent themselves to some unpleasant conclusions by nineteenth century thinkers that are worthwhile discussing. Death is central to both systems. In s-curves, death is the only way to relieve the pressure caused by growth in population outstripping the increase of agricultural production. In natural selection, when resources are limited, death is inevitable for those individuals who are outcompeted for resources in their ecosystem and their extinction is a necessary part of evolution. In both systems, the well-being of the individual is endangered, the well-being of the whole is impossible, the suffering of many is inevitable, and death is, from a certain point of view, desirable, because without it neither evolution nor agrarian civilisations could sustain themselves. The problem was that Malthus continued to apply his stark principles to human populations in his own period. Malthus had no way of knowing he was living in the middle of a transformation – an explosion of production no less significant than the transition from Palaeolithic foraging to the greater productivity of agriculture. The carrying capacity was being raised by new technologies, and the recurrent population crises of the agrarian era seemed to be coming to an end. Industrial production was increasing the productivity well ahead of prices, thus raising the carrying capacity and rendering Malthus’s principle of population less and less applicable. Unfortunately, Malthus and his adherents continued to apply his logic to industrial society for much of the nineteenth century even while his ‘iron law’ of population continued to rust. The Malthusian mindset impeded calls for the improvement of working conditions, the introduction of modest welfare services, and efforts to better the condition of the poor because it suggested that the death of the poor and starving was part of a ‘natural’ process that should not be tampered with. 206

Big history and historiography

The writings of Malthus were taken up by many nineteenth century thinkers to justify lack of support for the poor, oppose social reform, avoid ‘encouraging’ the tendency toward overpopulation, and view the suffering and death of the poor as inevitable or even a positive thing. This hard-nosed ideology created needless suffering in the nineteenth and twentieth centuries, since for one of the first times in all of human history, the carrying capacity was rising fast enough to support a rapidly increasing population. This tendency also provoked opposition to Malthusianism, particularly in the writings of Karl Marx, who stated that food shortage was the result of social structure not of ecology, something that echoed down the years to the debate between neo-Marxist historians and the neo-Malthusian annalistes in the 1970s and 1980s.

Ricardo identifies how overpopulation widens the wealth gap David Ricardo, a contemporary of Malthus, further developed the concept of demographically driven long cycles and one of his greatest contributions for our purposes was that he showed how the upper echelons of society profited from circumstances of overpopulation and stagnating production.19 Population growth might decrease wages, which increases profits, because more labourers competing in the workplace mean their employers will not have to pay them as high. If one presumes that there are no improvements in agricultural production (and indeed the rate of innovation in the agrarian era was relatively slow), the amount of profit and capital increases steadily with the growth of population until all available land is brought under cultivation. Thereafter increasing production for more people requires more capital put into either labour numbers or land/infrastructure development, eventually leading to a diminished return for a proportionally similar input. As the input of a landholder on his property increases for the same return, he raises the rents on his land to compensate. This dynamic is accentuated by overpopulation, high property prices, and a general shortage of land. At the same time, increased population, increased demand, and the increased cost of producing the same output raises the price of food, which greatly profits the landholder, much to the detriment of the labourer. As a result, the interests of the landlord are ‘always opposed to the interests of every other class in the community’.20 Unfortunately, this sort of trend does not create a wealthy nation, as Ricardo points out. It creates a wealth inequality gap. A nation’s wealth is based not on the price of goods but on the amount of goods that are circulating through a market. This is why Ricardo disagreed with Malthus about imposing the Corn Laws, and suggested a form of free trade to lower prices on goods and thereby to enhance the standard of living of the average subject. I do not wish to wade into the morass of ­Ricardian economics (including the idea of   ‘natural wages’) and while Ricardo vastly developed the Malthusian question in other respects, by far his greatest contribution for our purposes is to point out how long cycles in population have a varying impact on different social orders. In times of low population, wages are high and prices are low and the worker lives well. In times of high population, the landowner/producer profits from the overabundance of cheap labour and the high price he can assign to essential goods. Thus in all long cycles, there is a class divide. He did not delve into the social dynamics of an actual population crash. 207

David Baker

Twentieth century economists search for long durée cycles In the twentieth century, the study of long cycles drifted to the Continent in the works of Nikolai Kondratiev, François Simiand, Jenny Griziotti-Kretschmann, and Wilhelm Abel. Nikolai Kondratiev was a Russian economist who in 1925 published a work that identified roughly 50–60 year cycles of alternatively rapid and slow growth.21 He soon fell afoul of the Soviet regime for his opinions, was successively fired, shipped off to a Gulag, and eventually executed for his pains.The cycles, according to Kondratiev, were divided into periods of expansion, stagnation, and decline. In this way, his cyclical model was very similar to the cycles that ecologists study in demographics. Although a model for economic cycles, it is still an example at an attempt at longue durée cycles. Observing how economies in the late eighteenth and throughout the nineteenth centuries seemed to vary in pace of growth over time, he assigned the cause to the shaky logic that at some point the market is saturated with a certain major product. This could occur for a variety of reasons later guessed at by a variety of theorists. Unfortunately, no sound logic was crafted to explain these economic fluxes, and today ‘Kondratiev cycles’ are not accepted by the majority of economists. However, the idea of a division of growth into phases of expansion, stagnation, and decline is another major contribution to theories of long waves when assigned to more demographic and structural mechanisms, rather than just economic ones. In the meantime, work continued to be conducted in Western Europe exploring long cycles. In 1932, François Simiand looked at price movements over the early modern period and identified a series of long waves of rise and decline in the cost of various goods.22 Simiand hypothesised that the driving force was a change in the availability of precious metal for currency, which then impacted the rate of inflation. At no point did Simiand or his followers give primacy to a strictly demographic driving force, and erred more on the side of monetarist explanations. At best they suspected some proximate change in the yield of the harvest drove the fluctuations in prices.23 These explanations were never fully satisfactory, even to their authors, and lacked a strong empirical base. Not long after, however, Italian historian Jenny Griziotti-Kretschmann produced a work that was strongly couched in empirical evidence.24 Her findings did not conform to the Kondratiev sequence of 50–60 year waves, did not rely on its theory of saturation, and also were at variance with ­Simiand’s explanation of the availability of precious metals. Instead Griziotti-Kretschmann assigned the fluctuations to proximate changes in the structure of the economic and political systems of the time. This unfortunately could not account for the fact that similar waves struck regions with vastly different economies and political systems.

Wilhelm Abel revives the Malthusian model The turning point in this period of research came with the work of  Wilhelm Abel in 1935, which looked at Western Europe from 1200–1900 and contained a wealth of evidence, time-series information, and clear illustrations of the fluctuation in prices of various goods, along with corresponding graphs on wages, rents, and population numbers.25 He also identified several long waves of prices: an increase c.1200–1300, a decline 1300–1450, an increase 1450–1600, followed by stagnation and decline 208

Big history and historiography

1600–1700, and an increase 1700 and erratic fluctuations during industrialisation in the nineteenth century, before declining by 1900. Abel was more successful than his contemporaries in making sense of the price waves because he placed an emphasis on grain prices, which as the most basic amenity took precedent over all other goods in an agrarian economy, while other authors had focused on a collection of items. Abel also measured prices not in currency but in kilograms of silver to avoid clouding the data. Abel did not assign the cause of these fluctuations to the availability of precious metals or currency. Instead he identified a correlation between population levels and food prices and an inverse correlation between those and wages. It appeared that an increased population raised the price of food due to increased demand and lowered that of wages due to an oversupply of labour. A strong demographic element closely tied to population growth seemed to have a strong influence over the fluctuation of these waves. At a stroke the Malthusian model was revived. Naturally, as pointed out above, the Malthusian model failed to make sense for the Industrial Revolution. But for the era of agrarian civilisations, a clear explanatory mechanism for longue durée cycles was found. The slow rise of the carrying capacity between the invention of agriculture and advent of industry was responsible for s-curves in the population every few centuries even while the world’s population grew as a broad trend. A population grew rapidly in the agrarian era, outstripped the yield of the land, and crashed. Meanwhile, at large scales of many centuries or millennia, the yield of the land was slowly raised as generations passed by due to improvements in technology and agricultural methods. This was a result of what we refer to in big history as collective learning.

Is the driving force of history demographics or the class struggle? After the Second World War, a number of Marxist historians took up the demographic question. In 1946, Maurice Dobb looked at the population decline in Western Europe c.1315–1450 and saw it as the result of an increasing number of elites proportional to the total population, who were growing more aggressive in their conspicuous consumption. The peasants were then squeezed so badly they started dying.26 All of this is built up into the Marxist historical concept of a ‘crisis of feudalism’ (echoed by Guy Bois) as part of reconciling historical facts to Karl Marx’s model of a transition from feudalism to capitalism. Dobb did not employ a pure Malthusian model, but preferred to attribute the driving cause to social structure. The intensification of feudalism, according to Dobb, led to the death of large numbers in the population, which in turn backfired by destroying the incomes of the landowning class and caused the disintegration of feudalism. Paul Sweezy, another Marxist who, writing in 1950 said that Dobb’s book on the transition from feudalism to capitalism was important because they were living in an age of ‘transition from capitalism to socialism’, nevertheless took Dobb to task on this theory. Sweezy asserted that Dobb does not adequately explain why feudalism became so intense at that point in time. He also doubted whether there was a significant growth in the numbers of the landowning class. Finally, Sweezy hypothesised that the growth in number of elites and their extraction of resources from the masses 209

David Baker

was due to the expansion of trade following what some medievalists call the ‘Great Leap Forward’ of the eleventh century.27 This implies a longer process of transition from feudalism to capitalism than some models that place it in the sixteenth or seventeenth centuries. Both Marxists treat population fluctuations as a mere symptom of the changes in social structure. They saw little connection between population fluctuations and changes in modes of production. Other Marxists did not find the demographic factor to be at all compatible with the model of a transition to feudalism. E.A. Kosiminsky, for instance, completely denied a long-term population decline 1350–1450, though this position is roundly contradicted by a vast mass of historical evidence.28 To varying degrees, at this early stage, Marxist historians did not assign much importance to the demographic fluctuations and long cycles in wages and prices identified by Abel. Their focus was primarily on the shifts in social structure that may or may not have been associated with them. Around the same time, however, economic historian Michael Postan, a British historian and refugee of the 1917 October Revolution, rejected the theory that price fluctuations were determined by the availability of precious metals. He assigned prime importance to the demographic factor. Working mostly with English sources, Postan’s attack concerned monetarist theories and he more or less reaffirmed the findings of   Wilhelm Abel, whereby there is a correlation between population numbers, prices, the real wage, rents, and so forth. He was more ambiguous about the role elites played in these cycles and what sort of dynamics they experienced. For the most part, however, Postan found that overpopulation had reduced living standards and raised prices at the end of the 1200s and early 1300s and identified a reversal of this trend following the Black Death and the Hundred Years War. He did not like to associate himself with Thomas Malthus, however, preferring either to identify with a ‘Ricardian model’, or, more emphatically, to assert the uniqueness of the new demographic position. While virulently anti-communist throughout his career, Postan did not explicitly reject the application of the Marxist historical model or the impact of changes in social structure on population, and once referred to Karl Marx as ‘that universal genius’.29 Emmanuel Le Roy Ladurie was less hesitant to associate himself with the theories of Malthus. Ladurie was the disciple of Fernand Braudel in the French Annales school, which also boasted scholarly greats like Lucien Febvre, Marc Bloch, George Duby, Ernest Labrousse, and Jacques Le Goff. The Annales school was frequently opposed to the philosophy of Marxist historians and rejected class conflict as central to historical change. The implications of Malthusianism in the medieval and early modern period were already troublesome to the interpretations of some Marxist historians, and when adopted by a traditionally hostile school it had all the makings of a bitter debate. Ladurie’s mentor, Fernand Braudel, taught him to look beyond largely ‘ephemeral’ social structures toward longer, deeper, hardly perceptible processes that make up the Braudelian concept of the longue durée.30 Ladurie’s work also came along in the middle of a trend toward demography in French historical studies.31 While Braudel gave birth to the concept of the longue durée, Ladurie articulated a plausible theory about its mechanics and driving force.

210

Big history and historiography

Ladurie’s doctoral thesis, Les paysans de Languedoc was published as a book in 1966 and in English in 1974. Wholeheartedly adopting the interpretation of Wilhelm Abel and applying it to the data on population figures, prices, and wages in the medieval and early modern period, Ladurie’s work was the first systematic example of how history could be written from the perspective of long demographic cycles. He demonstrated, using numerous historical examples, how a population expands ­beyond subsistence levels, degenerates into famine, plague, and war, until the population is low enough to return to prosperity, a high standard of living, and renewed growth. He also pointed to the cyclical nature of this process and also noted how Malthus was writing at the end of this period of cycles. Ladurie’s work sent shockwaves into the English-speaking world.32 It changed the way many historians viewed history, not only the demographic question, leading to a flood of related works on population, prices, and wages, the foremost of which was the vastly extensive work by Jacques Dupâquier.33 Ladurie also polarised the debate. It did not help that Ladurie adopted an increasingly fanatical Malthusian tone and was very hard in his rejection of key Marxist concepts, for instance, ‘it is in the economy, in social relations, and even more fundamentally, in biological facts, rather than the class struggle, that we must seek the motive force of history’ and that structural factors like class were ‘meek’ before the great Malthusian ‘forces of life and death’.34 With language like that, it is no wonder that Marxist historians, at their intellectual and scholarly height in the 1970s, got their backs up. For the neo-Malthusian faction evolving under Ladurie, class was largely excluded from the question of a broader historical evolution through long cycles. For the Marxist historians, for whom the class struggle was central, this was intolerable.

The Brenner Debate It was these febrile conditions that kicked off what is known as ‘the Brenner Debate’. In a 1976 edition of Past and Present, Robert Brenner, a fairly devoted and doctrinaire Marxist historian who never, unlike many of his generation, dropped or moderated his adherence to Marxism after the Cold War, attacked what he called ‘demographic determinism’ as obscuring the ‘real’ processes that drive historical change, i.e. class.35 For Brenner, it is not an exaggeration to say class structure was all-consuming and always had to be analysed in order to understand long-term economic development. The class struggle had outcomes that directly determined how an economy evolved from feudalism to capitalism. This is why Brenner objected to demographic explanations since after the Black Death, Europe took different courses. Western Europe lost serfdom, Eastern Europe regained it. England made a gradual move to constitutional monarchy, France moved to absolutism. Such was his reasoning. Brenner nevertheless tipped his cap to Ladurie for producing something with compelling logic that ‘seems almost foolproof ’. But Brenner pointed out that the Malthusian model was not enough to explain why the population stayed low for a 100 years after the Black Death. According to Malthusian theory, overpopulation indeed could have provoked population decline, but after a drop-off as catastrophic as the Black Death, one should have seen

211

David Baker

the population rapidly rebound due to the sudden drop in food prices and hike in wages due to a labour shortage. Indeed these things occurred, but the population remained low. For instance, the population of France had a population of 18 or 19 million before the successive waves of the Black Death but did not begin to recover from a low point of 10 million until after 1450, and only reaching 18 or 19 million again c.1550.36 Thus, Brenner concluded, the continuing stagnation of Europe in the fifteenth century had to be explained by analysis of class structure and the overexploitation of the peasantry by the elite. Brenner went on to assert that peasants had no incentive to abandon traditional life ways and were forced into capitalism by the exploiting classes. It must be noted, however, that Brenner drew fire from other Marxists for both factual and theoretical inaccuracies. Foremost among his Marxist critics was Guy Bois, who congratulated Brenner for his ‘courageous attack’ on the ‘Malthusian orthodoxy’ that was ‘crushing Marxist historiography in its tentacles’ but accused him of making up his mind about Marxist ‘generalisations’ and then imposing them irrespective of what was shown by the historical source material. Bois also questioned Brenner’s ‘ideological motivations’ behind his ‘unbalanced injection’ of class conflict as the determinant of historical development. Bois argued that you cannot explain the Industrial Revolution in the nineteenth and twentieth century purely by class structure, so why presume you can for the medieval and early modern period?37 He accused Brenner of oversimplification, particularly in relation to the differences between England and France, treating them as success and failure stories. Bois claimed that capitalism was a by-product of feudalism as a whole (ignoring Brenner’s own example of   West and East). Bois had, of course, recently completed his own masterful work on French population and prices in Normandy.38 The book was extremely well supported with empirical evidence, birth rates, population estimates, lists of worker’s wages in Rouen, and showed many demographic fluctuations after the fashion of Abel and Ladurie, but used a Marxist interpretation whereby class struggle and exploitation was the driving force of human historical change. Nevertheless, demography played a very central role in Bois’s interpretation. He admonished Brenner that the Malthusian model was a useful tool that should not be abandoned by the Marxist school purely because it appeared to overshadow its core principles. One had to adapt and incorporate. In fact, Bois advocated the wholesale removal of population dynamics from the Annales school and its annexation to the Marxist one. Unfortunately in his work and in subsequent commentary, Bois never fully managed to distinguish why his analysis was Marxist and not ‘neo-­Malthusian’ with a simple Marxist veneer. And indeed many of his conclusions and much of his research can be employed to follow the Malthusian dynamics of what looks and feels like the longue durée without even taking time to explain away the Marxist aspects of his analysis. In many ways it reads and feels like an annaliste history, as Ladurie himself was quick to point out.39 In contrast, the response of the ‘demographic camp’ to Brenner was fairly muted. Michael Postan, for instance, remained respectful of Marxist historiography and went no further than to claim Brenner ‘misrepresented’ his views and that demography was not omnipresent in all economic and social activity according to his model. Postan found the accusation that he ignored social factors was unwarranted. On the contrary, 212

Big history and historiography

Postan claimed, the Marxist view was too narrow. There is more to the feudal system than class conflict, he said. He accused Brenner of assigning the landlord too much power over the peasantry and pointed out they too were afflicted by the processes of population collapse, referring to their impoverishment and the high wages they had to pay after the Black Death right up to the sixteenth century. In relation to Brenner’s objection that East and West parted ways after the population decline of the fourteenth century, Postan pointed out that similar processes do not need to produce identical results. He also rejected the title ‘Malthusian’ and claimed that he could not be described as such unless all demographic theories could be described as Malthusian. He preferred to be called ‘Ricardian’, but even that, he said, did not quite fit the demographically driven theory that was coming into being.40 Indeed we still lack an accurate label for it. Ladurie, in his own response, was perfectly comfortable being classed as a ‘neo-­ Malthusian’ but claimed that this was not a Malthusian model, because the theory had advanced since then, as indeed it had. He denied that his model excluded class structure, but in fact incorporated it. He accused Brenner of completely ignoring the work of Wilhelm Abel. He accused Brenner of oversimplification, for relegating the surplus extractors and ruling classes into one group. He also cited the work of Guy Bois on Normandy, which highlighted many of the variables of Ladurie’s own theories and he claimed that the fact Bois called himself a Marxist added all the more objectivity to the corroboration of his theory. Ladurie also stated that Brenner greatly underestimated the role of epidemics in the fourteenth and fifteenth century, and assigned more credit to the exploitation of the upper classes as a cause of population decline than to the devastation of pathogens. Ladurie also denied that the presence or absence of serfdom confirms or disproves the validity of demography playing an enormous role in medieval history. He pointed out that the intensification of labour services in the thirteenth century directly coincides with predictions of a neo-Malthusian model in times of overpopulation, and he pointed out that the French system was not so very different from England as to be more authoritarian in later periods, since in France too serfdom diminished. The rise of serfdom in Eastern Europe, Ladurie said, was also not contradictory since there were also holdovers in Burgundy and the Franche-Comté, and it would appear serfdom was not dependent on demographic processes to thrive or decline. France did not escape Malthus in the seventeenth century like England, Ladurie claimed, not because of any significant institutional differences, but because agricultural reform and the raising of the carrying capacity only took place from 1720 (and not after 1800 as Brenner suggested). Ladurie accused Brenner of being too disparaging of French agriculture in the eighteenth century, which indeed underwent many significant improvements that for the first time in French history raised the carrying capacity from a peak of about 20 million people (attained in the early fourteenth, mid-sixteenth, and seventeenth centuries) to nearly 30 million by the end of the 1700s. Indeed Brenner’s desolate view of eighteenth century French agriculture is extremely antiquated. Historical research, including that explored later in this work, has shown that eighteenth century agriculture in France was extremely innovative. Finally, Ladurie said that surely the Marxist route of peasant disenfranchisement á l’anglaise was only one path to modernisation and capitalism. Peasant economies were perfectly capable of supporting the transition to capitalism, Ladurie pointed out, as seen in Holland, Belgium, South France, Northern Italy, Japan, and Catalonia.41 213

David Baker

The unexplained problem of population depression While Postan and Ladurie’s responses, in addition to the critique of Bois, diminished the credibility of many of Brenner’s objections to the demographic thesis, they still failed in one significant respect. None of them could account for the fact that the population remained low for another 100 years after the Black Death (known as a population depression). Indeed the population in many areas dropped even lower, as Guy Bois illustrated for Normandy. Brenner responded to his Marxist and non-­ Marxist critics with a lengthy article in which article failed to establish any new objection or convince many people of his own position.42 But the crucial blow was already struck. The demographic school could not adequately account for the population depression in Western Europe c.1350–1450. Ladurie and Postan were unable to mount an effective response or to adequately incorporate the social side of the equation in a mechanistic way and soon even the Marxist school gradually fell into decline and disrepute in the late 1980s and 1990s. There the matter rested. A second wave of criticism of the demographic camp came from monetarist and neo-classicist economic historians in the 1990s and 2000s. Prominent among them were David Weir, George Grantham, and Jan de Vries.They rejected both the Marxist and Malthusian models as somewhat antiquated. In the spirit of much history of the post-Cold War era, these historians gave up on the idea of a meta-theory for history since human activity was just too complicated to be condensed to a set of broad patterns or scientific-sounding laws. While little time was spent discrediting the Marxist school, since they largely took Marxism’s disgrace for granted due to political events, the neo-classicists brought up all the old arguments against the Malthusian case: that it fails to account for social complexity and diversity of economic and political structures and that it fails to account for sustained population depressions. None of them clearly articulated anything to replace those theories, nor do they seem to feel any need to, dwelling, as deconstructionists of all stripes often tend to do, on the ‘complexities’ of the situation. Grantham was known for putting forward notions of ‘cliometrics’ as a promising new discipline, whereby economics could be the driving force of historical interpretation. Jan de Vries, a Dutch scholar, put forward the idea of ‘the industrious revolution of the seventeenth century’ being responsible for the industrial revolution, claiming that the downturn of the economy in the seventeenth century provoked people to work harder.This idea has managed to hold on to credibility, but it does not fully account for regional diversity outside of Holland and England where the same pinch was felt with no indication of a shift, or an empirically supported transition between the ‘industrious’ to Industrial revolutions that was any more clearly indicated by the evidence than the old Marxist transition from feudalism to capitalism.43 The work of the neo-classicists and monetarists also largely ignored the findings of economists François Simiand, Jenny Griziotti-Kretschmann, and the decisive proof put forward by Wilhelm Abel. The papers cited here also have a tendency to set up a straw man target in Ladurie’s l’histoire immobile by pointing out the numerous instances of growth and innovation in pre-industrial Europe. Unfortunately, the term as Ladurie used it does not imply that no growth or innovation occurred, but rather than the carrying capacity that existed before the Industrial Revolution meant that 214

Big history and historiography

waves of growth were followed by periods of decline – one step forward, one step back – rather than complete stasis. Nevertheless, the decline of meta-theories in general after the Cold War and the general disparagement of Ladurie’s school meant that as far as mainstream economic historiography was concerned, the Malthusian vision of long cycles was effectively dismantled. The population fluctuations of the medieval and early modern era were moreover assigned to more proximate causes and no broad or repeating patterns were recognised to exist. Thus a void was created where the old Malthusian and Marxist meta-theories once were, which these post-Cold War monetarist and neo-classicist historians did not expect (or want) anyone to fill, except perhaps with a gradual accumulation of many short-chronology pieces of specialised research.

The return of meta-theory and an answer to the depression problem However, in the very same works in which the neo-classicists were tearing strips off the Malthusian theory, most of these historians were suitably impressed by the ‘demographic-structural’ model put forward by social theorist Jack Goldstone in 1991.44 Jack Goldstone not only was reintroducing the Malthusian mechanism into economic history (he calls himself ‘post-Malthusian’) but also the concept of long cycles and repeating patterns. Once more the idea of long cycles was revived. Furthermore, Goldstone’s idea of population growth kicking sociopolitical instability into social structures in cyclical patterns contradicts the rejection of broad patterns and ­meta-theories, the assertion that human activities are ‘too complex’ to interpret in broad patterns, and the reliance on more proximate causes for economic change. Goldstone took Ladurie’s thesis about overpopulation leading to cycles of rise and decline and used social structure to explain why the population was often held low by the political instability that followed the decline of the incomes of the elite. Instead of overpopulation being the direct cause of state collapse, it provoked the sociopolitical instability that was responsible for it. Goldstone’s focus was more on the cause of revolutions, but here at last was a closer answer to why populations remained in depression long after the initial fourteenth century crash than the Marxist or Annales schools could ever conceive. Another admirable attempt at sketching the ‘long cycles’ was by David Fischer. His focus was more on the role of price movements rather than illustrating a purely demographic process. Fischer did tie his price movements, however, very closely to population fluctuations, and in this respect is the successor of Wilhelm Abel. He identified the same waves that Abel did and Fischer also identified another wave of prices in the twentieth century and warned of impending crisis at some point in the twenty-first century.45 Finally, Ian Morris has in two books addressed the effect of population on human social development, created a rough metric for social development, which roughly accords with the rise and fall of population and sociopolitical stability charted by the others mentioned here.46 Of the writings on long cycles of the post-Cold War era, the most effective has been those of Jack Goldstone. Building on his ideas, an ecologist Peter Turchin came up with a synthetic theory that revived many of Ladurie’s ideas and reconciled them with the Marxist emphasis on power relations and surplus extraction. Suddenly there was no contradiction between the demographic and the structural. The ideas of 215

David Baker

Marxist and Malthusian schools in this regard were reconciled. Population pressure exerts a powerful influence over sociopolitical instability, historical events, and the ebb and flow of state power.The dynamics are so constant that they can be identified in cycles of a few 100 years of expansion and contraction.47 The cycles can be divided into two phases and four sub-phases, harkening back to Kondratiev: expansion and stagnation-inflation in the good times (imperiogenesis) and crisis and depression in the bad (imperiopathosis).48 The major variables in these cycles are population growth, prices, real incomes, elite numbers, the wealth inequality gap, and social cohesion, a borrowing from Ibn Khaldun’s asabiyyah, which can be defined simply as the capacity of a society to cooperate in an efficient, effective, and unified fashion.49 However, Turchin stressed that this in no way attempts to preach a mono-causal version of history. Far from it, a number of exogenous variables such as geopolitics, diseases, and climate change also apply. Initially, when a population is low, there is plenty of food, a relative labour shortage, and so food prices are low and wages are high. Most of the population of a nation enjoy a contented, prosperous, and perhaps even steadily improving standard of living. High living standards translate into political stability. A rebellion is seldom waged on a full stomach. As a result of these high living standards, the population tends to grow. Eventually a population approaches its carrying capacity. There are shortages of food and an oversupply of labour. Prices rise, wages drop, and the standard of living declines. Unless population pressure is relieved by agricultural innovation or territorial expansion, this can lead to a disastrous crisis. As the crisis point is approached, the average person is paid less and has to pay ever more for the basic essentials. Famines increase in severity, the susceptibility of people to disease also increases, as does the possibility of widespread epidemics. It also increases the level of civil unrest among the masses. At the same time as the crisis point is approached, it is a veritable ‘golden age’ for the elite. Landowners pay lower wages and charge higher rents. Middling landowners are forced off their farms and land coalesces in the hands of the few. The inequality gap widens. Elite incomes grow. Contented in this golden age, they are more likely to support the government in the defence of the status quo. But the significant fact is that elite numbers and appetites grow. Then the crisis point is reached. People starve, social cohesion collapses, the number of people living at subsistence level grows, grain reserves disappear, diseases ravage a malnourished landscape, there are rural and urban uprisings and, ultimately, people die. As the general population shrinks, the elites, cushioned by their status and their wealth, do not die at the same rate. The social pyramid becomes immensely topheavy. This is called elite overproduction. It makes up a period of stagnation in the common sphere and inflation of elite numbers and has come to be known as the ‘stagflation’ period. These elites, numerous as they are as a proportion of a dwindling population, begin to see their incomes shrink. And they do not like it. The result has been called ‘intraelite competition’ or ‘intraelite conflicts’ or put simply, elite infighting. Factions form, both against the government and against each other. While the government’s tax revenues shrink due to depopulation, it is increasingly unable to quell, confront, or control or buy off these factions. As a consequence, there is a period of intense and 216

Big history and historiography

bloody conflict, much more violent than coups in times of prosperity. The situation explodes into civil war. Around the same time the nation can become extremely vulnerable to aggressive external invasion. Additionally, elites may latch onto popular discontent and become mass leaders, and this renders popular movements, which in the past might have been easily suppressed, more potentially threatening, violent, and destructive. As a result, the first crisis, spurred mainly by demographic causes, is followed by a second crisis or ‘depression’ which is largely manmade. That is why a society following initial depopulation does not immediately rebound. The social side of the equation holds recovery down, and this can last for decades. Eventually, however, a population does rebound. Elite numbers are reduced. Low numbers in the general population combined with high wages and low food prices lead to another period of expansion, peace, and stability. The new golden age, however, comes at the tremendous cost of the preceding period of starvation and bloodshed (Table 9.1). Peter Turchin has already assigned chronologies to a number of full secular cycles: for England 1150–1485 and 1485–1730, for France 1150–1450 and 1450–1660, Table 9.1  Predictions for the theory of secular cycles, in any given phase Predictions

Expansion

‘Stagflation’

Crisis

Depression

Population Elite Numbers Social Cohesion Instability Rural Settlements Cultivated Land Free Land Peasant Land Land Prices

Increasing Low Increasing Low Increasing Increasing Abundant High Low, increasing Low High Low High

Deceleration Increasing High Low, increasing Slow increase Slow increase In short supply Low High

Decreasing High, faction Collapse High Decline Decline Increasing Low, increasing Declining

Decreasing/stagnation Decreasing Revival/Relapse High, declining Lack of increase Lack of increase Abundant High Low

Increasing Declining High Declining, poverty Declining Increasing Increasing Increasing Increasing Increasing Increasing Sluggish recovery High

High Increasing Declining Subsistence

Decreasing High Low Variable

Nonexistent High High Declining High Many High No recovery

Variable High, declining Declining Low/Local Decreasing Decreasing High, declining Variable

Crisis

Fragmentation

Grain Prices Real Wages Rents Consumption Grain Reserves Urbanisation Artisanship Trade Usury Large Estates Inequality Epidemics Internal Peace

High Low Low Low/Local Absent Few Low Quick recovery Increasing

(Continued) 217

David Baker

Predictions

Expansion

‘Stagflation’

Crisis

Depression

Coin Hoarding State Finances

Decreasing Increasing

Increasing Bankruptcy

High Poor

Taxes

Increasing

Crisis

High/Collapse

Ideology

Optimistic

Low High expenditure Heavy burden on peasantry Social Pessimism

Pessimistic ideology, cult of death

State Policy

Laissez-faire Interventionist

Popular movements for social justice Social reforms, revolutions

Weakening of state

Based on Turchin and Nefedov, Secular Cycles, 33–34.

for Rome 350–30 BC and 30 BC–285 AD. Nefedov has done the same for Russia 1460–1620 and 1620–1922. There is still a great deal more to be done, however. Turchin and Nefedov study each of these cycles in a very short space. There is understandably not room for much more than a survey in their book. What is more, in spite of the plausibility of the theory,Turchin is an ecologist, not a historian. He tends to gloss over historical trends that do not fit his model and assigns dates that accord better with major political events rather than real changes. For instance, his treatment of the Western European medieval cycles dates their starting point from 1150, even though demographic recovery probably began at least a century before then. His treatment of the Roman Empire glosses over upheaval and demographic decline in Italy in the first century AD. Turchin himself acknowledges that the dating of cycles is not very precise. It is questionable whether he has even very approximately assigned the correct dates to each major phase: expansion, stagflation, crisis, and depression, or whether this is sacrificing empirical accuracy for theoretical symmetry with wellknown events. In addition, Turchin does not adequately account for the influence of exogenous factors. Kohler et al. test his theory for the population growth and violence seen in Pueblo societies and found that generally the model holds true except when influenced by climate, external war, and so forth. They also point out, very rightly in my opinion, that a complex phenomenon such as war cannot have a single cause.50 Yet if you read Turchin’s treatment of the French cycle from c.1150–1450 you almost gain the impression that the Hundred Years War was included in the symptoms of an imperiopathosis phase, even though he claims the invader, ­England, was undergoing the same phase at the time. External warfare really ought to be counted as an exogenous factor and human agency ought to be given more credit – in regard to the Hundred Years War, this included dynastic feuds, diplomatic quarrels, and ­centuries-old territorial claims. The core of the theory is plausible but it still requires the constant testing of both ecologists and historians. At any rate, this is the theory of secular cycles, a reconciliation of the Malthusian and Marxist schools of the 1970s and 1980s that employs both population change and social structure as the driving forces of history. 218

Big history and historiography

Reconciling Marx and Malthus to describe the longue durée The idea that history moves in cycles or waves of some kind is not a new idea. The idea goes back many centuries. Modern population theorists like Malthus and ­Ricardo, economists like Wilhelm Abel, and historians like Le Roy Ladurie have all illuminated possible mechanisms for these waves and have gone to pains to show their inner workings.The resulting historical debate in the latter half of the twentieth century fell between the ‘neo-Malthusians’ and Marxists, a divergence that harkened back to the diverse philosophies of Malthus and Marx themselves. The ‘Brenner debate’ was never really resolved to anyone’s satisfaction. In fact, both schools of thought on the question of long cycles and ‘driving forces of history’ fell by the wayside as the concept of meta-theory itself became rather unfashionable.Two related attempts have been made to revive the theory and reconcile the Malthusian and Marxist camps by Goldstone and Turchin. Neither of these scholars are historians.The former is a social theorist and the latter is an ecologist with a background in the sciences. The question now is whether such population-based theories of long cycles can find a home again within academic history. There is also the question of whether the implications of s-curves in the agrarian period can be connected to population development beyond the 10,000 year period between agriculture and industry. Are the mechanisms explored by Goldstone and Turchin merely linked to a particular epoch, or are they linked to the entire domain of human history? Indeed do they perhaps link to mechanisms that exist outside of human history that stretch into the biosphere or even the cosmos?

The ‘plus longue durée’ of big history When it comes to demographic cycles, it is clear that human behaviour is more complex than most animals that ecologists study in nature. Our history is not just influenced by traditional Malthusian dynamics but instead our complex social structures and hierarchies create conditions that can hold a human population low long after an initial collapse. Furthermore, a rise of sociopolitical instability in those social structures can often be more destructive than more ‘natural’ population decline. These influence micro-historical events. No matter how complex those human cycles are, however, they find their ultimate foundation in population dynamics. Underpopulation creates conditions of prosperity and social stability, overpopulation tends to provoke the opposite. And since human historical cycles may be influenced to a great degree by population dynamics, this means that much of human history is dependent on the state of the ‘carrying capacity’. As long as the species has existed, Homo sapiens has needed to innovate to be able to extract more and more energy from the environment and sustain ever larger populations. The alternative is periods more familiar to other living species, where the ecosystem is overburdened and exhausted leading to population declines (a process which, as Darwin discovered from reading Malthus, drives natural selection). It is here that Ladurie and Malthus link up with Darwin. Humans, however, possess the ability to accumulate more innovation with each passing generation than is lost in the next, a concept we in big history call collective learning. Over many centuries or millennia, total population levels do not remain 219

David Baker

stagnant. Although in the agrarian period collective learning was not fast enough to stave off longue durée cycles of overpopulation and decline, ultimately collective learning raised the carrying capacity and the overall trend of the human population of the Earth has been upward. Some of that knowledge improved agricultural techniques, which produced more people, who produced more knowledge, which raised the carrying capacity. Collective learning thus explains the steady curve of human population growth from the beginning of our species 300,000 years ago. Useful techniques were selected and improved upon by countless generations, and in this sense, collective learning forms a part of the wider scheme of cultural evolution, wherein humans generate many variations of ideas and apply them to the extremely wide range of viable selection paths that form human culture. As a result of this, human populations today harness more free energy density (or complexity) than anything else that we know of in the Universe.

Widening the lens: collective learning and population From the origins of collective learning in the Palaeolithic, it is clear from the rising carrying capacity and increase in cultural variants and innovations that collective learning has great bearing on the narratives that were explored by Braudel and Ladurie. Nowhere is this more relevant than the discussion of population cycles. Our starting point is the low watermark of the human species from the Toba event 74,000 years ago when a volcanic eruption reduced the population to a few 1,000, to our increased innovation and migration out of Africa, and the origins of agriculture through accumulated innovations in plant and animal domestication. From there we explore how innovation gradually raises the carrying capacity allowing the population to grow in the long term, though not without many cycles of rise and decline in the agrarian period.What Ladurie explored in early modern France was a microcosm of the population dynamics that reigned throughout the era of agrarian civilisations. It is the impact on human population dynamics where cultural evolution has had the most perceptible impact on human history: innovation produces more people, who produce more innovation. Eventually, there is so much generation of new ideas that the likelihood of a useful one is greatly enhanced. These ideas are selected and accumulated along with other ideas. Gradually, this raises the human population to impressive heights and the Darwinian algorithm in a new form of cultural evolution goes into overdrive. The inception of the current arc of complexity is easily spotted. Around 74,000 years ago there was a catastrophic eruption at mount Toba, on the island of Sumatra, part of what is now Indonesia. It was worse than anything in recorded history. The eruption drastically lowered temperatures on Earth for several years.51 Genetic studies show that the resultant decline in flora and fauna upon which humans could predate had reduced the population to near extinction. It is likely that in the aftermath of a period of starvation, on the entire face of the Earth there were scarcely more than 10,000 (and perhaps as few as 1,000) human souls, which, as an aside, is what makes our long history of racism so abhorrent and absurd, particularly those ideological impulses inspired by Darwinism.52 Here is a low watermark for the current trend of human population dynamics. Evidently, the starvation did 220

Big history and historiography

not last long. In approximately the same amount of time that separates us from the dawn of agriculture, the human species had recovered and c.60,000 years ago some of them migrated out of Africa across the world. By 30,000 years ago, the foraging human population had risen to half a million. By 10,000 years ago, the innovation of ­hunter-gatherer bands had allowed them access to almost every environment on Earth, from Eurasia to Australia to the Americas. We must remember that the carrying capacity for a foraging band is quite low and they need a vast area to supply relatively small numbers. Nevertheless, by the dawn of agriculture, the ranks of our species had swelled to 6 million people, approaching the full capacity for supporting ­hunter-gatherers of which the entire surface of the Earth is capable.53 Innovations began to mount up. Earliest recorded evidence for herding goats and sheep in Southwest Asia is 11–12,000 years ago, and 1,000 years later, we have evidence for the farming of wheat, barley, emmer, lentils, and pigs. By 8,000 years ago, East Asia had begun using millets and gourds, and the Americas had domesticated llamas and maize. By 6,000 years ago, Southwest Asia had domesticated dates and the grapevine, while East Asia had domesticated water chestnuts, mulberries, water buffalo, and that mainstay of all Asian crops: rice.54 All of a sudden, much larger numbers could be supported over a much smaller land area. Agrarian civilisations brought about a greater degree of connectivity, faster population growth, and a new rapid pace for innovation. Suddenly there were a lot more minds to generate ideas and a lot less space between those minds in order to conference. Agricultural efficiency gradually improved, and practices slowly spread to new regions. From the upper limits of the carrying capacity for foragers, the population increased nearly tenfold by 3000 BC to 50 million people, and it took only another 2,000 years to increase this number to 120 million.55 But there was a problem. Collective learning tends in the long run to cause demographic growth, but it does not automatically guarantee that the right innovation will arise as a population approaches its carrying capacity, or the numbers of humans that can be supported in a given environment with a certain arsenal of technology that can produce a certain amount of productivity. The tinkering of ideas in cultural evolution is random, after all. For nearly 10,000 years, the growth in the carrying capacity of agriculture was sluggish compared to population growth, and so there was a series of miniature waves of population collapse and recovery throughout the period of agrarian civilisations. Not all learning was lost in each decline, so innovation accumulated in each cycle. The next expansion period brought renewed accumulation and thus gradually raised the carrying capacity. But it was not fast enough. The demographic time scale was short because human populations grew so quickly, while the accumulation of new ideas took longer.Thus you have the lag between the introduction of agriculture and the Industrial Revolution by 10,000 years in which population pressure assiduously bore down upon the rate of growth. If collective learning can keep producing more to keep ahead of subsistence and starvation, the population continues to grow. If not, then the population temporarily bows to population pressure. As we shall see, this causes sociopolitical instability and sometimes periods of lengthy depression, elite infighting, civil war, and susceptibility to foreign invasion. This is the stage of the most risk, where learning can actually be lost, thus reducing the carrying capacity for when growth does resume. If, however, minimal 221

David Baker

or no learning is lost in those periods of decline, the carrying capacity continues to grow regardless of whether the population is in expansion, stagflation, crisis, or depression. Gradually the population grows as a long trend, regardless of the waves of mini s-curves when the population grows too fast, hits the carrying capacity, and temporarily recoils. Richerson, Boyd, and Bettinger hypothesise that after a rapidly growing population has hit the carrying capacity, it enters a ‘steady state’ population growth rate, where growth bows to population pressure but is gradually raised by innovation (Figure 9.1).56 For most of the Holocene, the ‘steady state’ growth rate was hardly steady. The gentle curve belies the brutal nature of the population ramming against the carrying capacity (Figure 9.2). If we take the average length of a secular cycle to be a few 100 years, and we allow for many variations, this graph would represent the life-cycle of the last few 1,000 years of an agrarian civilisation between its adoption of agriculture and the introduction of heavy industry. The population losses of these longue durée cycles in periods Initial Growth Rate

Steady State Pop Growth Rate Pop Size

Figure 9.1  Relationship between population growth and carrying capacity (Richerson, Boyd, Bettinger, 219).

Exponential Pop Growth

Pop Size

Agriculture

Industry

Figure 9.2  R  elationship between s-curves and carrying capacity. The asterisk (*) marks a period of severe population decline where learning is lost (Richerson, Boyd, Bettinger, 219). 222

Big history and historiography

of decline are not equal. Some disintegrative phases can be exacerbated by plagues, famines, or extended periods of civil warfare. Furthermore, the rate of growth is not constant either, since each innovation does not exert exactly the same amount of influence on the carrying capacity. Some innovations like the adoption of a better design of plough may have a slight improvement on total food production, whereas the adoption of legumes in a four-crop rotation can have a profound impact on the carrying capacity. The single cycle surveyed in Ladurie’s Peasants of Languedoc forms just one cycle in a long stream of cycles. And within each cycle comes changes to average standards of living, the wealth inequality gap, and the frequency of sociopolitical violence – and ultimately the rise of factions and the flow of micro-historical events.

Widening the lens: complexity So far we have gleaned that micro-historical events can be shaped by what phase of a population cycle a historical society is undergoing. These population cycles are shaped by the carrying capacity. The carrying capacity is determined by collective learning. And in turn, collective learning is a complexity generator that evolved in Homo sapiens. But it goes even deeper than that. A split second after the Big Bang, between 10−36 and 10−32 seconds, the Universe inflated from the quantum scale (much smaller than an atom, perhaps much smaller than a proton in that atom, or a quark in that proton) to the Newtonian scale (about the size of a baseball). During that time, quantum fluctuations that still happen all the time in nature were suddenly writ large in the Universe, something that is still reflected in C ­ osmic Background Radiation.57 It is these fluctuations that disturbed the perfect homogeneity of the Universe, creating tiny gradients of energy. The unequal distribution of energy allowed for energy flow, in accordance with the 2nd law of thermodynamics. It was from these flows of energy that all complexity in the Universe is created, sustained, and increased. Within tiny pockets of energy flow like ours, the density is temporarily increasing from the inanimate Universe, to biological life, to human culture (Table 9.2). Table 9.2  Amount of free energy running through a gram per second, and the australopithecine and human free energy rate density is determined from the average energy consumption of an individual, Chaisson 2010: 28 & 36 Generic structure

Average free energy rate density (erg/s/g)

Galaxies Stars Planets Plants Animals Australopithecines Hunter-Gatherers (i.e. 250,000–10,000 y/a) Agriculturalists (i.e. 10,000–250 y/a) Industrialists (i.e. 1800–1950) Technologists (i.e. present)

0.5 2 75 900 20,000 20,000 40,000 100,000 500,000 2,000,000

223

David Baker

All complexity needs more energy flows to keep on going. All complexity must confront scarcity. Otherwise stars burn out, life starves, and societies collapse. Some complexity evolves methods of coping. Biological life has natural selection to better extract energy from the environment. Humans evolved the capacity for collective learning to cope with this scarcity still further. Ultimately these two things have raised complexity to astounding heights. All of this should be familiar to a reader of big history. But its reach goes deep in ways that the scholars examined above probably did not expect. It is thanks to this wider need to sustain our own complexity that humans evolved collective learning. And it is thanks to collective learning that we see these unusual cycles that govern micro-historical events appear in conventional human history. In many ways, the civil wars, the societal collapses, the famines, are all manifestations of the temporary or permanent failure of a human society to find enough energy to sustain its own complexity.The Yorkists vs. the Lancastrians in the Wars of the Roses, for instance, are a symptom of exhaustion creeping up on a complex system. The inability of Late Medieval English society to channel more energy flows from the Sun into more crops to support a burgeoning population, thus being hit by famines, plagues, and the elite competition over the resources that remained, spawned the aforementioned civil war. All the turmoil and events of human history are symptoms of the wider trend of intensifying energy and the occasional exhaustion of those flows. Moreover, to widen the lens still further, all of big history can be summed up in one single sentence: ‘the movement of energy from where there is more to where there is less’. Without this movement, no complexity could have been created or sustained. Big history would have stopped before it began. Had the Universe started perfectly homogenous, there would have been no events, no change, no complexity, and ultimately no history.This obviously includes the micro-historical events that are driven by population, collective learning, and the continuum of life and chemistry from which they sprang. It is in that way that big history does not just explore broad trends over 13.8 billion years. It directly influences in a very mechanistic and predictable way the micro-historical events of the past 5,000 years of conventional history. It connects directly with the meta-theories that were devised and discarded in the twentieth century and links up directly with the socio-demographic cycles proposed by Goldstone and Turchin. It is in that sense that the very large reaches out across billions of years to grasp hands with the very small. Even while a historian investigates the particular, they must be conscious of the thermodynamic totality in which they are working.

Notes 1 Fernand Braudel, The Mediterranean and the Mediterranean World in the Age of Philip II, vol. 1, trans. Sian Reynolds (Berkeley: University of California Press, 1996), 21. 2 Fernand Braudel, ‘History and the Social Sciences: The Longue Durée’ Annales. ­Histoire, Sciences Sociales 13.4 (1958) 726. 3 Braudel, ‘The Longue Durée’, 730. 4 Emmanuel Le Roy Ladurie, The Peasants of Languedoc, trans. John Day. Urbana: ­University of Illinois Press, 1976), 3–4. 5 Ladurie, Peasants of Languedoc, 289. 224

Big history and historiography

6 Emmanuel Le Roy Ladurie, ‘History that Stands Still’ in The Mind and Method of the Historian, ed. Emmanuel Le Roy Ladurie, trans. Sian Reynolds and Ben Reynolds (Brighton: Harvester Press, 1981), 1–27. 7 Ladurie, ‘History that Stands Still’, 3. 8 Ladurie, ‘History that Stands Still’, 7. 9 Ladurie, ‘History that Stands Still’, 9–10. 10 Plato, The Republic, trans. Allan Bloom (New York: Basic Books, 1991 [c.380 BC]), II:372–373, Aristotle, A Treatise on Government, trans. William Ellis (Charleston: Forgotten Books, 1947 [c.330 BC]), II:6. 11 P.-T. Ho, Studies on the Population of China: 1368–1953 (Cambridge: Harvard University Press, 1959), W.-L. Chao and S.-C. Hsieh, History of the Chinese Population (­Beijing: People’s Publisher, 1988), and C.Y. Cyrus Chu and Ronald Lee, ‘Famine, Revolt, and Dynastic Cycles: Population Dynamics in Historical China’ Journal of Population ­Economics 7 (1994) 351–378. 12 Tertullian, De Anima, ed. Jan Hendrik Waszink (Leiden: Koninklijke Brill, 2010 [c.200 AD]), 30. 13 Estimate of world population from Massimo Livi-Bacci, A Concise History of World Population, trans. Carl Ipsen (Oxford: Blackwell, 1992), 31. 14 Peter Turchin and Sergei Nefedov, Secular Cycles (Princeton: Princeton University Press) 233–236. 15 Niccolò Machiavelli, Discourses on Livy, trans. Leslie Walker (London: Penguin Books, 1984 [1517]), II:5. 16 Ibn Khaldun, Muqaddimah, trans. Franz Rosenthal (Princeton: Princeton University Press, 1967), 2:272–278. 17 Thomas Malthus, An Essay on the Principle of Population: As It Affects the Future Improvement of Society with Remarks on the Speculations of Mr. Godwin, M. Condorcet, and Other Writers (London: J. Johnson, 1798). 18 Charles Darwin, The Autobiography of Charles Darwin, ed. Francis Darwin (London: John Murray, 1887), 82. 19 David Ricardo, An Essay on the Influence of a Low Price of Corn on the Profits of Stock: Showing the Inexpediency of Restrictions on Importation with Remarks on Mr. Malthus’ Last Two Publications (London: John Murray, 1815). 20 Ricardo, An Essay on the Influence, 20. 21 Nikolai Kondratiev, The Major Economic Cycles (Moscow: Voprosy Koniunktury, 1925). 22 François Simiand, Les fluctuations économiques à longue période et la crise mondiale (Paris: Félix Alcan, 1932) and Recherches anciennes et nouvelles sur le mouvement général des prix du XVIe au XIXe siècle (Paris: Domat Montchrestien, 1932). 23 C.-E. Labrousse, Esquisse du mouvement des prix et des revenus en France au XIIIe siècle (Paris: Dalloz, 1932). 24 Jenny Griziotti-Kretschmann, Il problema del trend secolare nelle fluttuazioni dei prezzi (­Pavia: University of Pavia, 1935) and ‘Richerche sulle fluttuazioni economiche di lungadurate’ Giornale degli Economisti 73 (1933), 461–508. 25 Wilhelm Abel, Agrarkrisen und Agrarkonjunktur (Berlin: Verlagsbuchhandlung Paul Parey, 1935) and the English reprint Agricultural Fluctuations in Europe: From the Thirteenth to the Twentieth Centuries, trans. Olive Ordish (New York: St. Martin’s Press, 1978). 26 Maurice Dobb, Studies in the Development of Capitalism (New York: International Publishers, 1946), 42–47. 27 Paul Sweezy, ‘The Transition from Feudalism to Capitalism’ Science and Society 14 (1950) 134–167, esp. 142–144 and Paul Sweezy, et al. The Transition from Feudalism to Capitalism (London: NLB, 1976), 38–39 & 106. 225

David Baker

28 E.A. Kosiminsky, Studies in the Agrarian History of England in the Thirteenth Century (­Oxford: Oxford University Press, 1956); for empirical contradiction one need look no farther than the work of fellow Marxist historian Guy Bois, La crise du féodalisme: économie rurale et démographie en Normandie orientale du début du XIVe siècle au milieu du XVIe siècle (Paris: Fondation nationale des sciences politiques, 1976) which shows an enormous drop of population 1350–1450, on the order of 50–75%. Numerous other empirical works on the late medieval population that contradict Kosiminsky are too numerous to list here, since virtually all of them written on the question do. Some of them are cited below. 29 Michael Postan, ‘Some Economic Evidence of Declining Population in the Later ­M iddle Ages’ Economic History Review 2 (1950) 4 and for distancing from Malthus, ­M ichael Postan and John Hatcher, ‘Population and Class Relations in Feudal Society’ Past and Present 78 (1978) 24–28, see also M. Postan ‘Moyen Âge’ in IXe Congrès international des sciences historiques, 2 vols. (Paris: A. Colin, 1950), ‘Medieval Agrarian Society in its Prime: England’ in Cambridge Economic History, vol. 1, ed. M. Postan (Cambridge: ­Cambridge University Press, 1966), and Essays on Medieval Agriculture and General Problems of the Medieval Economy (Cambridge: Cambridge University Press, 1973). 30 Braudel later took on the mantle of demography in his own treatments, Fernand Braudel, The Identity of France: People and Production, trans. Sian Reynolds (Glasgow: Collins, 1990). 31 Ladurie’s work came along in a general climate of work on demography in French historical studies, such as Robert-Henri Bautier, ‘Feux, population et structure sociale au milieu du XVe siècle: L’example de Carpentras’ Annales 14 (1959) 255–268, E. Baratier, La démographie provençale du XIVe au XVIe siècle (Paris: Ecole Practique des HautesÉtudes, 1961), Guy Fourquin, Les campagnes de la region parisienne, à la fin du Moyen Age (Paris: Presses Universitaires de France, 1964), Marc Bloch, French Rural History: An Essay on its Basic Characteristics, trans. Janet Sondheimer (London: Routledge and Kegan Paul, 1966), Yves Durand, ‘Recherches sur les salaries des maçons à Paris au XVIIIe siècle’ Revue d’Histoire Économique et Sociale 44 (1966) 468–480. 32 Emmanuel Le Roy Ladurie, Les paysans de Languedoc (Paris: SEVPEN, 1966), and the English translation Ladurie, Peasants of Languedoc, see also The French Peasantry: 1450–1660, trans. Alan Sheridan (Aldershot: Scolar Press, 1987), The Royal French State (­Oxford: Blackwell, 1994), and with Michel Morineau, Histoire économique et sociale de la France, de 1450 à 1660 (Paris: Presses Universitaires, 1977), with Joseph Goy, Tithe and Agrarian History from the Fourteenth to the Nineteenth Centuries: An Essay in Comparative History, trans. Susan Burke (Cambridge: Cambridge University Press, 1982). 33 Jacques Dupâquier et al. Histoire de la Population Française, 4 vols. (Paris: Presses Universitaires de France, 1988), see also Micheline Baulant, ‘Le prix des grains à Paris de 1431 à 1788’ Annales, Histoire, Sciences Sociales 3 (1968) 520–540, ‘Le salaire des ouvriers du bâtiment à Paris de 1400 à 1726’ Annales, Histoire, Sciences Sociales 26 (1971) 4­ 63–483, J.C. Russell, ‘Population in Europe 500–1500’ in The Fontana Economic History of ­Europe, ed. Carlo Cipolla (London: Collins, 1969), M. Belotte, La region de Bar-sur-Seine à la fin du Moyen Age: du début du XIIIe siècle au milieu du XVIe siècle, etude economique et sociale (Lille: Université de Lille, 1973), A. Croix, Nantes et le pays nantais au XVIe siècle: Étude demographique (Paris: SEVPEN, 1974), J. Jacquart, La crise rurale en Ile-deFrance: 1550–1670 (Paris: Armand Colin, 1974), Pierre Charbonnier, Une autre France: La seigneurie rurale en Basse Auvergne du XIVe au XVIe siècle (Clermont-Ferrand: Institut d’Études du Massif Central, 1980), H. Neveux, Vie et déclin d’une structure économique: les grains du Cambrésis, fin du XIVe – début du XVIIe siècle (Paris: École des hautes-études en sciences sociales, 1980), Elisabeth Carpentier and Michel Le Mené, La France du 226

Big history and historiography

34 35

36 37 38

39 40 41 42 43

44 45 46

XIe au XVe siècle: population, société, économie (Paris: Presses Universitaires de France, 1996), William Jordan, The Great Famine: Northern Europe in the Early Fourteenth Century (Princeton: Princeton University Press, 1996), David Fischer, The Great Wave: Price Revolutions and the Rhythm of History (Oxford: Oxford University Press, 1996), Robert Allen, ‘The Great Divergence in European Wages and Prices from the Middle Ages to the First World War’ Explorations in Economic History 38 (2001) 411–447. Emmanuel Le Roy Ladurie, ‘L’histoire immobile’ Annales E.S.C. 29 (1974) 675 & 689. Robert Brenner, ‘Agrarian Class Structure and Economic Development in Pre-­ Industrial Europe’ Past and Present 70 (1976) 30–75 and reprinted in T. Ashton and C. Philpin (eds.) The Brenner Debate: Agrarian Class Structure and Development in Pre-­ Industrial Europe (Cambridge: Cambridge University Press, 1985), from which the socalled ‘Brenner Debate’ draws its name. Peter Turchin and Sergey Nefedov, Secular Cycles (Princeton: Princeton University Press, 2009), 240–242. Guy Bois, ‘Against the Neo-Malthusian Orthodoxy’ Past and Present 79 (1978) 60–69, which despite its name is mostly devoted to quarrelling between the two Marxists. Guy Bois, La crise du féodalisme: économie rurale et démographie en Normandie orientale du début du XIVe siècle au milieu du XVIe siècle (Paris: Fondation nationale des sciences politiques, 1976), and the English version, The Crisis of Feudalism: Economy and Society in Eastern Normandy, c.1300–1550 (Cambridge: Cambridge University Press, 1984), see also La grande depression médiévale, XIVe-XVe siècles: le precedent d’une crise systémique (Paris: Presses Universitaires de France, 2000). Emmanuel Le Roy Ladurie, ‘En Haute-Normandie: Malthus ou Marx?’ Annales: Histoire, Sciences Sociales, 33 (1978) 115–124. Postan and Hatcher, ‘Population and Class Relations in Feudal Society’, 24–37. Emmanuel Le Roy Ladurie, ‘A Reply to Professor Brenner’ Past and Present 79 (1978) 55–59. Robert Brenner, ‘The Agrarian Roots of European Capitalism’ Past and Present 97 (1982) 16–113. David Weir, ‘Life Under Pressure: France and England 1670–1870’ Journal of Economic History 44 (1984) 27–48, George Grantham, ‘Contra Ricardo: On the Macroeconomics of pre-Industrial Economies’ European Review of Economic History 2 (1999) 199–232, Philip Hoffman and Jean-Laurent Rosenthal, ‘New Work in French Economic History’ French Historical Studies 23 (2000) 439–453, George Grantham, ‘Explaining the Industrial Transition: A non-Malthusian Perspective’ European Review of Economic History 12 (2008) 155–165, Jan de Vries ‘The Economic Crisis of the Seventeenth Century After Fifty Years’ Journal of Interdisciplinary History 40 (2009) 151–194, Anne McCants ‘Historical Demography and the Crisis of the Seventeenth Century’ Journal of Interdisciplinary History 40 (2009) 195–214. Philip Hoffman has also tried to paint a portrait of steady growth in agricultural productivity by using a questionable system of measurement that combines prices, wages, and leases, rather than counting what was actually produced per worker or hectare, Growth in a Traditional Society: The French Countryside, 1450–1815 (Princeton: Princeton University Press, 2000), 81–86 & 130. Jack Goldstone, Revolution and Rebellion in the Early Modern World (Berkeley: University of California Press, 1991). For instance see the McCants article, 203–204. David Hackett Fischer, The Great Wave: Price Revolutions and the Rhythm of History (­Oxford: Oxford University Press, 1996), 181–203. Ian Morris, Why the Rest Rules for Now: The Patterns of History and what they Reveal about the Future (New York: Farar, Straus, and Giroux, 2010) and The Measure of Civilization: How Social Development Decides the Fate of Nations (Princeton: Princeton University Press, 2013). 227

David Baker

47 The average appears to be roughly 300 years for a full cycle, but depends greatly on specific conditions. 48 Peter Turchin, Historical Dynamics: Why States Rise and Fall (Princeton: Princeton University Press, 2003), War and Peace and War: The Life Cycles of Imperial Nations (New York: Pi Press, 2006), with Nefedov, Secular Cycles. 49 Peter Turchin has tried to popularise the term ‘asabiya’ for social cohesion, a term used by North African fourteenth century social theorist Ibn Khaldun, a partial inspiration for Turchin’s theory. 50 Timothy Kohler, Sarah Cole, and Stancea Ciupe, ‘Population Warfare: A Test of the Turchin Model in Pueblo Societies’ in Pattern and Process in Cultural Evolution, ed. ­Stephen Shennan (Berkeley: University of California Press, 2009) 290–291. 51 Michael Rampino and Stephen Self, ‘Volcanic Winter and Accelerated Glaciation following the Toba Super-eruption’ Nature 359 (1992) 50–52. 52 Martin Williams, Stanley Ambrose, Sander van der Kaars, et al. ‘Environmental Impact of the 73ka Toba Super-eruption in South Asia’ Palaeogeography, Palaeoclimatology, Palaeoecology 284 (2009) 295–314, Michael Rampino and Stanley Ambrose, ‘Volcanic Winter in the Garden of Eden: The Toba Super-Eruption and the Late Pleistocene Population Crash’ in F. McCoy and W. Heiken (eds.), Volcanic Hazards and Disasters in Human Antiquity (Boulder: Geological Society of America, 2000), 78–80, Stanley Ambrose, ‘Late Pleistocene Human Population Bottlenecks, Volcanic Winter, and Differentiation of Modern Humans’ Journal of Human Evolution 34 (1998) 623–651. 53 Livi-Bacci, A Concise History of World Population, 31. 54 Neil Roberts, The Holocene: An Environmental History (Oxford: Blackwell, 1998), 136. 55 J.R. Biraben, ‘Essai sur l’évolution du nombre des hommes’ Population 34 (1979) 13–25. 56 Peter Richerson, Robert Boyd, and Robert Bettinger, ‘Cultural Innovations and ­Demographic Change’ Human Biology 81 (2009) 211–235. 57 The account of these events is given a decent treatment in many works, for instance, David Christian, Maps of Time: An Introduction to Big History (Berkeley: University of California Press, 2004), 24–27, John Barrow, The Book of Universes: Exploring the Limits of the Cosmos (London: W.W. Nortion and Company, 2011), Cesare Emiliani, The Scientific Companion: Exploring the Physical World with Facts, Figures, and Formulas (New York: John Wiley, 1995), 82, Stephen Hawking, The Universe in a Nutshell (New York: Bantam, 2001), 78, and many, many more.

References Abel, Wilhelm. Agricultural Fluctuations in Europe: From the Thirteenth to the Twentieth Centuries. Trans. Olive Ordish. New York: St. Martin’s Press, 1978. Allen, Robert. ‘The Great Divergence in European Wages and Prices from the Middle Ages to the First World War’ Explorations in Economic History 38 (2001) 411–447. Ambrose, Stanley. ‘Late Pleistocene Human Population Bottlenecks, Volcanic Winter, and Differentiation of Modern Humans’ Journal of Human Evolution 34 (1998) 623–651. Aristotle. A Treatise on Government. Trans. William Ellis. Charleston: Forgotten Books, 1947 [c.330 BC]. Baratier, Edouard. La démographie provençale du XIVe au XVIe siècle. Paris: Ecole Practique des Hautes-Études, 1961. Barrow, John. The Book of Universes: Exploring the Limits of the Cosmos. London: W.W. ­Nortion and Company, 2011. Baulant, Micheline. ‘Le prix des grains à Paris de 1431 à 1788’ Annales, Histoire, Sciences Sociales 3 (1968) 520–540. 228

Big history and historiography

Baulant, Micheline. ‘Le salaire des ouvriers du bâtiment à Paris de 1400 à 1726’ Annales, Histoire, Sciences Sociales 26 (1971) 463–483. Bautier, Robert-Henri. ‘Feux, population et structure sociale au milieu du XVe siècle: L’example de Carpentras’ Annales 14 (1959) 255–268. Belotte, Michel. La region de Bar-sur-Seine à la fin du Moyen Age: du début du XIIIe siècle au milieu du XVIe siècle, etude economique et sociale. Lille: Université de Lille, 1973. Biraben, Jean-Noël. ‘Essai sur l’évolution du nombre des hommes’ Population 34 (1979) 13–25. Bloch, Marc. French Rural History: An Essay on its Basic Characteristics. Trans. Janet ­Sondheimer. London: Routledge and Kegan Paul, 1966. Bois, Guy. ‘Against the Neo-Malthusian Orthodoxy’ Past and Present 79 (1978) 60–69. Bois, Guy. The Crisis of Feudalism: Economy and Society in Eastern Normandy, c.1300–1550. Cambridge: Cambridge University Press, 1984. Bois, Guy. La grande depression médiévale, XIVe–XVe siècles: le precedent d’une crise systémique. Paris: Presses Universitaires de France, 2000. Braudel, Fernand. ‘History and the Social Sciences: The Longue Durée’ Annales. Histoire, Sciences Sociales 13.4 (1958) 726. Braudel, Fernand. The Identity of France: People and Production. Trans. Sian Reynolds. Glasgow: Collins, 1990. Braudel, Fernand. The Mediterranean and the Mediterranean World in the Age of Philip II. Vol. 1. Trans. Sian Reynolds. Berkeley: University of California Press, 1996. Brenner, Robert. ‘Agrarian Class Structure and Economic Development in Pre-Industrial Europe’ Past and Present 70 (1976) 30–75. Brenner, Robert. ‘The Agrarian Roots of European Capitalism’ Past and Present 97 (1982) 16–113. Carpentier, Elisabeth and Michel Le Mené. La France du XIe au XVe siècle: population, société, économie. Paris: Presses Universitaires de France, 1996. Chao, W.L. and S.C. Hsieh. History of the Chinese Population. Beijing: People’s Publisher, 1988. Charbonnier, Pierre. Une autre France: La seigneurie rurale en Basse Auvergne du XIVe au XVIe siècle. Clermont-Ferrand: Institut d’Études du Massif Central, 1980. Christian, David. Maps of Time: An Introduction to Big History. Berkeley: University of ­California Press, 2004. Chu, C.Y. Cyrus and Ronald Lee. ‘Famine, Revolt, and Dynastic Cycles: Population ­D ynamics in Historical China’ Journal of Population Economics 7 (1994) 351–378. Croix, A. Nantes et le pays nantais au XVIe siècle: Étude demographique. Paris: SEVPEN, 1974. Darwin, Charles. The Autobiography of Charles Darwin, ed. Francis Darwin. London: John Murray, 1887. de Vries, Jan. ‘The Economic Crisis of the Seventeenth Century After Fifty Years’ Journal of Interdisciplinary History 40 (2009) 151–194. Dobb, Maurice. Studies in the Development of Capitalism. New York: International Publishers, 1946. Dupâquier, Jacques et al. Histoire de la Population Française. 4 vols. Paris: Presses Universitaires de France, 1988. Durand, Yves. ‘Recherches sur les salaries des maçons à Paris au XVIIIe siècle’ Revue d’Histoire Économique et Sociale 44 (1966) 468–480. Emiliani, Cesare. The Scientific Companion: Exploring the Physical World with Facts, Figures, and Formulas. New York: John Wiley, 1995. Fischer, David. The Great Wave: Price Revolutions and the Rhythm of History. Oxford: Oxford University Press, 1996. 229

David Baker

Fourquin, Guy. Les campagnes de la region parisienne, à la fin du Moyen Age. Paris: Presses Universitaires de France, 1964. Goldstone, Jack. Revolution and Rebellion in the Early Modern World. Berkeley: University of California Press, 1991. Grantham, George. ‘Contra Ricardo: On the Macroeconomics of pre-Industrial Economies’ European Review of Economic History 2 (1999) 199–232. Grantham, George. ‘Explaining the Industrial Transition: A non-Malthusian Perspective’ European Review of Economic History 12 (2008) 155–165. Griziotti-Kretschmann, Jenny. Il problema del trend secolare nelle fluttuazioni dei prezzi. Pavia: University of Pavia, 1935. Griziotti-Kretschmann, Jenny. ‘Richerche sulle fluttuazioni economiche di lungadurate’ Giornale degli Economisti 73 (1933), 461–508. Hackett Fischer, David. The Great Wave: Price Revolutions and the Rhythm of History. Oxford: Oxford University Press, 1996. Hawking, Stephen. The Universe in a Nutshell. New York: Bantam, 2001. Ho, P.T. Studies on the Population of China: 1368–1953. Cambridge: Harvard University Press, 1959. Hoffman, Philip. Growth in a Traditional Society: The French Countryside, 1450–1815. Princeton: Princeton University Press, 2000. Hoffman, Philip and Jean-Laurent Rosenthal. ‘New Work in French Economic History’ French Historical Studies 23 (2000) 439–453. Jacquart, Jean. La crise rurale en Ile-de-France: 1550–1670. Paris: Armand Colin, 1974. Jordan, William. The Great Famine: Northern Europe in the Early Fourteenth Century. Princeton: Princeton University Press, 1996. Khaldun, Ibn. Muqaddimah. Trans. Franz Rosenthal. Princeton: Princeton University Press, 1967. Kohler, Timothy, Sarah Cole, and Stancea Ciupe, ‘Population Warfare: A Test of the Turchin Model in Pueblo Societies’ in Pattern and Process in Cultural Evolution, ed. ­Stephen Shennan. Berkeley: University of California Press, 2009. Kondratiev, Nikolai. The Major Economic Cycles. Moscow: Voprosy Koniunktury, 1925. Kosiminsky, Evgeniĭ Alekseevich. Studies in the Agrarian History of England in the Thirteenth Century. Oxford: Oxford University Press, 1956. Labrousse, C.-E. Esquisse du mouvement des prix et des revenus en France au XIIIe siècle. Paris: Dalloz, 1932. Ladurie, Emmanuel Le Roy. ‘History that Stands Still’ in The Mind and Method of the Historian, ed. Emmanuel Le Roy Ladurie. Trans. Sian Reynolds and Ben Reynolds. Brighton: Harvester Press, 1981. Ladurie, Emmanuel Le Roy. ‘A Reply to Professor Brenner’ Past and Present 79 (1978) 55–59. Ladurie, Emmanuel Le Roy. The French Peasantry: 1450–1660. Trans. Alan Sheridan. ­A ldershot: Scolar Press, 1987. Ladurie, Emmanuel Le Roy. ‘En Haute-Normandie: Malthus ou Marx?’ Annales: Histoire, Sciences Sociales, 33 (1978) 115–124. Ladurie, Emmanuel Le Roy. Peasants of Languedoc. Trans. John Day. Urbana: University of Illinois Press, 1976. Ladurie, Emmanuel Le Roy and Goy, Joseph. Tithe and Agrarian History from the Fourteenth to the Nineteenth Centuries: An Essay in Comparative History. Trans. Susan Burke. ­Cambridge: Cambridge University Press, 1982. Ladurie, Emmanuel Le Roy and Morineau, Michel. Histoire économique et sociale de la France, de 1450 à 1660. Paris: Presses Universitaires, 1977. 230

Big history and historiography

Livi-Bacci, Massimo. A Concise History of World Population. Trans. Carl Ipsen. Oxford: Blackwell, 1992. Machiavelli, Niccolò. Discourses on Livy. Trans. Leslie Walker. London: Penguin Books, 1984 [1517]. Malthus, Thomas. An Essay on the Principle of Population: As It Affects the Future Improvement of Society with Remarks on the Speculations of Mr. Godwin, M. Condorcet, and Other Writers. London: J. Johnson, 1798. McCants, Anne. ‘Historical Demography and the Crisis of the Seventeenth Century’ Journal of Interdisciplinary History 40 (2009) 195–214. Morris, Ian. The Measure of Civilization: How Social Development Decides the Fate of Nations. Princeton: Princeton University Press, 2013. Morris, Ian. Why the Rest Rules for Now: The Patterns of History and what they Reveal about the Future. New York: Farar, Straus, and Giroux, 2010. Neveux, H. Vie et déclin d’une structure économique: les grains du Cambrésis, fin du XIVe – début du XVIIe siècle. Paris: École des hautes-études en sciences sociales, 1980. Plato. The Republic. Trans. Allan Bloom. New York: Basic Books, 1991 [c.380 BC]. Postan, Michael. Essays on Medieval Agriculture and General Problems of the Medieval Economy. Cambridge: Cambridge University Press, 1973. Postan, Michael. ‘Moyen Âge’ in IXe Congrès international des sciences historiques, 2 vols. Paris: A. Colin, 1950. Postan, Michael. ‘Medieval Agrarian Society in Its Prime: England’ in Cambridge Economic History, vol. 1, ed. M. Postan. Cambridge: Cambridge University Press, 1966. Postan, Michael. ‘Some Economic Evidence of Declining Population in the Later Middle Ages’ Economic History Review 2 (1950) 4. Postan, Michael and John Hatcher. ‘Population and Class Relations in Feudal Society’ Past and Present 78 (1978) 24–28. Rampino, Michael and Stephen Self, ‘Volcanic Winter and Accelerated Glaciation Following the Toba Super-eruption’ Nature 359 (1992) 50–52. Rampino, Michael and Stanley Ambrose, ‘Volcanic Winter in the Garden of Eden: The Toba Super-Eruption and the Late Pleistocene Population Crash’ in F. McCoy and W. Heiken (eds). Volcanic Hazards and Disasters in Human Antiquity. Boulder: Geological Society of America, 2000. Ricardo, David. An Essay on the Influence of a Low Price of Corn on the Profits of Stock: Showing the Inexpediency of Restrictions on Importation with Remarks on Mr. Malthus’ Last Two Publications. London: John Murray, 1815. Richerson, Peter, Robert Boyd, and Robert Bettinger, ‘Cultural Innovations and Demographic Change’ Human Biology 81 (2009) 211–235. Roberts, Neil. The Holocene: An Environmental History. Oxford: Blackwell, 1998. Russell, J.C. ‘Population in Europe 500–1500’ in The Fontana Economic History of Europe, ed. Carlo Cipolla. London: Collins, 1969. Simiand, François. Les fluctuations économiques à longue période et la crise mondiale. Paris: Félix Alcan, 1932. Simiand, François. Recherches anciennes et nouvelles sur le mouvement général des prix du XVIe au XIXe siècle. Paris: Domat Montchrestien, 1932. Sweezy, Paul. ‘The Transition from Feudalism to Capitalism’ Science and Society 14 (1950) 134–167. Sweezy, Paul, et al. The Transition from Feudalism to Capitalism. London: NLB, 1976. Tertullian. De Anima. ed. Jan Hendrik Waszink. Leiden: Koninklijke Brill, 2010 [c.200 AD]. Turchin, Peter. Historical Dynamics: Why States Rise and Fall. Princeton: Princeton University Press, 2003. 231

David Baker

Turchin, Peter. War and Peace and War: The Life Cycles of Imperial Nations. New York: Pi Press, 2006. Turchin, Peter and Nefedov, Sergey. Secular Cycles. Princeton: Princeton University Press, 2009. Weir, David. ‘Life Under Pressure: France and England 1670–1870’ Journal of Economic History 44 (1984) 27–48. Williams, Martin et al. ‘Environmental Impact of the 73ka Toba Super-eruption in South Asia’ Palaeogeography, Palaeoclimatology, Palaeoecology 284 (2009) 295–314.

232

10 BIG HISTORY AND CRITICAL THEORY Science, history and why theory matters David Blanks Why does theory matter? Since you’ve found your way to this chapter, you already have an idea of what big history is. You might be taking a course in it right now, or teaching one, or perhaps you’re doing some research on some aspect of big history. A few of you will be fullblown practitioners. In all cases you have decided that you want to learn about, or reflect upon, or perhaps argue about the relationship between big history and theory, which is commendable: most students of history shy away from theory. Indeed, maybe you are reading this under duress. It would be much more fun to get down to the business of doing the real stuff instead of wasting time on what seems more like philosophy than history. And yet, just as a musician needs to understand chords and scales, or an artist colors and composition—as a physicist needs math and chemists an understanding of the periodic tables—anyone who is serious about delving into any field owes it to himself or herself to spend at least some time examining its methods, its structure, its underlying assumptions, its antecedents, its history—its ideology even. Without some sort of interpretive framework, without understanding what we might call a discipline’s “imaginative vision,” it is difficult to know how to think about what you are doing, or how to understand what others are trying to achieve when they do it.1 And there are all sorts of problems that need our attention: Who is big history for? What is the relationship between history and science? Is scientific history even possible? If so, what should it look like? And what questions should it answer? Is it the business of science to supply society with myths? Theory matters because we need it to help us answer these and similar questions. Before we proceed, though, let’s be clear on what we are talking about. Scientific theories, such as the theory of natural selection, or the general theory of relativity, are designed to explain and predict natural phenomena. Big history uses this concept in its own way, such as the Goldilocks principle, or the theory of collective learning. Some big historians also look for a “grand unifying theory of the past,” which is an attempt to find an overarching explanation for increasing complexity that ties 233

David Blanks

together the natural and social sciences and the humanities.2 There is thought to be a “gap” between the culture of science and the culture of the humanities, commonly called the “two cultures problem,” and it is hoped that discovering a theory of how everything works will bridge the gap and make our understanding of the world whole again.3 But in this essay I am using the notion of theory in a different sense. We can think of the theory of big history as the philosophy of big history. It is a means of trying to understand approach and process: methods, structure, assumptions, and antecedents. It includes the sorts of things you can read about in the journal History and Theory: the critical philosophy of history, the speculative philosophy of history, historiography, the history of historiography, historical methodology, etc.4 The theory of science, or the philosophy of science, works in much the same way and is intended in the same way here. Simply put, theory, in the present context, asks how we know things and explores the underlying logic of a given discipline. Reading that last paragraph, it is easy to see why many historians are put off by theory. It is, after all, a form of philosophy. It’s not what historians do. And they really shouldn’t be expected to. It is another field altogether. As Peter Novick pointed out, for their part, most philosophers are lousy historians. So it is somewhat unfair for philosophers to criticize the methodologies of historians in this regard, and rather unseemly to suggest that someone’s views are a reflection of his or her background, prejudices, or psychic needs. To talk of epistemology loads the dice against historians. It is like a sportswriter criticizing their batting in the annual history department softball game.5 And yet the problem of the interpretive framework is a real one. Fair or unfair this is the work of the philosopher of history and the philosopher of science and it must not on any account be avoided. Far from creating uncertainty, doubt, or confusion, taking a theoretical approach to big history is better science than ignoring it. As the cognitive linguists George Lakoff and Mark Johnson put it, “Science cannot maintain a self-critical stance without a serious familiarity with philosophy and alternative philosophies. Scientists need to be aware of how hidden a priori philosophical assumptions can determine their scientific results.”6 So instead of getting bogged down in what those who do not like theory disparagingly call “jargon,” let’s try using a metaphor instead, and see if that can help us out.7 Imagine big history as a large house. There are rooms for physicists and geologists, chemists, biologists, social scientists, and yes, artists and musicians too. They live and work there together and share a space which represents a grand narrative that combines their areas of expertise to create a modern origin story based upon the best available empirical evidence and scholarly methods. But hidden inside the walls and under the floorboards of that house are the electrical and plumbing systems upon which they depend. The inhabitants take these for granted and none has been trained as an electrician or a plumber—which is fine until the power goes out or the hot water stops working.When this happens they will need to call in a specialist, someone who understands a building’s internal workings.This is when they will need a theorist. The influential American historian, Hayden White, used to say that “Those historians who draw a firm line between history and the philosophy of history fail to recognize that every historical discourse contains within it a full-blown, if only implicit, 234

Big history and critical theory

philosophy of history.”8 What he meant by this was that there are always already assumptions, myths, metaphors, ideologies, and such lying beneath the floorboards of any discipline, and that it does no good to pretend that they do not exist. In fact this can be dangerous, especially when these untended ideologies suddenly and publically manifest themselves in ways that are harmful, such as when they are used to justify imperialism or ethnic cleansing or worse. There are of course less deadly dangers—such as when drawing too tight of a boundary around a discipline leads to various forms of exclusion,—which is something else we need to watch out for in big history. As theorist William Katerberg observed, “The question is not whether any given version of Big History is mythic, but what the unspoken or directly stated mythic meaning is in the narrative.”9 Which is another way of saying that telling a story is never neutral or objective. Meaning does not emerge from the empirical evidence all on its own. One cannot, as some big historians claim, remove oneself from the equation by taking academic distance from the subject.10 This is theoretically naïve. “Big ideas are a matter of faith, not of science.”11 As my favorite philosophical plumber, Mary Midgley, puts it, this is one of those cases “where people who refuse to have anything to do with philosophy have become enslaved to outdated forms of it.”12 Let’s begin by reflecting upon the ways in which practitioners of big history describe what they are doing and then gently pry up some of those floorboards to see what’s underneath. I do this—and this bears repeating—not to dismantle the discipline but to help make it more intellectually robust. The claim that theory is relativism and nihilism and that it means that one can write absolutely anything is willfully ignorant nonsense. Understanding theory makes for solid history and for solid science. *****

Who is big history for? It is not uncommon at big history conferences to hear people give presentations that begin with a recounting of their personal journey to the field. They talk about how fulfilling the experience is, how big history answers questions that they had been wondering about for a long time. You see this in journals and textbooks too—which is rather unusual for an academic discipline.13 Katerberg likens these mini-­autobiographies to conversion narratives, “a genre characterized by stories of awakening, enlightenment and wonder, of being lost and then now found, and setting on a new path, often with a mission.”14 My own “conversion” took place in 2005 on the porch of a hotel café in an incongruous Alpine village in the Middle Atlas Mountains. It was my first World History Association conference, and it was there, in Ifrane, Morocco, that I met Craig ­Benjamin and David Christian. And although I didn’t know it at the time, as I stepped off that porch it was in truth onto a new intellectual path of enlightenment and wonder that has led to places that none of us imagined all those years ago. I wouldn’t go so far as to say that I was “lost and then now found,” but I was certainly intrigued, so much so that in 2007 I invited David to the American University in Cairo as a Distinguished Visiting Professor. He helped me set up the first big 235

David Blanks

history course in the Muslim world, and because by then I had done a considerable amount of reading on the subject, he also helped me work through many of my questions. But not all of them. And it was during these conversations, in a palm-shaded courtyard on the now famous Tahrir Square, that I came to realize that although David was my teacher, and still is, we did not necessarily agree on everything about big history, especially about its imaginative vision.15 My opinions have evolved considerably since those first, tentative conversations over ten years ago, but what has remained, and grown, is the awareness that not everyone in this field thinks the same way. There are tensions and contradictions that run through big history conceptually, methodologically, and ideologically. This is true for all disciplines, but particularly awkward in one that likes the idea of having a grand unified theory (Chaison, Spier) or a unifying paradigm (Christian, Brown, et al.), but doesn’t like the idea of reflecting on what that means.16 Thus, in trying to understand big history’s interpretive framework, you might want to ask such things as: Where does the idea of a unified theory come from? Is it possible? Is it desirable? And are there other ways of investigating and narrating the history of everything that do not rely upon a single perspective and a single methodology?17 Leaving these problems for you to solve, I would like to look at some examples of the tensions within the discipline as a way of illustrating how theory can be of use to us. Then I will make some observations about the relationship between history and science, and the relationship between myth and meaning. Perhaps this is more of a confession than a conversion, but for my part, as a consequence of having lived and worked for so long in Egypt, a former British colony that is both predominantly Arab and predominantly Muslim, I am particularly sensitive to expressions of cultural imperialism which, from my perspective, manifest themselves in big history in two interconnected ways.18 First, from the International Big History Association’s (IBHA) earliest days as an organization, at the World History Association meeting at Capital Normal University in Beijing in 2011, then at the first IBHA conference at Grand Valley State University in 2012, then at Dominican University in 2014, the University of Amsterdam in 2016, and Villanova University in 2018, there have been tensions between what some have labeled the “scientists” and the “mystics.”This is not something you will be able to see if your total exposure to the field of big history is through a textbook and a course. Doing theory will enable you to examine a problem from a variety of perspectives, and to use a variety of methods and models, all at the same time. Ultimately, though, it will require you to choose among them—and to know why you made the choice that you did. At the risk of being accused of being overly sensitive, it seems to me that since big history “went public,” beginning in 2010, it has suffered from an internal contradiction, which is the effort, on the one hand, to create and to teach a modern mythology, and the desire, on the other, to justify itself by insisting on its scientific credentials. But once a discipline is out there—once you have TED talks and courses and the Big History Project and a professional organization and a journal—you can no longer put the genie back into the bottle. As a consequence, although big history was born in opposition to religion, there are many now, both within the organization and without, who see big history somewhat the way Georg Wilhelm Friedrich Hegel (1770–1831) might have—as proof of divine providence.19 236

Big history and critical theory

And here is my second example of the tensions and contradictions inherent in this synthesizing genre of history writing.20 When we examine its foundations, we learn that big history’s claim to universality is not nearly as solid as its architects make it out to be. The principle textbook in the field, Big History: Between Nothing and Everything (2014), tells readers that “big history differs from traditional origin stories in that it is universal.”21 It’s a recurring theme throughout the literature. Spier, for example, argues that big history is for everyone—unless you are religious, he says—in which case you can’t come in.22 But this only serves to reinforce my point.The vast majority of people in the world are not modern and not scientific and will most certainly not throw over their traditional origin stories for this new, western one. For that matter, most people in the west won’t do it either, especially in America, which is known uniquely among western countries for believing in God, and for not believing in the theory of evolution.23 What does this do to the idea that this story is for everyone? This will only happen if you believe as Voltaire and Comte did that religion will eventually be entirely replaced by science. Big historians will respond that what they mean is that this is a global story based upon new scientific discoveries and sound scientific principles. They will say that this is the future and that we need this new way of thinking to preserve the planet. Implicit in this argument, however, is the expectation that eventually everyone will modernize, that it is just a matter of time. Indeed if you believe in the precepts of increasing complexity, emergent properties, thresholds, ever increasing energy density flow rates, and ever increasing collective learning, then it is very difficult not to come to the conclusion that this is the direction in which the world is headed, and that this is proven scientifically.24 As a result of economic development, political maturity, education, and continuing scientific advances, one day soon nearly everyone will come to accept the scientific creation myth, leaving but a few old believers whom we might consult now and again on questions of how to live well and how to live sustainably.25 But our philosophical plumber will point out that this interpretive framework is not global or universal at all but rather a product of its own cultural heritage. In big history’s case that cultural heritage comes quite explicitly from eighteenth- and ­nineteenth-century Europe, that is, from the Enlightenment project, and from romantic science.26 All of its ideas about the unity of the natural world, its secularizing tendencies, its universalizing, can be found there. And in this regard it is of some interest that today big history has a tendency to emphasize how new it is, and to downplay its connections to earlier modes of thought. There are a variety of reasons for this, not the least of which is that if it is going to cast itself as a modern epic, then it must accomplish heroic intellectual feats. More immediately, however, consciously or unconsciously, big history distances itself from its eighteenth end nineteenth century roots because it wants to get away from earlier Eurocentric histories and to be a new myth that will unite all peoples everywhere. But to try to do this by enforcing methodological purity, that is, by making determinations as to what is and what is not science, and by closing the door to those who do not meet these culturally determined criteria, it runs the risk of negating its own program.27 ***** 237

David Blanks

What is scientific history? In the previous section I alluded to the German philosopher, Hegel, who laid out his theology in The Phenomenology of Spirit (1807), Science of Logic (1812–1816), and Lectures on the Philosophy of Religion (1832). I was making the claim that people with opposing views on religion can nonetheless share an acceptance of the validity of science and scientific history, but whereas one group will see the big history narrative as proof that there is no such thing as God, the other will interpret it as a confirmation of God’s existence. It’s a classic problem in historiography—one where having the best available empirical evidence doesn’t help very much. You know who else believed in God? Leopold von Ranke (1795–1886). Ranke might well be thought of as the “father of American historiography.”28 Also ­German (American intellectuals idolized the Germans in the late nineteenth century), Ranke advocated an exhaustive search for primary sources, careful research, a strict adherence to the facts, an avoidance of bias, and the highest standards in historical scholarship; in short, he advocated the very methods and attitudes that big history insists upon and upon which it stakes its claim to being a scientific creation myth for all peoples. Except the God part—which has been quietly left out. So what? Why does this matter? The point is that insistence on method does not guarantee a successful, unified conclusion. Even in those cases where we agree on the methods, at some stage in the investigation we still have to choose a side. Not choosing is not an option. Or, rather, not choosing is an option—but it is also still taking a position. Most instances of ideological difference are not nearly as dramatic as the question of whether there is a supreme being, but because of the nature of our discipline, it must be asked: “Is God big enough for big history?”29 Thinking theoretically reminds us that like history science is also a contested discourse.30 But what an odd thing to say. Isn’t science, just, you know, science? Actually, no, science has a variety of aspects that are context dependent—which is why it is misleading to act as if following a rigorous, time-tested scientific method is the surest way to understand the world around us. Historians of all people should see this readily. Today religious thinkers and secular thinkers are battling for control of the s­cientific discourse, but it hasn’t always been that way.31 We need to think only of Bacon and Newton to remind ourselves that before the Enlightenment religion wasn’t seen as “infamous” at all but rather as an ally of science.32 Indeed, science was thought to provide evidence of the divine plan, and even a marginally unorthodox scholar like the German astronomer Johannes Kepler (1571–1630) were shunned by Catholics and Protestants alike. It was only much later that books were written with titles such as The History of the Conflict between Religion and Science (1874) or A History of the Warfare of Science with Theology in Christendom (1896).33 By then science had become secularized, and Christianity was seen as the enemy. This is still true. Religion is thought by many to exist in opposition to science.We all know that story. And yet there has been another shift as well. In the nineteenth century the sciences worked with the humanities to bring forth a modern, scientific 238

Big history and critical theory

worldview. But today, although many in scientific history take an anti-religious stance, they now place the humanities in the enemy camp too, especially history and philosophy which, from their perspective, pose a threat to science’s certainties. The science that was brought out in support of the Christian faith, and the one that came after that saw the humanities as its ally, were clearly not the same as the one that battles both religion and philosophy today. These are three different sciences, with different agendas, and different socio-historical contexts, different psychologies even. Again, Mary Midgley: The “science” that excluded Kepler’s doctrine of gravitation and enthusiastically accepted theism cannot be the same thing as today’s science which reverses those positions. And these various conceptions cannot help having a social meaning. To take sides for or against various elements in society in these various wars is inevitably to choose a particular notion of that society. To oppose some particular current ideology is to espouse and express a different one—an alternative idea of life as a whole, a distinctive view of what is important in it. Such opposition cannot just be something internal to science. It is not ‘value-free.’ Asking for more science and less of something else is itself a social and political move. This move can be quite legitimate but it must not be mistaken for a part of a pure, mysteriously objective science which stands outside of society. Past changes should surely make us think carefully about why we are now inclined to think of particular attitudes as demands of science. That is why the history and sociology of science are not luxuries but essential tools for any attempt to grasp its role in our lives.34 (emphasis mine) Another thing the history and sociology of science teaches us is that even within the broad category of science, in any given era there are varying methods, approaches, and underlying assumptions that are appropriate to different scientific fields. This is another sense in which science is not one thing. The hope that we can find a single, vast, ordered pattern of the universe, and that science could become some sort of universal language, was born during the halcyon days of the Scientific Revolution. The still surviving idea, that there is a single “scientific method” for all the sciences, is a “relic of this dream.”35 Novick argues that of all the humanities disciplines that took their modern form in the nineteenth century, “no group was more prone to scientific imagery, and the assumption of the mantle of science, than the historians.”36 (He’s talking about the Americans.) From the beginning, however, there were disagreements about method, which meant ideological differences, which meant competing material interests. There were those who understood the scientific method to mean discovering laws of history. This was the “Baconian” view of science,37 an approach which insisted on empirical data and the avoidance of unfounded hypotheses. Against them were those who understood the scientific method to be based upon critical thinking alone, even when that meant relying upon experience and the deductive method to interpret data that might be ambiguous. In a twist that mirrors the American interpretation of Ranke, those historians who were particularly given to a rigid materialism and the mocking of theory were also the ones who interpreted Darwin most literally, despite 239

David Blanks

the fact that Darwin himself rejected an undigested inductive reasoning in the same way that Ranke would have.38 “How odd it is,” Darwin wrote, “that anyone should not see that all observation must be for or against some view if it is to be of any service!”39 In Guns, Germs, and Steel: A Short History of Everybody for the Last 13,000 Years (1997), Jared Diamond reflects upon the plurality of scientific methods today. In a chapter called “The Future of Human History as a Science,”40 Diamond defends history, laying out the reasons he believes that it is scientific, not in the nomothetic (law-giving) sense of the so-called “hard sciences,” but in the sense that its methodology aims at carefully describing historical phenomena, making comparative analyses, and conducting “natural experiments,” which means taking advantage of unpredictable events such as an epidemic or revolution where nature provides a laboratory and scientists can observe the results. It works the same way as studying the effects of natural selection or the formation of new stars and galaxies. To make absolute claims, as some big historians do about what the scientific method is and must be, is misteaching. Good scientists do not engage in this kind of boundary work. They understand that different types of problems require different approaches. Although I have been trying to avoid difficult theoretical language, on this point Nasser Zakariya is worth quoting in full: When big, deep, and evolutionary histories argue that they can orient or reform historical understanding writ large, they tend to present historical scholarship as less problematic than it is, whether it is the nature of historical aggregation (how one history may or may not relate to another), the nature of historical objectivity, the critical analysis of scientific ideas inherent to the history of science, or the character of scientific truth and error. These are problematics that the universal historians largely reject.41 (emphasis mine) Science and history are not disciplines at war with one another.To characterize them as two distinct cultures is unnecessary. I think of them as sub-disciplines that occupy overlapping ranges of ideas and methods on the spectrum of natural philosophy. But that might be too much for most people. Let’s just say that the antithesis between science and the humanities is a false one. They are related aspects of a single cultural whole that share the same set of methods, myths, and metaphors. The real gap is not between science and the humanities, but between a secular universalism and a religious one. *****

Is it the business of science to supply society with myths? Big history, especially in the way it is taught, ignores these considerations. It acts as if there is a single scientific method that everyone has agreed upon. Much like positivism in the nineteenth century, it establishes a hierarchy of scientific endeavor that begins with physics and works its way through geology, chemistry, biology, and the 240

Big history and critical theory

social sciences. The humanities are grudgingly admitted on occasion, but they are lodged in the attic where it is hoped they will remain out of sight and make as little noise as possible.The exception of course is scientific history, which is understood, in a way, to be even more sciency than science. And bigger. And better. Spier excitedly suggests that “the study of history should be regarded as both the queen and king of the sciences.”42 McNeill is positively giddy: A historical worldview of enormous scope and grandeur has engulfed the no less grand, but now parochial, Newtonian world machine. … One may wonder if we are not once again making the world over in our own image, as older animistic and religious worldviews surely did. … [It is time for historians] to connect their own professional thinking and writing with the revised scientific vision of the nature of things. If we succeed in doing so, the convergence of the sciences will be complete, and human history will begin to play a part in shaping, and being shaped by, the scientific understanding of the world around us.43 All of us stretch our point from time to time, but the greater danger is that over-­ simplifying the story, re-mythologizing science, and packaging scientific history for public consumption leaves little room for critical thinking. In the long run this approach cannot help but diminish students’ capacity to judge public policy and make informed decisions. If what we are having them do is memorize the eight (or is it nine?) thresholds, then they will never develop the capacity to distinguish between science that makes sense and science that does not. Let’s not get carried away. My other favorite philosophical plumber, Gaston Bachelard (1884–1962), observed that “no one can say they have a scientific mind unless they are certain that at every moment of their thinking life they are reconstructing all their knowledge.”44 Like Einstein and Kafka before him, Bachelard began his career as a state employee, working, first, as postmaster in Bar-sur-Aube in the picaresque Champagne region of northeastern France, joining later the faculty at the University of Dijon, and eventually becoming the inaugural chair of the history and philosophy of science at the Sorbonne. Bachelard was a dedicated teacher and a great supporter of science education, but he felt that the way that science was being taught in the France of his day, especially the way it was being made to conform to common sense, was a violation of basic scientific truths. He called this cleaned up version of science an “epistemological obstacle” because it got in the way of real thinking. False hypotheses, wrong turns, contingent events (luck), the psychology of practitioners, the role of government bodies, self-interest: all the things that Bachelard’s admirer, Michel Foucault would later pick up on: all these Bachelard felt needed to be discussed in the classroom to make better educated citizens and better practicing scientists.45 Epistemological obstacles are deeply internalized ways of thinking about reality that obstruct greater clarity of thought. Big history has a lot of them. Entrenched cultural metaphors (the Arrow of Time, the Goldilocks Principle, the Selfish Gene); ideas left over from scientific work that has now outlived its value (reductionism, materialism); the tendency to canonize as necessary truths the contingent features of earlier historical periods of thought (unified theory, consilience); the tendency, in general, 241

David Blanks

towards a Whiggish interpretation of history: all form part of big history’s interpretive framework and prevent the discipline from running at full capacity as a research field. And there it is. If you shine your flashlight (theory) over in this direction, you’ll see it right down there between those floor joists (method): the greatest obstacle to big history as a research field is the way we teach it. So where does that leave us when it comes to the question of meaning? Some have tried to argue that scientific history has no meaning. Because it is science, it is claimed, it deals with facts, not values.46 And if you have not already run across this famous quotation by physicist Steven Weinberg, you soon will: “The more the universe seems comprehensible, the more it also seems pointless.”47 What these scholars believe is that our existence is the result of a series of incredibly improbable accidents such as the massive asteroid or comet strike in the Yucatan Peninsula 66 million years ago that made the rise of the mammals possible, or the period of global warming about 12,000 years ago that led to the emergence of farming and the earliest human civilizations. Because these events were the result of chance, and there was no design or intention, it follows therefore that the story uncovered through science and scientific history doesn’t mean anything. Of course most of the people in the world do not believe this, but that does not make the argument wrong. It is wrong—but not for that reason. Arguably, yes, nature in and of itself has no purpose or value, no ethics, no morals, no design, and no designer. However, as soon as we observe it, test it, try to become one with it, or try to control it, it most certainly has meaning, because humans invest it with meaning. Every time we tell a story about how we got here it has meaning, because we make choices about which parts interest us, because our ways of thinking are always already shaped by the culture we grew up in and the limitations of the language we speak, because we are “by nature” pattern-seeking creatures. It is not possible to tell a story without some sort of interpretive framework or imaginative vision even if we are not aware that we have one. People would not “convert” to big history if it didn’t mean something to them. Walter Alvarez, the geologist who developed the hypothesis that it was an asteroid strike that led to the extinction of the dinosaurs,48 fell in with big history in the 1990s while searching for a way to combine the histories of the cosmos, earth, life, and humanity into a single interdisciplinary field.49 Alvarez is a thoroughly scientifically oriented scholar, one of the organizers of the 2010 meeting at the Geological Observatory in Coldigioco, Italy, that led to the formation of the IBHA, and not at all prone to the flights of speculative fancy that one sometimes encounters in big history circles.Yet even he finds meaning in this story. Our appearance on this planet and in the universe is for Alvarez an “improbable journey,” which means, among other things, that it was not providential. This alone is a philosophical statement. The improbability that Alvarez chooses to highlight (one might after all highlight continuities rather than discontinuities), also means that the emergence of complex organisms like humans was not predictable, which in turn suggests that (1) even within the laws of nature there are contingent events (like the asteroid); (2) although laws can be established to explain the natural world, it is not possible to form similar laws for human history; (3) we have free will; and (4) there is a moral to this story, namely, that humans are collectively responsible for the future of the planet. 242

Big history and critical theory

David Christian’s position is similar—and both are typical of mainstream, moderate views within the IBHA. This group believes that the best way to guarantee the health and well-being of human society and of the plants and animals is to teach people how deeply connected we are to the history of the planet, the solar system, and the universe.The best way to teach them this is through scientific history. Science shows that it is our responsibility to make sound decisions about the environment so that we can guarantee a secure and healthy future for our descendants. Big history is the best guide we have on how to do this because it is based not on superstition but on evidence and the logic of science.Therefore big history replaces religion for those who share this set of assumptions about truth and reality. It is a modern creation myth derived from science. So, yes; it is the business of science to supply society with myths, they would say, because this is what the evidence tells us.50 *****

A bit of theoretical refitting Whatever your motivation for finding your way to this chapter in the first place, you have soldiered through to the end.You’ve examined all the critical aspects of the discipline’s construction and thought through the key theoretical questions.What are your ideas about big history now? Do you feel less confused? Or more confused? Or more or less confused? The reason I ask is that I am hoping you will see by now that your relationship with big history, and big history’s future standing as a scholarly field, depend upon your willingness, and the willingness of teachers and students and the interested general public to live with ambiguity.51 I wish I could tell you that having sifted through these various philosophical problems everything will now be clear and you can get on with the business of doing the real stuff. But I can’t. It doesn’t work that way.What happens now is that you need to make a choice. But don’t worry.This is a good thing. And I’ll let you in on a little secret: physicists, geologists, chemists, and biologists have to live with ambiguity too. Well, maybe not chemists, but other scientists do. And that’s okay. Remember Bachelard’s dictum: “No one can say they have a scientific mind unless they are certain that at every moment of their thinking life they are reconstructing all their knowledge.” Recall too that scientific history is a discourse and that there will always be a range of interpretations across which the concept moves. The key to not losing your way is to know why you think the way you do, why you have chosen the methodology you use, why you think one way and not another, and how all these viewpoints are related. In short, it is incumbent upon each of us to examine the interpretive framework of the problem we are working on, to understand our own imaginative vision, and to decide where we are going to locate ourselves. And this is what you will have to decide. Within the range of possibilities, which is the best one suited for you? This will be difficult to accept for true believers committed to a single scientific method, and this is also what makes big history’s ideological position such a delicate balancing act. It brings back to me an incident from the 2016 IBHA meeting in Amsterdam when at lunch I floated to one of the executive board members the idea that a unifying 243

David Blanks

paradigm might not be necessary and was told in response that I was “at the wrong conference.” Not everyone will have a sense of humor about these things because it cuts too close to the bone. Questioning the fundamentals of their belief system pisses them off. But if you want to be serious about this, then you will have to take a deep look at your own beliefs and ultimately you will have to make a decision. Will you insist on a unified paradigm? Or are you willing to work with a diverse, pluralist approach to knowledge? And here’s the tricky one. Do you believe that the way we teach big history must be consistent with the way we research it? Talk of running a discipline at two different levels feels awkward. It seems somewhat elitist for one. But the truth is that we are doing this already whether we recognize it or not. The discourses in play at professional conferences are worlds apart from those we deploy in the classroom. And in any case, this is how science works as well: the diagrams of the atom and the solar system found in textbooks do not remotely correspond to reality. Physicists and astronomers know this, but they assign those textbooks anyway and teach those courses the way their teachers taught them. Working at two levels is a time-honored and broadly accepted practice and need not cause us too much hand wringing. On the contrary, not doing it is the real obstacle. To remain in the realm of simple science when it comes to education is consistent with the moral imperative to make responsible choices about the future and consistent with the dictates of evidence and reason; however, at the point at which we want to critique popular culture, we must of necessity enter into the realm of difficult science. The only other viable alternative is to bring the theory into the classroom. This would require introducing into the big history curriculum historiography as well as the history and philosophy of science. It would require questioning everything. It was exactly what Gaston Bachelard was advocating for science education in mid-­twentieth century France—and some within big history have been making the same argument. What do you think? Would not the big history classroom be the ideal place for talking about how we know things and for debating what they mean? Peter Novick observed that in times of ideological consensus, truth needs to be one thing, but that in more contentious times, a pluralist, perspectival orientation is needed in order to maintain professional civility between competing schools.52 A pluralist ­position—I think I want to call this ideodiversity—is every bit as crucial to the future of our planet as biodiversity,—and for the same reasons. It benefits rather than threatens the big history agenda.Why alienate a sizable majority of the world’s students and scholars? They already share the same social and ecological goals, and they could be partners in our global endeavor.Why exclude them in the name of insisting on a narrowly defined science?53 Let them in and allow the process of natural selection to eliminate outlandish ideas from the meme pool. Just as a central research goal for big history research should be fostering discussions about the ways in which philosophy, theology, political theory, and aesthetics shape big history, perhaps the teaching goal should be for big historians to have “thoughtful conversations that make room for people with diverse points of view and to develop curricula that does the same in schools and universities.”54 The unified, objective, scientific foundation of big history is a myth. I do not mean this in the old-fashioned sense of myth as something that is untrue, but rather in the anthropological sense of myth as guiding principle. Myth-making is a vital human 244

Big history and critical theory

function. We must have myths to explain ourselves to ourselves. As much as we are pattern-seeking creatures we are likewise story-telling creatures. Symbolic language makes possible symbolic concepts and symbolic concepts make possible symbolic worldviews. When myths emerge from within a culture they do so through an organic process that can sometimes give birth to entire systems of thinking. In the present case the concept of scientific objectivity has brought forth a worldview that serves the emotional, psychological, and communal needs of its students, its teachers, its practitioners, and segments of the educated public. Faith in a unified, objective science is big history’s imaginative vision. It integrates and stabilizes professional activity, verifies procedures, purges preconceptions, insures solidarity, defends the borders, and guards against lawlessness and chaos.55 But it can’t fix a leaky pipe. What theory helps us to see is that while we have a choice about which myths or visions we wish to use to help us understand the physical world, we do not have the choice of understanding it without using any myths or visions at all.56 Big history, like all imaginative visions is a partial truth, a dream that can help us shape our enterprises, but one that will mislead us if we trust it on its own.57 In his 1978 book On Human Nature, the sociobiologist E. O. Wilson argued that myths are narratives by which a “tribe’s special place in the world” is explained in a way that is consistent with the audience’s understanding of that world.58 This is precisely how big history operates—except the big history tribe is global, and therefore the myth must accommodate a wide variety of beliefs. Sociobiology, also known as behavioral ecology, seeks to explain the evolution of animal behavior by means of naturally occurring genetic processes.59 When extended to human populations, especially in popular science, it can become markedly deterministic, as when someone says that humans are “hard-wired” for aggression. In his most recent book, Wilson describes his approach as interdisciplinary and scientific, as being based upon the best available empirical evidence and scholarly methods. It is, he writes, science committed to fact without reference to ideology or religion. It bridges the gap between science and the humanities. It aims to understand human society. Ultimately, it aims to change it for the better.Wilson hopes that adopting his scientific worldview will help save the environment and bring an end to sectarian conflict.60 Big history shares these hopes. And it shares the methodology. The year after Maps of Time (2004) was published, Christian explicitly cited Wilson’s Consilience:The Unity of Knowledge (1998) as a model.61 So too last year in the first number of the Journal of Big History, he wrote that “Big history represents an attempt at what E.O. Wilson has called ‘consilience,’ a return to the goal of a unified understanding of reality, in place of the fragmented visions that dominate modern education and scholarship.”62 In many ways, big history shares Wilson’s entire program, which amounts to a form of popular science writing that he called the “evolutionary epic.”63 In big history, we call it the “modern, scientific creation myth” or the “scientific origin story.” Following the success of On Human Nature, the notion of the evolutionary epic found traction among a much larger genealogy of popular science writers including Brian Swimme and Thomas Berry, Connie Barlow, Russ Genet, Ursula Goodenough, Loyal Rue, Mary Evelyn Tucker, and the big historians Christian, Brown, and Spier.64 If you are unfamiliar with Swimme and the others, they fall into the “mystic” camp, and the point I am making is that the mystics and the scientists are not as far apart as the 245

David Blanks

scientists sometimes try to claim. Christian, Brown, Goodenough, Rue, and Swimme, along with John Mears and Craig Benjamin, all made the trip to (or contributed to the proceedings of) a conference on the evolutionary epic that Cheryl and Russ Genet organized in Hawaii in January 2008. In the edited volume, which includes essays grounded in scientific materialism, essays on science as story-telling, ecological morality tales, essays that I would describe as “New Age,” essays on intelligent design, and at least one paean to E. O. Wilson, Christian writes: “The evolutionary epic (a.k.a. ‘big history’) can be defined as a particular way of thinking about, or imagining, or describing, the past. It is a historical genre.”65 Coming back the other way, ­Goodenough and Tucker, along with John Haught, spoke at the 2018 IBHA Conference at Villanova University. If it is difficult to establish and maintain disciplinary boundaries, it is because people are travelling back and forth across that border all the time.66 Like the evolutionary epic, big histories follow an established pattern.67 First, there is a survey of a general problem related to science. Commonly this concerns the perceived “two cultures” divide. The main part of the work is then dedicated to an excited description of all the new science.68 In big history’s case this means radiometric dating (the “chronometric revolution”), the discovery of DNA, the discovery of cosmic background radiation, and plate tectonics. As noted above, it also means downplaying its ties to earlier universal histories. This is typical of this genre, which has to be “new” in order to be persuasive. In the conclusion of the evolutionary epic (a.k.a. big history), there is always a return to the philosophical problem posed at the beginning, and a call for a new morality, conclusions which we now know do not flow easily from the facts themselves. These meanings only appear when those facts are narrated/interpreted by us. As Ian Hesketh puts it, the meaning of big history derives not from the empirical evidence but from the aesthetic and moral choices that big historians make—thus the idea of myth, the epic mode of its plot, and the futuristic, moralistic conclusions.69 This makes big history “less a contribution to historical knowledge than it is a narrativization of one or another worldview.”70 Wilson thinks that “the evolutionary epic is probably the best myth we will ever have.”71 Big history agrees with this. But theory tells us that it’s not as simple as doing the science, finding the facts, and coming to a consensus about what they mean. If the discipline is going to last, we need to understand how our house is constructed, how it works. And we need to build it big enough to accommodate everyone. In this spirit, and informed by theory, I close by offering my own definition. Big history is a methodologically reflexive, scientific approach to the entirety of the material and human past that is interdisciplinary and open-ended. This means we can share broad assumptions about how the world works while disagreeing about what it means.

Notes 1 See Sharlene Sayegh and Eric Altice, History and Theory (Boston, MA: Pearson, 2014). 2 Fred Spier, The Structure of Big History: From the Big Bang Until Today (Amsterdam: ­A msterdam University Press, 1996). 3 C. P. Snow, The Two Cultures and the Scientific Revolution (Cambridge: Cambridge University Press, 1959), and David Christian, “Bridging the Two Cultures: History, Big History, and Science,” Historically Speaking 6, no. 5 (May/June 2005): 21–26.

246

Big history and critical theory

4 Daniel Little, “Philosophy of History,” Stanford Encyclopedia of Philosophy, https://plato. stanford.edu/entries/history/#Aca and History and Theory: Studies in the Philosophy of History, www.historyandtheory.org/journal.html. 5 Peter Novick, That Noble Dream: The “Objectivity” Question and the American Historical Profession (Cambridge: Cambridge University Press, 1988), 11, 12, 15. 6 George Lakoff and Mark Johnson, Philosophy in the Flesh: The Embodied Mind and Its Challenge to Western Thought (New York: Basic Books, 1999). See also Lakoff and ­Johnson, Metaphors We Live By (Chicago: University of Chicago Press, 1980) and Jeremy Lent, The Patterning Instinct: A Cultural History of Humanity’s Search for Meaning (New York: Prometheus Books, 2017). 7 My inspiration comes from Mary Midgley, The Myths We Live By (Abingdon, Oxon: Routledge, 2011). First published in 2004. 8 Hayden White, Tropics of Discourse (Baltimore, MD: Johns Hopkins University Press, 1978), 126–127. 9 William Katerberg, “Myth, Meaning and Scientific Method in Big History,” Origins V, no. 12 (December 2015): 3. 10 Fred Spier, “On the Pursuit of Academic Research Across All the Disciplines,” Journal of Big History 1, no. 1 (2018), 20-39; and Fred Spier, “Big History is Not an All-­ encompassing World View,” Origins VI, no. 2 (February 2016), 3–5. 11 Graeme Donald Snooks, The Collapse of Darwinism and the Rise of a Realist Theory of Life (Oxford: Lexington Books, 2003), 7. 12 Midgley, Myths, p. 33. 13 Cynthia Stokes Brown, Big History: From the Big Bang to the Present (New York: W.W. Norton & Company, 2007), xi–xv; Fred Spier, Big History and the Future of Humanity (Oxford: Wiley-Blackwell Publication, 2010), ix–xv; David Christian, “The Case for Big History,” Journal of World History 2, no. 2 (Fall 1991): 223–238; Fred Spier, “Unexpected Goldmines,” International Big History Association, www.youtube.com/ watch?v=WZBeJQD3NJQ; and Walter Alvarez, A Most Improbable Journey: A Big History of Our Planet and Ourselves (New York: W.W. Norton & Company, 2017). 14 William Katerberg, “Is Big History a Movement Culture?” Journal of Big History 2, no. 1 (Spring 2018): 63–72, quote on p. 63. 15 Mary Midgley, Science and Poetry (London and New York: Routledge, 2001), 48–50. 16 See, for example, Elise Bohan, “Empiricism Is Not a Dirty Word,” Origins VI, no. 1 ( January 2016): 3–6. 17 Freud believed that to assert that the only possible source of knowledge about the universe was the “intellectual manipulation of carefully verified observations” was “empty and misleading.” It ignores revelation, intuition, inspiration, in short, “all the spiritual demands of man and all the needs of the human mind. “A Philosophy of Life,” New Introductory Lectures on Psycho-analysis (Hogarth Press, 1933), www.marxists.org/reference/ subject/philosophy/works/at/freud.htm. In this regard it is of some interest that Gaston Bachelard (1884–1962), whom I discuss below, employed this idea of Freud’s as a stepping stone towards his own notion of “epistemological obstacles.” See Georges Canguilhem, “Gaston Bachelard, psychanalyste dans la cite scientifique?,” Il Protagora: rivista di filosofia e cultura 24 (1984): 19–26. 18 David Blanks, “Cosmic Evolution in the Cradle of Civilization” in From Big Bang to Galactic Civilizations: A Big History Anthology, Vol. II, Education and Understanding: Big History around the World, eds. Barry Rodrigue, Leonid Grinin and Andrey Korotayev (Delhi: Primus Books, 2016), pp. 295–317 and “Towards A Theory of Big History,” Origins IV, no. 4 (April 2014): 2–6.

247

David Blanks

19 John F. Haught, The New Cosmic Story: Inside Our Awakening Universe (New Haven and London: Yale University Press, 2017); Ursula Goodenough, Sacred Depths of Nature (New York: Oxford University Press, 1998); and Cheryl Genet, Russell Genet, Brian Swimme, Linda Palmer, Linda Gibler, ed., The Evolutionary Epic: Science’s Story and Humanity’s Response (Santa Margarita, CA: Collins Foundation Press, 2009). 20 Nasser Zakariya, A Final Story: Science, Myth, and Beginnings (Chicago and London: University of Chicago Press, 2017). 21 David Christian, Cynthia Stokes Brown, and Craig Benjamin, Big History: Between Nothing and Everything (New York: McGraw Hill Education, 2014), 4; and David Christian, “The Return of Universal History,” History and Theory 49, no. 4 (December 2010): 6–27. 22 Spier, Big History and the Future of Humanity, 210. “From a very detached point of view, one may argue that, in principle, there is no reason why scientific principles ought to be applied to analyzing data in the present to reconstruct an account of events that may once have happened. One may, for example, decide to accept literally what sacred texts have to say regarding the past. This may not be scientific in the current meaning of the term, but I cannot see any reason why this would be an issue as long as one does not care about science.” 23 David Masci, “For Darwin Day, 6 Facts about the Evolution Debate,” Pew Research Center, www.pewresearch.org/fact-tank/2017/02/10/darwin-day/ (accessed October 5, 2018). 24 Herbert Butterfield, The Whig Interpretation of History (New York: W.W. Norton & Company, 1965). 25 David Christian, Origin Story: A Big History of Everything (New York: Little, Brown and Company, 2018), ix–x. 26 Zakariya, A Final Story, 419–430; Ian Hesketh, The Science of History in Victorian Britain: Making the Past Speak (London: Pickering & Chatto, 2011). 27 Thomas F. Gieryn, “Boundary-Work and the Demarcation of Science from Non-­ Science: Strains and Interests in Professional Ideologies of Scientists,” American Sociological Review 48, (December 2016): 781–795. On the pitfalls of trying to decouple large-scale history from Eurocentrism, see Barbara Weinstein, “History Without a Cause? Grand Narratives, World History, and the Postcolonial Dilemma,” International Review of Social History 50 (2005), 71–93. 28 The following discussion draws heavily on Novick, That Noble Dream, especially Chapter 1, “The European legacy: Ranke, Bacon, Flaubert,” 21–46. See also Georg G. ­Iggers, “The Image of Ranke in American and German Historical Thought,” History and Theory 2, no. 1 (1962): 17–40. 29 William Grassie, “Is God Big Enough for Big History” in A 21st Century Debate on Science and Religion, eds. Shiva Khalil, Fraser Watts, and Harris Wiseman (Cambridge: Cambridge University Press, 2017), pp. 72–87. 30 Fritjof Capra and Pier Luigi Luisi, The Systems View of Life: A Unifying Vision (­Cambridge: Cambridge University Press, 2014); David Faust, The Limits of Scientific Reasoning (­M inneapolis: University of Minnesota Press, 1984); and Hayden White, The Content of the Form (Baltimore, MD: The Johns Hopkins University Press, 1987), 26–57. 31 My thinking on this follows Mary Midgley, Science as Salvation: A Modern Myth and its Meaning (London and New York: Rutledge, 1992). 32 In the 1760s Voltaire (1694–1778), referring to the Catholic Church, took to signing his letters “Écrasez l’infâme!” A propos to big history, Voltaire also believed that religion would eventually be replaced by science. 33 Adam Frank, The Constant Fire: Beyond the Science vs Religion Debate (Berkeley and Los Angeles: University of California Press, 2009), 29–30. 248

Big history and critical theory

34 35 36 37 38 39

40 41 42 4 3

44 45 46 47 48

49 50

51

5 2 53 54 55 56 5 7 58 59

Midgley, Science and Poetry, 66–67. Midgley, Science as Salvation, 61. Novick, That Noble Dream, 33. Novick, That Noble Dream, 33. Novick, That Noble Dream, 33–40. Personal letter to Henry Fawcett, 18 September 1861, Darwin Correspondence Project, University of Cambridge, DCP-LETT-3257, www.darwinproject.ac.uk/letter/DCPLETT-3257.xml. Also cited by Novick, That Noble Dream, 35. Jared Diamond, Guns, Germs, and Steel: A Short History of Everybody for the Last 13,000 Years (London: Vintage Books, 2005), 405–425. First published in 1997. Nasser Zakariya, “Is History Still A Fraud?,” Historical Studies in the Natural Sciences 43, no. 5 (November 2013): 641. Fred Spier, Big History and the Future of Humanity, 5. William H. McNeill, “Passing Strange: The Convergence of Evolutionary Science with Scientific History,” History and Theory 40, no. 1 (February 2001): 5. See also “History and the Scientific Worldview,” History and Theory 37, no. 1 (February, 1998): 1–13. Gaston Bachelard, The Formation of the Scientific Mind: A Contribution to a Psychoanalysis of Objective Knowledge, trans. Mary McAllester Jones (Manchester: Clinamen Press, 2002), 19. Gaston Bachelard, Le Matérialisme Rationnel (Paris: Presses Universitaire de France, 1953). Spier, “Big History is Not an All-encompassing World View.” The First Three Minutes: A Modern View of the Origin of the Universe, updated edition (New York: Basic Books, 1993), 154. First published in 1977. This hypothesis was established by Alvarez, and his father, Luis Alvarez, who won a Nobel Prize in Physics in 1968, along with nuclear chemists Frank Asaro and Helen Michel. Walter Alvarez, A Most Improbable Journey: A Big History of Our Planet and Ourselves (New York: W. W. Norton & Company, 2017). See, for example, David Christian, “From Mapping to Meaning” in Creation Stories in Dialogue: The Bible, Science, and Folk Traditions, eds. R. Alan Culpepper and Jan G. van der Watt (Boston, MA: Brill Press, 2016). Novick makes some interesting observations on the degree to which the “tolerance for ambiguity” was present among American historians in the mid-twentieth century. This gets into questions of cognitive style, sensibility, and personality—all highly relevant for big history. Not everyone has what the English Romantic poet John Keats called “Negative Capability”: that is the capability of being in uncertainty or doubt. Novick, Noble Dream, 274–275. Novick, Noble Dream, 5. Blanks, “Cosmic Evolution,” passim. Katerberg, “Myth,” 6. These concepts are drawn from Eliade, Malinowski, Durkheim, and Lévi-Strauss in Novick, Noble Dream, 1–17. Midgley, Science as Salvation, 13. Midgley, Myths, xiii. E.O. Wilson, On Human Nature, (Cambridge MA: Harvard University Press, 2004), 189. (First published in 1978). E. O. Wilson, Sociobiology: The New Synthesis (1975). For a fuller explication, see especially Catherine Driscoll, “Sociobiology”, in The Stanford Encyclopedia of Philosophy (Spring 2018 Edition), ed. Edward N. Zalta. https://plato.stanford.edu/archives/ spr2018/entries/sociobiology/. 249

David Blanks

60 E. O. Wilson, The Meaning of Human Existence (New York and London: W.W. Norton, 2014). 61 David Christian “What is Big History?” The Journal of Big History 1, no. 1 (2017): 4–19. 62 Christian, “What is Big History?,” 4. See E.O. Wilson, Consilience: The Unity of Knowledge (London: Abacus, 1998). 63 Ian Hesketh, “The Story of Big History,” History of the Present: A Journal of Critical History 4, no. 2 (Fall 2014): 171–202 and Ian Hesketh, “The Recurrence of the Evolutionary Epic,” Journal of the Philosophy of History 9 (2015): 196–219. See also Martin Eger, “Hermeneutics and the New Epic of Science” in The Literature of Science: Perspectives on Popular Science Writing, ed. Murdo William McRae (Athens: The University of Georgia Press, 1993), pp. 186–209. 64 Hesketh, “Story of Big History,” 182 and Hesketh, “Return of the Evolutionary Epic,” 198. 65 David Christian, “The Evolutionary Epic and the Chronometric Revolution” in The Evolutionary Epic, eds. Genet et al. (Santa Margarita, CA: Collingswood Foundation Press, 2009), pp. 91–99, quote on p. 91. 66 Also note Katerberg’s observation that “the centrality of both impulses—‘scientific’ and ‘religious’—is evident in publications such as The Evolutionary Epic (2009) and at Big History conferences, where papers, panels and book tables of a sort common to scholarly meetings sit alongside displays and presentations of educational material for children and people at spiritual retreats. Tension between these impulses was evident at the International Big History Association (IBHA) conferences at Grand Valley State University in 2012 and at Dominican University in 2014, where there were notes of disquiet, criticism, and even occasional disgust at expressions of religion and spirituality” (“Myth, Meaning,” 3). 67 Egers, “Hermeneutics,” especially 189–192. 68 Hesketh, “Story of Big History,” 182–183. 69 Hesketh, “Story of Big History,” 182–183. 70 Allan Megill, “’Big History” Old and New: Presuppositions, Limits, Alternatives,” Journal of the Philosophy of History 9 (2015): 306. 71 On Human Nature, 201.

250

11 BIG HISTORY, MORALITY AND RELIGION Cynthia Stokes Brown

What do religion and morality look like when viewed through the lens of big ­history? When did they emerge in the cosmic story? What are we trying to look at when we aim our lens toward morality and religion? What relative influence has religion had in human history compared with other forces? These are some of the questions I want to address in this chapter. Big history is one of the current terms for telling the whole story from the Big Bang to the present. On this scale, many aspects highly significant to human lives tend to disappear from view—for instance, art, music, literature, dance, and sports. On this largest scale the lens of big historians usually focuses on energy flows, ­political/social organization, trade, environment, and resources. Religion hovers on the periphery of these major themes; it merits some attention because it is closely connected to political/social cohesion and organization and because, at least in the past, it has been a major avenue through which humans formed their conception of the world. For these reasons this paper is offered as a supplement to other accounts of big history. In it I will use the lens of big history, but will zoom in a bit to bring religion into focus. Discussing religion in a scientific way is fraught with difficulty, since most religions, certainly those since the development of agriculture, point to something beyond empirical knowledge of nature, something supernatural or transcendental. Yet scientists, by methodological choice, cannot evaluate anything beyond the natural, empirical evidence available. Big history is a scientific origin story and like science, is limited to what can be known through natural, empirical evidence. It can examine religious practices and describe their history, but it cannot make a judgment one way or another about supernatural claims. In this way, big history is by definition methodologically materialistic or naturalistic, to use the language of philosophers. Materialism is the assumption that all reality consists of what we call matter/energy, two forms of the same thing. Naturalism is the assumption that all reality is natural, but may consist of some additional as yet unknown components (Goetz and Taliaferro, 6). Big history can be used 251

Cynthia Stokes Brown

as a foundation on which to add further metaphysical, i.e., beyond nature, beliefs, but it cannot offer those itself. Hence, in this chapter, I will not attempt to evaluate the validity of religious content, but will simply describe a selection of religious beliefs, practices, and institutions through time, suggesting their interplay with other cultural forms. In a single chapter, this can be done only in a large-scale, summary way. Consistent with scientific methodology, I shall attempt to do this from as neutral, objective point of view as possible, without taking sides on the issues raised, and without making judgments about particular religions or religion in general.

Definitions Since the beginning of civilizations, morality and religion have been so closely intertwined that we have difficulty thinking of them as separate phenomena.Yet recent studies from ethnology have shown that other animals, at the very least mammals, display behavior that can be called moral, or at least proto-moral, while not displaying behavior that can be called religious. Therefore, a distinction must be made between these phenomena. I am using morality to mean the suite of other-regarding behaviors that enable social animals to function as groups, with some flexibility and choice of behaviors. These behaviors include three large constellations—cooperation, empathy, and ­fairness—that umbrella even more nuanced emotions and behaviors.These behaviors reach their most complex level in humans, developing gradually through natural selection. Biologists do not yet widely accept the idea of morality in non-human animals, but biologists and philosophers working together are making a strong case for it (see Bekoff and Pierce 2009; Churchland 2011). A definition for religion is more difficult to attain. Nearly all known human societies have had “religion” or something like it. It seems to fill some universal human need and/or spring from some common characteristic of the human mind. It is a complex phenomenon consisting of multiple aspects—experiences, behavior, practices, doctrines, and institutions. Religion can even be defined in such a way that any worldview that has utmost importance in one’s life can be seen as a religion, which could be capitalism, Marxism, science, big history, or even loyalty to the local sports team. I will survey several contemporary definitions of religion and then will synthesize them into a formulation as simple as possible to use in this paper. To begin with etymology, the Latin roots of the word “religion” indicate a meaning of “to bind back together.” William Grassie, who promotes the scientific study of religion, says that currently there is no agreement on a definition of religion. He points out that at the annual meetings of the American Academy of Religion (AAR) there are hundreds of interest groups with little commonality. Yet he believes that whatever religion is, humans have a propensity for it and that it serves diverse human purposes. Studying it in a scientific manner free of prejudices and prejudgments proves quite rare, since many social scientists are motivated either by an atheistic anti-religiosity or by a bias as committed believers wanting to validate a religious worldview (2010: 14–15). 252

Big history, morality and religion

As a first example of a possible definition, here is one from Nicolas Wade (1942–), an English science writer who has worked for the New York Times, and Nature and Science magazines. In 2009, he defined religion as: …a system of emotionally binding beliefs and practices in which a society implicitly negotiates through prayer and sacrifice with supernatural agents, securing from them commands that compel members, through fear of divine punishment, to subordinate their interests to the common good. (2009: 15) Sociologists and anthropologists tend to omit the supernatural from their definitions. The U.S. sociologist, Robert Bellah (1927–2013), defined religion in at least two ways in 2011. One of them paraphrases the U.S. anthropologist, Clifford Geertz (1926–2006): “…religion is a system of symbols that, when enacted by human beings, establishes powerful, pervasive, and long-lasting moods and motivations that make sense in terms of an idea of a general order of existence” (2011: xiv). The Dutch chemist, anthropologist, and big historian, Fred Spier (1952–), elaborated Geertz’s definition in 1994 by saying that religion is: A complex of empirically neither verifiable or falsifiable representations and practices with the aid of which powerful, pervasive and long-lasting moods and motivations are established in people, with conceptions of a general order of existence, clothed with such an aura of factuality that the moods and motivations are often regarded as uniquely realistic. (1994: 17–18) Generally the academic study of religion by historians, sociologists, and anthropologists focuses on religion as an institution. In contrast, the British theologian and philosopher of religion, John Hick (1922–2012) focused on the inner experience of religion. He believed in 2010 that inner religious experience provides “occasional fortunate glimpses of a reality beyond the physical… we have…to postulate an ultimate transcendent Reality, whose nature is beyond understanding in human terms (transcategorical), and which is being humanly apprehended and responded to within different cultural contexts as the various world religions” (2010: xiii–xiv). Since the lens of big history is focused on the bigger picture, I cannot here delve more into personal religious experience.Yet I acknowledge the importance of it in human history. Intense research is currently being focused on individual consciousness, which is still mostly a mystery; perhaps soon there will be more answers to report. Loyal Rue, a U.S. philosopher with a position called religious naturalism, defined in 2005 the essence of religion as: “a myth, a narrative integrating ideas about how things ultimately are and which things ultimately matter” (2005: 143). In this chapter, I will use Rue’s definition with short additions: “Religion is a human, socially-enacted myth or narrative integrating ideas about how things ultimately are (facts) and which things ultimately matter (values).” I chose Rue’s definition because it excludes the supernatural and focuses on the narrative story. It 253

Cynthia Stokes Brown

combines what reality is (religion) and what matters most (morality). I have added “socially enacted” to indicate that one individual’s belief in a story does not constitute a religion.

Pre-human morality (13.8  billion–200,000 years ago) To begin, we must consider the relationship of religion to morality. Most definitions of religion, including the one we are using, suggest that religion and morality are intertwined, that religious belief suggests, promotes, and supports moral behavior. Yet as we go far back in time looking for the emergence of morality and religion, we encounter an unexpected phenomenon—at least precursors of morality emerge in other animals long before humans appear on the scene. Religious behaviors (dance, trance, rituals, sacrifices, prayer) appear later as behaviors distinct to modern humans, or at least to later hominins. How far back can we find evidence for moral behavior? If we consider cooperation as moral behavior, then according to a few physicists cooperation goes back to the formation of the first stable subatomic particles—protons and neutrons—shortly after the Big Bang. Their formation can be seen as a cooperative act of their three constituent quarks mediated by gluon color forces. Viewed in this way, cooperation can be seen as the key principle in the evolution of the universe, though that is not yet the consensus view of mainstream science (David Hookes). Biologists are in more agreement that cooperation has been a basic part of the evolution of living organisms from the beginning of life. Biologists in the last century and a half have emphasized competition in the struggle for existence, perhaps influenced by industrialization and the general cultural milieu. (However, Darwin himself recognized the importance of cooperation as part of evolution.) Yet since the groundbreaking ideas of the U.S. biologist, Lynn Margulis (1938–2011), many biologists today recognize that cooperation is a significant component in the struggle for existence. See Margulis and Sagan (1986) and Nowak and Highfield (2011) for a presentation of these ideas to the general public. The emerging consensus on the balance of cooperation and competition in evolution goes something like this. As individual organisms compete to survive they find it helpful to cooperate with other individuals in the group, especially against outsiders. The basic formula seems to be: cooperate with insiders, while competing with outsiders. Competition and cooperation seem to be complementary aspects of the same process (Spier 2013). Even single-celled organisms without a nucleus (prokaryotes) find cooperation helpful. They group into mats, sticking together without further interaction. The oldest evidence of this are the fossils of stromatolites that are 3.45 billion years old found on the western coast of Australia. Stromatolites are still forming today in a few shallow and tropical waters, where tiny bacteria find that they can cling to the rock without being washed away by the tides more effectively in mats than separately. It is still possible to watch prokaryotes cooperating in ways like they did more than two billion years ago. About the time (2.5–1.5 billion years ago) that prokaryotes were developing ways to use oxygen (respiration), eukaryotes began to emerge. These single-celled 254

Big history, morality and religion

organisms were much larger, had enclosed nuclei, and had a skeleton of protein fibers. They also contained organelles, or separately functioning structures, like organs in a body. One organelle, the mitochondria, is the site of aerobic respiration that fuels the cell’s metabolism. Another organelle, in plants, is the chloroplast that conducts photosynthesis. Biologists put so much emphasis on competition in nature that it took until the 1970s for them to be convinced that eukaryotic cells must have formed by a symbiosis of early types of prokaryotic cells. In other words, the earliest cells cooperated and eventually formed a permanent symbiosis in which one kind of cell enclosed two other kinds to form a new kind. As Lynn Margulis surveyed evolution, she concluded: …the view of evolution as chronic bloody competition among individuals and species, a popular distortion of Darwin’s notion of ‘survival of the fittest,’ dissolves before a new view of continual cooperation, strong interaction, and mutual dependence among life forms. Life did not take over the globe by combat, but by networking. Life forms multiplied and complexified by co-opting others, not just by killing them. (Margulis and Sagan 1986: 28–29) Once multi-cellular life developed some 1000 million years ago, cooperation increased dramatically. Since it depends on sensation and communication, cooperation could increase as living organisms developed hearing, sight, and smell. By the time of the dinosaurs the first parental caring for offspring, such as bringing food to their offspring, may have appeared. This hypothesis is based on fossilized baby duck-billed ­dinosaurs found in 1978 in Montana by Jack Horner (1946–) and named “­Maiasaurus,” Greek for “good mother lizard” (Horner 2001: 82, 102–103). Some degree of parental care has also been found among fish, squid, crocodiles, and ­rattlesnakes (Bellah 2011: 69, quoting Hrdy 2009). However, most scientists would not call animal behavior other than in mammals “moral” because it seems too rigid and dictated by instincts. To be called moral, animals need to exhibit emotional complexity, enough flexibility to make some choice, and a particular set of cognitive skills. The animals that have been studied sufficiently to be able to conclude that they exhibit moral-like behavior are the great apes, some monkeys, wolves, coyotes, hyenas, dolphins and whales, elephants, rats and mice ­(Bekoff and Pierce 2009: 9). Examples of mammals behaving in moral-like ways abound. For instance, hungry rats will refuse to pull a lever for food if it also delivers a shock to their littermates. Dolphins routinely show kindness and generosity in reciprocal relationships. Wolves hunt in long-lasting packs, taking down elk too large for any single wolf. They take turns on the carcass, with high-ranking members eating first. Skeptics of animal cooperation don’t want to call this cooperation—who knows how much consciousness and intention go on in a wolf ’s mind? (Bekoff and Pierce 2009: 21, 52, 64). The empathy and caring of elephants seems even clearer. Appearing about 60 ­million years ago, elephants have evolved into living in tightly bonded matriarchal groups. Layers of extended family care for the young. The attachment between 255

Cynthia Stokes Brown

mother and infants permits the development of neuro-physiological structures that allow the expression of normal elephant social behavior, which seems to include helping injured members and showing empathy for dying and dead elephants. When elephants experience culling of their group, loss of numbers, and fragmentation, their social structures break down and young males often display unpredictable social behavior and aggressiveness. This has happened with terrible frequency as the number of wild elephants has declined from an estimated ten million in the early 1990s to only about half a million in 2009 (Bekoff and Pierce 2009: 105–06). The American ecologist and psychologist, Gay Bradshaw, finds elephants so similar to humans that she founded the field of trans-species psychology (Bradshaw 2009). Once we come to our nearest relatives, the two remaining species of chimpanzees (common chimps and bonobos), we find levels of caring, empathetic, reciprocal behavior that is exceedingly familiar to us. Chimps exhibit frequent expressions of caring and helping, including embracing, kissing, patting, and holding hands. Since male chimpanzees form no permanent bonds with their mates and offspring, the females carry out childrearing mainly on their own and often form lifelong bonds with their offspring. Older siblings adopt younger ones if a mother dies and may even adopt a non-sibling without a mother or older siblings. In zoos a chimpanzee will sometimes fall into the moat surrounding its enclosure, and the others will rush to the rescue, a risky business since chimps cannot swim (Goodall, 1999a, 1999b, Chap. 10; Waal 2005). After her first ten years of studying chimpanzees, Jane Goodall (1934–) portrayed them as very much like humans, although somewhat nicer. Eventually, however, she witnessed brutal attacks by males against a female of a neighboring community, attacks by a high-ranking female against newborns of her own community, and a fouryear war when one group split away from the other, ending only with the deaths of the split-away group except for young females (1999b, Chap. 4). This behavior of killing within one’s own species or even one’s group is rare among animals. Most animals have limits to violence, whether encoded in their genes and/or learned from the group behavior. Fighting over females, resources, and territory occur, but rarely do animals seek to exterminate each other. Common chimpanzees and modern humans, however, seem to lack the gene(s) that would prevent them from killing large numbers of their own species. Perhaps the predisposition to immoral behavior, as well as to moral behavior, is encoded in our genes (Spier 2013: 12). Modern humans developed from a common ancestor with chimpanzees and bonobos. The line split about 5–8 million years ago, with at least some 18–25 species appearing before Homo sapiens (modern humans) emerged about 200,000 years ago. By about 15,000 years ago only H. sapiens remained. During their 5–8 million years of evolution hominins gradually acquired the full range of moral and immoral behavior that characterizes modern humans.This behavior, much more flexible than in any other species, is made possible by greater mental abilities for recognizing motives and feelings in others and by greater capacity for communication and collective learning. Life for the first hominins who came down out of trees to live on the grassy savannah required and encouraged increasing cooperation. On the savannah early ape people (Australopithecus) were prey; as many as ten different kinds of big cats prowled

256

Big history, morality and religion

the territory. Females needed males for protection; bipedalism enabled people to move faster and seeing further (Waal 2005: 107). Bigger brains began to appear already in various Homo species, but rapid brain growth began once Homo erectus appeared about 1.8 million years ago. Their brain size was about 70% of modern humans, and their pelvises had narrowed and flattened compared with those of chimpanzees to provide the base for standing and running. These changes in pelvis size reduced the diameter of the birth canal, at the same time that brain size was increasing. As a result, babies were born earlier in order to get out. This meant increased helplessness and the need for intense care. Homo erectus are believed to have achieved pair bonding, in which females gave up some of their sexual freedom in exchange for protection and assistance, while males gave up some of theirs to increase the chances for survival of their offspring. (See Bernard Chapois, Primeval Kinship: How Pair Bonding Gave Birth to Human Society.) Pair bonding may have become possible after adult males learned to form a coalition to keep any alpha male from dominating, as argued by Christopher Boehm in Hierarchy in the Forest:The Evolution of Egalitarian Behavior. In another leap of cooperative behavior Homo erectus probably began to use fire for protection, warmth, and for cooking food, which could then include a higher proportion of meat and tubers. Cooking may have fueled brain expansion and certainly promoted social discussion around the fireplace, as argued by Richard Wrangham in Catching Fire: How Cooking Made Us Human. Fossilized bones indicate that H. erectus had the three semi-circular canals in the inner ear that provide balance for running and dancing. When did hominins begin rituals of dance and music? We can only guess. The earliest musical instruments yet found date to about 42,000 years ago. For the hypothesis that singing preceded language, see Mithen, Singing Neanderthals. The big question underlying human moral behavior is: When did full human language ability arise? When did humans reach their full capacity for what Terrence Deacon calls symbolic language? The answer eludes us, since there is no direct evidence for how our ancestors spoke. Members of the species H. erectus are believed to have used a proto-language of simple verbs and nouns. Sometime, probably within the last 200,000 years, gene mutations probably appeared that enabled humans to use grammatical rules properly to express much more explicit information. This is inferred from studying an English family who seemed unable to use grammar properly; they were found to have a mutation on a single gene known as FoxP2, a gene shared by Neanderthals but not by great apes (Christian, Brown, and Benjamin 2014: 90). After examining hominin development, one must conclude that evolution provided modern humans with genes for much of their cooperative, moral behavior, as well as for their aggressive, immoral behavior.The former appears greatly to outweigh the latter—even warfare requires elaborate teamwork. In addition to their gene-controlled behavior, modern humans have cultural, learned behavior that is enabled by their large brains, with learned behavior able to override gene-controlled behavior. The human ability to accumulate and pass down learning seems to most big historians to be their unique feature. Hence, human behavior is a highly complex mix of both genetic and cultural behavior. This mix may

257

Cynthia Stokes Brown

be visualized by thinking of the three layers of our brains. Fred Spier describes it like this: The most inner part, the reptilian brain, may still contain ancient reflexes that are mostly determined by genes. The second layer, the paleo-mammalian brain, may be more flexible, and may be an area where cultural learning is functioning to some extent. The neo-mammalian neocortex on top, in its size unique for modern humans, represents the layer of the human brain that is most susceptible to cultural learning. If this picture is correct, human actions may be determined by very complicated interactions of genes and culture mediated through three different portions of the brain. (2013: 13)

Paleolithic/tribal morality and religion (195,000 years ago) Two hundred thousand years of time—with no writing before five thousand years ago—how can we ever know what our human ancestors were thinking during the first 95% of human history? This conundrum has not stopped scholars from investigating and speculating. Any attempt to describe the origin of human morality and religion is bound to be highly speculative and full of debatable assumptions because the archeological evidence (bones, tools, paint, and drawings) is inconclusive. For years most of the excavations were done in Europe; as a result archeologists hypothesized that some kind of revolution occurred about 40,000 years ago in Europe, in which modern human behavior appeared, even though the biological evolution of H. sapiens had taken place much earlier. This incongruity could not be explained based on European evidence. Once excavations began in Africa, two influential archaeologists, Sally McBrearty and Allison Brooks, could argue that elements of modern behavior can be found in African sites that predate 40,000 years ago. For example, bone points, fragments of red ocher, eggshell beads, and intentional images dating from 100–70,000 years ago were found at Blombos Cave in South Africa. McBrearty and Brooks concluded that modern behavior expanded fitfully as needed over at least 100,000 years. Even the evidence in Europe is highly ambiguous. For instance, consider the socalled “Venus” figurines that are often considered icons of the European Paleolithic. These figurines are found plentifully in Europe from about 26–25,000 years ago. But what did they mean to the people who made them? Archaeologists can only speculate, as they have in a variety of ways. They have seen the figurines as talismans to induce pregnancy, as a general symbol of womanhood, as playing a role in social negotiations near the height of the last ice age in Europe, and as a self-expression of pregnant women, not to mention less lofty speculations, such as that the figurines served as dolls for children or sex objects for adolescent men (Pettitt 2005: 164–165; Brown 2012: 61–62). The study of groups of people who still live in Paleolithic lifeways provides another kind of clue to human thinking during the Paleolithic. Appealing as it is to make inferences from our knowledge of these groups, they do not provide highly 258

Big history, morality and religion

reliable evidence, since these people have also evolved in the last 15,000 years. They have been influenced by agricultural and industrial peoples, and they have been shunted to less fertile land, making their continuance precarious. Nevertheless, many of them have managed to carry forward some of their basic ways of thinking since Paleolithic times. (One example will be discussed later.)

Hypotheses about the origin of religion Various philosophers, anthropologists, sociologists, psychologists, and theologians have offered theories to explain the origin of religion, since it is a complex phenomenon encompassing many aspects—intellectual, experiential, ritualistic, aesthetic, and institutional. Often scholars focus on one aspect of religion to the exclusion of other aspects. Some examples will follow. The French father of sociology, Emile Durkheim (1858–1917), argued that religion is not about explaining natural phenomena, or coping with insecurities, or receiving divine messages; rather, it is primarily about promoting group survival. It emerged as a strategy for legitimating and maintaining the social order.To Durkheim “sacred” meant whatever was of vital interest to the group, as distinct from what was not. For him, religion’s central activity was ritual, carried out to reinforce group solidarity (Rue 2005: 147–148; Haidt 2012: 317). The science writer, Nicolas Wade, also stresses the effectiveness of religion as a means of social cohesion and a way for egalitarian communities to govern themselves. This point of view sees religion as being adaptive, meaning that religious groups were selected by natural selection, since cohesive and cooperative groups generally defeated groups of selfish individuals (Wade 2009: Chap. 3). This is sometimes called multilevel selection, or selection by groups as well as by individuals, an idea currently in vigorous debate. Some contemporary evolutionary biologists and psychologists see religion as non-adaptive, not evolved by natural selection, but as a by-product of other features that are adaptive. For example, the English evolutionary biologist, Richard Dawkins (1941–), hypothesizes that religion is a by-product of other traits that are useful. As an example, he notes children’s propensity to believe what their parents tell them. Dawkins thinks that natural selection builds child brains to believe their elders, since trusting obedience is valuable for survival and reproduction. Believing parents’ religious ideas seems to Dawkins a misfiring of that useful trait (2006: 200–221). The U.S. evolutionary psychologist and philosopher, Daniel Dennett (1942–), ­argues that the human brain is primed by evolution to ascribe action to agents. It tends to ascribe agency to anything that moves and project it onto supernatural agents who cause everything. Thus religion is an accidental by-product of the brain’s agency-detecting abilities (2006). The account that follows is based largely on the naturalist theory of religion developed by Loyal Rue. It honors the complexity of religion, both the individual experience it provides and the social cohesion it promotes, without including any supernatural elements. Other ideas will be included as contrasts or supports to this view. 259

Cynthia Stokes Brown

The gradual emergence of morality and religion Back in the mists of time, about 200,000–50,000 years ago, early modern humans lived in small bands of 10–40 individuals. They hunted and gathered and increased their capacity for symbolic language. They made music, danced in rhythm, and went into trances, sometimes from the repetitive beat and sometimes with the aid of hallucinogenic plants. In their trances and in their dreams they seemed to be in another world, another realm. They were aware that in death something seemed to vanish from the body of the person. By about 100,000 years ago, archeological evidence shows that humans were burying their dead with goods, indicating some belief in existence after death (Rue 2005: 151, quoting Mithen 1996: 178). In the setting of hunting/gathering life, it seems plausible that early humans practiced a campsite morality. Small groups needed minimal social organization. They needed no explicit moral rules because the group was small enough to operate mostly on evolved behavior. Kinship ties governed most interactions, plus reciprocal altruism, i.e., tit-for-tat or “I’ll do you a favor in exchange for a future one.” Life was based on the immediate gratifications of hand-to-mouth survival. To understand other people, early modern humans began to be able to read other human minds, to imagine what another human was thinking. Psychologists call this skill “theory of mind.” It is plausible to conclude that early modern humans used this ability to imagine minds/spirits in other animals and in natural phenomena. They saw everything as alive, with a spirit as well as a material presence, even rocks, trees, and wind. This sort of worldview is often called animism, a term invented by Sir Edward Tylor (1832–1917), an English anthropologist, in his studies of Paleolithic people. In groups of people thinking animistically, certain ones called shamans were able to enter trances more easily than others. They were considered special communicators with the spirits. Shamans could then assist individuals coping with diseases and fears, by soliciting aid from the powerful spirits/beings. By the definition of religion that we are using, following Rue’s ideas, animism would not yet be considered a religion. It had not yet developed a narrative. It had rituals—music, dance, trance, shaman, but it did not yet have a story of how the group came to be and how people should behave. Episodically a group would encounter another group. These meetings must have been meaningful, a chance to trade, to find new partners, and to conduct rituals.Yet these meetings must have also been brief; food supplies would run out, and there were no common rules for behavior. Trouble would often break out. Groups would then try to work out rules for behavior to prevent trouble at the next encounter, since the benefits of such encounters were appealing and useful. In this way the need for a group origin story arose, a story to tell about who they were, where and how they came to be, and why they should follow the new rules. ­Gradually, we can imagine, such stories and rules were in place in many areas of the world, as groups became larger and encountered each other more often. Clearly, there must have been much variation, depending on the environment and the density of population. Some groups would be loosely organized, individualistic and secular, while others were more tightly organized and ritualistic. Anthropologists 260

Big history, morality and religion

believe that, rather than using the term “gods,” people related to (not worshipped) powerful beings or spirits located in their vicinity (Bellah 2011: 137, 141). To illustrate these abstract generalizations, we can look at some examples of current people still living in hunter/gatherer lifeways. But which ones? Often studied are the !Kung in southern Africa and the Aborigines in Australia. Because information about these groups is easily available, I will use as examples the Navahos and Pueblos in the southwestern U.S. as described by the U.S. sociologist, Robert Bellah (1927–2013). The Navajos are descendants of southern Athabaskans, who left the area of the MacKenzie Basin in sub-arctic Canada about 1000 CE and arrived in their present location in southwestern U.S. about 1500, not long before the Spanish arrived. They brought a generic shamanism considered common among North American hunter/ gatherers. Bellah describes a shaman as “an individual who seeks or is sought by a powerful being for direct experience through which some of the being’s power becomes available to the shaman, usually for the purpose of curing” (2011: 163–164). Navajos are still tribal semi-nomads, but they have acquired some cultural traits of agricultural societies from their neighbors, the Pueblos. Navajos acquired sheep and horses from the Spaniards, which enabled them to follow semi-nomadic ways. They live in extended families and gather in larger groupings for temporary purposes, such as rituals. The earliest records of their myths and ceremonies date from the late nineteenth century, and variation in documentation enables many conflicting interpretations. Navajo ceremonies are called “songs.” They are led, not by shamans, but by “singers,” who are believed to have been originally learned the songs by direct contact with powerful beings.The rituals are carried out when needed to cure a person, or to deal with a situation, not in a regular fashion according to the calendar. The central ritual is called “Blessingway.” It includes stories of events just after the emergence of the original people on the present earth-surface. The original people are believed to be present in the ritual, and participants can become one with them. The stories are portable, not tied to one place, and contain no explicit moral code but tell of the relationships necessary for life. In them the wind, air, or breath animates all things and is the means by which people are connected to all beings. Pueblo beliefs provide a contrast to Navajo ones, in that Pueblos live in settled villages dependent largely on produce from their surrounding fields. They have moved from nomadic lifeways to agricultural ones. Their rituals are organized by priestly societies that hand down the teachings; their major rituals are based on the calendar and are linked to the growing seasons of their chief crop, corn. Their origin story is much more elaborate than that of the Navajos and is special to each village, which is considered the center of the world. The Pueblos have a more anthropomorphic pantheon of gods as powerful beings, but they are not yet worshipped and sacrificed to, but rather invoked and identified with. Certainly by the definition being used here, religion and morality have developed in both Navajo and Pueblo societies, since they have constructed narratives that tell what is real and what matters most.The practices of the Navajo are an example of the shamanism/animism practiced much earlier in North American hunting/gathering 261

Cynthia Stokes Brown

societies. Since studies of other current hunting/gathering groups suggest similar practices, we can generalize similar practices to most hunting/gathering societies. For most of our human history, people must have believed in and practiced some form of animism/shamanism that included dancing, singing, and going into trances. These beliefs gradually evolved into stories that explained what was real and what was important.When people began to settle in villages supported by agriculture, their moral and religious beliefs shifted significantly. (For an argument that different levels of energy capture push people toward different interpretations of core human values, see Morris 2015. He argues that the values held by foragers were egalitarian and violent relative to those of farmers, who were more hierarchical and less violent. In contrast, fossil fuel societies are more egalitarian and less tolerant of violence relative to earlier societies.)

Agrarian civilizations (3500 BCE–1000 CE) After the emergence of modern humans some 200,000 years ago, the next threshold in human life occurred about10,000 years ago, when people learned to domesticate certain plants and animals and settled into village life, dependent on their crops and herds. Eventually people produced enough surplus food to allow some villages to expand into towns and then into cities, where some people could specialize in occupations other than farming. With city life came the characteristics of “agrarian civilization”—unequal hierarchies, development of states, rule by kings and elites, coerced tribute, standing armies, specialization of labor, and usually the development of writing. As people settled down in villages they had to learn new forms of self-discipline. They could no longer act primarily on genetically controlled instincts. For instance, they had to resist eating all the available food, because some seeds had to be saved for planting in the spring. They had to adapt to the seasons and learn not to slaughter their animals but to use their milk and eggs and their strength/energy for plowing. People had to learn to restrain themselves in the moment and plan ahead for the future (Spier 1994: 36–37; Christian 2004: 255). Agricultural people also had to learn to resolve internal conflicts; once they settled down it wasn’t so easy to move away after arguments over wives or land. They also suffered more illness than during hunting/gathering days, often caught from their animals. To confront these new challenges, people developed community rituals that helped them determine when to plant and how to resolve conflicts, how to seek healing and appeal to the spirit world for assistance.

Chiefdoms The first step toward the emergence of agrarian civilizations occurred when the egalitarian structures of hunting/gathering society began to change with the beginning of settled life. As food surpluses began to occur in some groups, the need for leadership arose. Leaders were needed for defending their productive land and its surpluses, for mediation with the powerful spirits, and for legal and administrative matters, like dividing up the surplus food. Some communities began to choose a single person 262

Big history, morality and religion

to make decisions on their behalf, a chief or Big Man. (Yes, this person was usually a man, as shown by bones and grave goods.) This chief may have been selected for his talent or as the eldest son of the senior lineage. He was selected by consent and could be easily removed, since he had no standing army but called on the men of his clan when he needed assistance. He engendered a sense of obligation from his group by holding elaborate feasts for everyone and giving away accumulated surpluses. Burial sites clearly reveal the existence of chiefdoms. For example, at Varna on the Black Sea Coast of modern Bulgaria a cemetery from about 4500 BCE revealed 211 graves, of which 170 had fewer than ten items buried with the body, 18 had a much larger collection, and one grave of a 40–50-year-old male contained more than 1000 objects, almost all made of gold (Christian, Brown, and Benjamin, 2014: 122–123). To learn how people in chiefdoms practiced religion, anthropologists study Polynesian societies. This area of the world still had chiefdoms in the nineteenth century because it was late in being inhabited by humans and was isolated from the practices of other places. Anthropologists have reconstructed ancestral Polynesian societies and have concluded that their basic social form was a simple chiefdom. One three-squaremile island named Tokopia, southeast of the Solomon Island group, can serve as an example of how people in chiefdoms practice religion, since this island has been well studied (Bellah 2011: 182–197, footnotes #20–26, 641–642). Tikopians have a longer sense of human history than do hunter/gather people generally. Tikopians keep track of at least ten generations and include a time when gods and men both walked on earth, but then the powerful beings/gods retreated to the spirit abode. In Tikopia the chief is also the high priest and only he, as the representative of his people, performs the rituals while others watch in an open-air temple. The rituals are no longer enacted collectively. Anthropologists call this “worship” rather than “invocation,” although the god being worshipped might be believed temporarily to enter the body of the chief. Tikopians believe their gods behave like their chiefs do, but in an invisible spiritual world. “They [the gods] encapsulated ideas about the structure of Tikopian society,” concludes their ethnographer, Raymond Firth (quoted in Bellah 2011: 188–189).

Early agrarian civilizations If the first step toward power and control was made from the bottom up, by consent of people who needed leadership (consensual power), the next step came from the top down, by those who acquired sufficient resources from the agricultural surpluses to impose their will (coercive power). Consent and coercion worked hand in hand as the process of urbanization developed. Early cities and states appeared independently in at least seven places in the world. The number of places varies, as do the terms used for these entities—archaic societies, early states, early civilizations, etc. They often arose in fertile river valleys—the Tigris/ Euphrates, the Nile, the Indus, the Chang Jiang (Yangtze), and the Huang He (Yellow), but also in other fertile places (Mexican basin, Peruvian mountains, sub-Saharan Africa). The characteristics of early agrarian civilizations also vary; not everyone had writing, for example, but in most of them written language developed as a function of 263

Cynthia Stokes Brown

bureaucracy and trading. The basic shared characteristics arose as city populations became dense enough from agricultural surpluses in the surrounding areas that kinship structures gave way to hierarchy and class stratification, along with specialization of occupations. Elites seized political power, military power greatly increased with standing armies, and elites exploited their populations with enforced tribute of labor and goods. The priestly class aligned closely with the ruling elites, often by kinship. These changes brought great achievements as well as great hardships. The imposition of political and military power could bring at least temporary peace to an area. Then agriculture could become more productive and markets and long-range trade, art, architecture, and literature could prosper. As human population increased, the struggle for fertile land and resources increased, leading to endless small-scale violence and warfare. The only way out of warfare occurred when a successful warrior could assume authority in ways that appeared legitimate. Usually he did this by claiming that he had unique access to and backing from the gods. Often in the development of early states the ruler/king claimed to be divine, to be a god himself, supported by his priests. One exception seems to have been the Chinese Zhou, where the king claimed to be the “Son of Heaven,” but not himself divine (Bellah 2011: 212). Anthropologists seem to agree that the primary political and religious innovation of early states was the invention of divine kingships. The leaders of emerging cities and states enlisted the assistance of the religious realms in order to establish their legitimacy. In doing so they established a dominant state religion that overshadowed and constrained other local religious practices (Bellah 2011; Spier 1994). The Inca state in Peru as described by Fred Spier can serve as a case study for how religion developed in an early agrarian civilization because it occurred about 1400–1532 ce rather than 5000 years ago. It cannot, of course, show us what happens in an expanding agrarian civilization, since Spaniards destroyed the Inca civilization at an early moment in its development. Early Inca rulers were a combination priest/chief who lived in a sacred homestead that also functioned as a temple. Later they moved out, leaving control of the sacred family land to close relatives. The rulers continued to live in the same city (Cuzco) and were linked by close kinship ties to the sun priests, although apparently never considered themselves divine. Yet the rulers performed many religious functions as an important source of power and prestige. The increased hierarchical representation of the supernatural world reflected the social structures of Inca state society, and the power of the sun priests rested mainly on functions they performed for rulers rather than on serving the needs of ordinary people (Spier 1994: 60–61, 96–98). As a result of the domination of powerful elites, of both rulers and priests, standards of moral conduct began to be imposed for their purposes—to legitimate their domination and to pacify their people with at least the appearance of justice and care. These standards became much more complicated as people lived in denser clusters, in more disease, poverty, and coercion, with loss of kinship groups and in closer contact with strangers speaking other languages. Writing developed in most early agrarian civilizations, though not among the Inca, who used a system of knots in ropes to record goods and ideas. Writing provided the opportunity for more reflection on moral and religious issues than did oral 264

Big history, morality and religion

traditions alone. Yet this took place only among the elites who had the opportunity of mastering the new literacy; it was not yet widespread enough to encourage largescale independent thinking. The inequality of early agrarian civilizations sometimes seems difficult to understand when compared with the egalitarian ways of hunting/gathering groups. Why would large numbers of people accept the rule of tiny elites who coerced them? The world historians, William H. and J.R. McNeill, address this question succinctly: …an encompassing process of trial and error rewarded all those changes in social organization, technique, and communication that enhanced deliberate control, both over natural resources and over concerted human effort. We are still caught in this historic process and unlikely to escape it, simply because most people, most of the time, prefer collective and personal wealth and power to poverty and weakness, even at the cost of subordination to rules and commands issued by distant strangers. (42–43)

Expanding agrarian civilizations The process that rewarded changes that increased control over natural resources and human effort continued, as early rulers learned how to expand their powers in the period from about 3000 BCE to 1000 CE. During this time, cities and states developed, and agrarian civilizations spread until they became the dominant type of human community, especially in Afro-Eurasia. Four large-scale trends characterized these expanding agrarian civilizations (Christian, Brown, and Benjamin 2014: 156): (1) Agrarian civilizations increased in size, power, and effectiveness. (2) Networks of exchange developed, especially the Silk Roads of Eurasia, which enabled increasing interconnections among civilizations. (3) Social and gender relations increased in complexity. (4) Yet the pace of change remained slow compared to modern times, with low rates of innovation and growth and Malthusian cycles of diebacks occurring after increased population growth. About the middle of this period of expanding agrarian civilizations, in the mid-first millennium BCE, there emerged in Eurasia individual thinkers who made a profound impact on their societies. Usually named are: Zarathustra, Buddha, Confucius, Lao-tse, the prophets in Israel, and the Greek philosophers. One of the first to draw attention to the creativity in this period was the ­German-Swiss philosopher, Karl Jaspers (1883–1969) who, in 1949, named it the “Axial Age.” He said that it extended from roughly 800–200 BCE and characterized it as the time when “Man, as we know him today, came into being.” Jaspers referred to philosophy appearing for the first time, to individuals thinking for themselves, with rationality pitted against mythical thinking. He could not be confident of the causes of these breakthroughs, but he guessed that wars, the breakdown of empires, and frequent misery favored spiritual creativeness (1953: 13–18). 265

Cynthia Stokes Brown

Since Jaspers, philosophers and sociologists have carried on much discussion about the so-called “Axial Age,” but no agreement has been reached about what characterized axial thinking or why it occurred.Various users define the term differently, but in general the term is used to the emergence of portable, universal religions, of increased self-awareness, of a search for transcendence, of the articulation of a “do unto others” morality, and of the beginning of scientific thinking. Big historians, and even most world historians, have avoided using the term “Axial Age.” David Christian gave it one paragraph in Maps of Time, noting that it was no accident that universal religions appeared in large empires controlling diverse people and at the hubs of trade routes and exchange networks (2011: 319). Fred Spier avoids the term but describes the emergence of moral religion, which he attributes to life in cities where people had to get along with strangers, some of whom were very different (2015: 248). Jaspers adopted the term, “Axial Age,” as a way to break out of the Christian use of Jesus’ birth as the axis, or turning point, of history. With the idea of the Axial Age, Jaspers encompassed all of Eurasia, a step beyond the more limited Christian outlook. But today as we think globally, the term “Axial Age” seems outdated, since it refers only to Eurasia. It elevates world religions over indigenous ones and religions from other places than Eurasia. The characteristics of religious and secular thinking that emerged in the mid-first millennium BCE in Eurasia were part of the expansion of agrarian civilizations there, but not the single axis of human history. Instead of using a single axis, big historians use some series of turning points in human history, with major ones occurring at the emergence of agriculture and at the industrial revolution. What, if any, cautious generalizations can be made about the development of religion and morality during the period of expanding agrarian civilizations? During this period, conditions developed that favored the emergence of portable, universal religions—beliefs that could spread beyond the culture that created them. Foremost among these conditions were vastly increased networks of trade and exchanges, spurred by the development of coinage in the Huang He Valley, the ­Ganges Valley, and the shores of the Aegean Sea during the mid-first millennium BCE (­Graber 2011: 242). Conditions also included more widespread literacy, even though it probably only reached about 10% in Greece, and perhaps 15–20% in China. When large empires broke down, as they periodically did, small states competed against each other for a new primacy. During these times, itinerant intellectuals could criticize prevailing structures and propound novel ideas free from bureaucratic constraints, ideas transmitted both orally and in writing (Brown, 134; Bellah, 269). In the first millennium BCE the human population of Earth at least doubled and then stayed about the same until 1000 CE (Christian, 143). Life in early cities proved immensely difficult for everyone other than the elites. Common people had to deal with poverty, disease, and strangers with strange behaviors. Moral codes were needed to provide guidelines and to wrap individuals inside a moral group, giving them an identity against competing outside groups. The new element in both religions and morality lay in the fact that neither was anymore tied to a specific location. Ideas and codes could be carried anywhere and embraced by anyone. The deities were thought to reside in a transcendent realm, no longer in specific mountains or waters (McNeill and McNeill, 60–61). 266

Big history, morality and religion

Urban populations needed continual replenishing, as death rates in cities usually exceeded birth rates (Christian, 330). The misery of urban life for commoners resulted in people looking for a better life in some other realm after death—or at least contact with some transcendent realm. Finding salvation and eternal life became key themes in many religions. The need to get along in the density of neighborhoods in urban life encouraged the development of sensitivity to the needs of other people. Many observers of world religions note the common theme of their moral codes, namely “doing unto others what you want done to you” or “what you do to others comes back to affect yourself.” Some current philosophers and theologians conclude from the common moral codes constructed during this period that there must exist an objective, a priori, moral truth—something like a transcendental, universal morality (Ronald Dworkin, Religion Without God). Many social, technological, economic, and demographic developments contributed to the emergence of the world religions. The ideas best adapted to specific circumstances survived and formed into religious practices and institutions. These new religions interacted with all the other forces to help their societies expand into flourishing agrarian societies (Personal exchange with Esther Quaedackers). Eventually rulers adopted some of the new religious ideas as state religions in order to legitimate their rule, as Confucianism in China in 206 BCE or as the Roman Empire accepted Christianity in 380 CE. What conditions pressed them to adopt new religions? Was it that the new ideas had spread around—people embraced them, and rulers thought they could get more loyalty from their subjects by adopting popular religions? When the new religion became the required orthodoxy, it suppressed other older remnants, such as Daoism in China and Shintoism in Japan (both never fully suppressed) and paganism in Europe. In general, urban religions tended to lose standards of conduct toward nature, as people became more and more divorced from direct experience of nature. The natural world gradually became considered inanimate and not sacred, as natural was distinguished from supernatural and deities were viewed as increasingly transcendent. This generalization seems less valid for East Asian and indigenous religions than for the Abrahamic religions (Trigger, 44; Spier, 2013, 4; Grim and Tucker, 23–24). The Chinese big historian, Sun Yue, points this out forcefully: …the ‘unity of Heaven and Humanity’ involves a totally different mode of thinking [from the dualistic pattern of contemporary science], the one that incorporates the whole humanity, the earth, and Heaven in a grand integrated scheme of Oneness. In other words, the Chinese answer to the problem of humanity and nature is that nature and humanity mutually shape and condition each other through numerous rituals, consciously instituted or unconsciously there, so as to maintain a harmonious sustainability. (116) Theologians tend to find divine explanations for the sudden, concurrent appearance during the mid-first millennium BCE of what became world religions. Some, such as John Hick, believe this phenomenon can only be understood on the assumption of 267

Cynthia Stokes Brown

transcendent, divine self-disclosure (Rue, 2005, 145). This contrasts with sociologists and historians, who usually see the appearance of world religions as a result of changing human needs and changing social environments.

Morality and religion in the modern state Agrarian civilizations based on monarchies continued to expand, except for periods when they contracted, right up to the emergence of the first modern states—the U.S. in 1776 and France after the French Revolution in 1789. The final colossal agrarian empire was that of the Mongols, ruled by Genghis Khan and his successors. The Mongolian empire held together about 60 years, from 1206 to 1265, and proved to be the largest contiguous land empire in history, only surpassed later by the world-wide, non-contiguous, colonial British Empire ruled by a modern state with a symbolic monarch. Once human history reached the threshold of the industrial revolution, beginning in Europe in the late eighteenth century, the nature of government began to change. The monarchical structure of agrarian civilizations began to give way to the more democratic structure of nations, or modern states. The power of the king, the priests, and the nobility began to decline. Industrialists gained enough wealth to demand a voice in government, and increasing populations required more central coordination. Increased central control could succeed only with increased loyalty from the populace, now called citizens rather than subjects. Legitimization of power changed from divine approval to public approval, which governments won through increasing services, such as infrastructure and education. Paradoxically, the more nearly democratic state often wielded more power over the lives of its citizens than did the monarchy it replaced, but it wielded more of that power in their service. (See Christian, Brown, and Benjamin, 252–253.) Along with these other characteristics, modern states are characterized by a separation of church and state—a withdrawal or ousting of religion as source of legitimization. This can be seen clearly in the stories of the U.S. and France, where the new structures were created from scratch in the former and erupted quite abruptly from traditional structures in the latter. As far as religion is concerned, the development of the modern state meant the demotion of religion as an equal partner with the state and as the state’s chief legitimization. Instead, as citizen loyalty became the chief legitimization for states, citizens demanded freedom of religious beliefs and state toleration of various faiths. This required a new kind of morality, one that flowed from natural civil ideals that were shared by everyone rather than from supernatural religious beliefs not shared by everyone.

Natural, human-based morality As new structures of government arose, which transformed agrarian civilizations and rejected alliances with religion, new ideas of morality arose concurrently. A major source of those ideas in Europe was a small group of European philosophers known

268

Big history, morality and religion

as the Radical Enlightenment. Their ideas, though still unrecognized by many, have recently been acknowledged by some historians as playing a central role in shaping the core values and ideals of the modern world (Israel, vii). Two of the most influential Radical Enlightenment philosophers were Paul Henri Thiry, baron d’Holbach (1723–1789) and Denis Diderot (1713–1784), editor of the celebrated Encyclopedie. Holbach used his assets to host a legendary salon in Paris, where these ideas could be discussed. He also wrote the influential System of Nature, published secretly in 1770, the first uncompromisingly atheistic book since antiquity (Blom, x; Israel, 56). Diderot edited the celebrated Encyclopedie of 28 volumes (Paris, 1751–1772), in which his stated aim was to change how people thought. The frontispiece showed a female, dressed only in transparent veils, representing truth, while conventionally dressed men, who represent reason and philosophy, are pulling off her veils (Wikipedia). These radical thinkers differed from moderate Enlightenment thinkers, such as Francois Voltaire (1694–1778) or Immanuel Kant (1724–1804). The radical thinkers relinquished the idea of God completely; they were convinced that the world had evolved through chance without any guiding intelligence or higher being. They believed in abolishing all the legal ways in which monarchs, priests, and nobility gained privilege. Holbach believed that everyone had an equal right to pleasure and happiness, but happiness pursued in a virtuous way, namely by pursuing the happiness of others as well as one’s own. If people pursued a virtuous happiness, he believed, a harmonious society would result naturally without the need for religiously based moral guidelines. He may have influenced Jefferson to include the phrase, “the pursuit of happiness” in the Declaration of Independence (Blom, x–xv; Spier, 2). These ideas found partial expression in the U.S. Declaration of Independence and Constitution and more strongly in the French Declaration of the Rights of Man, confirmed under Napoleon. As industry developed in Europe, standards of living rose and cultural change, including literacy, occurred swiftly. A radical shift in consciousness seemed to occur. People showed a new capacity for empathy and an aversion to cruelties previously accepted—to slavery, torture, and severely harsh punishments. Citizens recognized in others their own feelings, and they asserted universal, equal, and natural human rights for everyone. For example, the Atlantic slave trade was legally dead by 1842, while Mauritania was the last country to abolish slavery in 1981. An estimated 20–30 million people remain enslaved in 2014, but in most countries today it is not recognized as legal. Many human advancements based on civic moral codes of freedom and equality followed in modern states—suffrage for all men, then for women; universal education; social programs; health insurance and old age pensions; and prohibitions on racial and sexual discrimination. The modern state had to earn the loyalty of its citizens at the same time that it defended the interests of its capitalist entrepreneurs. It emphasized the first three of the cross-cultural moral foundations outlined by the U.S. social psychologist, Jonathan Haidt (1963–): caring, fairness, liberty, loyalty, authority, and sanctity (The Righteous Mind: Why Good People Are Divided by Politics and Religion).

269

Cynthia Stokes Brown

The current situation When something new, such as the modern state, emerges, older ways of doing things persist in the midst of the new. There are now about 196 nations in the world, but some 50 are still fully authoritarian regimes, such as Chad, Saudi Arabia, Iran, and North Korea (www.en.wikipedia/wiki/DemocracyIndex, compiled by Economist Intelligence Unit). Even the so-called democracies have developed aristocracies of such wealth that they mock the assertion of equal participation. At the same time, some modern nations are finding it difficult to hold together. As globalization reduces national loyalties and as unsolved problems increase, states ­ zechoslovakia, are breaking up into older ethnic and religious groupings—consider C Yugoslavia, Serbia, Georgia, Sudan, Syria, Libya, and Argentina. Even the Scots are seeking separation from Britain. Some of these states were established without much regard for ethnic groupings in the first place. As national identities become less important, people seem to choose either to be world citizens or to revert to more regional and local identities. In the realm of global religion, what is the current situation? It is difficult to tell. At present, there seems to be a rise in religious fervor in some places and a decline in others.There may be a widening split between religious and secular groups in many places. The rise in religious fervor has occurred notably in post-Communist counties, in a rush back to traditional beliefs and practices that had been prohibited under Communism, a rush encouraged by the missionaries arriving from the proselytizing religions, especially Christians, Muslims, and Mormons. Religious fervor has also increased in post-colonial countries, also promoted by proselytizers. At the same time, religious belief in the more developed countries and in academia has declined, at least in the U.S. and Europe, where polls have taken place. Studies have shown that as many as 90% of the members of the U.S. National Academy of Science are non-theists [Natalie Anger, “My God-Problem—and Theirs,” The American Scholar 72 (Spring 2004) 131–134, cited by Haught, 25]. It seems clear that secularization has correlated with increased education and security and with the expanding explanatory power of natural science (Raskin, 102). A U.S. defender of theism, John Haught, is worth quoting, as he describes how he believes that the philosophy of naturalism (nature is enough—no supernatural exists) has taken over academia worldwide: Naturalism is now so entrenched in science and philosophical faculties around the globe that it constitutes one of the most influential “creeds” operative in the world today. Scientific naturalists are still a small minority in the world’s overall population, but their influence is out of all proportion to their numbers. Generally speaking, their beliefs quietly determine what is intellectually acceptable in many of our universities. Naturalism has now spread from science and philosophy departments into social studies and humanities. Even departments of religion are not immune. (25) No one really knows how many people worldwide regard themselves as religious or not. Many polls are not reliable; figures seem to vary according to the beliefs of the 270

Big history, morality and religion

reporters (see Wuthernow for an analysis of this). Grim and Tucker claim that 97.7% of the world’s people in 2012 were adherents to religion (28). Adam Gopnick, writing for the New Yorker, says that the polls most generous to non-believers find about 20–30% of people in the U.S. are not religious or not very religious and about 50% in Europe (Gopnik, 107). The Pew Forum on Religion and Public Life reports that 84% of the world’s people in 2010 identified with a religious group and 16% were unaffiliated. Of the unaffiliated, 700,000 are in China, more than twice the U.S. population (www.pewforum.org/global-religion-landscape.exec.aspx). Jared Diamond reminds us that people in poorer social strata, regions, and counties tend to be more religious; he believes they need more comforting, as the comforting function of religion increases in societies where more bad things happen to people. He believes that in nations with a per capita gross domestic product (GDP) under $10,000, 80–90% of people say that religion is an important part of their daily lives. In nations with more than $30,000 per capita GDP, only 17–43% say that. Perhaps in many countries, as in the U.S., there is a widening gap and polarization between the religious and non-religious sectors, especially as the gap in income between the wealthy and others increases (353–355). No world religion has managed to curb people’s desires for material consumption (Rue, 2000, 37). Human desires for consumption have led humanity to the limits of the planet’s resources. The moral codes of religion from the era of agrarian civilizations have focused people’s attention on how to behave toward other people rather than toward other species and the environment. Likewise, the newer moral civic codes from the era of modern states have focused primarily on human rights and have reflected the tremendous affluence that modern people have enjoyed from burning fossil fuel. Both kinds of moral codes seem in need of revision as the future approaches. Into this situation a new origin story has recently arisen. A narrative constructed from the findings of both science and the humanities, this story is being called by various names—“Universe Story,” “Epic of Evolution,” “Journey of the Universe,” “Big History.” This is the project that E. O. Wilson (1929–) imagined when he began to envision sociobiology and called for consilience, or the unity of knowledge; he coined the term “Epic of Evolution.” The new story has subtle variations in meaning depending on the version. The “Universe Story” and “Journey of the Universe” emphasize the creativity of the universe and the creative response of humans.The “Epic of Evolution” features evolution as the underlying theme, while “Big History” hypothesizes an underlying pattern of increasing complexity based on increasing energy flows. Yet all versions of the new story have a similar fundamental moral meaning—that human behavior is causing mass extinctions and climate change and that humans must create a massive transformation in their way of living in order to sustain planetary conditions conducive to human life.

The future No one can predict the future. The current pace of innovation and change is more rapid that at any time in human history; unforeseen contingencies can be expected. 271

Cynthia Stokes Brown

All versions of the new story indicate that we have moved into a critical period in human and Earth history, a new threshold as many big historians call it, a transformation as extensive as that from foraging to farming or from farming to civilization. This transition, already underway, will involve vast changes in human consciousness and in social arrangements, as humans create values coherent with emerging realities. We face a limited supply of the non-renewable natural resources on which our civilization is based. The human explosion from two to over seven billion people in the last hundred years is threatening the planet’s ecosystems and other species. Our current use of fossil fuels for energy is not sustainable indefinitely; we are using up a limited supply of fossil fuel, and we are changing our climate. Some transition to a new energy source must be made, while at the same time the human population must stabilize and reduce. This transition will occur; the question is how much of it will be under human control. Three basic possibilities can be sketched; here I follow the scenarios suggested by Paul Raskin, president of the Tellus Institute in Boston. (1) Conventional Worlds: Humans can continue to muddle through, using prevailing structures. One possible variant would feature free markets and deregulation as drivers of development. Another variant would feature coordinated government actions to reform modern capitalism. Neither of these variants seems likely to deal effectively with the global crisis. (2) Barbarization: Instability swamps free markets and/or government policies. Many scenarios are possible as the norms of civilization dissolve. Two possible ones are: elites retreat into enclaves, leaving the masses in poverty, while some kind of global authoritarian regime imposes order, or complete breakdown ensues, with institutional collapse and abrupt human dieback. Some humans survive to devise a much simpler way of life within planetary limits. (3) Great Transitions: Humans use the opportunities of this transitional period to create a planetary civilization, unified but containing many variations. It would feature democratic global governance, economics geared to the well-being of all, and environmental stewardship. Raskin gives this global civilization the name “Earthland” and imagines that it might come into existence over the next 50 years, propelled by a mass global citizens’ movement (26–32). Sustainability is reached quite suddenly on the long-term timescale, and humans learn to cooperate globally to deal with their challenges. What can we predict about how religion and morality might evolve in each of these three scenarios? In the scenario of Conventional Worlds, we can foresee increased conflicts and misery, contributing to a resurgence of traditional religions and continued polarization with non-religious groups. Many find scientific naturalism not appealing when material conditions are declining. In the scenario of breakdown, we can follow Loyal Rue, who has imagined what the moral and religious outcome might be after a massive loss of human population. He believes that after much death and devastation, there is reason to hope that the surviving humans would learn to live sustainably and would formulate a new 272

Big history, morality and religion

nature-centered meta-myth to provide an eco-centric rather than egocentric morality (2005, Chap. 11). In the scenario of Great Transitions, we can imagine a change in values from consumerism and individualism to fulfillment and solidarity with other humans and the biosphere. Reformed world religions might help evoke a commitment from humans to care for their planet and its life, as the Big History story might in secular settings. Of course, the reality of climate change and the rising costs of non-renewable natural resources will help create this commitment. The ethics and morality of how to subsume human desires to environmental limits are beginning to be imagined. Is the idea of Great Transitions a naïve hope of dreamers or is there reason to believe that the new story might help to provoke a new ethical commitment? We can point to some hopeful signs: •





There is a strong movement within world religious groups to connect/ re-­connect with the natural world.The term, “religious ecology,” is beginning to be used frequently, as more people realize that the engagement of religious communities is necessary if humanity is to sustain itself and the planet. John Grim and Mary Evelyn Tucker make this argument in Religion and Ecology; they state that the aim of religious ecology is “to retrieve, re-examine, and re-construct these human-Earth relations that are present in all the world’s religions” (42). Sometimes religious communities can unite for action more easily than can public entities paralyzed by opposing factions. Increasingly, scientists are willing to consider moral and religious issues, as they did before the break between religion and science in the seventeenth century. They are using the term “religious naturalists” to describe their position of complete devotion to the scientific method while also discussing moral issues and underscoring the awesomeness and mystery of creation. Strong voices of religious naturalists include biologist Ursula Goodenough, physicist Chet Raymo, and philosopher Loyal Rue. In global academia the “Big History” version of the new story is gaining traction. The International Big History Association (IBHA) is dedicated to keeping the story scientifically accurate and to promoting the teaching of it in schools worldwide from kindergarten through graduate programs (see www.ibhanet. org). Montessori teachers have been teaching Maria Montessori’s version of cosmic education in the elementary grades for over 60 years; they are doing so now with renewed commitment (www.deeptimejourney.org). Bill Gates funded the development of a high school curriculum, completed in August 2013, which is being taught in over 600 schools in over 15 countries (www.bighistory project.com). In 2014 McGraw-Hill published the first university level textbook; see Christian, Brown, and Benjamin. These efforts indicate the willingness of many people worldwide to build an international intellectual foundation for cooperation.

If  humanity is entering a period in which it must make a common struggle for its survival, what are the biological and cultural chances that we will cooperate rather than compete against those we define as outsiders to our group? Can we learn to 273

Cynthia Stokes Brown

define everyone as insiders? Can we cooperate on a large-scale for our common survival when enough people realize that the planet itself is a limited resource? Can global belief in the scientific origin myth hold us together? Both tendencies—­ competition and cooperation—seem deeply rooted in our evolution, and both seem likely to be operating in the future. Both will play a role in whatever the future holds in store for us. Humans have difficulties holding our desires in check. Perhaps that is why we devise moral codes and gods to enforce them—to articulate those behaviors we wish for but can’t easily attain. It is likely that we cannot reduce our population and move to renewable energy only by telling new stories. Sometimes natural reality itself must change our attitudes, and afterwards we will write new stories and new moral codes. We are enmeshed in a complex process of determinism, contingency, innovation, competition, and cooperation. We humans are deeply connected to all of reality. Becoming aware of our interconnectedness with the universe and our planet seems our first step toward devising actions appropriate to our time.

Bibliography Bekoff, M. and Pierce, J. (2009) Wild Justice: The Moral Lives of Animals. Chicago and ­London: University of Chicago Press. Bellah, R. H. (2011) Religion in Evolution: From the Paleolithic to the Axial Age. Cambridge and London: Belknap Press of Harvard University Press. Blom, P. (2010) A Wicked Company: The Forgotten Radicalism of the European Enlightenment. New York: Basic Books. Boehm, C. (2011) Hierarchy in the Forest: The Evolution of Egalitarian Behavior. Cambridge, MA: Harvard University Press. Bradshaw, G. (2009) Elephants on the Edge: What Animals Teach Us About Humanity. New Haven, CN: Yale University Press. Brown, C. S. (2007) Big History: From the Big Bang to the Present (2nd ed. 2012). New York: New Press. Chapois, B. (2008) Primeval Kinship: How Pair Bonding Gave Birth to Human Society. ­Cambridge, MA: Harvard University Press. Christian, D. (2004) Maps of Time: An Introduction to Big History (2nd ed., 2011). Berkeley, CA: University of California Press. Christian, D., Brown, C.S., and Benjamin, C. (2014) Big History: Between Nothing and ­Everything. New York: McGraw-Hill. Churchland, P. S. (2011) Braintrust: What Neuroscience Tells Us About Morality. Princeton and Oxford: Princeton University Press. Dawkins, R. (2006) The God Delusion. Boston: Houghton-Mifflin. Deakin, T. W. (1997) The Symbolic Species: The Co-Evolution of Language and the Brain. New York and London: W. W. Norton & Co. Dennett, D. (2006) Breaking the Spell: Religion as a Natural Phenomenon. New York: Penguin. Diamond, J. (2012) The World Until Yesterday: What Can We Learn From Traditional Societies? New York: Viking. Dworkin, R. (2013) Religion without God. Cambridge, MA and London: Harvard University Press. En.wikipedia.org/wiki/Democracy_Index; compiled by Economist Intelligence Unit. Goodall, J. (1999a) Forty Years at Gombe. New York: Stuart, Tabori & Chang. 274

Big history, morality and religion

Goodall, J. (1999b) Reason to Hope: A Spiritual Journey. New York: Warner Books. Goetz, S. and Taliferro, C. (2008) Naturalism. Grand Rapids, MI, and Cambridge: William B. Eerdmans Publishing. Goodenough, U. (1998) The Sacred Depths of Nature. Oxford and New York: Oxford ­University Press. Gopnik, A. (2014) “Bigger than Philosophy: When Did Faith Start to Fade?” New Yorker, Feb. 17 & 24: 107–111. Graber, D. (2011) Debt: The First 5000 Years. Brooklyn, NY: Melville House. Grassie, W. (ed.) (2010) Advanced Methodologies in the Scientific Study of Religion and Spirituality. Philadelphia, PA: Metanexus Institute. Grim, J. and Tucker, M. E. (2014) Religion and Ecology. Washington, DC: Island Press. Haidt, J. (2012) The Righteous Mind: Why Good People Are Divided by Politics and Religion. New York: Vintage Books. Hrdy, S. B. (2009) Mothers and Others: The Evolutionary Origins of Mutual Understanding. Cambridge, MA: Harvard University Press. Haught, J. F. (2006) Is Nature Enough? Meaning and Truth in the Age of Science. Cambridge: Cambridge University Press. Hick, J. (2006) The New Frontier of Religion and Science: Religious Experience, Neuroscience and the Transcendent. New York: Palgrave Macmillan. Re-issued 2010. Hookes, D. “Cooperation: The Key Principle in the Evolution of the Universe.” Unpublished Manuscript. Horner, J. (2001) Dinosaurs under the Big Sky. Missoula, MO: Mountain Press Publishing. Israel, J. (2010) A Revolution of the Mind: Radical Enlightenment and the Intellectual Origins of Modern Democracy. Princeton, NJ: Princeton University Press. Jaspers, K. (1953) The Origin and Goal of History. Trans. Michael Bullock. New Haven: Yale University Press. First German ed. 1949. Margulis, L. and Sagan, D. (1986) Microcosmos: Four Billion Years of Microbial Evolution. Berkeley: University of California Press. McBrearty, S. and Brooks, A. S. (2000) “The Revolution that Wasn’t: A New Interpretation of the Origin of  Modern Human Behavior,” Journal of Human Evolution 39: 453–563. McNeill, W. H. and McNeill, J. R. (2003) The Human Web: A Bird’s-eye View of World History. New York and London: W. W. Norton & Co. Mithen, S. (1996) The Prehistory of the Mind. London: Thames and Hudson. Mithen, S. (2005) The Singing Neanderthals: The Origins of Music, Language, Mind and Body. London: Weidenfeld & Nicolson. Morris, Ian. (2015). Foragers, Farmers, and Fossil Fuels: How Human Values Evolve. Princeton and Oxford: Princeton University Press. Nowak, M. A. and Highfield, R. (2011) Super Cooperators: Altruism, Evolution, and Why We Need Each Other to Succeed. New York: Free Press. Pettitt, P. (2005) “The Rise of Modern Humans,” in C. Scarre (ed.) The Human Past: World Prehistory & the Development of Human Societies, London: Thames and Hudson, 124–173. Raskin, Paul. (2016) Journey to Earthland: The Great Transition to Planetary Civilization. ­Boston: Tellus Institute. Raymo, C. (2008) When God Is Gone, Everything Is Holy: The Making of a Religious Naturalist. Notre Dame, IN: Sorin Books. Rue, L. (2000) Everybody’s Story: Wising Up to the Epic of Evolution. Albany, NY: State ­University of New York. Rue, L. (2005) Religion is Not about God: How Spiritual Traditions Nurture Our Biological Nature and What to Expect When They Fail. New Brunswick, NJ and London: Rutgers University Press. 275

Cynthia Stokes Brown

Rue, L. (2011) Nature Is Enough: Religious Naturalism and the Meaning of Life. Albany, NY: State University of New York Press. Spier, F. (1994) Religious Regimes in Peru: Religion and State Development in a Long-term Perspective and the Effects in the Andean Village of Zurite. Amsterdam: Amsterdam University Press. Spier, F. (2010) Big History and the Future of Humanity (2nd ed., 2015). Walden, MA: Wiley-Blackwell. Spier, F. (2013) “How Can We Understand the Emergence of Morality in Big History? A First Exploration.” Unpublished manuscript. Sun, Y. (2014) “Chinese Traditions and Big History” in L. Grinin, D. Baker, E. ­Quaedackers and A. Korotaev (eds.) A Teaching & Researching Big History: Experiencing a New Scholarly Field, Volgograd: Uchitel. Trigger, B. G. (2003) Understanding Early Civilizations: A Comparative Study. New York: Cambridge University Press. Waal, F. de. (1996) Good Natured: The Origins of Right and Wrong in Humans and Other Animals. Cambridge, MA and London: Harvard University Press. Waal, F. de. (2005) Our Inner Ape: A Leading Primatologist Explains Why We Are Who We Are. New York: Riverhead Books. Waal, F. de. (2006) Primates and Philosophers: How Morality Evolved. Princeton, NJ: Princeton University Press. Wade, N. (2009) The Faith Instinct: How Religion Evolved and Why It Endures. New York: Penguin Press. Wrangham, R. (2009) Catching Fire: How Cooking Made Us Human. New York: Basic Books. Wright, R. (1995) The Moral Animal: Why We Are the Way We Are: The New Science of ­Evolutionary Psychology. New York: Vintage Books. Wuthernow, R. (2015) Inventing American Religion: Polls, Surveys, and the Tenuous Quest for a Nation’s Faith. Oxford: Oxford University Press. www.bighistoryproject.com. www.ibhanet.org. www.deeptimejourney.org. www.pewforum.org/global-religion-landscape.exec.aspx (retrieved 4.10.14) www.pewforum.org/2012/12/18/global-religion-landscape-exec/ (retrieved 5. 20. 14) Many thanks to those who engaged with me in developing these ideas: Frances Berry, Lauren Mezey, Phil Novak, Esther Quaedackers, Scott Sinclair, Fred Spier, Harlan ­Stelmach, Davidson Loehr, and David Blanks.

276

PART III

Little big histories

12 A CASE FOR LITTLE BIG HISTORIES1 Esther Quaedackers

Little big histories are studies that connect a specific subject to all the major phases of big history. They can be written about almost anything, from seemingly simple parts of our everyday world such as a cup of tea or a brick, to more intricate concepts, such as scientific theories or historical developments. Little big histories link these subjects to processes such as the formation and development of our Solar System and the Earth, the origin of life and the emergence of human agriculture. At first sight, trying to connect such subjects and processes may seem rather quirky. But when you think about it, little big histories can help answer some of the most fundamental questions we can ask about our world. They can help do so for two reasons. First of all, little big histories tell us how specific and often relatively small subjects fit into the biggest picture of the world that we have, and as a result, can be used to prod and test that picture. Secondly, and at this point in time perhaps most importantly, little big histories can help us understand more fully why their subjects are the way they are. I would like to explain that last point first.

The nature of things Many phenomena that exist on Earth today or that we know have existed in the past are many things at once.2 They are many things at once because they are complex, because they are part of something complex or because they are both.Take, for example, a brick, perhaps one that is part of the building you currently find yourself in or near to.3 This brick may not seem very complex at first. But when you take a closer look, it quickly becomes clear that there are many levels of complexity present in the brick, with more complex levels stacked onto less complex ones. For instance, at its lowest level of complexity, a brick consists of a collection of interacting quarks and electrons. But of course a brick is more than that. Quarks and electrons are bound together by the strong force and by the electromagnetic force into protons, neutrons and eventually atoms, such as oxygen, silicon, aluminum and iron.4 So a brick is also a collection of these and a few other interacting atoms. In turn, these atoms form three-dimensional networks of oxygen-silicon tetrahedra, often interspersed by 279

Esther Quaedackers

metals and other atoms that we know as silicate minerals. This means a brick is also a collection in such interacting minerals. When forming minerals out of atoms and atoms out of quarks and electrons things become more complex, because at each level, the number and variety of building blocks that can be used and the number and variety of interactions between these building blocks that become possible becomes greater.5 So even though a brick may seem relatively simple, it consists of what is already a rather complex assembly of minerals. But a brick is more than that. A brick is also part of an even more complex world. It is the product of life that at some point in history started to build and create its own building materials, including bricks. It is part of a building tradition in which people use bricks to separate themselves and their belongings from the natural and social environment. And it is part of a culture in which certain meanings are attached to bricks and in which these meanings evolve over time.6 Even though many phenomena are many things at once, most people coming from specific academic and professional disciplines only look at one or a limited number of the things something is. When considering a brick, material scientists often just look at its mineral properties, architects often just focus on its use, and art historians often just study the development of its meaning. This is understandable, given the tremendous amount of knowledge about almost anything that is available today. The availability of that knowledge ensures that learning all there is to know about a subject on one level of complexity can take years. This in turn can make trying to learn about a subject at many levels of complexity at once seem daunting. After all, trying to do so also means risking missing or misunderstanding important details, simply due to a lack of time to study every level of complexity exhaustively. Not trying to do so, however, also comes with certain risks. Not looking at the many different things a subject is at once makes it much more difficult to see how all these different things can influence each other. Such influences are everywhere, including in and around bricks. For instance, as a result of the abundance of silicate minerals on the surface of the earth, it is easier to actually shape the world through building for animals such as ourselves that are good at turning them into building materials.This connection between the physical and chemical, biological and cultural aspects of a brick becomes particularly interesting once you realize that because silicate minerals on the surface of the earth often come in grains that are too small or boulders that are too big for most animals to handle, there are not that many animals that are able to efficiently use them as or develop them into building materials. Particularly amongst larger animals, the way humans are able to turn small mineral grains into building materials such as bricks is exceptional. Party as a result of that, we are such proficient builders, whereas most other animals are not. Adding ideas about such connections to the detailed knowledge we already have about aspects of a bricks such as its mineral composition, its use and the history of its meaning can help us understand it more fully. Developing such a fuller understanding is particularly important when it comes to subjects that, unlike bricks, are vital to our survival. After all, when studying systems like cells, cities, societies or climates, missing how the different levels of complexity such systems consist of or are part of influence each other makes it harder to predict, anticipate and perhaps even modify their behavior. This is what people who study 280

A case for little big histories

such systems have realized for years and why many of them have advocated the development of interdisciplinary ways to approach their subjects and that may help them to tease out these connections.7,8 There are several different ways in which people are trying to develop such interdisciplinary approaches. The simplest situation in which an interdisciplinary perspective can emerge is a situation in which knowledge from different disciplines just clicks after it has been brought together. Sometimes the connections simply appear. But often, they do not. In such situations, scientists and scholars often apply specific theories or use specific methods to study their subjects. For instance, people who work in systems theory or complexity science both, in sometimes overlapping ways, model how interactions between different parts of a system can lead to characteristics that are otherwise hard to understand. And there are other theories and methods that can be used.9 The little big history method is one of them. As mentioned, little big histories connect their subjects to all the major phases of big history. By doing so, they can connect the different levels of complexity their subjects consist of or are part of in their own unique way. They can use big history as a framework for studying how during certain phases the higher levels of complexity emerged from and started to interact with the lower levels of complexity and therefore how they are causally connected. In other words, they can be used for studying how interactions between the different levels of complexity their subjects consist of or are part of developed over time and as a result, lead to a fuller understanding of how these subjects came to be the way they are today. This distinguishes little big histories from most other interdisciplinary perspectives, which often focus more on how things work or worked at a specific moment in time instead of how things came to be over time. Little big histories are therefore complementary to such perspectives and are both informed by them and can inform them.

The history of little big histories Little big histories are relatively new to this interdisciplinary scene.This is not surprising, since they are based big history, which itself is a relatively new field. Even though it has existed in its current academic form since the 1980s, up to the beginning of the twenty-first century, only a handful of people worked in it.10 Since the beginning of the twenty-first century, coinciding with the rise of the internet that made it much easier to access all kinds of information from outside one’s own discipline, big history has become much more popular. This increase in popularity spurred the first works that I would now call little big histories. Most notably, such works include Jan Zalasiewicz’ 2010 book The Planet in a Pebble and Neil Shubin’s 2013 book The Universe Within.11 Both books link their respective and relatively small subjects, a pebble and the human body, to all major phases of big history. Both books do so in order to understand more fully why pebbles and human bodies are the way they are. The authors just do not call their works little big histories, nor do they explicitly look for new connections between their subject and the various periods in big history that are relevant for these subjects. Instead, it seems like these authors more intuitively sensed that writing about their subjects in a little big history-like way would stimulate their readers to appreciate their subject’s richness. Zalasiewicz expresses this by referring to the opening lines of 281

Esther Quaedackers

William Blake’s famous poem Auguries of Innocence, which I feel perfectly captures the spirit of little big histories: To see a world in a grain of sand, And heaven in a wild flower. Hold infinity in the palm of your hand, And eternity in an hour.12 This intuitive appeal of little big histories, which allowed people to see the big past in something relatively small, was also the reason why I personally started to experiment with them. It was the reason why in 2007 I started asking the students in my big history courses to choose a subject at the beginning of the course and to connect that to an aspect of every class. My colleague Fred Spier started to call the results little big histories.This term stuck as other universities and high schools that were teaching big history adopted this simple but effective assignment.13 While students kept discovering new aspects of their subjects they had never thought of before and while books like The Planet in a Pebble and The Universe within affirmed the richness of the approach, I started to reflect on the academic value of little big histories. What kind of knowledge did they reveal exactly? And was that knowledge new or not? I had started to realize that, as explained above, little big histories should be able to tease out connections between the different levels of complexity a subject consists of or is part of, and that few disciplinary or interdisciplinary perspectives were able to do that in the same way. But that was theory. How would this work in practice? I therefore started to experiment with little big histories, and unlike authors like Zalasiewicz and Shubin, explicitly started to look for novel ideas that could be generated by writing them. In order to do so, I embarked on studying the little big history of a building, Tiananmen, in Beijing. Around the same time, my colleague Jonathan Markley started working on a little big history of grass.14 And a bit later, colleagues Craig Benjamin and Olga García-Moreno and her team also started writing little big histories. Little big histories from both these authors have been included in this book.15 Their results and my own have confirmed that little big histories can add new ideas to our existing knowledge. Quite often they add such new ideas in the form of fundamental questions that somehow have not been asked before. They also provide glimpses of answer, even though definitive answers usually remain beyond the scope of a little big history. But in a time when we already seem to know so much about so much, perhaps new questions are more exciting than definitive answers. In any case, these questions and preliminary answers have convinced me it is time to make a case for little big histories.

An exemplary little big history of a brick I think the best way to do so is by not just talking about the potential of little big histories to help us understand more fully why something is the way it is, but by also demonstrating that potential by providing an outline of a little big history. On the next pages I will therefore describe what an outline of all little big history of a brick could look like. I chose to write about a brick, because it is a subject that 282

A case for little big histories

almost all people are familiar with, perhaps so familiar that it probably does not seem very complex, let alone wondrous, to most. It is therefore a subject that can help me demonstrate that little big histories can raise new questions, even about seemingly mundane subjects. In this little big history, I will connect my brick to three major phases in big history: the history of the non-living world, the history of life and the history of human culture. It is important to note that there is no consensus amongst big historians that these are the most important phases. In fact, there seem to be almost as many subdivisions of big history into particular epochs, thresholds or parts as there are big historians.16 These three phases are simply the ones that seem most useful to me, as each new phase ushers in the development of levels of complexity that are governed by new rules, such as those dictated by evolution or cultural learning, which have shaped little big history subjects in their own unique ways.17 I will connect the brick to these phases by formulating a question about how certain characteristics of the brick originated during each of these phases and subsequently developed over time and by partially answering that question. For instance, I will connect the brick to the history of the non-living world and beyond by exploring why certain types of atoms ended up in bricks. I will connect the brick to the history of life and beyond by exploring why certain animals build with brick-like materials. And I will connect my brick to the history of human culture by exploring why humans use bricks.These questions are not chosen entirely randomly. During this little big history of a brick, it will become clear they are derived from each other. But before I start exploring that history, there are some general remarks about little big histories I would like to make. The formulated questions already reveal two important properties of little big histories. Perhaps needless to say, while they allow us to understand why their subjects are the way they are more fully, little big histories of course do not allow us to fully understand their subjects. They can add to our understanding of subjects by exploring how aspects of different levels of complexity a subjects consists of or is part developed in conjunction, but they can never cover these levels of complexity in their entirety. For instance, exploring how a brick came to be the way it is on a physical and ­chemical level can be done by looking into why certain elements ended up in the brick, but also why looking into how particular bricks came to be reddish-brown or porous. And it can be done in many other ways. There are usually so many questions that can be asked and answers that can be given that it is impossible to take all of them into consideration in one chapter or even a whole book. This means that a little big history is always limited in its own way. It also means it is possible to write several complementary little big histories on the same subject. A second property of little big histories that is revealed by these questions is that they can be described as layered and as is depicted in Figure 12.1. They do not need always to be: there are often so many kinds of connections that can be found and kinds of questions that can be formulated within the context of a little big history that they can be structured in a variety of ways. But when systematically thinking about how the different levels of complexity that a subject consists of or is part of came to be, it can be useful to both go back to the emergence of such an aspect and to trace its development up until today. Doing so for aspects characterized by different levels of complexity means that after tracing the development of a less complex 283

Esther Quaedackers

How did the subject become the way it is on a cultural level? How did the subject become the way it is on a biological level? How did the subject become the way it is on a physical and chemical level? Big Bang

Now Time

Figure 12.1  A possible layered structure of a little big history.

aspect up until the present, it is necessary to jump back in time to the emergence of the more complex aspect and again, trace its development to the presence. For example, for the question why certain elements ended up in bricks, it is necessary to explore how these elements became available for building on earth in the first place and it is necessary to explore why life in general and humans in particular favored these types of atoms. Subsequently addressing the question why certain animals build with brick-like materials, and not just certain atoms, requires jumping back to the history of life.

Why did certain elements end up in bricks? I will start this little big history of a brick by exploring why certain elements ended up in the brick. This implies investigating why bricks are mainly made of oxygen, silicon, aluminum and iron, in that order of abundance, and not others. A first preliminary answer to that question seems simple. Perhaps we create bricks out of these elements simply because they are commonly available to us. After all, the composition of bricks very much reflects the composition of the crust of the planet we find ourselves on. Like bricks, that crust is made of oxygen, silicon, aluminum and iron, in that order of abundance.18 The crust is made of those elements first of all because stars produced them in relatively large quantities. For instance, oxygen is a main product of nuclear fusion processes that take place inside stars. Silicon and iron are forged in somewhat lesser quantities, but are still the 8th and 9th most abundant elements in our Solar System.19 Stars did produce more common elements, such as carbon, nitrogen and neon, but these did not end up in our planet’s earliest crust in large quantities.20 Instead, while oxygen, silicon, aluminum and a few other elements clumped together into a planet orbiting our fledgling Sun, these more common but also more volatile elements were blown away from the early Earth or ended up in the our planet’s oceans and atmosphere. It may be the case that we create bricks out of oxygen, silicon, aluminum and iron just because they are the most abundant elements that were left in the Earth crust after a series of cosmological and geological processes. We subsequently used that crust, simply as it is, as the raw material for our bricks. This preliminary answer has a few problems though. First of all, the fact that we make bricks from elements that are common does not have to mean that we use these 284

A case for little big histories

elements because they are common. But it is relatively easy to find arguments in favor of a causal relation between the use and commonness of elements. Most importantly, common elements and materials are normally cheap, as it does not require too much energy to find them. And when building materials are cheap, building can compete with other strategies that animals have developed to deal with certain problems. When building is sufficiently cheap, building can win such a competition and can be favored over, for example, growing protective structures or developing defensive behavior.21 Secondly, if we do make bricks out of oxygen, silicon, aluminum and iron because these elements are common this does not have to mean this is the only reason we use these elements. Some of them may also have, and indeed do seem to have, certain characteristics that make them very suitable for building in general and for making bricks in particular. This becomes clear when comparing the rocky material that bricks are made of other materials that are common on Earth, such as the gasses in our atmosphere and the liquids in our oceans. It is not impossible to build with such materials. There are a few frogs and fish that build bubble and foam nests out of air and liquid secretions.22 And it could be argued plants use the pressurized liquids inside their cell walls to give certain plant parts sufficient strength. But both bubble and foam nest and cells constantly need to be maintained, not in the least because liquids and gasses have a tendency to try to spread out and escape. Since most animals do not seem to be willing to invest a lot of energy into constant maintenance, they therefore often opt for building with solids. And unlike most other common elements, silicon, as it turns out, is particularly good at forming sturdy solids, at least at temperatures and pressures that are present on Earth. Silicon is particularly good at forming such solids because it is able to form four covalent bonds with other atoms. That means that unlike most other elements that can only form three or less covalent bonds or form different types of chemical bonds, silicon can serve as the basis for large 3D network molecules. In such molecules, each atom is kept firmly in place by the numerous bonds, which makes them both solid and sturdy. There is only one type of atom that can also form four covalent bonds and somewhat similar types of network molecules. That type of atom is carbon.23 It is therefore not a coincidence that carbon serves as the basis for another types of material that is also commonly used by life and general and human in particular to build with: the 3D matrix of cellulose, hemicellulose and lignin that gives wood its strength. This raises the question why we do not use wood to create bricks. Part of the answer seems to relate back to the commonness of carbon that is rarer on earth than silicon. In addition, and in contrast to silicate minerals, the carbon in wood needs to be actively extracted from the atmosphere and converted into cellulose, hemicellulose and lignin by life. But there are also practical considerations that have prevented humans and other animals to create wooden bricks.24 Wood needs to be cut into bricks whereas geological processes have already broken down many crustal rocks into smaller fragments. Under the influence of, for instance, water, temperature changes and biological activity, tiny cracks can emerge in silicon-based network molecules. Due to the nature of the network, these cracks redirect and concentrate stress in specific places in the network molecule. At such places, the bonds between 285

Esther Quaedackers

atoms may be unable to withstand the stress and can break down. As a result, rocks fracture and crumble.25 In contrast, life invented all kinds of clever tricks to prevent the same from happening in wood.26 Because rocks crumble, silicate mineral on the surface of earth often come in smaller pieces or even in tiny grains, such as the ones that are present in sand and clay.This can be advantageous, as it is much easier to pick up rocks or scoop up sand or clay than it is to chisel out rocks. It is also easier to carry around smaller pieces or portions of silicate minerals. And it can be relatively easy to shape them into any desired shape. That is, if you can somehow glue them together again. That is of course what humans do when they create bricks. But as it turns out gluing silicate minerals together again has proven to be rather difficult for most other animals. Before I get into that and venture in more detail into the history of life, I would like to point out that the story about characteristics of silicon adds another dimension to the origin story of the physical and chemical aspects of a brick. After all, together these stories demonstrate that we happen to live on a planet where some of the most common elements are suitable for building, and particularly suitable for making bricks. As mentioned, in order to make bricks, it is necessary to glue fragmented silicate minerals back together. In order to glue them back together, it is necessary to find or create glue. I suspect that there is no such glue present on earth. I think so because, to my knowledge, no animals have discovered such a naturally occurring glue yet.27 Given the creativity of evolution, that makes me suspect it does not exist. That means that animals have had to develop their own ways of gluing silicate mineral fragments together. I think there are a few major strategies that animals have used while doing so. Some animals, for instance, certain species of wasps, martins and swifts, collect or create mud, which of course consists of silicate mineral fragments and process this mud in very specific ways. They do so by making rapid movements when adding new mud to their structures. These movements liquefy both the new and older mud and allows them to vibrate together, exchange water and occupy the small air spaces in between them. When the water evaporates from the mud, it hardens into a solid material that contains few gaps in the structure that could become the starting point of fractures.28 Certain species of swallows and swifts also use an additional strategy to glue their mud together. They intersperse layers of mud with fibrous materials such as grass or horse hair.29 This gives their structures additional strength, much like the adding steel cables to concrete gives reinforced concrete its strength. Other animals, always relatively small ones, glue mineral grains together with the aid of their own secretions or excretions. Such secretions or excretions include the silk that, for instance, caddis larvae and polychaete worms use to construct cases out of rock fragments, and the saliva and feces that ants and termites use to construct their nests and mounds.30 This is never done by larger animals. Perhaps this is the case because building larger structures requires relatively large amounts of supporting building materials, much like large animals such as elephants need bones that are much more massive than those of say a mouse. That means building larger structures made of silicate mineral fragments require relatively large amounts of glue-like secretions or excretions. This may be too expensive for larger animals to produce.31 The costs of producing enough 286

A case for little big histories

glue may also be confounded by the fact that the metabolic rate of larger birds and mammals is generally lower than that of smaller animals, which may make it even more difficult to produce enough.32 Anyway, it seems that as for larger animals it is more difficult to glue fragmented silicate minerals together with the aid of secretions or excretions, there are fewer opportunities for them to build with such minerals. Larger animals indeed hardly ever build with such fragments, despite their commonness on earth.33 Humans are an exception. Exploring the question why, unlike most other large animals, humans do build by gluing fragmented silicate minerals such as clays and sand back together requires venturing into the history of human culture. In that culture, all kinds of knowledge can accumulate more rapidly than it can in the brains or even cultures of other animals.34 And indeed it is possible to see that happening when it comes to the use of sand and clay for building. Once people discovered a way to utilize these materials, they often preserved it and improved upon it. As a result, humans were not just able to rapidly invent and refine building strategies that helped them glue fragmented minerals back together, they were also able to combine them into a whole repertoire of ways to efficiently build with them. It is striking that this repertoire includes all the tricks that had been invented by other animals to build with rock fragments before. For example, people learned long ago that it is possible to create sturdy blocks or bricks that could be used for building by mixing earth with water, loading the damp earth into boxes and tamping it.35 This is actually quite similar to the way certain wasps, swifts and swallows build with created or collected mud, as both processes ensure that when the water evaporates from the damp earth or mud, mineral grains are glue together into masses that contain few gaps or cracks that can lead to fractures. Like certain swallows and swifts, people also learned long ago that blocks or bricks made of damp earth could be reinforced by adding fibrous materials such as barley straw or other organic materials.36 Like many smaller animals, people learned to produce their own glue. They managed to circumvent the barriers that prevent other larger animals from produce such glues by producing them externally, instead of internally. For example, like ants and termites, humans have made extensive use of excrements as a glue. They just haven’t used their own excrements, but that of their domesticated animals, just as cow dung.37 More importantly, they learned to glue silicate minerals together with the aid of fire. Fire itself can function as a kind of glue, as sufficiently high temperatures irreversibly change the structure within a block of mineral grains in ways that glue these grains together.38 This is what happens when people fire bricks.The use of fire also enabled humans to create true glues, for example, when they learned to heat limestone to turn it into slaked lime. Slaked lime has been used for many centuries as a basis for mortars that were used to glue dried or fired bricks into walls and roofs.39 In sum, certain types of atoms, most notably silicon atoms, ended up in bricks because, due to cosmological and geological processes, they are abundant on the surface of the earth, and because they can form sturdy solid building materials, as long as animals find a way to glue fragmented silicon mineral back together. Unlike other large animals, humans have found such ways, which enabled humans to use a common and suitable building material to transform the world around them. Of course, this raises many new questions that cannot be answered within the context of this little 287

Esther Quaedackers

big history, for instance, about the possibility to build with gasses or liquids, about the exact circumstances in which is it worthwhile to glue mineral fragments together into building, and about the striking similarity between the strategies that human developed to glue mineral grains together and the strategies other animals came up with. Little big histories often seem to raise more questions about a single subject than they can answer, and as a result could easily be used as starting points to generate research agendas for interdisciplinary research groups. For now, I will limit myself to exploring one of these questions in greater detail; the question why for certain animals it is worthwhile to build with brick-like materials, by which I mean fragmented silicate minerals that have been glued together into certain shapes, whereas for others it is not. This requires jumping to the next layer of this little big history, and back in time from the history of human culture to the history of life.

Why do certain animals build with brick-like materials? The fact that silicon minerals are common and can form sturdy building materials if the material fragments can be glued back together raises the questions of why all animals have not developed ways to do the latter. After all, this would have enabled them to build relatively cheaply, with materials that can be found everywhere and that do not require a lot of maintenance. I already provided a small part of the answer, by arguing that for most large animals, it may be too costly to produce glue. But also for many small animals, building with silicate minerals fragments is apparently still too costly. These animals do not build with them, and instead developed other building strategies or completely different ways to deal with the problems that building can solve. So when is building with brick-like materials worthwhile? I would like to argue that there are two situations in which this is the case. While some aspects of building with silicate mineral grains are cheap, not all of them are. Most notably, assembling a building out of such grains can be quite costly. One of the reasons assembling such a building can be so costly is that fitting the mineral grains together may require a large brain. And brain are energy guzzling organs: our brain consume up 20% of our energy budget while they only make up 2% of our body weight.40 In response, some animals have developed standardized ways to assembly mineral grains into buildings. These animals build by following a set of simple rules and repeating that over and over again. Animals do not really need to think about such rules, and they can even become hard-wired. As a result, standardized building practices do not require advanced cognitive skills and relatively large brains.41 Here is an example. There are certain kinds of caddis larvae that construct cases by weaving mineral fragments together with the aid of silk. These larvae have standardized their building practices by using a three-tiered selection process. First, they search for sand grains and pick up anything they can handle. Secondly, they try to turn the grains around in their legs and reject them if that does not work as it should. Thirdly, they try to fit them into their partially build cases and reject them if that does not work as it should. At every step, they automatically select for grain types that are of a more or less standard size, which can always be put together in a similar way.42 Assembling the case therefore doesn’t require much creativity. Moreover, both the production of the silk that binds the grains together and the selection process 288

A case for little big histories

have become subject to natural selection, which has led to the evolution of an efficient binder and building practice.43 Again there are parallels with human building, and particularly with the human use of bricks. After all, it is much cheaper to build with pieces of rock that have the same size and can always be fitted together in the same way than it is to build with irregular pieces of rock. The latter requires more creativity, and therefore more time and more specialized, experienced craftsmen. This was the case in, for instance, Inca society, where the most prestigious buildings were made out of huge natural boulders that were pieced together by highly specialized and appreciated craftsmen.44 In contrast, building simple brick walls could and can be done by most people. Moreover, cultural evolution has been able to exert its pressure on such practices, for instance, by making sure that the most efficient brick materials, sizes and other characteristics eventually became the most widespread. Perhaps there is a whole evolutionary history of bricks waiting to be uncovered, undoubtedly with its own internal contradictions, and with similarities but also many differences compared to the evolution of standardized building practices in other animals. A second situation in which building with brick-like materials is worthwhile is a situation in which the solid, sturdy and durable nature of such materials is particularly important. For animals that can use their building for prolonged periods of time, it may be wiser to invest a larger amount of energy into their structures once than to invest smaller amounts of energy into, for instance, building made out of less durable plant parts that need to be renewed all the time or into constant fleeing. Yet using a built structure for a prolonged period of time is something that few animals are able to do. It has been suggested that at least in birds, this is the result of the rapid buildup of ectoparasites in nests that are reused.45 In addition, it has been suggested that ­chimpanzees build a new nest each night to prevent the buildup of not just arthropod parasites but also of harmful bacteria.46 And it could be argued that using a building for a longer period of time can also make its inhabitants more vulnerable to predators; that once they have found the building, they can just wait for the inhabitants to appear. I therefore do not think it is a coincidence that many animals that do use their buildings for longer periods of time and, related to that, build them out of brick-like materials, evolved in relatively safe places or from ancestors that already possessed advanced defensive capacities. Examples of such animals include beavers that during the winter stay in the middle of frozen ponds and certainly the wasps, ants and bees that all evolved from primitive wasps that had just developed the capacity to sting.47 The accumulation of parasites, bacteria and predators is not the only disadvantage that comes with using a building for a longer period of time. Buildings, particularly those made out of brick-like materials, are normally fixed in place because of their weight. This means that for animals that do not live in an environment characterized by a dense concentration of food and potential mates,“commuting” from and to their building is necessary. I therefore again do not think it is a coincidence that another group of animals that use buildings made out of brick-like materials for longer periods of time descended from animals that happened to live in very rich environments where animals did not have to travel far in order to obtain food, and which can sustain large groups of animals. A notable example is mound-building termites that are descended from ancestors that lived in logs they ate.48 289

Esther Quaedackers

Are there once again parallels with human history here? Did early humans perhaps also evolve new defensive strategies, like the ancestor of certain wasps, ants and bees did, when they moved into the African savanna? Could tools and new types of coordinated behavior be seen as such strategies, as has been argued by some?49 Did they also start to live in environments that were able to support denser populations, like the ancestor of termites did? Did such environments perhaps not become richer themselves but also richer to humans as they learned to tap into new energy sources with the aid of tools and fire, and new ways in which they roamed and looked at the landscape?50 Scientists and scholars cannot confirm such patterns yet. They can confirm that when humans finally did become sedentary and started to use their built structures for prolonged periods of time, in many parts of the world they also started to build these structures out of tamped earth or tamped earth bricks. There is another similarity between the development of the building practices of humans, wasps, ants, bees and termites.While these animals, including humans, started to use their buildings for prolonged periods of time and while some of them started to construct them out of brick-like materials, they started to live in very large groups. Certain wasps, all the ants, certain bees and all the termites even became eusocial, meaning they started to live in colonies in which “colonial members belong to two or more overlapping generations, care cooperatively for the young and are divided into reproductive and non-reproductive (or at least less-reproductive) castes.”51 Even though how this happened is still a topic of debate, scientists agree that living together in a permanent, sturdy, defensible nest has been an important prerequisite for the evolution of eusociality.52 The capacity to build with brick-like materials has been a contributing factor in this process. Has or is building in such a way also contributing to the evolution of eusociality in humans? I think perhaps to some extent, as it also allowed people to live together in safe compounds where, for instance, grandmothers or eunuch could help take care of the young. But human history is more complex than the history of eusocial insect societies. Perhaps as a result of cultural evolution outpacing biological evolution, humans living in large groups have retained a sense of individuality that they sometimes want to express.This had important implications for why and how humans use bricks. I will explore this question in greater detail in the last part of this exemplary little big history. But first I would like to point out that again this little big history raises many questions that cannot be answered right now. For instance, how similar or dissimilar is the evolution of bricks to the evolution of standardized building practices that have been developed by other animals? How important has building, and in particular building sturdy structures from brick-like materials, been for the development of human social structures? Answers to these questions could provide us with a fuller understanding of what a brick is, but would need to be picked up by groups of researchers who could explore them more extensively.

Why do humans use bricks? The parts of the history of the inanimate, biological and cultural world that have been surveyed so far tell us that we build with bricks because doing so can be relatively cheap. It also tells us that perhaps partly as a result of building with brick-like 290

A case for little big histories

materials, people have come to live in groups of individuals. So how do you, as an individual, stand out in such groups? Building can be a good way of doing so. After all, buildings are generally honest signals of a certain strength, wealth, power or status, as individuals who do not have any of those usually cannot construct buildings or have them built.53 On top of that, they are usually very visible to other members of the community the builders belong to. This is not something that just humans have discovered. Animals such as bower birds and cichlids also build structures that are meant to stand out and, in their cases, advertise their biological fitness to potential mates.54 Building with bricks, however, complicates this strategy, exactly because doing so can be relatively cheap. What has this meant for the reasons humans have had to use bricks? Of course there are a number of simple ways in which people have can build with bricks and that still allowed them to show off. For instance, people have used bricks for load bearing bulk of buildings, which they subsequently covered with more special materials. This has been done everywhere, from houses made of brick walls that have been coated in stucco, paint or wallpaper, to the Taj Mahal, the bricks of which have been completely covered in stunning white marble.55 People have also created intricate masonry structures from bricks that were not very expensive themselves, but clearly expensive to assemble into all kinds of exceptional patterns because doing so required a lot of specialized expertise. While doing so, they have started to use the building blocks that were once created to build efficiently to do something much less efficient in order to stand out. This is a beautiful example of how layers of meaning can accumulate in human culture and sometimes start to contradict each other. On top of such layers, perhaps an even more subtle layer of meaning was added when people started to use bricks in order to stand out by not standing out too much. Chinese architectural history contains many examples of elites that in many ways built much like the common people, constrained by many rules and customs that prevented them from distinguishing themselves too much. Even the halls of the imperial palaces were not unlike the halls of the common people: they were built according to the same plans, using the same technologies and with similar materials. They were just bigger and more richly decorated. For instance, the three main halls of the Ming palace that were built between 1407 and 1420 CE were paved with unglazed bricks.56 But whereas these bricks may not have seemed very exceptional at first, they were made of a very special clay that had been processed for months and as a result, was turned into a very dense, strong and durable stone.57 Why would the emperor invest so much effort into creating such bricks that at first sight seemed just a nicer, more polished version of what common people used?58 Part of the answer may come from the other side of Eurasia, where during roughly the same period relatively independent warriors were developing into courtiers that resided at the court of Versailles. At this court, all the members became increasingly dependent on each other. As a result, in order to survive and even flourish in this court, they constantly needed to consider what other members of the court were doing, thinking and feeling before acting themselves. It has been described in detail how, as a result, members of the court started to constrain and refine their own behavior, in order to not accidentally antagonize one of the many others on which they 291

Esther Quaedackers

depended. For instance, physical violence was replaced by intrigue, and other shows of status also become more muted and less visible.59 A similar process may have shaped the Chinese court, and in fact, Chinese society at large. I cannot go into the details here, but there are several indications that, due to the nature of the Chinese agricultural sector, large groups of people became highly interdependent early in Chinese history.60 Perhaps partly as a result, Chinese people in general and the Ming emperor in particular constrained themselves while building, and often decided to use seemingly common bricks, as a way to not stand out too much. Of course, to well informed groups of elites it must have been clear how special the particular bricks that were used by the emperor were. To them, the use of the bricks must have seemed proper and a sign of respect for the cultural traditions that were shared by all people, while at the same it conveyed to this more limited and powerful group that the use of bricks was not a sign of a lack of wealth and power. Similar processes influenced the way many people think about bricks today. As a result of ongoing specialization, people in modern societies have become even more interdependent than people were further back in history, and therefore have to take into consideration how many groups of people might perceive their actions. This has stimulated them to act and build in more neutral ways.61 This of course does not mean that people do not want to stand out anymore: in many situations and for many reasons, people still feel they need to distinguish themselves from others. And they still do so, but in subtle ways. Bricks can be useful to give subtle meaning to buildings, because their commonness makes them an inoffensive choice but also because they can be glued together in a wide variety of ways. It is possible to create bricks with all kind of different structures, densities, colors and finishes and assemble them into a large diversity of shapes. This is something that can be seen in, for instance, the rows of canal houses that have been built in Amsterdam, mostly since the seventeenth century and often by wealthy merchants. Many of them are brick buildings with rather specific width and heights, but all these bricks buildings have been built out of slightly different colored bricks, have been subtly decorated with distinctive masonry patterns and have different kinds of roof lines.They express both a kind of down to earth attitude but also small differences that existed between their builders. Like the previous parts, this part of my little big history can only touch upon a few processes that made a brick into what it is today, and while doing so, raise more questions than it answers, for instance, about the development of neutral building practices and the exact role bricks played in that in various parts of the world. But again, these questions could be picked up by groups of researchers. And if answered, they would provide us with a fuller understanding of how bricks came to be the way they are. Best of all, answers to these question would be related to answers to the questions raised in the previous parts of this little big history. After all, questions about the neutrality of building with bricks are closely connected the ways in which animals in general and human in particular learned to use brick-like materials and to the origin and characteristics of the elements these materials are made of. The fuller understanding provided by this little big history would therefore not just mean a richer understanding, but also a more integrated one, as I argued it would be in the beginning of this chapter. 292

A case for little big histories

The big picture At the beginning of this chapter, I also pointed out something else. I pointed out that in addition to providing us with a fuller picture of their subjects, little big histories could also help us to improve our understanding of the biggest picture of the world that we have. Such big pictures are constructed with the aid of models, into which each little big history subject should fit. If they fit, then eventually this could lead to a model of big history that is grounded in lots of detailed and interconnected evidence and would allow people to zoom in on and out of subjects as they pleased. If they do not or not easily fit in, then perhaps the models should be nuanced or revised, or it should be concluded that the model simply cannot be applied everywhere. Of course, all of this works best if there are many little big histories available that can be used to test and prod the big picture. That is not the case right now. Still it is possible to at least try see how this little big history of a brick aligns with our current big history models. The model that is most widely used today revolves around the idea that throughout big history, complexity could emerge when certain amounts of energy flowed through matter within certain Goldilocks conditions that were about right.62 This can also be recognized in this little big history of a brick. After all, energy flowing through stars and the Earth’s mantle and crust have allowed for the formation of silicon and for the development of silicon-based minerals. Animals invest energy while gluing these minerals together into brick-like materials. And ­humans started to utilize the energy released by burning fuels, and the energy produced by domesticated animals and other people to produce bricks and assemble them into building. All of these developments were made possible by certain Goldilocks conditions, such as an environment in which solid rocks crumbled into small pieces which subsequently could be glued together again, conditions that allowed animals use their buildings for longer periods of time and therefore made producing brick-like materials worthwhile, and societies in which people feel the need to distinguish themselves from others, but should not distinguish themselves too much. Energy flows through systems are not featureless entities though. They have more characteristics than just the amount of energy that flows through matter and are directed in certain ways by life and human culture. Life can choose to direct its energy flows into building with brick-like materials or it can choose not to. And complex life and particularly human culture can direct their energy in a variety of ways that can sometimes seem to contradict each other. For instance, human cultures invest energy into creating and building with bricks and into showing off while doing so, but they also invest in the development of rules and other constraints that tell people not to show off too much. Exploring how and why life and human culture direct their energy flows, and how that contributes to the stability and further development of a system, could perhaps help to further refine the model of big history that is currently most widely used. This last conclusion is interesting for academics working in the field of big history, but perhaps less relevant for a broader audience. What is important for this larger group is the demonstration that little big histories can help test and prod the big picture while adding levels of details.The idea that grand narratives are unable to deal with many of such details, particularly the ones that do not directly conform to the 293

Esther Quaedackers

narrative, has been one of the main points of criticism of such metanarratives since the 1970s.63 Little big histories can help to refute such criticism and therefore also help to make a case for big history itself.

Notes 1 Readers familiar with the history of big history publications may recognize that this title is a reference to Christian, “The Case for Big History.” Christian’s article was one of the first to introduce the concept of big history to the world and to argue why it was worthwhile studying big history. This chapter aims to do the same for little big histories, which is why the reference seemed appropriate. 2 This thought was very much inspired by a conversation with Armando Menéndez Viso. I am very grateful that he put this idea in my head, because it allowed me to explain logically what I intuitively felt little big histories added to our understanding of subjects. 3 I will use this example throughout this chapter, both because a brick represents one of the simplest units of building, the subject I am specialized in, and because I was inspired to do so by a passage from Pirsig, Zen and the Art of Motorcycle Maintenance: An Inquiry into Values. On page 194, Pirsig describes how a student suddenly started to see all kinds interesting aspects of a brick that she had never considered before when she was forced to write a piece about just one brick. This chapter is somewhat similar in spirit. 4 The metal content of a brick of course varies a bit and depends on where exactly the soils where the brick was made of came from. The rather basic composition of bricks mentioned here was characteristic of bricks that were used throughout much of Chinese history. See Needham, Science and Civilisation in China Volume 5: Chemistry and Chemical Technology Part 12: Ceramic Technology, 49; Chen et al., “Variations in Chemical Compositions of the Eolian Dust in Chinese Loess Plateau over the Past 2.5 Ma and Chemical Weathering in the Asian Inland.” 5 This definition of complexity is based on Spier, Big History and the Future of Humanity, 48. 6 This last argument ensures that the idea that many things are many things at once also applies to things that we know have existed in the past, because our knowledge of them is also part of our complex and ever-changing culture. 7 In the Netherlands, a growing greater demand for interdisciplinary perspectives is apparent from the development of numerous study programs at honors colleges and interdisciplinary institutes that teach students to look at important current and future challenges from a wide variety of disciplinary perspectives. There is also some tentative evidence for such a growing demand from elsewhere. See Frodeman, Klein and ­Pacheco, The Oxford Handbook of Interdisciplinarity, 11–13. 8 Frodeman, Klein and Pacheco, 431.For the sake of simplicity, I am not making a distinction between interdisciplinarity and transdisciplinary here, as is sometimes done. 9 For an overview, start at Frodeman, Klein and Pacheco, 431. 10 Other forms of big history may be much older. See amongst others Spier, Big History and the Future of Humanity, 18–28. For an overview. 11 Zalasiewicz, The Planet in a Pebble: A Journey into Earth’s Deep History; Shubin, The Universe Within: A Scientific Adventure. 12 Blake, The Pickering Manuscript. 13 I have written about this history of this assignment more extensively in Simon, ­Behmand and Burke, Teaching Big History, 15%; Quaedackers, “A Short History of Little Big Histories.” 14 Markley, “A Child Said, ‘What Is the Grass?’: Reflections on the Big History of the Poaceae.” 294

A case for little big histories

15 García-Moreno and her team also published another little big history. See G ­ arcía-Moreno et al., “A Little Big History of Iberian Gold.” 16 See, for instance, Chaisson, Epic of Evolution: Seven Ages of the Cosmos; Benjamin, Brown and Christian, Big History: Between Nothing and Everything; Alvarez, A Most Improbable Journey: A Big History of Our Planet and Ourselves. 17 See for a further elaboration Spier, Big History and the Future of Humanity, 48–52. 18 Haynes, CRC Handbook of Chemistry and Physics, 97th Edition, 14–18. 19 McSween and Huss, Cosmochemistry, 92. The attentive reader may notice that I am not mentioning aluminum here. Aluminum is made in stars in much smaller quantities than silicon and iron, but is more common in the Earth’s crust as it floated upward during the formation of our planet. 20 The only exception is magnesium, which is more common than silicon and iron in our Solar System, but not in the Earth’s crust, as most of it sank down towards the core of our planet during its formation. 21 This argument is complicated by the fact that some animals, including humans, sometimes also use exceptionally rare elements to build or to decorate their buildings with. I will get back to that later. 22 See, for instance, Bastos, Haddad and Pombal, “Foam Nest in Scinax Rizibilis (­A mphibia: Anura: Hylidae)”; Jaroensutasinee and Jaroensutansinee, “Bubble Nest Habitat Characteristics of Wild Siamese Fighting Fish.” 23 There are a few other elements that commonly form covalent bond with four other atoms as well, perhaps most notably phosphorus and sulfur when they combine with oxygen to form phosphates and sulfates. Yet tetrahedra consisting of phosphorus and oxygen and of sulfur and oxygen do easily not link up with the aid of covalent bonds. They therefore cannot form the same types of large network molecules as silicon and carbon can. 24 At least during most of history. Today building systems based on wooden bricks, such as Brikawood, do exist. 25 See for a more elaborate description Gordon, The New Science of Strong Materials: Or Why You Don’t Fall through the Floor, 25–26%. 26 See Gordon, 35–41%. 27 It could be argued that there is one exception: the frost that is used by beaver to turn the mud that covers their dams and lodges into an extremely hard material. See MüllerSchwarze and Sun, The Beaver: Natural History of a Wetlands Engineer, 57. 28 Eberhard, “The Natural History and Behaviour of the Wasp Trigonopsis Cameronii Kohl (Sphecidae)”; Hansell, Built by Animals: The Natural History of Animal Architecture, 74–75. 29 Hansell, Bird Nests and Construction Behaviour, 67. 30 Hansell, Built by Animals: The Natural History of Animal Architecture, 41, 76; Curry, Grassland Invertebrates: Ecology, Influence on Soil Fertility and Effects on Plant Growth, 302; ­Garnier-­Sillam and Harry, “Distribution of Humic Compounds in Mounds of  Some Soil-Feeding Termite Species of Tropical Rainforests: Its Influence on Soil Structure Stability.” 31 It has been argued that this may be the case for secreted building materials that can bear load in tension and also for secreted nests. See Hansell, Ruxton and Ennos, “Collected and Self-Secreted Building Materials and Their Contributions to Compression and Tension Structures”; Hansell, Built by Animals: The Natural History of Animal Architecture, 179. The same argument could also be applied to producing glue. 32 See for a good overview of such scale related issues Schmidt-Nielsen, Scaling: Why Is Animal Size So Important? 33 I am not talking about burrowing here, which many larger animals do. Burrowing, however, does not require glue. And there are reasons to believe that burrowing is 295

Esther Quaedackers

34 35 36

37 38 3 9 40 41 4 2 43

44 45 46 47

48 4 9 50

51 52

53 54

55 56

actually more efficient for larger animals. For such animals, this may therefore be a good alternative to building with silicate mineral fragments. This is a point that has been frequently made by David Christian, first in Christian, Maps of Time: An Introduction to Big History. See, for instance, Needham, Science and Civilisation in China: Volume 4: Physics and Physical Technology, Part 3, Civil Engineering and Nautics, 38–39. See, for instance, The Getty Conservation Institute, 6th International Conference on the Conservation of Earthen Architecture: Adobe 90 Preprints, 67; Yang et al., “Traditional ­Mortar Represented by Sticky Rice Lime Mortar—One of the Great Inventions in Ancient China.” The Getty Conservation Institute, 6th International Conference on the Conservation of Earthen Architecture: Adobe 90 Preprints, 124. Needham, Science and Civilisation in China Volume 5: Chemistry and Chemical Technology Part 12: Ceramic Technology, 57–58. Allen and Iano, Fundamentals of Building Construction: Materials and Methods, 302. Raichle and Gusnard, “Appraising the Brain’s Energy Budget.” This point is beautifully made in Hansell, Built by Animals: The Natural History of Animal Architecture, 58–91. Hansell, 72–73. This selection process is actually best exemplified by the evolution of spider silk. See, for instance, Blackledge et al., “Reconstructing Web Evolution and Spider Diversification in the Molecular Era”; Opell, “Economics of Spider Orb-Webs: The Benefits of Producing Adhesive Capture Thread and of Recycling Silk.” Dean, A Culture of Stone: Inka Perspectives on Rock, 77–78. Hansell, Bird Nests and Construction Behaviour, 142–43. Thoemmes et al., “Ecology of Sleeping: The Microbial and Arthropod Associates of Chimpanzee Beds.” See for details about the life and evolution of these animals Müller-Schwarze and Sun, The Beaver: Natural History of a Wetlands Engineer, 57; Kirkendall, Biedermann and Jordal, “Evolution and Diversity of Bark and Ambrosia Beetles”; Branstetter et al., “Phylogenomic Insights into the Evolution of Stinging Wasps and the Origins of Ants and Bees.” Korb and Heinze, Ecology of Social Evolution, 172. Kortlandt, “How Might Early Hominids Have Defended Themselves Against Large Predators and Food Competitors?” Demographic trends during the Paleolithic remain contested. There is clearer evidence for a more intensive use of the land. See, for instance, Potts, “Variables Versus Models of Early Pleistocene Hominid Land Use”; Zeder, “The Broad Spectrum Revolution at 40: Resource Diversity, Intensification, and an Alternative to Optimal Foraging Explanations.” Wilson and Hölldobler, “Eusociality: Origin and Consequences.” Nowak, Tarnita and Wilson, “The Evolution of Eusociality.” This is a rather contested paper: also see the responses to it. Yet none of the responders disagree on the importance of a permanent, sturdy, defensible nest. The importance of such honest signals has amongst others been stressed in Zahavi, “Mate Selection—A Selection for a Handicap”; Veblen, The Theory of the Leisure Class. Diamond, “Experimental Study of Bower Decoration by the Bowerbird Amblyornis Inornatus, Using Colored Poker Chips”; York et al., “Evolution of Bower Building in Lake Malawi Cichlid Fish: Phylogeny, Morphology, and Behavior.” See Tillotson, Taj Mahal, 108. They were called bricks, but they actually look more like tiles. 296

A case for little big histories

57 Needham, Science and Civilisation in China Volume 5: Chemistry and Chemical Technology Part 12: Ceramic Technology, 516–17. 58 For the various floor that were used in dwellings throughout China see Knapp, China’s Old Dwellings. 59 Elias, The Civilizing Process: Sociogenetic and Psychogenetic Investigations, 365–414. 60 This becomes clear from the way farmers and other groups in Chinese society were and felt dependent on each other. For instance, rice growers generally are and feel more interdependent than grain growers. See Talhelm et al., “Large-Scale Psychological Differences Within China Explained by Rice Versus Wheat Agriculture.” Chinese farmers lived closer together, worked closer together and lived closer to cities and interacted more with city dwellers than farmers in, for instance, the west. See Needham and Bray, Science and Civilisation in China Volume 6: Biology and Biological Technology Part 2: Agriculture, 8, 64–68. And in Chinese culture, farmers were seen as a more important part of society than elsewhere. See Zhang, An Introduction to Chinese History and Culture, 209–14. 61 I have described this process more extensively in Quaedackers, “Architectuur, Macht En Moderniteit: De Oprichting van Het Hoofdgebouw van de Technische Hogeschool Eindhoven.” 62 These ideas have been developed in Chaisson, Cosmic Evolution: The Rise of Complexity in Nature; Spier, Big History and the Future of Humanity. 63 This criticism was most famously formulated in Lyotard, The Postmodern Condition: A Report on Knowledge.

References Allen, Edward, and Joseph Iano. Fundamentals of Building Construction: Materials and Methods. John Wiley & Sons, 2011. Alvarez, Walter. A Most Improbable Journey: A Big History of Our Planet and Ourselves. W. W. Norton & Company, 2016. Bastos, Rogério P., Célio F.B. Haddad, and José P. Pombal. “Foam Nest in Scinax Rizibilis (Amphibia: Anura: Hylidae).” Zoologia (Curitiba) 27, no. 6 (2010): 881–86. Benjamin, Craig, Cynthia Brown, and David Christian. Big History: Between Nothing and Everything. McGraw-Hill Education, 2013. Blackledge, Todd A., Nikolaj Scharff, Jonathan A. Coddington, Tamas Szüts, John W. Wenzel, Cheryl Y. Hayashi, and Ingi Agnarsson. “Reconstructing Web Evolution and Spider Diversification in the Molecular Era.” Proceedings of the National Academy of ­Sciences 106, no. 13 (2009): 5229–34. Blake, William. The Pickering Manuscript. Kessinger Publishing, 2004. Branstetter, Michael G., Bryan N. Danforth, James P. Pitts, Brant C. Faircloth, Philip S. Ward, Matthew L. Buffington, Michael W. Gates, Robert R. Kula, and Seán G. Brady. “Phylogenomic Insights into the Evolution of Stinging Wasps and the Origins of Ants and Bees.” Current Biology 27, no. 7 (2017): 1019–25. Chaisson, Eric J. Cosmic Evolution: The Rise of Complexity in Nature. Harvard University Press, 2002. ———. Epic of Evolution: Seven Ages of the Cosmos. Columbia University Press, 2007. Chen, Jun, Zhisheng An, Lianwen Liu, Junfeng Ji, Jiedong Yang, and Yang Chen. “Variations in Chemical Compositions of the Eolian Dust in Chinese Loess Plateau over the Past 2.5 Ma and Chemical Weathering in the Asian Inland.” Science in China Series D: Earth Sciences 44, no. 5 (2001): 403–13. Christian, David. Maps of Time: An Introduction to Big History. University of California Press, 2004. 297

Esther Quaedackers

———. “The Case for Big History.” Journal of World History 2, no. 2 (1991): 223–238. Curry, Jim P. Grassland Invertebrates: Ecology, Influence on Soil Fertility and Effects on Plant Growth. Springer, 1993. Dean, Carolyn J. A Culture of Stone: Inka Perspectives on Rock. Duke University Press, 2010. Diamond, Jared. “Experimental Study of Bower Decoration by the Bowerbird ­A mblyornis Inornatus, Using Colored Poker Chips.” The American Naturalist 131, no. 5 (1988): 631–53. Eberhard, William G. “The Natural History and Behaviour of the Wasp Trigonopsis ­Cameronii Kohl (Sphecidae).” Ecological Entonomy 125, no. 3 (n.d.): 295–328. Elias, Norbert. The Civilizing Process: Sociogenetic and Psychogenetic Investigations. John Wiley & Sons, 2000. Frodeman, Robert, Julie Thompson Klein, and Roberto Carlos Dos Santos Pacheco. The Oxford Handbook of Interdisciplinarity. 2nd ed. Oxford University Press, 2017. García-Moreno, Olga, Luís Erick Aguirre-Palafox, Walter Álvarez, and William Hawley. “A Little Big History of Iberian Gold.” Journal of Big History 1, no. 1 (2017): 40–58. Garnier-Sillam, E., and M. Harry. “Distribution of Humic Compounds in Mounds of Some Soil-Feeding Termite Species of Tropical Rainforests: Its Influence on Soil Structure Stability.” Insectes Sociaux 42, no. 2 (1995): 167–85. Gordon, J. E. The New Science of Strong Materials: Or Why You Don’t Fall Through the Floor. Penguin, 1991. Hansell, Michael H., Graeme D. Ruxton, and A. Roland Ennos. “Collected and Self-­ Secreted Building Materials and Their Contributions to Compression and Tension Structures.” Biological Journal of the Linnean Society 112, no. 3 (2014): 625–39. Hansell, Mike. Bird Nests and Construction Behaviour. Cambridge University Press, 2000. ———. Built by Animals: The Natural History of Animal Architecture. Oxford University Press, 2007. Haynes, William M. CRC Handbook of Chemistry and Physics, 97th Edition. CRC Press, 2016. Jaroensutasinee, M., and K. Jaroensutansinee. “Bubble Nest Habitat Characteristics of Wild Siamese Fighting Fish.” Journal of Fish Biology 58 (2001). Kirkendall, Lawrence R., Peter H. W. Biedermann, and Bjarte H. Jordal. “Evolution and Diversity of Bark and Ambrosia Beetles.” In Bark Beetles: Biology and Ecology of Native and Invasive Species, edited by Fernando E. Vega and Richard W. Hofstetter, 85–156. Academic Press, 2015. Knapp, Ronald G. China’s Old Dwellings. University of Hawaii Press, 2000. Korb, J, and J Heinze. Ecology of Social Evolution. Springer, 2008. Kortlandt, Adriaan. “How Might Early Hominids Have Defended Themselves Against Large Predators and Food Competitors?” Journal of Human Evolution 9, no. 2 (1980): 79–94. Lyotard, Jean-François. The Postmodern Condition: A Report on Knowledge. University of Minnesota Press, 1984. Markley, Jonathan. “A Child Said, ‘What Is the Grass?’”: Reflections on the Big History of the Poaceae.” World History Connected 6, no. 3 (2009). McSween, Harry Y., and Gary R. Huss. Cosmochemistry. Cambridge University Press, 2010. Müller-Schwarze, Dietland, and Lixing Sun. The Beaver: Natural History of a Wetlands Engineer. Cornell University Press, 2003. Needham, Joseph. Science and Civilisation in China: Volume 4: Physics and Physical Technology, Part 3, Civil Engineering and Nautics. Cambridge University Press, 1971. ———. Science and Civilisation in China Volume 5: Chemistry and Chemical Technology Part 12: Ceramic Technology. Cambridge University Press, 2004. Needham, Joseph, and Francesca Bray. Science and Civilisation in China Volume 6: Biology and Biological Technology Part 2: Agriculture. Cambridge University Press, 1984. 298

A case for little big histories

Nowak, Martin A., Corina E. Tarnita, and Edward O. Wilson. “The Evolution of Eusociality.” Nature 466, no. 7310 (2010): 1057–62. Opell, B. D. “Economics of Spider Orb-Webs: The Benefits of Producing Adhesive ­Capture Thread and of Recycling Silk.” Functional Ecology 12, no. 4 (n.d.): 613–24. Pirsig, Robert M. Zen and the Art of Motorcycle Maintenance: An Inquiry into Values. Harper Collins, 2009. Potts, Richard. “Variables versus Models of Early Pleistocene Hominid Land Use.” Journal of Human Evolution 27, no. 1–3 (1994): 7–24. Quaedackers, Esther. “A Short History of Little Big Histories.” BHP Teacher Blog (blog), 2018. https://blog.bighistoryproject.com/2018/03/01/a-short-history-of-little-big-histories/. ———. “Architectuur, Macht En Moderniteit: De Oprichting van Het Hoofdgebouw van de Technische Hogeschool Eindhoven.” Eindhoven University of Technology, 2006. Raichle, Marcus E., and Debra A. Gusnard. “Appraising the Brain’s Energy Budget.” Proceedings of the National Academy of Sciences 99, no. 16 (2002): 10237–39. Schmidt-Nielsen, Knut. Scaling: Why Is Animal Size So Important? Cambridge University Press, 1984. Shubin, Neil. The Universe Within: A Scientific Adventure. Penguin, 2013. Simon, Richard B., Mojgan Behmand, and Thomas Burke. Teaching Big History. University of California Press, 2015. Spier, Fred. Big History and the Future of Humanity. John Wiley & Sons, 2015. Talhelm, T., X. Zhang, S. Oishi, C. Shimin, D. Duan, X. Lan, and S. Kitayama. “LargeScale Psychological Differences Within China Explained by Rice Versus Wheat Agriculture.” Science 344, no. 6184 (2014): 603–8. The Getty Conservation Institute. 6th International Conference on the Conservation of Earthen Architecture: Adobe 90 Preprints. Getty Publications, 1991. Thoemmes, Megan S., Fiona A. Stewart, R. Adriana Hernandez-Aguilar, Matthew A. ­Bertone, David A. Baltzegar, Russell J. Borski, Naomi Cohen, Kaitlin P. Coyle, ­A lexander K. Piel, and Robert R. Dunn. “Ecology of Sleeping: The Microbial and Arthropod Associates of Chimpanzee Beds.” Royal Society Open Science 5, no. 5 (2018): 180382. Tillotson, Giles. Taj Mahal. Profile Books, 2010. Veblen, Thorstein. The Theory of the Leisure Class. Oxford University Press, 2007. Wilson, Edward O., and Bert Hölldobler. “Eusociality: Origin and Consequences.” Proceedings of the National Academy of Sciences of the United States of America 102, no. 38 (2005): 13367–71. Yang, FuWei, BingJian Zhang, ChangChu Pan, and YuYao Zeng. “Traditional Mortar Represented by Sticky Rice Lime Mortar—One of the Great Inventions in Ancient China.” Science in China Series E: Technological Sciences 52, no. 6 ( June 1, 2009): 1641–47. doi:10.1007/s11431-008-0317-0. York, Ryan A., Chinar Patil, C. Darrin Hulsey, Onyemaechi Anoruo, J. Todd Streelman, and Russell D. Fernald. “Evolution of Bower Building in Lake Malawi Cichlid Fish: Phylogeny, Morphology, and Behavior.” Frontiers in Ecology and Evolution 3 (2015). Zahavi, Amotz. “Mate Selection—A Selection for a Handicap.” Journal of Theoretical Biology 53, no. 1 (1975): 205–14. Zalasiewicz, Jan. The Planet in a Pebble: A Journey into Earth’s Deep History. Oxford University Press, 2010. Zeder, Melinda A. “The Broad Spectrum Revolution at 40: Resource Diversity, Intensification, and an Alternative to Optimal Foraging Explanations.” Journal of Anthropological Archaeology 31, no. 3 (2012): 241–64. Zhang, Qizhi. An Introduction to Chinese History and Culture. Springer, 2015.

299

13 THE LITTLE BIG HISTORY OF THE NALÓN RIVER, ASTURIAS, SPAIN Olga García-Moreno, Diego Álvarez-Laó, Miguel Arbizu, Eduardo Dopico, Eva García-Vázquez, Joaquín García Sansegundo, Montserrat Jiménez-Sánchez, Laura Miralles, Ícaro Obeso, Ángel RodríguezRey, Marco de la Rasilla Vives, Luis Vicente Sánchez Fernández, Luis Rodríguez Terente, Luigi Toffolatti and Pablo Turrero Asturias is a small region (ca 10,600 km2 and 106 inhabitants) in the North of Spain in Southwestern Europe.This region of the Iberian Peninsula is quite special due to the climatic conditions caused by the proximity of the Cantabrian Mountain Range to the coastline, with peaks reaching altitudes higher than 2,500 meters above sea level. This green and mountainous region has had its own and particular cultural evolution within the North of Spain, with its own language, gastronomy and historical events. Since the comparatively late conquest by the Romans, around the first century before the current era (BCE), triggered by the discovery of rich and numerous gold deposits in NW Iberia (García-Moreno et al., 2017), the region has shown mining activity up to the present day. However, in this little big history, we will not deal with gold, but with another chemical element that is much more common than this precious metal, namely carbon. The Nalón River is the largest river of Asturias. It has a basin of 4,900 km2 (­Figure 13.1), a length of 145 km and a mean volume flow of 81 m3/s. Its major tributary is the Narcea River, 123 km long, while other tributaries are the Nora and Caudal rivers.The Nalón River can serve as a perfect subject for a little big history, as we can trace its evolutionary processes in the different regimes – cosmos, Earth, life and humanity – by examining its history.

Why is the story of the Nalón River related to the element carbon (C)? We will try to understand the answer to this question in the following pages.This relation between carbon and the Nalón River will be studied from the first human impacts on the landscape to the most recent ones related to the exploitation of carbon coal deposits. While doing so, the reader will be able to understand the formation of 300

The little big history of the Nalón River

Figure 13.1  Iberian Peninsula. Inset: Asturias and the Nalón River basin. Image from Google Earth.

the natural landscape, the importance of the carbon cycle in that process and the formation of the coal deposits related to that cycle. Following, the most recent of these environmental changes will be discussed in order to frame the settlement of the first humans in this region. Then, a vision of human evolution through the study of the diseases will bring us to the recent past, when coal mining was the activity that shaped this particular human society.The present day and a gaze to the near future will again have some relation with C as the reader will recognize at the end of this chapter. The main characters of this story are the coal deposits, consisting mainly of ­carbon, which are exploited along the Nalón Carboniferous basin (formed approx. 300 ­million years ago (Ma)). This mining activity has shaped both the culture and the landscape of this region until today. In Asturias, talking about the Nalón River and its basin means talking about coal mining. Humans have left their footprint on the planet in every phase of our history, modifying the landscape. Earth science researchers analyze these modifications, the result of the historical process of interactions between society and nature (e.g. Jiménez-­Sánchez et al., 2011). This footprint can also be seen in study of the ecosystems (­McDonnell and Pickett, 1990). Based on empirical evidence, we know that the human presence 301

Olga García-Moreno et al.

changes natural ecosystems and modifies their biological/geological balance: everything becomes dependent on human population density and its cultural phases. Since the advent of agriculture and animal husbandry (between 10,000 and 14,000 BCE), humankind has altered not only the landscape in a fundamental way but also the ecosystems. Agriculture and livestock are based on the controlled breeding of plants and animals selected from their wild forms. This means less diversity (monoculture, prevalence of certain species over others), edaphological changes (loss of organic soil and mineral contents) and sometimes even desertification. Furthermore, in the evolution of human culture, specialized production was always accompanied by urbanization. We can track these changes following the historical evolution of the Nalón River.The following timeline illustrates this evolution (Table 13.1): The main human-mediated change in the Nalón River basin occurred during the industrial revolution. Industrial development was supported by coal mining as a source of energy, and agriculture in the region involved the use of pesticides and artificial fertilizers. Coal mining in the Nalón River basin caused a large economic and demographic growth that went hand in hand with a huge urban and industrial development. However, many mines discharged pollutants in the river, contributing to a drastical change in the water quality. Yet the rural society of the Nalón River basin also tried to preserve its ancestral traditions and local culture. Many farmers in the basin refused to use modern fertilizers and continued to use traditional manure, while seeking to fight plagues with folklore remedies (Dopico et al., 2009).Traditional crops were also preserved all along the

Table 13.1  Timeline for the Nalón River basin human history TIMELINE for the Nalón River 1 Lower Palaeolithic, more than 200,000 BCE: first traces of human presence. Since 127,000 BCE occupied by the Acheulean groups. 2 Middle Palaeolithic, around 90,000 BCE: Neanderthal remains and Mousterian archaeological record. • Upper Palaeolithic, around 37,000 BCE: Sapiens individuals. Presence of rock and portable art. • Neolithic, around 6,000 BCE: Cattle farming and agriculture. Necropoles. • First millennium BCE: the first castros (aggregations of circular dwellings) appear. 3 Late first century BCE: the Romans endeavor to invade these gold-bearing territories. • 409 CE:  Vandals and Swabians arrive in the Iberian Peninsula. • X to XV centuries CE: feudal kingdoms develop in Asturias. • XV to XVIII centuries CE: Asturias is an agricultural and livestock region with an elite composed of nobles and clergy. 4 Late XVIII century CE: industrial development begins, to fade in the mid-1980s of the XX century CE. 5 X to XV centuries CE: feudal kingdoms develop in Asturias. 6 XV to XVIII centuries CE: Asturias is an agricultural and livestock region with an elite composed of nobles and clergy. 7 Late XVIII century CE: industrial development begins, to fade in the mid-1980s of the XX century CE. 302

The little big history of the Nalón River

basin. Consequently, the levels of heavy metal contamination by artificial ­fertilizers in the areas where traditional agriculture was preserved, such as the upstream zones, were much lower than in most other rivers of the region. This rich ensemble of culture, tradition and profound contact with nature is quite unique and compensates in many ways for the impacts of mining and other industries. The currently declining economic situation also affects the basin, yet the spirits are high. Some recent initiatives are ecologically friendly, such as sheltering rescued bears in protected refuges, kayaking in reservoirs, promoting sustainable sport fishing and developing green trekking routes in the gorgeous mountains that surround the basin. Zooming out again, we can see that nowadays ecologically sustainable activities are considered the only sensible alternative for future developments. Perhaps the Nalón River basin inhabitants have known it since their very first arrival in the Palaeolithic.

Human footprint in the Nalón Landscape We can explore the role of landscape transformations since the Industrial Revolution in the Nalón River basin, in order to understand its different spatial-temporal scales, by examining historic aerial photography, which has proven to be a valuable source of spatial change information (Fernández García, 2004). To do this, six empirical examples of transformations will be addressed at different scales, aiming to draw a detailed picture of collective learning processes and the increase of complexity. Before doing so two things should be pointed out. First, the population distribution is not uniform along the river banks, while there is a higher population density in the coal mining area of Langreo. Second, the current socio-economic structure is characterized by the context of a crisis resulting from the gradual abandonment of traditional industries such as mining and steel in favor of a diversified economy that is, however, unable to offer the same number of jobs as the preceding economic model. Only the new ecological activities mentioned above appear to offer a positive future for the economy in the region. Figure 13.2 illustrates these changes of the land using aerial photography.  The human foot print in the Nalón basin landscape will be summarized in the following paragraphs.

Alluvial plains The fertility of alluvial soils on the floodplains arises from the fine particulate organic matter that is deposited there during flooding. These conditions make floodplains ideal places to grow crops. As a result, the Asturian landscape was shaped thanks to traditional agriculture, first with crops such as cereals and beans.With the changes introduced in the eighteenth century, the agricultural landscape shifted to a model dominated by beans, corn and potatoes, while using crop rotation practices. Plots were organized in elongated narrow strips of land. A group of such plots that are fenced off to keep the cattle out is called ería. It is a collectively managed space with individual use. Following European agrarian policies, the rural economy became dairy oriented. This transformed the landscape in the 1970s by replacing croplands by grasslands while abolishing the former way of parceling. Over the last decade, a new form of intensive agriculture has emerged in the floodplains of Pravia, namely kiwi plantations, thus erasing the traces of the past. 303

Figure 13.2  A  nthropic changes through time in landscapes in the Nalón River basin. Orthophoto mosaic derived from the photogrammetric flights of the Serie B 1956–1957, Agriculture Ministry 1983–1986 and the campaign of Plan ­Nacional de Ortofotografía Area (PNOA) 2011.  Bottom images: The Nalón River and its natural areas (left), Samuño mining ecomusem (right).

The little big history of the Nalón River

From traditional farming to industrial land use Olloniego – Tudela Veguín comprises another alluvial plain, but one with a different fate: the traditional agricultural landscape changed into an industrial park. In this example, regional land planning played an important role in changing the agricultural landscape into an industrial one. Several decisions, such as river canalization and a new road crossing the floodplain, were the triggers for this huge transformation.

Spatial distribution of woodlands Asturias has a temperate, warm and rainy climate that provide very favorable conditions for deciduous forest development. The spatial distribution of the woodlands is linked to several factors, both natural and human. Humidity, orientation, slope and altitude are the main natural factors that explain the spatial distribution of beech forests in the upper course of the river, or of chestnut trees in the middle course. However, the presence of oak trees in the mountains is an effect of human actions that have replaced the earlier vegetation due to activities such as mining.The vegetation near the river mouth has been altered entirely ever since eucalyptus plantations, a fast-growing source of wood, have eliminated native species.

Integrated urban regeneration During the 1980s, the Langreo municipality, situated in the middle of the river ­basin, was faced with the closure of the mines as well as with several environmental problems such as water pollution. In order to solve these problems the municipality ­developed an urban regeneration program consisting of the construction of a water treatment plant, the restoration of the riverbed, an in-fill (integrated planning) operation aiming to merge the different parts of the city, and a program to improve public services (Fernández García, 1996). In addition, new companies using advanced ­technologies settled in this area taking advantage of the European funds for redeveloping coal mining regions.

From rural landscape to hydroelectric dam The water demand from the central area of Asturias is served by the water reservoir of the upper Nalón River. In 1978 a hydroelectric dam was built for this purpose in Tanes and Rioseco. In consequence, the environmental impact resulting from the artificial lake completely erased the traditional rural landscape on the valley floor.

Protected natural areas and industrial heritage The landscape is considered of common value for the whole society. This is why the number of protected natural areas has increased in the recent past. Something similar has been occurring with the industrial heritage. The Nalón River basin has it all: natural and cultural heritage.The Redes Natural Park (Biosphere Reserve) is covered by forests, mainly beeches, which create a spectacular landscape in autumn owing to the different colors of the leaves before they fall. Samuño Valley is one such example 305

Olga García-Moreno et al.

of mining heritage. The intensity of the exploitations and the different technologies have left a remarkable heritage in the present landscape, which allows us to consider the valley as both an ecomuseum and a mining park. Landscapes always change, because they are the expression of the dynamic interaction between natural and cultural forces in the environment (Antrop, 2000, 2005). The spatial transformations carried out in the recent past as well as those made in the present will determine the future of the Nalón basin. Such an approach attempts to break the usual narrative of facts arranged on the same timeline by opening the analysis to the spatial dimensions (Massey, 1999).

Natural evolution of the Nalón Landscape At this point, to understand the landscape we should consider the influence of the geological evolution of this region. We have seen how human activities have modified the natural landscape that has been shaped by the geological processes.The natural landscape, here as well as in any other place on Earth, is the result of internal processes caused by plate tectonics, and external processes caused by the interaction of the geosphere with the hydrosphere, atmosphere and biosphere. In the case of the Cantabrian Mountains, in particular in our region of interest the Nalón River basin, the first event of internal ­ aleozoic era and external processes that shaped the landscape took place during the P (542–251 Ma). By the end of the Paleozoic, there was a process of elevation forming a great mountain range. This was caused by the collision of the continent that existed at that time – called Gondwana – with another continent. This collision formed the ­Pangea continent, in what geologists call the Variscan orogeny. Orogenesis is the geological process of mountain range formation caused by the interaction of tectonic plates. This Variscan orogenic process produced the strong deformation of previously formed rocks, which included limestones, coal layers, shales and sandstones. Earlier, sedimentation had taken place on the bottom of an ancient continental shelf including areas close to the ancient continent, giving rise over time to limestones and, wherever the remains of plants were buried in the sediments, to coal deposits. The mountain chain that was formed in this process was possibly even taller than the Himalayas of today. The second event started at the beginning of the Mesozoic Era (252–66 Ma), when the crust was submitted to stretching, or extensional processes, and the Variscan mountain chain was completely eroded. At that time the current Nalón area (we cannot speak yet about the Nalón River until much more recent times, as we will see below) was again submerged under the sea. It was during this Mesozoic extension that most of the deposits of metallic minerals of the Cantabrian Mountains emerged. By the end of the Mesozoic, a second orogenic event complicated the bedrock structure in the Nalón basin further. This complex structure is very important to understand the history of coal mining in this region. During this second orogenic event, related to the Alpine Orogeny, the Cantabrian Mountains were uplifted and a fluvial network started to flow into the Cantabrian Sea. At that time, the Nalón River basin acquired the general shape that we know today, even though it would still be modified to some extent through geological time. The Alpine Orogeny in this region is related to the opening of the Bay of Biscay. In relation to this process, the Iberian small plate moved towards France, triggering a collision that formed the Pyrenees as well as the elevation 306

The little big history of the Nalón River

of the Cantabrian Mountains. All of this occurred around 30 million years ago. The movement of large faults generated a huge relief with an E-W orientation, greater than the current Cantabrian Mountains. Various external processes eroded this enormous relief. At the beginning, the fluvial network was settled, and the rivers would circulate towards the Cantabrian Sea following the shortest way, from South to North, following the slope of the cordillera itself. However, these first Cantabrian rivers would also settle in zones with some substrate weakness, such as rocks with lower resistance or faults. This is the case of the Nalón River, which at some points takes advantage of one important and currently active fault, the Cantabrian or Ventaniella fault. Knowing the geological history of the Nalón River is important to explain why it is not economically viable anymore to extract coal from these mountains, as well as the reason causing the end of the mining activities in the region. The complex geological structure of this coal area is in stark contrast with the more simple geology of other regions such as the coal mining region of South Africa, where the geological processes that shaped the structure of the coal layers were much simpler.The technical challenges involved in extracting coal from a geologically complex structure are with no doubt greater (thin coal layers, difficult to follow, etc.) and generate more costs than the easy to extract coal of, for instance, the Richards Bay Port Terminal, which is one of the currently leading export coal terminals in the world. In addition, the e­ nvironmental policies in the different countries would influence the costs and feasibility of coal extraction in a certain place. These facts can be applied to other coal regions in Europe, such as the UK, although the effects of the second process of mountain building, the Alpine Orogeny, did not have any such strong effects there, if any. All of this helps us to understand how the geological history of a certain region provides opportunities for, and poses clear constraints to, economic and social evolution of a society whose economy is based on the exploitation of natural resources.

The carbon cycle and the formation of coal deposits The carbon cycle, or the way that this element is exchanged among the atmosphere, hydrosphere and geosphere, begins in the early Earth and is soon affected by biological activity. The evolution of life has been modifying the equilibrium of the carbon cycle for the last 3,800 million years. However, in recent times a new agent has been altering this cycle: anthropological activity over the last 250 years has introduced important changes in the carbon cycle. A good example for understanding this great change, which is affecting Earth’s climate and may be causing irreversible consequences for the future of humanity, can be found in the history of the Nalón River. Through billions of years of evolution, living organisms have been using carbon for their life cycles as part of their internal biochemistry: in fact, all building blocks of life consist of carbon-based biomolecules that all living organisms make themselves or obtain from other lifeforms. The continuous and sometimes catastrophic geological processes that shaped the history of our planet also molded the evolution of life. It is beyond the scope of this chapter to describe the different hypotheses for the origin of life and its evolution. Our case study will focus on a particular time of the Paleozoic Era. As described above, it is possible to trace the history of the Nalón region back to the Paleozoic Era (starting 540 Ma).The pre-Paleozoic history of this region is difficult 307

Olga García-Moreno et al.

to assess, as its geological record only comprises rocks of Paleozoic age, although some older rocks can be found to the West, in the Narcea area. The beginning of the ­Paleozoic Era is marked by the so-called Cambrian explosion, the momentous period in the evolution of life when organisms acquired hard skeletons, which presumably fossilized easier than the soft bodies of Precambrian organisms. From the existing fossil record, it can be shown that the Nalón basin occupied a southern position in the early Paleozoic, actually close to the South Pole (and therefore having a colder climate), and that it later drifted northwards, reaching equatorial regions in Carboniferous times. Because the Nalón basin was a prominent coal hub, we will focus on the evolution of plants whose remains produced the coal layers, as well as, more in general, on the carbon cycle. At the end of the Silurian (420 Ma) plants conquered the land.They soon diversified and expanded over the continents. Pteridophytes, vascular plants that reproduce via spores, dominated the continent. Yet the Nalón area at that time was under the sea, so we do not find such fossils in the rock record. During the end of the Lower Devonian (~400Ma), there was an important development of coral reef life in the region, as at that time the climatic conditions of our area of study were similar to those of the tropics today. Devonian limestones – sedimentary rocks that form under water – found today in the northern portion of the Nalón basin preserve beautiful examples of the great variety of organisms that once formed these coral reefs. Later on, during the Carboniferous period, a wider carbonate platform developed in this area. Today, such platforms are currently forming carbonate sediments, for example, in the Bahamas banks, which offer a good comparison with our area of study at those earlier times. The formation of carbonate sediments is an important CO2 sink in the carbon cycle. Carbon is fixed by organisms from the air to form sediments or is directly precipitated from seawater under the proper conditions of temperature and depth.These sediments transform into the sedimentary rock called limestone by diagenetic processes (processes that involve an increase in pressure and temperature deeper under the surface of the Earth and, in doing so, convert those sediments into sedimentary rocks).These limestones were subsequently deformed and faulted during the Variscan orogeny, as it explained above, and later by the Alpine orogeny. Today, they form the highest reliefs in the region.We will come back again to these limestone reliefs while discussing the caves used by the first humans inhabiting the region. The calm and stable environment needed for the formation of the carbonate plat­ arboniferous form and the formation of limestones was disrupted at the end of the C (300 Ma). The formation of Pangea was then taking place, which led to the closure of seas like the Rheic Ocean where our carbonate platform was developing. In our region, a new relief was forming to the West, while the sedimentation in the basin changed from carbonate to detrital sediments that resulted from the erosion of this new relief that was being uplifted as part of the Variscan orogeny. As a result, the geological record of this age in the area consists of sandstones and shales formed from those detrital sediments. Together with the continental sediments transported from the West, mainly by the ancient rivers, the remains of the dominant living organisms were buried there. The mountains that were eroded and were, in consequence, feeding the sedimentary basin by their erosion were covered by dense forests. Enormous tree-sized plants and ferns grew and reproduced easily because the O2 concentration in the atmosphere at that time was greater than the current concentration. 308

The little big history of the Nalón River

Figure 13.3  Recreation of Carboniferous period forests by D. Álvarez Laó.

The fabulous vegetation that formed the forests developed so well because the region was at the Equator at that time (Figure 13.3). The remains of these forests buried together with sand and clays formed the coal layers that are now interbedded in the sedimentary rocks.This carbon sink in the carbon cycle was very important in Carboniferous times. Tons of plant remains were buried in the bedrock over 300 million years, thus storing the great amounts of energy that they had accumulated by the process of photosynthesis. This energy that stayed in the geosphere over millions of years is now suddenly released by the anthropogenic exploitation of coal. In only 250 years, more than 32,000 metric tons of carbon have been emitted to the atmosphere worldwide due to the burning of fossil fuels: coal, gas and oil (data from the World Resources Institute (2017) including the changes in land use).According to the Intergovernmental Panel on Climate Change (IPCC) report in 2007, the burning of fossil fuels is the main factor (75%) altering the carbon cycle and thus causing climate change, while the rest would be due to changes in land use. Both CO2 and CH4 are emitted by coal burning and are important carbon-based greenhouse gases.

Recent environmental changes in the Nalón River basin The Quaternary, the period that spans from the last 2.58 million years to the present, is characterized by significant climate changes, from warm and wet to severe cold and dry episodes.These climate fluctuations involved deep transformations in the environment, concerning both flora and fauna. They were not related to human activities, as is the case with current climate change. Humans are recent actors in our story, and the first human societies (hunter-gatherers) presumably caused few changes in the environment. The causes of those natural climate changes in the Quaternary are related to cyclical fluctuations of the Earth’s orbit and the tilt of its axis, known as Milankovitch cycles.There are other natural causes for climate changes on Earth as well, for instance, related to variations in the activity of the Sun (van der Plicht et al., 2004). 309

Olga García-Moreno et al.

The Quaternary environmental record has been well preserved in the Nalón River basin thanks to the great expanses of limestone rocks and the cavities that formed within those rocks. Again, the element carbon allows us to understand this, as limestones are rocks made of calcium carbonate (CaCO3), which can be easily eroded by water. These rocks of the Devonian and Carboniferous ages in the Nalón River basin allowed the development of numerous caves and sinkholes. In their turn, these cavities provided excellent conditions for the preservation of faunal and floral evidence over thousands of years: constant temperature and humidity, low energy sedimentation (that allows the slow stacking of layer upon layer of sediments), lack of organic activity that could otherwise destroy that evidence, etc. As a result, a number of caves from the Nalón River basin have provided good Quaternary sedimentary sequences with plenty of fossil bones, pollen, charcoal as well as evidence of early human activity (stone tools, Paleolithic art and even human bones). In most cases, the Paleolithic humans who lived in this basin thousands of years ago were responsible for the bone accumulations. But in other cases, carnivore activity (mostly hyenas) transported large amounts of bone fragments into the caves. Finally, a few other fossil assemblages, mainly at high altitudes, originated as a result of natural traps, for instance, sinkholes into which animals fell accidentally. This faunal and floral record shows clear evidence of the occurrence of temperate and cold episodes in the basin. During the temperate phases, the faunal spectrum was dominated by a wide variety of herbivore and carnivore species, summarized in Table 13.2.

Table 13.2  Large-mammal spectrum during the Late Pleistocene in the Nalón River basin Herbivores

Carnivores

Woolly mammoth Straight-tusked elephant Red deer Reindeer Roe deer Wild boar Chamois Ibex Aurochs Steppe bison Wild horse Woolly rhinoceros Steppe rhinoceros Cave hyena Cave bear Brown bear Wolf Lion Leopard Eurasian lynx 310

Mammuthus primigenius Palaeoloxodon antiquus Cervus elaphus Rangifer tarandus Capreolus capreolus Sus scrofa Rupicapra pyrenaica Capra pyrenaica Bos primigenius Bison priscus Equus ferus Coelodonta antiquitatis Stephanorhinus hemitoechus Crocuta Crocuta Ursus spelaeus Ursus arctos Canis lupus Panthera leo Panthera pardus Lynx lynx

The little big history of the Nalón River

The vegetal landscape, inferred from pollen and charcoal analyses, was dominated by deciduous forests, mostly oak trees, birches, hazel, etc., but also by pine forests. Cold phases were characterized by the occurrence of different herbivores, while the carnivore spectrum was virtually the same. Typical cold-adapted species such as woolly mammoth (Mammuthus primigenius), woolly rhinoceros (Coelodonta ­antiquitatis) (Figure 13.4) and reindeer (Rangifer tarandus) were present during these periods. Among them, other species already mentioned, such as red deer, steppe bison, horse, chamois and ibex, also occurred during the cold phases. The flora was dominated by steppe-like environments, with a predominance of herbaceous plants (mainly grasses) and the scarce presence of trees (mostly conifers). The scenery reminds us of a type of “savanna” but then developed under cold conditions: herds of large mammals like mammoths, rhinos, bisons and horses living among great predators as hyenas and leopards, all roaming a landscape dominated by herbaceous vegetation with few trees. The faunal record not only consists of fossils but also Paleolithic art: there are many findings of parietal and portable (“plaquettes”) art in the basin representing red deer (probably the favorite prey of the Paleolithic hunters in the valley), bison, horse and, in a few cases, mammoth and woolly rhino, among other species. Salmonids (the Atlantic salmon Salmo salar and the brown trout Salmo trutta) have been fished throughout most, if not the entire history of the human occupation of the basin. They are still being fished today. The Nalón River was one of the refuges for these species during the cold periods that led to glaciation in most of their current locations. A characteristic of the cold periods at that time in the Nalón basin was the formation of glaciers at high altitudes. As a result glaciers contributed to shaping the landscape of the Nalón River basin. Thus, in the headwaters of the Nalón River, located to the South, glacial landforms and deposits are well preserved at altitudes above 1,100 m altitude. We know that glaciers developed in this area, reaching their largest extension at around 33.5 ka BP ( Jiménez-Sánchez et al., 2013).

Figure 13.4  W  oolly rhinoceros lived in the Asturias area during cold phases in the Quaternary, by D. Álvarez-Laó. 311

Olga García-Moreno et al.

After glacial retreat, the evolution of the landscape was mainly conditioned by rivers and landslides. At the present day, excessive rainfall triggers landslides and floods, thus creating risks for the most populated areas of the Nalón basin, in particular in the areas with coal mining activity. Today, the basin landscape is very similar to how it was during these previous temperate phases, except for the changes produced by human activity discussed above. Autochthonous forests are still abundant in the middle and lower altitudes of the area. Among the herbivores, red deer, roe deer, wild boar and chamois are still frequent, but large ungulates such as rhinoceros, bison and wild horse disappeared a long time ago. Domestic ungulates such as cattle, goat, sheep and domestic horse now occupy their ecologic niche. Fortunately two big predators, the brown bear and the wolf, still live in the basin, although under increasing stress due to their troubled relationship with human activity, which has been, without a doubt, the major force driving the environmental change as it occurred during the last part of the Quaternary in the Nalón River basin.

Humans in the Nalón and the evolution of their health complications Since ~1.2 million years ago the human population in the Iberian Peninsula has consisted of four Homo species: H. antecessor, H. heidelbergensis, H. neanderthal and H. sapiens (AMH, or Anatomically Modern Humans). In the Cantabrian region, and specifically in Asturias, there is an archaeological record of material objects of the last three species mentioned above, yet only fossil remains from Neanderthal and AMH individuals. Most of the archaeological record testifies of the presence of humans in the region since at least 127,000 years BP, with Acheulean artifacts largely present in river terraces, followed by Middle and Upper Palaeolithic sites, some of them of worldwide relevance: La Viña, Las Caldas, La Lluera or Peña de Candamo, usually accompanied by a huge archaeological record of material remains as well as impressive rock and portable art (Figure 13.5) (AA.VV. 1990, 2007). The archaeological record continues with Neolithic and Bronze Age discoveries such as the great megalithic structures of La Cobertoria as well as the copper mines

Figure 13.5  P  eña de Candamo Cave. Niche: Horses and aurochs. Photography: Javier Fortea. 312

The little big history of the Nalón River

and the cemetery of El Aramo. There is also significant architectural evidence of the Iron Age, the so-called castros (for instance, San Martín or Llagú), that extends to Roman times. Good examples of this period are La Carisa camp and route, and the interesting Valduno baths. The Middle Ages are represented by castles (San Martín, Doña Paya), towers (La Buelga, El Condado), monasteries (Cornellana), Collegiate churches (Salas) and complexes such as Olloniego. Special attention must be paid to the churches belonging to the named and unique Prerrománico such as Bendones, San Pedro de Nora, S­ antianes or Tuñón, but also to the Romanesque remains of Sograndio and Ciaño. Finally, there is the weapons factory at Trubia, founded in 1794 and still in use (AA.VV. 1990). Thus, the discussion above makes clear that the Nalón basin exhibits a great many examples, sometimes extraordinary, that belong to all historical periods. In addition, according to the particular technology, economy, social values, symbolism, scientific and cultural knowledge characteristic of the different periods involved, the population organized their own cultural heritage, their relationship with nature and, as a consequence, their way of living. During their evolution, it is evident that humans suffered from diseases, since pathology is inherent to life. The conjectures for reaching some conclusions about how humans got sick come from paleopathology studies. We can trace the intricacies of the first inhabitants’ health and way of life by studying primary sources (anatomic materials, generally bones) and secondary sources (artistic representations, votive offerings, coins, etc.), which have yielded some interesting findings. Marks found in teeth illustrate that these early humans were mainly right handed, and that they used their teeth as a tool to hold, cut, etc. This is why they had severe dental wear and osteoarthritis of the temporomandibular joint. We can say that they used sticks for their dental hygiene by the wear marks in the dentin, and that they rarely suffered from caries. The fossil record tells us that most of the deaths were of women during childbirth (Bermúdez de Castro, 2002). Another curiosity is the record of cranial trepanations. We will not go through the history of all human diseases. But we would like to highlight the big change in the way of life of the Nalón inhabitants when coal mining activities started in the region. People stopped farming because they could get salaries from the mines and slowly improved their way of life. Even though different diseases related to coal dust in the air made their appearance, their life expectancy increased. Compared to the diseases in agriculture societies, such human afflictions became very different after the Industrial Revolution. Pollution and bad hygienic conditions gave rise to tuberculosis. Many miners suffered silicosis, another lung disease caused by the inhalation of siliceous powder during coal mining. Even today, now the decline of coal mining activities is evident, the main hospital in Asturias has a world reference unit about this disease. Alcoholism was also an important health issue at that time.

Coal mining Black coal is the main type of coal deposit in Asturias, but the other three types (peat, lignite and anthracite, all classified according to their carbon content), are also found. Black coal is the only type of coal mined in the Nalón basin. The start of 313

Olga García-Moreno et al.

scientific research on Asturian coal and its usefulness goes back to the eighteenth century, namely by G. Casal, Conde de Toreno and by Jovellanos (Adaro Ruiz, 2003). ­Jovellanos was a major figure of the Enlightenment in Spain. He founded the Real Instituto de Mineralogía y Náutica (The Royal Mineralogy and Nautical Institute). At that time he was one of the main promoters of the use of coal as a resource. The industrial use of coal meant a radical change in the economy as well as in the way of life in the region. The change from a farming to a mainly industrial economy has influenced the history and the characteristics of Asturias and the Nalón basin up to the present day. Coal mining prompted changes in land use, including the design and building of new infrastructures for transporting coal from the mines to the seaports, from where it was exported to the consumers. Because of coal mining, in the nineteenth century also the construction of new communication lines was initiated with both the coast and the rest of the Peninsula. At the beginning of this development, coal mining was mainly a small-scale craftlike activity. In the early nineteenth century, coal was exploited in what in Asturias is called chamizos, or small open pits, while was transported by oxen and mules along dirt roads. The Nalón River was the main highway for transporting coal, in small boats or barges called chalanas. These artisanal pits only exploited the most accessible coal layers using only very basic technology. Gradually, the technical knowledge increased. The mines became larger and the companies that owned them started to exploit deeper and more complex layers. To do so, it was very important to educate professionals (engineers and technicians) who, in some cases, were sent to different European countries to acquire that knowledge. In addition, the geological information about the area increased, while professionals such as Wilhem Schultz developed detailed topographic and geological maps. Also important for the mining activities were the changes in policy: in 1825 the Ley de Minas (Law of mining) was approved, in which coal received a special treatment. As a consequence, changes in the economy and way of life were underway in the area. A new railroad infrastructure replaced the primitive barges shuttling between the mines and the San Esteban de Pravia port, while the road infrastructure was improved as well. The first miners in Asturias were called “mixed laborers,” because they worked both at the mines, with very low efficiency, and at their own lands. They earned very low salaries and still lived in misery due to the low output in the mines. The general life expectancy increased only later. The golden age of coal mining in Asturias lasted up to the 1970s. While at the beginning the coal was mostly exported, soon industries were established in the region to take advantage of this energy resource. Iron and steel industries began to operate close to the coal mining. These industries needed coal to make steel, and it was better to have the blast furnaces close to the coal sources. Iron can also be found relatively close by. It was mined in Asturias and not very far from there in other places in Northern Spain as well. Supported by some foreign investment, Duro Felguera (and the Fábrica de Mieres in the Caudal Valley) ruled the economy of this region for a considerable period of time. In 1961, the peak production of Asturian coal reached 7.9 million tons. The history of coal mining is also marked by the foundation of the state institution HUNOSA in 1967 that brought together almost all the mining 314

The little big history of the Nalón River

companies in Spain with the objective of taking advantage of achieving synergies between all the firms involved. However, soon after that, in the 1970s, the industrial and mining activities in the area started to decline. There are several reasons for this, ranging from geological and technical reasons to economic changes. The coal layers in the basin are relatively thin (between 0.8 and 1.8 m) and often situated in a subvertical position, which makes it difficult to extract the coal because underground mines are needed to extract it. In addition, the reserves are small compared to other mining areas worldwide, while the cost of transporting this competing coal also dropped. The total amount of coal production since the beginning to 2012 is almost 500 million tons. Also hazards related to its exploitation, such as the presence of firedamp (gas in coal mines that explodes on contact with air) influenced the decline of coal mining. All these factors have made Asturian coal non-competitive compared to the black gold from other countries that have more favorable reserves and types of deposits. Other industries were installed in the Nalón basin also related to the element carbon: an important limestone quarry was developed where they mixed Portland cement for all of Spain; the first cement industry established in Spain is located in Tudela Veguín, in the Nalón riverbed. It has been active since 1905. Even with all the inconveniences caused by the mining and industrial activities, we have to acknowledge that many good things were related to the opening of the coal mines and our regional cultural development: infrastructures such as the road Gijón – León, built in 1832, which connects Asturias with the rest of the Peninsula, or the Langreo – Gijón railway, built in 1856, one of the first railroads in Spain. From the first part of the twentieth century to the 1960s towns in the Nalón basin doubled or trebled their population. The imprint of coal mining activities in the cultural sphere is prominent, with movies such as “Las aguas bajan negras” (Waters flow in black, 1948, J. L. Sáenz de Heredia) and the folklore and other cultural traditions, including music and costumes.

The future and the very beginning: carbon from the stars to the Nalón basin and back We have reviewed the processes that explain how coal was formed, how the element carbon is concentrated in Earth as well as its biogeochemical cycle. However, we have not mentioned yet how this element was formed in the stars and how it got concentrated on our planet. This should have been the beginning of our little big history. Nevertheless, we decided to link this very beginning to the future, because looking at the stars means not only looking to the past but also to the future. We are the only species on Earth that looks at the sky asking fundamental questions about our origin and our future. Moreover, all the environmental changes that we are causing are almost negligible for the Earth as a planet (in the sense that there have been many other periods of climate change in the history of our planet). Yet they are of extreme importance for the fate of our species. This is so true that scientific studies of other possible habitable planets and exo-terrestrial resources are cutting-edge research agendas of all the main international space agencies (NASA, ESA, etc.). To perform these studies, new technologies need to be (and currently are) developed in 315

Olga García-Moreno et al.

the field of aerospace exploration. Most of these technologies are based on new materials needed to build new satellites and spacecrafts with appropriate properties that allow them to be able to go further in that exploration. The element carbon is very closely related to these new materials. Graphene, nanocarbon fibers and ­carbon-based nanomaterials are widely used all in these technologies. In this respect, in the Nalón basin, the Nanomaterials and Nanotechnology Research Centre (National Research Council – University of Oviedo – Principality of Asturias) combines high quality and internationally competitive interdisciplinary research with scientific and technological demonstrations. Some of them are related to the aerospace sector. For most of them, carbon-based nanomaterials are used. One of these applications is the development of new materials that serve as blanks in mirrors for the most recent generation of satellite and land telescopes (Figure 13.6). According to the currently accepted standard Hot Big Bang model, only the lightest elements – hydrogen, helium and a very small percentage of lithium – were produced during the first minutes, right after the initial very fast expansion of the Universe. Later on, due to its continuous expansion and consequent cooling of the plasma that filled it, the Universe experienced the recombination of electrons and protons, thus forming the first atoms, some 380 thousand years after the Big Bang. At that time, the Universe became transparent to radiation, which as a result could move freely in space without further interactions with matter. Thus, as it is usually stated, matter “decoupled” from radiation. All the ordinary matter (mainly hydrogen and

Figure 13.6  S park Plasma Sintering equipment (left) in the Nanomaterials and Nanotechnology Research Centre (upper right) and nanostructured materials made therein (lower right). The research center building is located on the former coal mining plot. Images by CINN. 316

The little big history of the Nalón River

helium) could then start to form cold clouds of matter under the slow gravitational attraction produced by the more dense matter regions that emerged. During this epoch the Universe was essentially “dark,” since there were no sufficiently energetic processes able to excite atoms and make them emit radiation. By the end of these so-called “dark ages,” some 200–400 million years later, those clouds very probably condensed into the first generation of (very massive) stars, compacted and built up by the action of gravity in higher-density local fluctuations of the overall matter-energy distribution. Those very massive stars (up to or even more than one hundred times the mass of our Sun) evolved on timescales of less than a few million years and went through different burning stages, in doing so producing heavier and heavier elements all the way up to iron, the most stable element in the Periodic Table. After forming an iron core, which cannot support any further fusion burning, these massive stars suddenly collapsed under their own weight and then blew off their outer layers in a type-II “supernova” explosion, seeding the Galaxy with many chemical elements, especially carbon, oxygen, silicon, sulfur, etc., that had been produced by those stars. During those explosions, heavier elements, between iron and uranium, were also produced, this time in the expanding envelope by neutron-capturing processes. Also these elements were blown into the interstellar medium.The ejected material, mixing with hydrogen and helium clouds, then served as the seeding material for new generations of stars, such as our Sun. Only much later did low- or intermediate-mass stars – which were undergoing much slower hydrogen burning over billions of years – finally evolve into their “Red Giant” phase and began to contribute to seeding the Galaxy. These stars, such as our Sun, cast off most of their elements – notably carbon and ­nitrogen – near the end of their lives in gassy exhalations called “planetary nebulae.” In this way, after various billions of years and different generations of stars, it is possible to obtain the carbon abundance that we currently observe close to the Sun in our Galaxy. As discussed before, after stars produced the observed abundances of the most common chemical elements, some 5.0–4.8 billion years ago, a large, diffuse and cold interstellar gas cloud (with dust grains inside) started to contract under gravity. As it contracted, the atoms and molecules inside it interacted with a higher collision rate. Thus, the cloud (or “nebula”) heated, flattened and started to spin faster, due to the conservation of the cloud’s total angular momentum. After tens to hundreds of ­million years, depending on the original density of the cloud environment, the cloud evolved into a spinning disk of gas and dust. While hydrogen and helium remained gaseous, other heavier materials condensed to form solid “seeds,” the first building blocks that led to “planetesimals.” As demonstrated by hydro-dynamical simulation studies, solid seeds collided and stuck together, larger ones attracting smaller ones by gravity, and thus growing in size. Terrestrial planets are too small to capture hydrogen and helium gas which, after about one hundred million years were stripped off from the central portion of the solar system by the wind of the new-born Sun. The debris left over from this formation process became the rocky asteroids and, farther away from the Sun, the (mostly icy) comets. This process finally led to the current solar system, with most mass concentrated into planets, after the end of the chaotic “collision” phase, some 4,550 million years ago, which is the age of the most ancient meteorites so far observed. 317

Olga García-Moreno et al.

In this epoch, the most-supported theory suggests a major collision of the Earth with Theia – a body the size of Mars – that stripped part of the Earth’s crust and outer mantle and led to the subsequent formation of the Moon. After this catastrophic event, due to the action of gravity, Earth experienced a segregation of the different materials inside it, leading to the formation of the nickel-iron core and of the ­silicate-rich mantle. In the meantime, without any more collisions from planetesimals and debris of relevant size, Earth’s crust started to solidify.This latter event marked the end of the Hadean or Hadic Eon, some 4,000 million years ago. Carbon, being a light element, was concentrated in the outermost parts of the Earth, mostly in the crust and the atmosphere, where it became involved first in its own geochemical cycle and, after the appearance of life, also the biogeochemical cycle. With the formation of carbon in the stars and the beginning of the carbon biogeochemical cycle, we end this little big history. Our main purpose was to emphasize the importance of coal mining for the Asturias region studied here from an interdisciplinary point of view, and at the same time, highlighting the effects of this human activity in not only the natural environment but also in its social evolution. Looking at the history of the Nalón River from such a broad scope could hopefully offer its inhabitants – as well as other populations living in similar mining regions – the special opportunity to reflect upon their future by using a better understanding of their history in a more positive and integrative way.

Acknowledgements We thank Fred Spier for thorough review of the manuscript and Esther Quaedackers for edits and discussions about this little big history.

References Adaro Ruiz, L. (2003) Jovellanos y la Minería en Asturias. Fundación Foro Jovellanos, Gijón, pp. 475. Antrop, M. (2000) Background concepts for integrated landscape analysis. Agriculture, Ecosystems & Environment, (77), 17–28. Antrop, M. (2005) Why landscapes of the past are important for the future. Landscape and Urban Planning, (70), 21–34. AA.VV. 1990. Historia de Asturias. Editorial Prensa Asturiana y Caja de Ahorros de Asturias, 4 tomos, Oviedo. AA.VV. 2007. Prehistoria de Asturias. Un legado artístico único en el mundo. Editorial Prensa Asturiana, Oviedo. Bermúdez de Castro, J. M. (2002) El chico de la Gran Dolina. En los orígenes de lo humano. Barcelona, Crítica. Dopico, E., Linde, A. R., and Garcia-Vazquez, E. (2009) Traditional and modern practices of soil fertilization: Effects of cadmium pollution in river ecosystems. Human Ecology (37), 235–240. Fernández García, A. (1996) La calidad ambiental como premisa del desarrollo urbano: propuestas y actuaciones en la Cuenca del Nalón (Asturias). Ería: Revista cuatrimestral de geografía (41), 249–258. Fernández García, F. (2004) La explicación del paisaje a través de la imagen. Ería (63), 115–119. 318

The little big history of the Nalón River

García-Moreno, O., Aguirre-Palafox, L. E., Álvarez, W., and Hawley, W. (2017) A little big history of Iberian Gold: How Earth processes concentrated the precious metal that played a critical role in the history of Spain and Portugal. Journal of Big History (1), 40–58. Jiménez-Sánchez, M., González-Álvarez, I., Domínguez-Cuesta, M. J., and Requejo Pagés, O. (2011) Enhancing rescue-archaeology using geomorphological approaches: Archaeological sites in Paredes (Asturias, NW Spain). Geomorphology (132), 99–110. Jiménez-Sánchez, M., Rodríguez-Rodríguez, L., García-Ruiz, J. M., Domínguez-Cuesta, M. J., Farias, P., Valero-Garcés, B., Moreno, A., Rico, M., and Valcárcel, M. (2013) A review of glacial geomorphology and chronology in northern Spain: Timing and regional variability during the last glacial cycle. Geomorphology (196), 50–64. Massey, D. (1999) Philosophy and politics of spatiality: Some considerations. The ­HettnerLecture in Human Geography. Geographische Zeitschrift (87), 1–12. McDonnell, M. J. and Pickett, S. T. A. (1990) Ecosystem structure and function along ­urban-rural gradients: An unexploited opportunity for ecology. Ecology (71), 1232–1237. van der Plicht, J., van Geel, B., Bohncke, S. J. P., Bos, J. A. A., Blaauw, M., Speranza, A. O. M., Muscheler, R., and Björck, S., (2004) Early Holocene solar forcing of climate change in Europe. Journal of Quaternary Science (19), 263–269. World Resources Institute. http://www.wri.org/blog/2014/05/history-carbon-­d ioxideemissions. Recovered on May 23, 2017.

319

14 SKETCH OF A LITTLE BIG HISTORY OF PRIVATE E.E. BENJAMIN AND THE GREAT WAR Craig Benjamin Introduction In 1980, while helping my grandfather Eric Esmond Benjamin move from his home in Queensland, Australia, into the home of his oldest son Keith, I discovered in a drawer under a bed in his guest room a plastic bag containing five old battered and stained diaries, collectively dated from March 7th, 1917 to January 14th, 1920.When I showed these to Eric, he told me he had almost forgotten about them, but they were indeed a personal record of his experiences as a soldier in the First World War, and of living and working in Britain in the 12 months that followed its conclusion in November 1918. I was vaguely aware that Eric had served on the Western Front during the last years of the war as a member of the Australian Imperial Forces (AIFs), but these handwritten diaries of his experiences were an unexpected find and a treasure trove for a future historian. For the past several years, I have thought about using these diaries, along with other sources including his surviving letters, to write a book that considers Eric’s experiences from a big history perspective. The publication of The Routledge Companion to Big History has provided me with just such an opportunity to write a brief ‘Sketch for a Little Big History of E.E. Benjamin and the Great War’. A little big history of any individual’s experiences in war could potentially incorporate many aspects of the big history narrative, including earth’s physical, chemical, and biological history, human migration and settlement, industrialization, technological innovation, European imperialism, and the continuing neo-colonial ties between settler colonies like Australia and modern European nations. Big history can be an effective tool for examining events and individual experiences that seem microscopic on the scale of the cosmos, showing how these connect to larger patterns and processes. In this chapter, I can offer only a brief sketch of Eric’s experiences and indicate places where a big history perspective would provide useful context in a longer book-length narrative.

In the beginning Every little big history has the same beginning, which is the Big Bang that created space and time 13.82 billion years ago. Some 200 million years later the forces of 320

Private E.E. Benjamin & the Great War

gravity worked on vast nebulae of hydrogen and helium to create the first stars and galaxies.The death of the largest of these first generation stars, sometimes in supernovae, created all the more complex elements that exist today in the universe, including many that would play a crucial role in the course of the First World War. One example is the element of Nitrogen – N7 – cooked up in the death of large stars billions of years ago but only discovered and isolated in 1772 by Scottish scientist Daniel ­Rutherford. A little more than a century later, German chemist Fritz Haber was able to use high pressure and a catalyst to synthesize nitrogen with hydrogen to make, amongst other products, weaponized gas. More about this has been discussed below. All little big histories must then account for the formation of the Sun and the planets and other objects orbiting it in our Solar System, with a particular focus on the forces that shaped the Earth, including a late heavy bombardment by iron-rich meteors that brought that element so crucial to modern war to the surface of the planet. A little big history of Eric Benjamin should next consider the tectonic forces that shifted continents across the surface of the planet, which by the early twentieth century had left Australia a long way from Europe. A geological perspective would also consider the forces of soil deposition, glaciation and erosion that created the particular geographical context of the Western Front, which during the Great War stretched for many hundreds of miles along the border between modern Belgium and France. The next chapter in the little big history of Eric would unfold the history of modern Australia, including both the penal servitude and free migration that brought millions of Europeans to the Great Southern Land, where they displaced indigenous populations of aboriginal peoples who had lived on the arid continent for 60,000 years. This chapter would include the circumstances behind the migration of Eric’s own ancestors from Europe, and the success Eric’s parents enjoyed after settling in Brisbane in Southeastern Queensland. Eric’s father, Herbert Asher Benjamin (1867–1947), was a senior civil servant employed by the Australian Post Office. Herbert also went to war in the late-nineteenth century, serving in the Boar War as an officer in the 5th Queensland Imperial Bushmen. Our family still possesses the magnificent Wilkinson stainless-steel sword that he used in South Africa between 1899 and 1902. With his father as inspiration, it is hardly surprising that young Eric, while still at school, heeded the call of King and country and became determined to enlist in the AIFs, to play his part the great adventure he was certain was awaiting him in Europe. Eric’s decision suggests that this chapter would need to conclude with a consideration of why so many young Australian’s were convinced that it was their patriotic duty to travel halfway round the world and risk their lives for the imperial power that had sent most of their ancestors to Australia as shackled convicts a century or so earlier.

Enlistment in the AIF: March 1917 This sense of duty and adventure inspired Eric to try and enlist in the AIFs as early as 1915, even though he was too young to do so. According to family ‘lore’ after keeping quiet about his age, he passed every test before appearing for his medical exam. The examining officer happened to be his family’s doctor; he looked up at Eric, blinked twice and told him to go home and get back to college. Eric’s mother was furious when she heard the news, but his father, the distinguished Boer War veteran, was 321

Craig Benjamin

more sympathetic. Eric did eventually successfully enlist on March 7th, 1917. He was given the serial number 14358, and classified as ‘Driver, Army Service Corps, ­Australian Imperial Force’. I have copies of all of his war records. Due to, as Eric would have seen it, the unhappy circumstance of his youth, he had already missed out on nearly three years of the Great War, although tens of thousands of other young Australian men had not been so unlucky, and were resting in graves in places like Gallipoli and Flanders. In his letters, it is obvious that Eric was excited to be finally getting out of sleepy old Brisbane and heading off to play his part in a great adventure. As a strong Christian, he also saw this as a moral obligation, as part of a crusade against the heathen Hun. On March 20th, 1917, Eric entered the military base at Enoggera, which his father Herbert Asher had been partly responsible for establishing in 1908, to begin his basic training. Several weeks later, on May 14th, Eric detrained to Sydney to begin driver training in Liverpool and notes in his diary that he was proud to walk up and down the parade at Manly Beach on June 4th on a day off.

To the Motherland: August to September 1917 At 3.30 pm on August 2nd, 1917, Eric Benjamin, aged 21, left Sydney on the troopship S.S. Miltiades.The Miltiades was a 7817-gross ton, 504 feet long ship with clipper bows, two masts, two funnels, and twin screws. She had been completed on October 1903 by Alex Stephen and Sons of Glasgow, for the Aberdeen Line, a product of the industrialization of the Glasgow shipyards that had occurred in the mid-nineteenth century. In 1915, she had been converted to trooping duties, and only resumed commercial service in 1920. After picking up New Zealand troops in Wellington, Eric and thousands of fellow ANZACs spent a month crossing the Pacific in convoy, before entering the Panama Canal on the 1st of September. The canal had only been opened 3 years earlier, after 23 years of difficult and deadly construction, and Eric was overwhelmed by its scale and dimensions, reminding us that, while those of us living in the twenty-first century have become somewhat inured to the staggering engineering and technological achievements of our species, just a century ago such achievements were celebrated as something quite extraordinary. Eric later described his passage through the Panama Canal in great detail in a letter to his parents, beginning with this sentence: ‘Saturday September 1st: Today we saw the result of the greatest engineering feat in the history of mankind; how one continent was divided from another’.1 With the menacing Atlantic crossing ahead of them, the ANZAC troopships remained in Colon Harbor for a week before sailing north up the east coast of North America to Halifax, Nova Scotia.The troops were allowed ashore in Limon Bay for a swim, which Eric enjoyed on September 8th, his 22nd birthday.The convoy of troopships, now escorted by a US cruiser, made slow going through the rough seas off the East coast of North America, arriving at Halifax in the evening of September 18th. Eric and his companions had now been on board the Miltiades for seven weeks with that single swim in Limon Bay their only shore leave. Eric wrote that the men were decidedly restless, and that three Australian soldiers had decided to desert in Halifax. What was now a large and important convoy of troopships and war ships departed Halifax Harbor on September 21st. All the ships were heavily armed, with new guns 322

Private E.E. Benjamin & the Great War

having been added even to the troopships in Halifax. Eric was relieved to note that the convoy also included the USS Charleston, a St. Louis-class Cruiser built at Newport News Virginia in 1902, armed with 14 guns capable of firing 6″ shells, and 18 firing 3″ shells; and HMS Carnarvon, a Devonshire-class armored cruiser that had taken part in the Battle of the Falkland Islands before being assigned to the North America Station in 1915. A week after setting out, on September 30th, the 8760 ANZAC troops on board the convoy of 13 troopships were joined by no less than 10 additional British and US destroyers, with a further three joining the following day, essentially providing one armed escort vessel for each of the troopships. Eric noted that the cruisers ‘never stop in one place very long, but keep on moving everywhere’, presumably on the lookout for the deadliest enemy that Atlantic possessed in 1917, German submarines.2 With its speed determined by the pace of the slowest vessel, the convoy steamed slowly across the Atlantic. Eric counted at least 10 lookouts on his ship alone, peering out for the wake of submarines. As the convoy steamed closer to the British coast the men were warned that they were, as Eric puts it, ‘in an extremely dangerous place, and we were all told to sleep in our clothes with our lifebelts on’.3 One night they were awakened by the pounding of guns, and all raced on board to see something large and dark-hulled pass close by, presumably a U-Boat. The Germans navy had been at forefront of submarine design since the nineteenth century, launching prototypes as early as 1850. In 1906 the German navy commissioned the Germaniawerft Company to construct a fleet of Karp-class submarines, but by the start of First World War, the Germans had only 24 submarines of various classes available, while the British and French navies between them had considerably more.The Germans quickly intensified their production of submarines however, and within a year had added at least another 15 to their fleet.4 Throughout the war, U-Boats sustained the bulk of German naval activity. Despite several notable successes in sinking enemy war ships, the U-Boats were most effective against merchant shipping. After a hiatus during 1916, the Germans had resumed their policy of unrestricted U-Boat warfare early in 1917. The Kaiser, urged on by military staff and facing an increasingly grim situation in Germany because of the allied blockade of German ports, signed the policy in January 1917 in the hope that it would help end the war quickly in Germany’s favor. The campaign was initially a tremendous success, with hundreds of thousands of tons of allied shipping sunk between February and April for the loss of only nine submarines. By the end of 1917, U-Boats had sunk a total of 2,439 ships.5 Some historians estimate that Britain’s supplies of wheat shrank to reserves of just six weeks because of these devastating losses.6 It was in response to the effectiveness of U-Boat warfare that the British instituted the sort of comprehensive convoy system that Eric and his colleagues now participated in as they steamed across the Atlantic in September 2017. Although they were still sinking ships at an alarming rate during the summer and fall, losses never approached those of the spring of 1917, and the Germans were also losing between five and ten U-Boats per month. By the end of the war, of the 360 German subs that had been built, 178 had been destroyed; but the U-Boat fleet had destroyed 11 million tons of shipping.7 On October 2nd the great convoy of troopships and warships steamed up the Firth of Clyde in Scotland, right past Alex Stephen’s and Sons shipyard where Eric’s 323

Craig Benjamin

ship the Miltiades had been constructed 14 years earlier. Eric noted in his diary the ‘enormous amounts of ship building going on; and torpedo boats and tanks being built by the dozen’.8 Eric was almost immediately put on a train after disembarking, and by next day had traveled almost the length of Britain to arrive at the No. 4 Park House Camp near the village of Tidworth, on a very cold day. Tidworth, on the eastern edge of the Salisbury Plain in Wiltshire, is listed in the Norman’s Doomsday Book, the great survey of much of England and Wales completed in 1068 under orders of William the Conqueror. In the nearly thousand years that had followed, Tidworth had enjoyed a long history as a military garrison town and continues to function as a major base for the British Army today. On October 9th, Eric was transferred from Driver to ‘Signaler Class’ in the artillery, where, for obvious and ominous reasons, severe shortages existed and fit, strong men with good eye-site who could run fast were sorely needed. After a medical examination, Eric marched 15 miles to a specialized artillery signalers camp near Heytesbury, where training in signaling, and gas instruction with mask and helmet, was intense. By coincidence, the great British war poet Siegfried Sassoon spent the latter part of his life in Heytesbury. Eric spent the next three weeks of his young life mastering the signals that would be required for his new role, using Lucas Lamps, one of the many inventions that appeared during the Industrial Revolution in Britain.The Joseph Lucas Ltd Company had been founded by an unemployed father of six in the 1850s, initially to manufacture a variety of pressed metal goods, and later to invent and perfect new lamps for shipping. In 1879, Harry Lucas designed a new type of oil lamp, initially for use on bicycles and automobiles, which was advertised as the King of the Road, a lamp that ‘will not blow out in the toughest gale’.9 By the end of World War One the Lucas Company had expanded to manufacture a wide range of automotive parts and was the principal supplier of such materials to BSA, Norton and Triumph. As Eric continued practicing signals throughout a chilly October, on the 20th, just across the English Channel the British were unleashing the latest technological innovation to be applied to warfare, an attack at Cambrai of 381 tanks. These crude, wedge-shaped armored vehicles armed with small caliber machineguns were celebrated in the press as a ‘uniquely British solution to the deadlock’, and although the initial assault ultimately made little difference to the stalemate, the events at Cambrai marked a decisive shift in strategy, the adoption of a ‘save the men and use the machines’ approach that would have profound ramifications for all future conflicts. In many of the military operations that Eric would participate in on the other side of the English Channel, tanks were destined to play a crucial role, a fact he often noted in his 1918 diaries. Eric spent the next couple of months finishing his training, as a bitter winter pounded the south of England and the troops on the Western Front, many of whom were now enduring their fourth and most miserable winter of the war, in more or less the same position they had been in four years earlier. As bad as conditions were in the winter of 2017/2018 however, they were not quite as severe as the previous winter, when many soldiers had lost fingers and toes through frostbite. On ­Wednesday ­January 16th, 1918, Eric Benjamin detrained for Southampton and boarded the steamer Caesarea. Early next morning he set foot for the first time on the soil of France, disembarking at Le Havre. 324

Private E.E. Benjamin & the Great War

Western Front: January and February 1918 After collecting his trench kit (three blankets, waterproof combination cape and groundsheet, soles for boots, waterproof dubbin, steel helmet and gas mask), Eric took up his official duties as a Signaler in the 6th Army Corps, Mobile, and was transported to the Corps’ HQ at Baillieul. En route Eric bumped into two colleagues from Brisbane; indeed this was something of a regular occurrence for Eric during the war, and it is quite extraordinary how many acquaintances he encountered in Britain and France. Eric’s initial impression of the front was mud! Despite the fact that mud was ubiquitous and unavoidable, the officers expected his boots to be shiny each morning. Actually, more than just his boots. Eric describes the situation thus: … the bandolier must be shiny, the brass buckles clean, and the brass buckle and stud on the SBR Respirator must be shiny … that is the way to win the war it seems – it is time that this stupid nonsense of showing was cut out and more attention given to increasing the efficiency of the troops. The men resent this extremely, more especially as they have had some hard fighting lately.10 Geology made the long line of battlefields that constituted the Western Front particularly susceptible to the formation of mud. The Front passed through a variety of landscapes, meandering across the countryside from the sand dunes and flat, reclaimed sea level land on the Belgian coast in the north, to the mountain peaks 4,500 feet above sea level in the Vosges mountain range at their southern end. The geology of the Western Front battlefields included sand, clay, chalk and rock, while the terrain included rivers, canals, valleys and cliffs, ridges and mountains, plains, forests and swamps. These geological and geographical conditions inevitably influenced military strategy, tactics, and the development of new weaponry and fighting techniques in the battles that were fought along the Western Front over a four year period.11 The Somme region, for example, consists of a rolling chalk upland, while Flanders is a clay plain composed of Paleogene clays with occasional sand units. Within both these areas, the occurrence of low (500 ft) hills and ridges provided the focus for many of the most famous attritional battles of the Western Front, as each side fought to gain the strategic advantage provided by the high ground. These same geological constraints also impacted trench construction, water supply and the efficacy of tunneling; and also the daily life experiences of soldiers based in the Somme-Flanders area, where the poorly drained soils were particularly susceptible to creating mud.12 In early February, Eric moved up to the front line with his artillery, below the Messines Ridge. On February 9th, after spending a sleepless first night out in a dugout the previous evening, the guns went into action and Eric faced mortal danger for the first time in his life as he signaled the outcome of the firing of his artillery by telephone. The long sloping ridge of Messines had been the focal point of bitter fighting in 1916 and was captured and lost several times by both sides. Next evening the assault continued, and Eric described it thus: ‘Tonight at 10.00 pm a terrific barrage was opened up by all guns in our vicinity – Fritz immediately sent up all kinds of star shells and made a grand fireworks display’.13 Eric assumed that some kind of infantry raid must have been imminent, and the German star shells were an attempt 325

Craig Benjamin

by the enemy to illuminate the battlefield. Later that evening they were informed by an officer that a highly successful infantry raid had indeed taken place that night. Somehow a small newspaper clipping concerning this event found its way into the back pages of Eric’s first 1918 diary. Faded and fragile, it tells its story: WELL DONE AUSTRALIANS! BAYONET FIGHT IN HUNS SECOND LINE! Haig to Australians: ‘Congratulations on successful raid carried out against enemy trenches west of Warneton on night of 10th. From FM Cutlack, with the Australians in France Wednesday Victorians last night made a highly successful raid on the enemy trenches, east of Messines. At 10.00 pm a heavy barrage was put down along a considerable length of front, combined with heavy counter-battery fire against enemy guns at the same time as a bombardment of enemy headquarters and dugouts behind Warneton. The attacking force entered the enemy line over a front of 400 yards and encountered a considerable garrison which they overcame by bomb and bayonet fighting. The attack was pursued against large numbers of a further garrison in the second line which the Australians entered with magnificent bayonet fighting. They say this is the first occasion for a long time on which the enemy met them with bayonet, but the Australians completely outfought him. Trenches and dugout behind the 2nd line were entered and many dugouts destroyed. The raiders withdrew after half an hour, having killed in infantry fighting alone 90 Germans, and brought back 31 prisoners, 3 machine guns, and one light trench mortar. The enemy counterattack was feeble and consisted mainly of an attempt to overtake and outflank our rear guards during the retirement.14

Western Front: March – May 2018 Eric spent the rest of February on various camp duties, including guard duty, stable cleaning and feed, and signaling. On February 21st, he was moved to a new position on the Kemmell Front, for further artillery signaling duty and to prepare camouflaged dugouts for the signalers in a region pock marked by an enormous number of old shell holes. On the 28th they were moved back to their old position on the Messines Front, where they laid out three telephone lines. From March 1st, Eric’s battery was continuously in action, again sending hundreds of rounds crashing down upon the gas works at Warneton, just over the Messines ridge, in support of further Australian infantry raids. Eric describes the firing of the guns late in the night of March 3rd: Bang, bang, bang everywhere … Bang, bang, bang everywhere – our Major blows his whistle at 11.40 pm and off we go – bang, bang, bang with lightening like flashes and a tremendous row as our messengers of death rush through that 326

Private E.E. Benjamin & the Great War

moisture laden air like so many express trains. The earth seems to be vomiting fire and the reports from the guns come in a tremendous long roll.15 The powerful guns Eric’s unit was using were British 60 pounders, 168 inches long and weighing close to 10,000 pounds. They had a muzzle velocity of over 2000 feet per second, a range of 15,000 yards, and fired a shell five inches long and 60 pounds in weight.16 They left a crater often a 100 yards wide and 20 feet deep! The development of these high-powered industrialized weapons would be another subject worthy of deeper exploration in a book-length little big history of Private Benjamin. Once the firing was all over and the infantry and artillery men had stood down, the signalers had to go out in the dark and retrieve the cable, which Eric described as ‘very difficult work especially over barbed wire entanglements’.17 On that same day, March 3rd, 1918, the new Bolshevik government in Russia signed the Treaty of Brest-Litovsk with the German government, which essentially took Russian forces out of the war and freed up some 70 German divisions from the Eastern front. After relocating these to Belgium and France as quickly as possible, the German high command could now focus all its attention on the Western Front. Even an individual soldier such as Eric, who because of his job as artillery signaler was often perilously close to enemy lines, could sense something was in the air. On Monday March 11th, Eric accompanied two officers to the forward Observation Post, from where he could see the Germans’ front lines and pill boxes: Plenty of movement by Hun – saw some Hun soldier with a lot of medals walking about – also saw other things of interest … Hun started shelling perilously close to our battery but thank heavens he never got closer than about 300–350 yards – pieces of material and dirt were flying around us … when I went to bed – feeling a little nervy.18 He was back on duty from 1.00 to 5.00 am on the 12th, where he encountered gas that resulted in vomiting. Later that day he noted: ‘a good bit of gas about causing irritation to the nose and violent sneezing – had to use gas mask for about an hour’.19 The Germans had developed weaponized gas first and were the first to use it, at Ypres in 1915 with devastating effect. The British government responded quickly by dispatching gas masks to the allied troops, and then by developing their own weaponized gas later that same year. Two types of gas were in regular use on the front: chlorine, phosgene and prussic acid gasses were designed to kill; and mustard gas which was intended to permanently disable. Ironically, the development of weaponized gas was a by-product of attempts to improve the productivity of agricultural land in the early twentieth century. Since the beginning of the agricultural revolution 10,000 years earlier, renewing the fertility of the soil had meant taking it temporarily out of cultivation by fallowing, or, following the secondary products revolution 6,000 years ago, enriching it with animal or human wastes. But stores of natural fertilizers were limited, and by 1900 even the rich South American deposits of guano (the waste products of birds), discovered in the early nineteenth century, had largely been worked out. A major breakthrough came in 1909 when a German chemist, Fritz Haber, discovered how to directly react 327

Craig Benjamin

nitrogen gas (which is quite inert and does not react well with other chemicals) with hydrogen gas to create ammonia. Now, in principle, there was no limit to the capacity of industrialized societies to fertilize their soils using nitrates made from artificially synthesized ammonia. Haber’s invention did more than anything else to expand food supplies in the twentieth century, which led to a dramatic increase in human populations. Haber was awarded the Nobel Prize in Chemistry in 1918 (actually awarded the following year) for his discovery. But ironically, during the Great War, Fritz Haber’s methods for synthesizing ammonia also made it possible to produce huge quantities of explosives, and Haber himself helped design some of the poison gasses used to such devastating effect on the major battlefields of Europe. Weaponized gas ultimately accounted for some 800,000 casualties between 1915 and 1918, in what was essentially a war within a war – a war between chemists! Eric and his comrades continued to experience desultory gas shelling for the next few days until, in Eric’s words, on the night of March 17th, all hell was let loose and the Hun sent over hundreds and hundreds of gas shell – they came into our position at 10.15 and kept up the awful bombardment until 11.30 pm – shells were falling around our dugout one eventually falling in the dugout adjoining our own wounded one signaler and severely gassed another.20 By dawn, 37 men had been badly gassed, and by dinner time Eric’s eyes were sore. By 3.00 am, he was unable to see out of his eyes, so he reported to a first aid post where his eyes were dressed before he was removed by ambulance to another aid post where ‘the place is full of gas patients’.21 Eric’s injuries, including eye and lung damage and a hacking cough, kept him away from the front for the next 12 weeks, several of which he spent at a large convalescence camp at Boulogne. During this three-month period the Germans used their new post-Brest-Litovsk numerical superiority to tear a tremendous hole in the allied line. By nightfall on March 21st during this Spring Offensive they had advanced some 40 miles into the Allied lines, the greatest single gain of the war thus far. By the 23rd the German advance threatened to engulf the forward hospital where Eric was recuperating, so the entire hospital was evacuated by train while German aircraft attempted to bomb it. As the Germans pushed forward, Eric found himself in continuous retreat, from Messines to Baillieu to Boulogne, suffering from blistered eyes, a bloody cough, and heart palpitations. On April 1st, he settled into a convalescent camp overlooking the Channel and was still there on April 9th when the Germans overran his old forward observations posts at both Messines and Baillieu. By the 30th the Germans had reached the River Marne, and Paris, less than 50 miles away, seemed in imminent danger of falling, but there the assault stalled. On the 30th of May 1918, Lieutenant-General Sir John Monash was appointed to command the Australian troops on the Western front, which had previously been under the command of British generals. Monash was a daring and dashing commander who, like the majority of Australian officers, preferred to lead by example. In his later book, The Australian Victories in France in 1918, Monash likened the preparation of battle plans to an orchestral composition, in that each instrument needed 328

Private E.E. Benjamin & the Great War

to be allotted its own part to ensure a harmonious and coordinated performance. As  Monash prepared to put new offensive compositions into play, Eric, who had gradually recovered throughout the month of May, was declared fit and able for duty on the 30th and ordered back to the front. It took Eric seven days to rejoin his unit, traveling by train, lorry, bus, automobile, cart, horseback, and on foot, essentially every mode of land transport then available to humans! The old unit HQ at Baillieu was now well and truly in enemy hands and was relocated at Morlancourt near Amiens. While Eric had been recovering, the Germans had pushed three huge bulges in the allied lines, but they had also stretched their men, ammunition and supplies to the limit in doing so. While allied commanders considered the best way of launching a counter offensive, and as American troops started arriving on the front in numbers, the war in the air reached its climax virtually right above Eric: Great aerial activity by Fritzy with enormous squadrons … Saw two airplanes fall, one of ours and one enemy … Saw aerial combat between three planes. Had the glasses on them but unable to determine their nationality. One of them suddenly burst into flames and began to fall headlong. Suddenly a burning mass fell from the plane, presumably the engine. The plane came down slowly and fell into our lines and I watched it burn to a cinder. It was a wonderful sight but an awful death for the occupants.22 The air activity Eric observed over the Western Front was probably the first time he and his colleagues had even seen aircraft flying. The first powered, controlled flight in Australia had only been made in 1910 by the visiting escapologist Harry Houdini, and the first Australian Flying Corps squadron had only gone into operation in 1916. Eric’s use of the word ‘wonderful’ should thus be taken literally, despite the grisly death of the pilot – to see machine’s flying in the air must have been a truly wondrous sight, even above the Western Front in May 1918! The Great War was thus the first major conflict in history that featured the largescale use of aircraft, which only came into military use at the start of the war. After three years of combat, experimentation and innovation, the planes that took to the air on both sides during the German Spring Offensive of 1918 were well-equipped with machine guns and experienced pilots, but the losses of aircraft throughout the spring were particularly heavy, due mostly to improved anti-aircraft fire. The month of April 1918 began with the consolidation of separate British air divisions into the new Royal Air Force, the first independent air military force not subordinate to its national navy or army. By the day the great German ace Manfred von Richthofen, the so-called ‘Red Baron’, was killed on April 21st, the allies had largely regained superiority in the air.

Western Front: June and July 1918 Australian troops on the ground were reinforced in June by 1000 infantry from the newly arrived American 33rd Division. The soldiers in this division had been members of the National State Guard of Illinois that had been activated in July 1917, 329

Craig Benjamin

trained at Camp Logan in Texas, and then sent to France. The symbolic date of July 4th was chosen by the Australian and British high command as the first day on which Americans and Australians would fight side by side. Eric had been co-opted into the infantry for this particular attack near Hamel, and after sheltering quietly in a forward trench with his comrades, he went over the top at 3.10 am and charged for a German trench. Germans in forward dugouts and observation posts quickly retreated back about 400 yards to a rear position, allowing the Australians and Americans to take possession of the German forward positions in 93 minutes, just three minutes longer than the meticulous Australian commander General Sir john Monash had estimated! The Germans shelled and bombed the occupied trenches throughout the day, but strong allied artillery prevented German infantry from attempting to recapture it. Enemy shelling continued throughout the day and night, but Eric and his colleagues held the trenches until 11.30 pm when reinforcements from the 60th Battalion arrived to take their place. Eric, along with two of his closest friends from Brisbane, Arthur Taylor and Sergeant Gridley, walked for six hours until finally tumbling onto their cots around 5.00 am. The officers, out of an unfathomable act of kindness, let the men sleep until noon! Eric sums up his 4th of July experiences thus: … the net results on our sector were a gain of 500 yards on a 1000 yards front with the capture of 61 prisoners and a number of machine guns – our casualties fairly light – 2 officers killed and 3 wounded – other ranks killed, wounded or missing about 50. After this experience in the infantry, Eric returned to his old job as a signaler for the artillery, and on July 25th helped send a deadly barrage into German lines in support of a daylight raid carried out by American troops near Ville-sur-Ancre. At one point his signal telephone line was broken, so, as Eric puts it, ‘I had to hop out in full view of Fritzy, but he must have been asleep because he took no notice of me thank Heavens’.23 Eric now worked closely with American forces through the rest of July and into August, because the Americans, with no artillery of their own, had to rely on Australian artillery support. July 28th, 1918, was a sad day for Eric. Two days earlier Eric had moved up to a new forward position at Villers Bretonneux. He was resting in an English trench when deadly German artillery fire killed two fellow signalers, Gridley and Linton. Arthur Linton’s body was uncovered by Eric and another comrade as they frantically dug into the ruined trench next to theirs. Two days later he was signaling for the ­artillery again, this time ‘in a rickety old mill’. On August 2nd, he was up all night on the telephone signaling from the forward Observation Post. His telephone line was disconnected by enemy shelling so he had to go out and repair it; and ‘after a very very long walk over trenches, barbed wire and mud found the damage and fixed it’.24 That night Eric had had enough, writing in his diary: Today marks the beginning of my second year of absence from dear old Australia. Time has flown by and many an old friend has gone west and still the abominable conflict rages on … I am one of many who will be more than thankful when that dim distant vision of Peace has become a reality … This day twelve months 330

Private E.E. Benjamin & the Great War

ago I sailed from Sydney into the unknown. May we be spared to spend our next August with our own folk over yonder.25

Western Front: August and September 2018 For the next few days the Allies prepared in secret to launch their counter attack to the German Spring Offensive of March 21st. Eric was only too aware of these preparations. On August 4th, which he noted in his diary ‘was the beginning of the fifth year of this cruel war’,26 Eric left the trenches and arrived at the wagon lines. On the way down he noticed an enormous amount of artillery being massed. On the night of the 7th Eric wrote that wagon lines moved late tonight to Fouilloy, close to the front line – bivouacked here for the night in open fields … everyone ready to move off at a moment’s notice – everything points to the imminent offensive being very quickly followed up by all arms – very large numbers of guns pulled in tonight, ours being right up in the support trenches.27 This was Eric’s experience of the eve of August 8th, a date that would ever afterwards be referred to as ‘the Black Day of the German army’. From Eric’s diary of the 8th: Punctually at 4.25 am the offensive opened. It seemed as if the heavens had opened up a tornado of fire – never had such a mass of artillery been concentrated in one place. 9.2s were alongside 8″ and 6″ and 18 pounders – on our left were Tommies and on our right Canadians – our own boys advanced very rapidly, and the whippets (tanks), armored cars and cavalry soon had the enemy in a state of confusion – we stood to at 5.00 am and moved off towards Warfusce where we bivouacked until about 2.00 pm – numerous prisoners were coming through – Germans were carrying in wounded (ours & theirs) on stretchers and sometimes without a guard … In the afternoon moved off again through Warfusce which was thick with traffic … town knocked to smithereens – high velocity shells playing on roads – bivouacked close to Bayouvillers – shelled out of our position and moved with lines nearer to village – guns in action all the time, we are continually being moved forward – late at night prisoners were still coming back – slept in the open tonight.28 The front of attack was 14 miles wide and was defended only by a skeleton ­German force. Following the intense artillery barrage, 456 tanks led the way, followed by 20,000 Canadian and Australian troops. Next day, August 9th, Eric wrote that ‘our boys made a wonderful advance and reached their objectives by 2.00 pm, a distance of 8 miles from the start’; but ‘On our left the Tommies could not get the C ­ hipilly Ridge; and the Canadians also had severe fighting’.29 The attacking soldiers had reached the old battlefields of the Somme battles of 1916, a wasteland or mud, rusted coils of wire, and crumbling trenches. So the pace of advance slowed, although German planes dropped enormous numbers of bombs on the allies. On August 10th, Eric noted: ‘Hun planes flying at a 331

Craig Benjamin

tremendous height dropped a load of bombs close by, one of which fell twenty yards from my dugout. Two Tommies and two Australians killed, several horses injured and one had to be shot’.30 But the offensive continued; on the 13th, Eric moved through Cayerise to Caise, passing enormous numbers of captured German munitions, although the enemy kept up a relentless and deadly bombardment of the front lines with artillery and aircraft. Despite this allied morale was high, as was camaraderie with fellow Commonwealth troops. Eric describes the Canadians as ‘a fine lot of fellows who fought over exceptionally difficult ground in advance’.31 Eric and the allied forces continued to move forward until, by September 5th, the Germans were back against the Hindenburg Line, their backs literally against the last line of defense, and German morale shattered. Eric moved constantly forward during the first two weeks of September through the blighted landscape: September 13th; moved forward again to a position between Flectin and Poeuilly. These two villages exist in name only now; there is not a stick of wood in the former, and all are covered in weeds. Four Hun bombers down.32 On September 18th the first assault on the Hindenburg Line was launched by 1,488 guns, followed up by infantry. Eric noted that large numbers of German prisoners came in, and they seemed very pleased to be captured: ‘Some of them strolled in on their own without an escort’!33 The Americans, with very limited experience, found themselves at the sharp end of much of the action in September, sometimes with tragic results. In the foggy early morning of Sunday September 29th, Eric noted ‘a fairly good advance was made by the inexperienced Yanks, assisted by our chaps, and the Hindenburg Line was penetrated’.34 But during the heavy bombardment that preceded the attack, the German troops had crawled deeper into their trenches, and the American troops, not realizing this, had run on in their enthusiasm without completely mopping up the trenches and dugouts that they had overrun.When Australian troops followed up the American advance two days later, they were sickened to discover large numbers of dead ­American troops hanging on barbed wire and machine-gunned in the back. Eric ascribed their deaths to ‘barbed wire and overeagerness on their part’.35 By the end of the war the US 33rd Division had suffered 6864 casualties, including 691 killed in action, and 6,173 wounded in action.36

Western Front: October 1918 On the first day of October Eric was detailed to accompany two officers on a reconnaissance of that portion of the Hindenburg Line now in Allied hands: We passed over the great Hindenburg Line with its rows upon rows of barbed wire, deep and wide trenches, concrete dugouts, and concrete machine gun posts; in fact everything that made it formidable was there. The country is bare and hilly giving beautiful fields for machine gun fire over the barbed wire … The capture of this Hun piece of work was really wonderful.37 332

Private E.E. Benjamin & the Great War

During October, Eric and his artillery unit pushed relentlessly forward, first to Levergies, then on the 7th to Ramicort where he saw ‘plenty of dead Tommies and Hun’.38 His artillery unit was under constant enemy bombardment during this forward offensive. October 4th: ‘Shelling, gas as well. One landed ten yards from our dugout’. October 5th: ‘Another lively day … I was out mending a line when two 5′9″s landed within thirty yards of me. No damage done’.39 October 6th: ‘Heavy shelling at intervals. Went out on O.P line tonight when the Hun again started to shell quite close to me. In fact, he seemed to chase me until I reached the battery’!40 On October 8th, Marchall Foch ordered General Haig to launch an all-out offensive. Eric played his part: At Zero Hour (5.00 am) we opened up a heavy barrage when the Yanks advanced. About 10.00 am we moved through the village of Montbrehain. The Hun was shelling the place as we came through at a gallop. I was on one of the wagons when a shell landed 100 yards from us. One horse was badly wounded and had to be shot.The leaders were unhooked and galloped off just before other shells fell – there were no casualties but we received a little gas which made our eyes sore. Later we came into action to the left of Brancourt … we did a fair amount of shooting at various targets.41 Inevitably, this constant push forward meant that Eric and his colleagues encountered a constant stream of refugees fleeing the front line. October 10th: ‘A lot of French civilians in Busigny (200 it was rumoured) who gave us a great welcome … their hatred of the Boche was intense’.42 October 11th: Many civilians from the bombarded areas came in today and presented a most pathetic sight – some were hatless, improperly dressed, ill shod, and many were in tears while some were hysterical. Some had baggage which they brought along in wheelbarrows.43 Next day ‘… one woman of about 30 flung her arms around me and cried ‘Bon Garcon’!44 From October 17th to 19th, Eric and his battery were again on action, supporting an infantry advance of Americans and British: I arrived at the guns at 7.00 am.Went forward on reconnaissance with the Major. Things were very warm plenty of gas about … Next day went forward again with the guns at 9.00 am and took up a forward position … I was sent away with a message and went back through Busigny through shell fire and delivered it. Returned to battery position and dug in for the night. Some shells came close during the night, one of which destroyed the wheel of an ammunition wagon. Raining at night and very cold.45 Eric didn’t know it yet, but this was to be his final active engagement of the war. On October 19th, he wrote: ‘Came down out of action this afternoon to our Wagon Lines in Busigny – raining again tonight – the enemy has withdrawn his guns a good 333

Craig Benjamin

distance’.46 Eric had fought his last battle. By the 27th, he was not feeling at all well. He bivouacked near the Hindenburg Line at Harmelas, but noted that at the end of the journey I was feeling very sick with a bad feverish head and a pain in the small of my back. I paraded to the doctor who took my pulse and temperature and immediately ordered me to hospital. A motor ambulance later called and took me to No. 58 CGS where I was put on a stretcher with three blankets47. Eric had caught the flu, actually the deadly influenza virus that was set to ravage the battlefields and so many other regions of the world during the remainder on 1918 and much of 1919. One measure of the devastation wrought by the virus is that of the 115,000 deaths United States forces incurred in the war, more than 50% were caused by the great influenza pandemic. No one knows for sure how many soldiers living in wet, cold, slimy trenches and dugouts died, but Eric was not to be amongst the victims. He was lucky – the influenza pandemic of 1918–1919 is today regarded as the most devastating epidemic in world history, infecting 20% of the world’s population and causing the deaths of somewhere between 20 and 40 million people. People were struck with illness on the street and died a rapid death. Even President Woodrow Wilson was suffering from the flu when he negotiated the Treaty of Versailles in 1919. Scientists today describe the conditions around the world in 1918–1919 as being ‘not so far removed from the Black Death in the era of bubonic plague in the Middle Ages’.48 The origins of the virus itself are still not clear to medical science, but it is thought to have originated in China as a result of a rare genetic shift of the influenza virus.The recombination of surface proteins within the virus created something new that no one had immunities to. Recently the virus was reconstructed from the tissue of a dead soldier and is currently being genetically characterized.49

Western Front: November 1918 Eric was moved gradually back from the front on a series of hospital trains until he arrived at the No. 1 Australia General Hospital at Rouen, where he slowly recovered. By the last day of October, he was ‘still in bed but feeling much better except for a pain at the base of the lungs’.50 By November 9th, he had improved to the extent that he was now classified as an ‘up-patient’ and was able to do odd jobs, including typing for the Major. November 11th dawned cross the battlefields, a cold but sunny day. At 10.15 am a message reached the headquarters of General Haig bearing the following news: ‘Hostilities will cease at 11.00 hours today, November 11th. Troops will stand fast on the line reached’. At that precise moment, the 11th hour of the 11th day of the 11th month 1918, a great silence descended on the battlefields of Europe. The news reached the No. 1 Australian Hospital at 12.30. Eric described the scene in his diary: We all went dotty. Sirens, whistles, church bells, guns, bands, in fact everything that could make a noise did so! We in hospital formed a procession with the allied 334

Private E.E. Benjamin & the Great War

flags in front (including an Aussie flag) and, led by the Australian 20th B ­ attalion Band, marched around the grounds … Instead of keeping in the grounds the boys marched downtown to Rouen, around the square, and home again … The French people were hilarious with joy. Some wept while others kept a straight and stern appearance … Kisses were being lavishly distributed also … It was hard to realize that after over four years of this awful strife hostilities are ceasing. We should all give thanks to heaven for this wonderful day, when Right triumphed over Might, and the greatest injustice the world has suffered was beginning to be righted.51 By the morning of November 30th, Eric was back in England and was transferred to a large hospital complex at Exeter, where he stayed until December 26th, convalescing with the aid of some very delightful young nurses with whom he enjoyed a ‘grand time under the mistletoe’ on Christmas Day!

1919 and Conclusion Eric remained in England for the next 12 months, promoted to the rank of ­ ­Temporary Sergeant, and employed as a typist by the team writing the six-volume O ­ fficial History of Australia in the War of 1914–1918, under the leadership of historian C.E.W. Bean. Eric also met and fell in love with a young Scottish woman who was the first female medical student admitted into the University of Edinburgh’s Medical School; but all these post-war experiences belong to another story! As this brief sketch hopefully suggests, taking a little big history approach to the extraordinary adventures of my grandfather 100 years ago might offer some new insights into the individual experience of the Great War. It also suggests that the field of Big History has much to contribute to the human part of the story of our cosmos, providing enhanced context and meaning to the extraordinary journey our species has taken since appearing on Earth some 250,000 years ago. This part of Eric’s journey ended thus: At 7.00 pm on November 29th, he sailed for Australia on board the S.S. Aeneas, and on January 1st 1920, he disembarked at Circular Quay in Sydney. Two weeks later, he was home with his parents in dear old Brisbane.

Notes 1 E. Benjamin, Private Correspondence, October 12th 1917. Copyright the Benjamin family. 2 Ibid. 3 Ibid. 4 P. Haythornthwaite, The World War One Sourcebook, London: Arms and Armour Press, 1993, p. 106. 5 Ibid., p. 107 6 J.H. Morrow, The Great War: An Imperial History, London: Routledge, 2005, p. 202. 7 Uboat.net: https://uboat.net/wwi/ 8 E. Benjamin, 1917 Diary. Copyright the Benjamin family. 9 See P.W. Card, Early Vehicle Lighting, London: Shire Publications, 1991. 10 E. Benjamin, 1918 ( Jan–June) Diary. Copyright the Benjamin family. 335

Craig Benjamin

11 See ‘P. Doyle (2014), ‘Geology and the war on the Western Front 1914–1918’, Geology Today 30, pp. 183–191. 12 See J. Gargani, O. Stab, I. Cojan and J. Brulhet. (2006), ‘Modelling the long-term fluvial erosion of the river Somme during the last million years’, in Terra Nova 18 (2): 118–129. 13 E. Benjamin, 1918 ( Jan–June) Diary. 14 Source and copyright unknown. 15 E. Benjamin, 1918 ( Jan–June) Diary. Copyright the Benjamin family. 16 Military factor.com: https://www.militaryfactory.com/armor/detail.asp?armor_id=787 17 E. Benjamin, 1918 ( Jan–June) Diary. 18 Ibid. 19 Ibid. 20 Ibid. 21 Ibid. 22 Ibid. 23 E. Benjamin, 1918 ( July–Dec) Diary. Copyright the Benjamin family. 24 Ibid. 25 Ibid. 26 Ibid. 27 Ibid. 28 Ibid. 29 Ibid. 30 Ibid. 31 Ibid. 32 Ibid. 33 Ibid. 34 Ibid. 35 Ibid. Oct 1st. 36 F.L. Huidekoper, The History of the 33rd Division A.E.F., Springfield, IL: Illinois State Historical Library, 1921. 37 Ibid. 38 Ibid. 39 Ibid. 40 Ibid. 41 Ibid. 42 Ibid. 43 Ibid. 44 Ibid. 45 Ibid. 46 Ibid. 47 Ibid. 48 The Influenza Pandemic of 1918: https://virus.stanford.edu/uda/ 49 Ibid. 50 Ibid. 51 Ibid.

336

PART IV

Teaching big history

15 THE BIG HISTORY PROJECT IN AUSTRALIA Tracy Sullivan

Big history is a modern scientific origin story, exploring 13.8 billion years of shared knowledge and challenging intellectual boundaries by proffering a framework to connect that knowledge. A by-product of integrating big history into a traditional secondary classroom is an amplification of this process, challenging the boundaries of traditional curriculum structures, the role of discipline trained teachers and approaches to knowledge in the classroom. One response to these challenges is to ask, ‘why is this necessary?’ To answer this question, we can look to the philosopher and educator John Dewey, who outlined his philosophy of education as, “the reconstruction of experience to add meaning and contribute to subsequent experience” (Dewey 1916:82). Based on Dewey’s philosophy, challenging the abovementioned boundaries is necessary to add meaning to students’ educational experiences in a way that contributes to their subsequent experience of the world they live in. For twenty-first century students, this means being able to function successfully as part of a globally connected society. Prior to the public launch of Tim Berners-Lee’s World Wide Web in 1990, information was sourced individually from books, CD-ROMS and physical archives; research, whether it be structured or out of personal interest, involved a physical visit to a location housing the relevant information; communication was predominately via telephone or possibly fax machine; global personal communication was slow via mail or expensive via telephone; school computers were housed in a single room and access limited; the workplace did not have the use of email, social networking or cloud-based file sharing. In this scenario, the organization of knowledge and school curriculum into disciplinary silos was logical. The time and effort required to locate and analyse information to investigate complex lines of inquiry in a school context was logistically improbable. Less than 30 years later the scenario in the majority of western countries includes individual networked devices for every student; career opportunities in interdisciplinary fields, such as bioengineering, environmental planning, international relations and nanotechnology; and daily social interactions negotiated through multifunction phones,Twitter and Facetime. In this scenario, traditional siloed curriculum structures don’t reflect how students experience knowledge in the 339

Tracy Sullivan

world beyond the classroom. One antidote to this misalignment is the implementation of interdisciplinary curriculums in school environments to demonstrate how deep understanding occurs when knowledge is connected. Big history is an example of such a curriculum. This chapter will present an overview of how big history is being incorporated into Australian secondary schools as an alternative to compartmentalized curriculum offerings. I will begin by providing definitions of knowledge and curriculum in the context of big history and outline a continuum of disciplinarity to structure discussion in the remainder of the chapter. For contextual clarity, this will be ­followed by a brief history of the development of the Big History Project (the online platform through which big history is currently delivered to Australian students) from its inception to its public release, discussing the challenges and inherent dangers in ­integrating an interdisciplinary course into a siloed curriculum structure. This account is based on my recollections as a member of the original Big History Project curriculum design team and leader of the Australian arm of the Big History Project from inception to 2017, along with informal interview testimony from Professor David Christian. This will lay the foundation for an examination of the pedagogical framework underpinning the delivery of big history as a vehicle to develop interdisciplinary thinking. I will conclude with a discussion of how transformative learning theory can be used to examine the potential for big history to deliver transformative learning experiences for students, and propose directions for further big history ­education research.

Defining key ideas: knowledge, curriculum and interdisciplinarity Before exploring big history in the context of Australian secondary schools, it is important to provide working definitions of knowledge, curriculum and interdisciplinarity.

Knowledge Big history is a framework to connect knowledge, but what is this thing called ‘knowledge’ big history claims to be connecting? There are many answers to this question evidenced by centuries of debate. In the context of how big history is implemented in Australian schools, this is important because a philosophy of knowledge will characterize subsequent learning experiences. Any formal learning experience is guided by two instructional questions – “what knowledge should students be taught? And what are the appropriate strategies to teach it?” (Smith & Lovat 1991:16). A brief examination of the philosophies of knowledge championed by Plato and Protagoras illustrates the linked nature of these questions. Adherents to Plato’s school of thought claim knowledge is fixed and unchanging, framing learning experiences as acts of retention, aimed at facilitating accurate transference and protection of knowledge. On the other hand, those who agree with Protagoras argue that knowledge is dynamic and constantly changing, framing learning experiences as acts of exploration and inquiry, aimed at understanding lived experience (Smith & Lovat 1990:56). If you are

340

The Big History Project in Australia

sympathetic to Plato’s philosophy of knowledge as fixed and unchanging, adopting a proactive instructional strategy of inquiry and exploration will not produce the ­desired outcome of accurate transference and protection of knowledge. Similarly, if you adhere to Protagoras’ theory of knowledge as dynamic and ever-changing, adopting an instructional strategy of factual retention will not produce the desired outcome of understanding lived experience. A discussion of the nuisances of varied philosophies of knowledge is a chapter unto itself, thus in this chapter, knowledge is referred to in the Protagorian sense. As dynamic and ever-changing, producing learning experiences that nurture an understanding of knowledge as evolving through human discovery, to understand the world around us.

Curriculum The two guiding instructional questions, “what knowledge should students be taught? And what are the appropriate strategies to teach it?” interact via curriculum. This is the mechanism for organizing knowledge and teaching strategies, for delivery to students. An examination of big history in Australian schools involves a discussion of three levels of curriculum-mandated curriculum, school curriculum and course curriculum. Using Smith and Lovat’s four aspects of curriculum – product, process, intention and reality (Smith & Lovat 1990:15) – the three levels of curriculum referred to in this chapter are: (1) Mandated curriculum Curriculum as the process of identifying ‘high status knowledge’ (Apple 1979:36) by curriculum authorities, who organize that knowledge into categories, with the intention of directing schools within their jurisdiction to deliver identified knowledge to students, to measure student learning and school performance. (2) School curriculum Curriculum as the process of schools determining how ‘high status knowledge’ is delivered in a school environment, with the intention of meeting knowledge and skills requirements prescribed by curriculum authorities, to facilitate the assessment and measurement of student learning. (3) Course curriculum Curriculum as the process of selecting, sequencing, organizing and structuring knowledge, resources and activities with the intention of directing student learning experiences in the classroom. In many cases the shape and structure of course curriculum is determined by the requirements of both mandated and school curriculum. Big History Project is an example of course curriculum. These classifications demonstrate the complex nature of curriculum and its varied contexts. In the case of big history in Australian secondary schools, using these classifications helps facilitate a discussion of the tensions that arise when course curricula, such as big history, don’t directly align to the organization of knowledge prescribed in mandated curriculum and school curriculum.

341

Tracy Sullivan

Interdisciplinarity Big history as taught in Australian secondary schools is interdisciplinary, but what does that mean? And how do we make that determination? Approaches to the organization and integration of knowledge can be viewed along a four-point continuum, positioning discrete disciplines at one end in the form of disciplinarity, and the transcendence of disciplines at the other in the form of transdisciplinarity (­Godinho & Shrimpton 2008:3; Thompson Klien 2010:17–27), see Figure 15.1 (Sullivan 2013:58). At the first point and disciplinary end of the continuum knowledge is housed in silos. Questions are framed and examined using the methods and lens of a single discipline, tending to be narrow in scope. The second point on the continuum is multidisciplinarity. Here disciplines are juxtaposed so questions and problems can be explored through a variety of disciplinary lenses, while the methods and forms of analysis remain discipline specific.This results in multiple views of the same problem or question. The third point on the continuum is interdisciplinarity. Here the interaction and integration of disciplines becomes proactive (Thompson Klein 2010:18). Disciplines come together and link knowledge from their respective fields making it possible to explore questions that can’t be answered within the confines of the separate disciplines.1 If you think of a question or problem as an impressionist artwork, when you stand very close you view the painting through a single brush

Figure 15.1  Disciplinary continuum.

342

The Big History Project in Australia

stroke seeing its distinctive texture, direction and colour – this is the disciplinary lens. If you take a step back, you start viewing the painting through its multiple brush strokes.You still see the distinctiveness of each stroke but when you put them together you see something more complex with different shades and structure – this is multidisciplinarity. Then you take another step back and you no longer see the brush strokes but the results of their interaction and integration in a clear picture – this is interdisciplinarity. The fourth point on the continuum is transdisciplinarity. This is a rapidly evolving term with a range of interpretations. Thompson Klein identifies four ‘trend lines’ in the interpretation of transdisciplinarity. The first is transcending disciplines in an intellectual expedition uncovering a system to unify all knowledge. The second is ‘transgressing’ disciplinary boundaries to develop new intellectual frameworks to examine questions, e.g. cultural studies. The third is related to expanding theoretical paradigms that push worldviews beyond disciplines, e.g. sociobiology. The fourth is tied to solving problems framed by real-life dilemmas, not disciplines, and providing active solutions to those problems, e.g. product development (Thompson Klein 2010:18). This chapter focuses on how secondary students interact with knowledge presented to them in big history, not what defines big history as a field of study. One could argue that big history as a field of study is transdisciplinary, because it reflects each of the four trend lines to a degree. This is not true of the pedagogical structure underpinning how big history is taught at a secondary level. Big history at a secondary level requires students to connect knowledge to answer questions that can’t be answered via a single discipline. This makes it interdisciplinary. As a modern scientific origin story big history seeks to answer the questions – “Who are we? Where did we come from? And where are we going?”  These questions are beyond the knowledge confines of a single discipline, and even if viewed from a multitude of disciplinary perspectives, this won’t provide a holistic answer. Chemistry can provide a partial insight into these questions, as can biology, physics, history and anthropology, but the answers would be pieces of a much larger story. When these disciplinary insights are intertwined and framed through the narrative of big ­history an integrated whole emerges inclusive enough to explore these big questions in their entirety. For example, to answer the question ‘where did we come from?’ we need to draw on knowledge from cosmology, astrophysics, astronomy, history, mathematics and physics. This is just to understand the origins of our universe, let alone how we came to inhabit planet earth, or how we evolved into modern humans. It is the questions posed by big history as an origin story of our time that defines its interdisciplinary scope. From an instructional perspective, much of the knowledge student’s encounter in big history has been acquired in their discipline-based school ­curriculum subjects. They bring this knowledge with them to big history course curriculum and add knowledge unique to big history, such as the concept of thresholds of increasing complexity, discussed later in this chapter (Christian, Stokes Brown & Benjamin 2014:6). They then connect this knowledge and place it within big ­history narrative to deepen their understanding of who they are, where they came from, and where they are going.

343

Tracy Sullivan

A brief history: Big History Project and the road to implementation in Australian schools At the time this article was written, big history was predominately taught in ­Australian  secondary schools via the Big History Project, the first attempt to create a course ­curriculum to teach big history at high school level. As the popularity of big history grows the number of options for teaching big history in the classroom is growing with it. In 2018 the Big History Institute, Macquarie University will launch Big History School, a comprehensive online resource refining and expanding the work of Big History Project to meet the needs of a range of schooling levels, across multiple regions, in multiple languages (Big History Institute website 2016). My big history journey began eight years ago, in late 2010. I was asked to locate two Australian secondary schools to participate in an experimental curriculum design project, and attend a scoping workshop for that project. The aim was to convert the successful undergraduate big history course, developed and taught by Professor David Christian at Macquarie University for over 25 years, into a course for year 9 high school students. This would become the Big History Project. In 2004 the Historical Society of  Booth Bay Harbour in Maine held a conference including two Big History panels of which Professor David Christian was a participant. In attendance was a representative of the Teaching Company, an organization that reproduces and publishes college level courses on CD and DVD for lifelong learners titled, The Great Courses. Hearing Professor Christian speak and recognizing the intellectual power of big history, the Teaching Company asked Professor Christian to work with them to create a series of 42 lectures on big history to add to their Great Courses repertoire. After three years of development, Big History: The Big Bang, Life on Earth, and the Rise of Humanity (Christian 2008) was released in 2008. Later that year, in November, Professor Christian received a phone call from the office of Bill Gates. Mr Gates had watched Big History: The Big Bang, Life on Earth, and the Rise of Humanity and was interested in speaking further with Professor Christian about big history and its educational potential2. This meeting took place at a hotel in La Jolla, San Diego, and is when the idea of an online big history course for high school students around the world was first proposed (David Christian 2017).Within five years of this meeting the Big History Project would be taught to secondary students in Australia and the US. In 2009, a follow-up meeting was held at bgC3 (Bill Gates private office) in S­ eattle with Professor Christian, Professor Marnie Hughes-Warrington3, Bill Gates and Larry Cohen (bgC3 Managing Partner), and plans were put in place for the scoping of the Big History Project. The concept development team at Intentional ­Futures4 (at that time consisting of Michael Dix, Greg Amrofell and Ian Sands) was brought onboard to work with Professor Christian (lead on course and curriculum design) and a team of curriculum experts including Professor Bob Bain, University of ­Michigan and myself to make the idea of a high school big history course a reality. 2010 was a pivotal year in the history of the Big History Project. Course development began in earnest and included the abovementioned scoping workshop at Macquarie University. Here the Intentional Futures team, Professor David Christian, Professor Marnie Hughes-Warrington and myself worked on a blueprint for the 344

The Big History Project in Australia

Big History Project. It outlined 20 units of study mirroring Professor Christian’s first year undergraduate big history course, as an interdisciplinary exploration of the history of the universe from the Big Bang to the present day, and the future. By the end of 2010, Big History Project LLC was incorporated as a private entity funded by Bill Gates, and Macquarie University had licenced intellectual property in its big history courses to enable Big History Project course development. Later, in March 2011, Professor David Christian delivered his hugely successful TED talk on big history ‘A History of the World in 18 Minutes’ which to date has over seven and a half million views (TED 2011). Here it was publicly announced that school students around the world would be able to study Big History online. It would take 2 years from this point for the Big History Project to be publicly available. A crucial decision was made from the outset that classroom teachers needed to play a key role in the development and refinement of the Big History Project course curriculum. Success was predicated on working closely with practicing teachers to ensure successful translation of the Big History Project course curriculum into the classroom.This needed to work for all schools and all students.This decision led to an extensive two-year classroom pilot period of gathering feedback and making refinements before the course curriculum was made publicly available. To ensure the Big History Project would work for all schools and all students, it was important the co-creating schools represented a diverse student population. Two Australian schools and five US schools5 were recruited to work with developers on creating the first version of the Big History Project, based on the October workshop blueprint, for trialling in their classrooms beginning in 2011. This was called the Big History Project small pilot and would be expanded to 75 schools the following school year for the large pilot. Together these were the first schools, teachers and students in the world to study Big History at the high school level. The initial Big History Project course curriculum was vastly different from the resource today. It was a website of 20 units comprised primarily of text documents and video, requiring further modification to meet the needs of a range of students. Big history is about breaking down barriers and crossing divides in knowledge. The same can be said of the development process of the Big History Project. Often there is an ideological and practical divide between classroom teachers and academics. While not always the case there can be a level of resistance to engaging with one another that is ultimately to the detriment of both. This divide can also apply to the private sector. When innovative teaching comes together with evidence-based scholarship and private sector expertise the results can be incredible. As the Big History Project evolves over time, and other big history course curriculums enter the secondary education space, ensuring the integrity of scholarship defining big history ­remains intact as part of these collaborations is essential. This means sourcing feedback and contribution from Big History teachers and Big History scholars alike. A point of contention in the Big History Project development process, and one that remains today, was the amount of time required to complete the course, as opposed to the amount of time schools had available in their school curriculum to deliver the course. Secondary school curriculum is a product of mandated curriculum and is traditionally organized in discrete subjects, delivered by teachers trained in those subjects. School curriculum time (the amount of face-to-face classroom time between students 345

Tracy Sullivan

and teachers) is divided between these discrete subjects at the individual school level. Subject areas comprising knowledge and skills deemed more important for students to learn are given a greater proportion of this time. School curriculum time is finite and naturally those representing discrete subjects work to attain the maximum amount of school curriculum time possible. When this allocation process is confronted with an interdisciplinary course curriculum that doesn’t align with a discrete subject offering, as in this case big history, the limitations of siloised mandated and school curriculum collides head on with the practicalities of school curriculum timetabling. In the case of the Big History Project course, it was estimated that 200 hours of school curriculum time would be needed to complete the course. In the context of Australian mandated curriculum requirements, this was prohibitive for almost all ­Australian schools. To address this, problem content was categorized as ‘core’ or ‘optional’ allowing schools to implement the course in less than 200 hours. In Australian we are fortunate to have space in the school curriculum for elective subjects, and an Australian Curriculum that allows for some flexibility in how to deliver mandated curriculum content. This meant there were two options available for teachers and schools to ­implement big history. The easiest was in the elective space, the other required experimenting with teaching mandated curriculum content through big ­history. The first option was chosen for the small pilot, and the Big History Project course was taught as an elective subject.6 This allowed Australian small pilot schools to deliver big history content in a format that would connect and reinforce knowledge students encountered in mandated subject areas without being subject to the reporting requirements of a mandated curriculum. This gave teachers and schools the flexibility to trial a previously untested course. At this point in time the small pilot schools had 100 hours of school curriculum time for big history. After a year of classroom trials, teacher feedback led to a major overhaul of the Big History Project course curriculum. As a result, the 20 unit course was condensed into 10 units, and a larger selection of activity based resources was created. Organization of the content in these ten units is the foundation of the Big History Project course today (Big History Project LLC 2016). In September 2012, the Big History Project small pilot transitioned to the large pilot and the number of Australian schools grew from 2 to 25. This expansion allowed the course to be tested across a wider range of school settings. It also broadened the areas ­ nglish, of teaching expertise of those delivering the course to include teachers of E ­Geography, Language and Religion (the small pilot consisted only of History and Science teachers). Big history is intrinsically designed to break down barriers between knowledge.This becomes a daunting proposition for a teacher who has been trained in a single subject area. Big history weaves its narrative through such a vast number of different knowledge disciplines that every teacher is out of their subject specific comfort zone at some point.The concept of teacher as content expert needed to be replaced by one of facilitator and lead-learner (Big History Project LLC 2016:17).Teachers needed to join students on a journey of knowledge discovery. This proved a powerful (and sometimes difficult) psychological shift for the traditionally trained teacher. The timing of the Big History Project pilot program in Australia proved fortuitous as it coincided with the implementation of the new Australian Curriculum. In 2008 as part of Prime Minister Kevin Rudd’s education revolution, the development of an Australian Curriculum for kindergarten to year 12 was announced (Parliament of 346

The Big History Project in Australia

Australia, Department of Parliamentary Services 2008). Prior to this individual states and territories set their own curriculum and reporting standards. In 2009 the Shape of the Australian Curriculum paper was approved by all state and territory Ministers for Education. From then on, it constituted the mandatory content outline to be delivered and assessed in all Australian schools (Commonwealth of Australia 2009). Stage one of implementation began in 2011 with English, Maths, Science and History (Commonwealth of Australia 2009:15). Although the timing created opportunities to re-examine existing curricula, the process shone a bright light on the complexities and obstacles for schools wanting to implement an interdisciplinary course curriculum using a siloed mandated curriculum model. While the Big History Project pilot program found success in the elective space, the perennial issue was, and is, how to utilize the second implementation option available to Australian teachers, and transfer a niche elective course into the mainstream curriculum without sacrificing the integrity of its content or structure. In the Australian experience, these issues became apparent through the process of attaining endorsement from the New South Wales Education Standards Authority (NESA)7 for the delivery of the Big History Project course curriculum to year 9 and 10 students in NSW.8 Even in the realm of New South Wales elective subjects the parameters outlining what can and cannot be studied are stringent.Through the innovate work of the small pilot teachers at Narara Valley High school, Bernie Howitt and Natalie Karazinov, a model was developed that maintained the integrity of big history’s content and structure whilst meeting the NESA requirements, but this was not without its challenges. Being an elective… it got approval in NSW as a philosophy course, a story in itself and one to do with curriculum politics. (Lane 2012) It became clear working on the NESA endorsement that teachers wanted to apply an interdisciplinary lens, however the endorsement requirements, and their subject based experience, pushed them to look at big history through the lens of their own expertise. For example, big history is framed as an historical narrative with the word ‘History’ in the title.This tempted the History teacher to position it as a History course to justify the allocation of school curriculum time. Big history is also grounded in modern day science, so the Science teacher felt the course could be offered as a science course. Ultimately, because there was more school curriculum time available for a History elective, it was decided this would be the best positioning option.The danger in these types of decisions is that once this compromise has been made, the stage is set to encourage further compromises that recreate the siloed, discipline-based structure of mandated curriculum, jeopardizing the integrity of big history and its potential to transform student learning. This dilemma is not restricted to the Australian secondary school landscape. The current Big History Project website defines Big History as: Big History is a social studies course that spans 13.8 billion years. It weaves insights from many disciplines to form a single story that helps us better understand people, civilizations, and how we are connected to everything around us. (Big History Project LLC 2016) 347

Tracy Sullivan

In this case the integration of multiple disciplines is acknowledged, however the Big History Project course curriculum as a whole is positioned as a social studies course. The intent is to communicate to US schools and teachers that they can use social studies school curriculum time to teach big history. At first this may seem a valid approach to increasing uptake of big history in schools, but the inherent danger is the misrepresentation of big history as belonging to a disciplinary category which recreates the disciplinary compartmentalisation that big history seeks to move beyond. Big history is neither ‘big picture’ history nor ‘big picture’ science. It is interdisciplinary. While this may pose challenges for aligning to mandated curriculum, and integrating into school curriculum, it is not an unachievable proposition. Big History as a course curriculum is an opportunity to challenge the siloisation of knowledge, create innovative approaches to organizing school curriculum, and demonstrate how an interdisciplinary course can be implemented successfully in secondary schools. This requires individual teachers and schools to contemplate existing mandated and school curriculum structures and imagine ways to reinvent them on a case-by-case basis. In some instances, teachers and schools may still attempt to fit Big History into a disciplinary category. However, even if this is the outcome, this approach has undoubtedly sparked meaningful discussion around interdisciplinary curriculum and the benefits of big history course curriculum for students. Fortunately, in the ­Australian context, there are innovative and pioneering teachers who see the benefit in looking beyond school curriculum compartmentalization and implement Big History as the interdisciplinary course it is. These schools are teaching Big History as an interdisciplinary elective, but also aligning big history course curriculum content to the Australian Curriculum. This is a process of Identifying mandatory Australian Curriculum content that overlaps with the Big History Project course curriculum and teaching this content through Big History. An example of a compromise between these two options is Somerville House in Brisbane. Somerville House was one of the Australian Big History Project large pilot schools. They began teaching Big History as an interdisciplinary elective to one year 9 History class, team taught by a History teacher, Kira Sampson and a Science teacher, Alanna Johnston. Based on their success the course was later expanded for delivery across the whole year 10 cohort as an elective subject linked to Australian Curriculum content. Overall, it was a fantastic experience teaching this course; we are both delighted to be teaching it again this year. It is also interesting in the midst of new ­Australian Curriculum units being published and taught for the first time, to teach a course that is international in its execution and global in focus. The current Year 9 students who chose the course seem to have benefited measurably and immeasurably. ( Johnson & Sampson 2013) Looking above the school level, just as teachers need to see demonstrated success of a course curriculum like Big History in their classrooms to justify widespread adoption across the school, so too do education authorities need to see demonstrated success in their schools to justify widespread adoption across their jurisdiction. It is not enough for those of us who’ve seen the positive impact of Big History in secondary schools 348

The Big History Project in Australia

to lament that education authorities have not formalized interdisciplinary course offerings in mandated curriculum structures. We need to demonstrate those positive impacts, and clearly explain the pedagogy that underpins them.

How it works: pedagogical framework for teaching big history and interdisciplinary thinking Before examining the pedagogical framework underpinning the teaching of Big History in Australia, it is important to outline the organization and structure of the Big History Project course curriculum currently used to teach it. The Big History Project is an online blended learning platform comprising ten units of study reflecting the big history narrative of thresholds of increasing complexity (Big History Project LLC 2016): (1) What is Big History? (2) The Big Bang (3) Stars and Elements (4) Our Solar System (5) Life (6) Early Humans (7) Agriculture and Civilisation (8) Expansion and Interconnection (9) Acceleration (10) The Future Each unit contains a range of resources including: lectures, readings, timelines, activities and infographics, supported with marking guidelines and lesson sequences. A crucial factor in the successful teaching of the Big History Project course ­curriculum, and any secondary course curriculum in Big History, is maintaining the pedagogical narrative of increasing complexity. This is achieved by teaching all ten units in sequential order. The level of depth in covering each unit is flexible and can be adjusted at the discretion of individual teachers. This general conclusion is evidenced in a series of case studies commissioned by Big History Project LLC in 2014/2015, that found ‘higher fidelity implementations are strongly correlated to positive student outcomes, including long-term learning and retention’ (Big History Project LLC 2016:20).  The course takes a long view lens, requiring content selection decisions that prioritize adherence to the narrative over depth of specialist knowledge.The scale and interdisciplinary breadth of the course can be confronting for the ­discipline-trained teacher. The aim of big history course curriculum is not to replicate the deep content knowledge that students encounter in other subjects, but to teach them to make connections between that knowledge and provide a framework within which to make sense of new, interdisciplinary concepts and ideas. Big history as taught in Australian schools is grounded in three core concepts and three core skills (Big History Project LLC 2016).These reflect the foundation big history skills and concepts as designed and taught by Professor David Christian at Macquarie University for over 25 years. 349

Tracy Sullivan

Core big history concepts Modern scientific origin story This concept frames big history as a modern scientific explanation of our origins as a species, and the world we live in. This concept positions big history as seeking answers to big questions as old as humans – Who are we? Where did we come from? Where are we going? What makes big history unique as an origin story is its grounding in scientific evidence (Christian, Stokes Brown & Benjamin 2014:4).The concept of modern scientific origin story and the questions it poses define Big H ­ istory as interdisciplinary, as they cannot be answered using knowledge from a single discipline.

Collective learning This concept explains the process of human knowledge accumulation that underwrites this modern day scientific origin story. Collective learning is the capacity, most e­ fficiently mastered by Homo sapiens, to pass more information to the next generation than is lost by the one before (Christian 2005:146; Baker 2014:42; Christian, Stokes Brown & ­Benjamin 2014:80; Baker 2016:77).This concept acts as the intellectual driver for understanding how what we know about ourselves, and the universe, developed over time. For example, to understand the current model of the universe big history course curriculum not only looks at our understanding of the universe today, but how that understanding has changed over time via the accumulation of evidence as a product of human discovery, examining the theories of Ptolemy, to Copernicus, to Newton and Hubble.

Thresholds of increasing complexity The big history narrative is scaffolded by the third big history core concept: t­hresholds of increasing complexity, or points in time where conditions are just right and a new form of complexity appears that did not exist previously (Christian, Stokes Brown & Benjamin 2014:4).This concept is a powerful pedagogical tool that ­provides the content backbone for teaching big history course curriculum at the secondary level. It also functions as an intellectual life preserver for students and teachers feeling adrift in a sea of content covering 13.8 billion years of shared knowledge. Big history isn’t a simple retelling of a scientific story of change over time, and by proxy an exercise in engaging students in the passive recitation of events. Combining the three big history core concepts cultivates student understanding of  how, propelled by the capacity for collective learning, humans have developed and refined scientific theories over time, bringing us to an understanding of the world and universe today. “I enjoy big history because it gives me a sense of understanding. It brings together different opinions, beliefs and values which together helps me understand that things change. The main thing that I enjoy about Big History is that it changes as new evidence is discovered changing what we understand” – Elisha, Year 9 Big History student, New South Wales, Australia. (Sullivan 2014:325) This understanding is cultivated through the practising of the three big history core skills. 350

The Big History Project in Australia

Core big history skills Interdisciplinarity As a modern scientific origin story, big history responds to the questions: Who are we? Where did we come from? And where are we going? The scope of these questions challenges students to hone their skills in thinking across and between ­disciplines, and to create interdisciplinary questions of their own (see Figure 15.2). Rather than the room they’re sitting in at a certain time of the school day determining the disciplinary knowledge they need to draw on, in Big History students are posed broad questions of inquiry that challenge students themselves to identify and select the appropriate disciplinary knowledge relative to the question they are exploring.

Thinking across multiple scales This involves students learning to fluidly shift perspective from, and between, spatial scales as small as an atom to as large as the universe, and from temporal scales as short as the fraction of a second after the Big Bang to the 13.8 billion years of the universe’ existence. By examining events, concepts and objects on a range of different scales, students learn to deepen their understanding of trends and processes, identifying a vast array of new patterns and connections. For example, if students are studying the Black Death in a traditional History classroom this would be viewed, at the largest scale, as an isolated event. On the temporal scale of several decades and the spatial scale of the European landmass, students would see patterns relating to how the plague spread and impacted humans in affected areas. In big history course curriculum this event, which would not even be visible on a universal scale, would be viewed on a much larger scale, possibly on the scale of human population growth since the appearance of Homo sapiens. This would involve shifting perspective to the temporal scale of 200,000 years and the spatial scale of planet earth. By expanding the scale on which this event is examined, students gain a vastly different perspective of this event and its significance.

Claim testing Claim testing is the most explicit of the big history core skills and involves the use of four metacognition tools. These metacognition tools are intuition, authority, logic and evidence (Big History Connecting Knowledge 2015). Students explore their own thinking by repeatedly asking which of these four types of justifications they use when deciding to accept or reject a claim of knowledge. This toolkit furnishes students with a vocabulary to explain their decision making processes to themselves and others. This is a practice in both critical and visible thinking. For example, if a student is told that 13.8 billion years ago everything in the universe was crushed into something smaller than an atom, they will have to decide whether they accept that claim or not.They then ask themselves why they accepted this claim or not. Did they do so intuitively? Or because they trust (or distrust) their teacher or textbook? Did they feel it made logical sense? Or was it based on the evidence presented? Perhaps it was a combination of these justifications. 351

Tracy Sullivan

The above skills and concepts intertwine to create a pedagogical framework for big history in the secondary context that is hierarchical in structure and cyclical in function (see Figure 15.2). Combined, the big history concepts and skills work together to heighten students’ awareness of the processes of knowledge creation, refinement and analysis across a myriad of disciplines. The big history pedagogical framework interacts with the course curriculum content at three levels (see Figure 15.2). At the top level, students connect with the big history core concept modern scientific origin story and the big questions, “Who are we? Where did we come from? Where are we going?” T   hese questions act as an anchor point as students move through the course curriculum content. By continually reconnecting with these questions, students are interacting with the content in a way that maintains an interdisciplinary focus. This level of the framework is also important for maintaining student engagement. Today’s students are learning pragmatists (Shaw 2009:13). They want to know why what they are being asked to learn is important, and how it’s relevant to their lived experience. Placing the concept of modern scientific origin story, and the big questions,

Figure 15.2  Big history pedagogical framework. 352

The Big History Project in Australia

at the top level of the pedagogical framework means that as students move through the course curriculum content, and cycle through the pedagogical framework, everything they learn in big history becomes an exploration of their lived experience. Examining “Who they are? Where they came from? And where they are going?” Once students have been introduced to the concept of modern scientific origin story and connected with the big questions “Who are we? Where did we come from? And where are we going?”  They move into the second level of the pedagogical framework (see Figure 15.2). Here students engage with the evolving big history narrative of thresholds of increasing complexity to explore answers to the big questions. The narrative structure of big history situates knowledge from a range of disciplines in a coherent chronology, whereby students feel intellectually supported in exploring responses to the big questions. Information in this narrative is not presented as a series of given facts but as the result of the ongoing process of knowledge accumulation over time by Homo sapiens. This is how the pedagogical narrative of thresholds of increasing complexity surfaces the core concept of collective learning. Students are challenged not only to examine how knowledge itself interacts to answer the big questions but how Homo sapiens interact with knowledge and each other to give us our present day, and constantly changing, scientific understanding of our world. As students engage with the pedagogical narrative, they’ reconfronted with a vast array of ideas and claims of knowledge requiring assessment and evaluation. At this point, students move into the third level of the pedagogical framework working with the claim testers (see Figure 15.2). Students employ the metacognition tool kit consisting of the four claim testers to assess and evaluate claims of knowledge encountered in the pedagogical narrative. Students make informed decisions to accept or reject a claim of knowledge encountered in the pedagogical narrative and then begin working upwards through the pedagogical framework. Students take the claim of knowledge they have decided to either accept or reject using the four claim testers, and position the critically assessed claim of knowledge in the appropriate place within the pedagogical narrative of threshold of increasing complexity. Students then return to the top level of the pedagogical framework and ask how this critically assessed and evaluated claim of knowledge contributes to their understanding of the modern scientific origin story big questions, “Who are we? Where did we come from? And where are we going?” Students repeat the above process continually as they work through the course curriculum, exploring the origins of our universe in the Big Bang, from multiple scales of time and space, and making their own informed predictions about the future of humanity, based on the patterns and trends that become visible as they examine a range of events and phenomena from different scales. To illustrate this process, we can use unit 5 in the Big History Project course curriculum, and the fifth threshold of increasing complexity in big history, life.We begin at the top level of the pedagogical framework focusing on two of the modern scientific origin story big questions, “Who are we? And where did we come from?” In the context of unit 5 or the fifth threshold on increasing complexity, and the second level of the pedagogical framework, answering the modern scientific origin story big questions directs us to an exploration of the origins of life. Here we are presented with claims of knowledge 353

Tracy Sullivan

from a range of disciplines including biochemistry and genetics (structure and role of DNA), and evolutionary biology (process of natural selection). We also deepen our understanding of the big history concept of collective learning by looking at how our understanding of the origins of life has developed over time through the accumulation of knowledge by human beings, such as Charles Darwin, and the development of the theory of natural selection, and the work of Crick, Watson and Franklin in understanding the relationship between DNA and life. Equipped with these newly acquired claims of knowledge, we move into the third layer of the pedagogical framework and use our metacognition toolkit, the claim testers, to critically evaluate and assess these claims of knowledge and make a decision if we are going to accept or reject them. Once we have a clear understanding of how and why we accept or reject these claims of knowledge, we move back up to the second level of the pedagogical framework and revisit the pedagogical narrative of thresholds of increasing complexity. We identify how our assessed and evaluated claims of knowledge contribute to our understanding of the origins of life, and then take our deepened understanding of the origins of life and place this into the broader context of the modern origin story big questions “Who are we? And where did we come from?” Throughout this process, we have also been practicing the big history skill of thinking across multiple scales, switching temporal scales from a human life time of discovery (Darwin, Crick, Watson and Franklin), to the existence of life on planet earth, and spatial scales from the size of a strand of DNA to the size of planet earth. This process is repeated through every threshold of increasing complexity in the pedagogical narrative. Continually working through the course curriculum using this pedagogical framework compels students to refine their critical and visible thinking skills while integrating knowledge from a range of disciplines to answer questions that cannot be answered within the confines of a single discipline. This is interdisciplinary thinking in action.

Looking ahead: big history, transformative learning theory and further big history education research The phrase, ‘transformative learning,’ is ubiquitous in today’s education dialogue and is often used in reference to big history. But what exactly do these words mean? And how are they enacted in big history at a secondary level? One option for quantifying how big history is ‘transformative’ would be to use Mezirow’s transformative learning theory. Current research in transformative learning theory focuses on adult learners; however this doesn’t mean transformative learning theory can’t be applied to the adolescent learner.Transformative learning theory originated in the work of Jack ­Mezirow in the 1970s. Mezirow conducted a national field study in 1975 and ­follow-up study in 1978, examining the experiences of adult women returning to study after extensive breaks in learning in US community colleges (Kitchenham 2008:105). This is the largest qualitative research study on transformative learning theory to date. ­Mezirow sought to identify factors that impeded or facilitated student success in these programs. He initially surveyed 83 women across twelve colleges and followed up with mail enquiries to 1072 students of which 846 responded. Mezirow concluded that respondents had undergone a “personal transformation” (Kitchenham2008:105). ­Mezirow 354

The Big History Project in Australia

expanded on his observations of these “personal transformations” to develop a model of transformative learning, defined as: “Learning that transforms problematic frames of reference – sets of fixed assumptions and expectations (habits of mind, meaning perspectives, mindsets) – to make them more inclusive, discriminating, open, reflective, and emotionally able to change” (Mezirow 2003:58). He also concluded that such frames of reference are better than others because “they are more likely to generate beliefs and opinions that will prove justified to guide action” (Mezirow2003:58). A frame of reference has two dimensions: (1) Habits of mind, these are habitual and orienting ways of thinking, feeling and acting that are influenced by assumptions making up a set of codes that can be cultural, social or educational. (2) Point of view, this is a mixture of belief, value judgment, attitudes and feelings that shape a student’s interpretation (Mezirow 1997:6). Mezirow later expanded the definition of transformative learning, including an outline of autonomous thinking, defined as “an understanding of the skills, and disposition necessary to become critically reflective of one’s own assumptions and to engage in discourse to validate one’s beliefs through the experiences of others who share universal values” (Mezirow 1997:6). Mezirow posits that autonomous thinking, and thus transformative learning, is founded in childhood and adolescence, but achieved in adulthood (see Figure 15.3).

Figure 15.3  Mezirow’s progression of autonomous thinking. 355

Tracy Sullivan

Mezirow’s developmental sequence works on the implicit assumption that adolescent scan only reflect critically on what they see, read and hear in single instances, via encounters with distinct pieces of information, at distinct times and in relation to distinct subjects. It does not allow for the possibility that adolescents have the capacity to analyse collective sources of information, compile critical reflections and apply them to their belief systems. Theoretically, using secondary big history course curriculum, and the big history pedagogical framework, it would be possible to investigate transformative learning theory and autonomous thinking in adolescents at a secondary level (see Figure 15.4). Based on anecdotal comments from Australian Big History students the definition of ‘frame of reference’ could be applied to two aspects of students’ learning experience: (1) Cultural and social ‘habits of mind’ engaging with big history’s modern origin story big questions ‘Where did we come from?’ ‘How did we get here?’ ‘Where are we going?’ “Big History was an incredible learning experience because it allowed me to gain knowledge about the question, how did humans get here today? And this is a question I have asked myself all my life.” – Emily, Year 9 Big History student, Brisbane, Australia. ( Johnston & Sampson 2013) (2) Educational ‘habits of mind’ engaging with big history as a vehicle that connects knowledge across disciplines. “Big History has explained how seemingly isolated events we are taught in school fit in the context of the universe, it’s history, and how this may have further implications in the future.” – Emily, Year 9 Big History student, Brisbane, Australia. ( Johnston & Sampson 2013) Overlaying the above ‘frames of reference’ on the big history pedagogical framework identifies key data collection points to facilitate the exploration of student learning in the context of transformative learning theory as they work through secondary big history course curriculum (see Figure 15.4). As students enter the first level of the pedagogical framework, they are confronted with questions calling them to acknowledge preexisting cultural and social habits of mind in relation to the modern origin story big questions – “Who are we? Where did we come from? Where are we going?” For some students, this may be the first time they’ve had cause to reflect on these questions.This signals the starting point for possible transformation in understanding their beliefs, value judgments and understanding of the world they live in and how it came to be. This constitutes an awareness of their ‘point of view’ as part of their preexisting ‘frame of reference’ (see Figure 15.4). Theoretically, on achieving a level of awareness of their ‘point of view,’ a transition can be made into the second level of the pedagogical framework, the narrative of thresholds of increasing complexity. Here students undertake the process of reconciling their existing ‘point of view’ with information presented via the pedagogical 356

The Big History Project in Australia

Figure 15.4  Big history pedagogical framework and transformative learning theory.

narrative. Students are compelled to ask themselves if this information aligns to their preexisting point of view, and if there are differences, what are they? The act of engaging with information in the second level of the pedagogical framework calls students to generate an awareness of their ‘habits of mind’: their habitual ways of thinking, and preconceived codes of how knowledge is organized and utilized to answer questions. If students are products of a traditionally structured, siloed curriculum, their accustomed ‘habits of mind’ reinforce a paradigm where knowledge is compartmentalized. This is an ineffective model for responding to big questions (see Figure 15.4). Awareness always precedes transformation. You cannot transform something you are not aware of. The first two levels of the big history pedagogical framework can be categorized as creating awareness of existing frames of reference. The third level of the pedagogical framework arms students with the analytical tools (claim testers) to assess and evaluate their preexisting ‘points of view’ and ‘habits of mind,’ to make choices to revise or reinforce those ‘frames of reference’. Theoretically, the repetitive process of cycling through the big history pedagogical framework has the capacity to transform how students view the world and their place in it, as well as to change how they use knowledge to answer complex questions (see Figure 15.4). 357

Tracy Sullivan

As a research field, transformative learning theory is currently limited by its research context/s and subject/s, specifically adult learners. Expanding this research field requires examination of contexts and learners beyond its current boundaries (Taylor 2000:285). Secondary big history course curriculum and the associated pedagogical framework outlined in this chapter provides such an opportunity. Secondary big history course curriculum offers a scaffold for potentially boundary-­breaking ­research into how adolescent learners enact the core elements of transformative learning theory and cultivate the capacity for autonomous thinking. Based on ­Mezirow’s sequence of development of autonomous thinking, this type of research could give insight into how adolescent students: (1) Become more aware, and critical, in assessing assumptions of others and those governing their own beliefs in relation to their understanding of, who we are as a part of an infinite universe, and how that understanding has developed and changed over time. (2) Become more aware and better able to recognize frames of reference and paradigms, and imagine alternatives. Such as identifying a range of origin stories over time, from a range of cultures, and making an informed decision as to which parts of these they accept, reject or redefine. (3) Become more responsible and effective in working with others to collectively assess reasons, pose and solve problems, and arrive at a tentative best judgment regarding contested beliefs about the major challenges that face the human race today. Understanding their complex history and imagining possible futures. At this stage, these hypotheses remain conjecture. Qualitative and quantitative studies with secondary Big History students, and course curriculum, will be required to determine the validity of these claims. The experience of teaching Big History in Australian schools, as outlined in this article, furnishes a rich environment to pursue such education research. Big History is a course curriculum for today. Never before in the history of human kind has there been more need to educate students in a way that allows them to ­interact with knowledge at the points where it connects, rather than within the imagined boundaries that separate it. The example of big history in Australian secondary schools demonstrates that interdisciplinary course curriculums, focusing on connection, can be implemented within mandated curriculum structures that emphasize separation. I began this chapter using Dewey’s philosophy of experiential education as a pretext for the claim, today’s students need an educational experience that will ­prepare them to live in a globally connected society. My hope is that the explanation of how big history functions in Australian schools in this chapter, has provided an insight into how big history is creating experiences for students to help them understand the increasingly connected world they live in. Experiences they can take with them, out of the classroom, and practice in their everyday lives as they encounter, process and build upon, the vast amounts of information presented to them every day. It has taken 25 years from its inception for big history to mature into a flourishing university level offering. For Big History and other interdisciplinary curricula to find enthusiastic embrace at a secondary level, it may be a comparable timeframe. Key to 358

The Big History Project in Australia

the national, and global embrace of  big history at the secondary level, is the iteration and proliferation of versions of big history course curriculums in a range of languages, mapped to regional curriculum standards, and embedding rigorous evaluation methods producing extensive region-specific student performance data. The early years of big history in Australian secondary schools tell a story of an innovative experiment in interdisciplinary education proving big history can work in a secondary classroom. The challenge now is to take the lessons of the previous eight years, refine and proliferate big history course curriculum offerings, to demonstrate that big ­history can not only work in a portion of secondary classrooms but every classroom.

Notes 1 A detailed discussion of the complexities of interdisciplinarity requires further examination than is presented in this chapter. The definition provided in this paper is using the term interdisciplinarity in its broadest form. For more detailed explanation on the functioning of interdisciplinarity, you can see either Lattuca L, Creating Interdisciplinarity: Interdisciplinary research and teaching among college and university Faculty, Vanderbilt University Press, Nashville, 2001 or Thompson Klein J, Crossing Boundaries: Knowledge, Disciplinarities and Interdisciplinarities, University Press of Virginia, Charlottesville, 1996. 2 Professor Christian’s 35 year tenure at Macquarie University was interrupted when he moved to San Diego and taught at San Diego State University for eight years. He returned to Macquarie University in 2009. 3 Professor Hughes-Warrington was employed by Macquarie University as Associate Dean Learning and Teaching in the Faculty of Arts at the time of this meeting. She would leave Macquarie University prior to intensive development of the Big History Project. 4 Big History Project would be the launch project for Intentional Futures www.intentional futures.com. Ian Sands is no longer working with Intentional Futures. 5 Narara Valley High School NSW, Nossal High School VIC, Lakeside School WA, Greenhills School MI, Northville High School MI, Rivers School MA, Brooklyn Latin NY, San Diego High School CA) (Big History Project LLC, 2016). 6 Many Australian schools continue to teach Big History as an elective, however, there are a growing number of schools working to integrate Big History as a standalone course integrating mandated content from a range of subject areas. 7 Prior to January 1, 2017, this organization was the New South Wales Board of Studies and Education Standards. 8 In New South Wales, endorsement must be sought from the Education Standards ­Authority for any non-mandated course to be recognized as part of required student learning up to Year 10.

Bibliography Apple, M. (1979) Ideology and curriculum, London, Routledge and Kegan Paul. Baker, D. (2014) ‘Standing on the shoulders of Giants: Collective Learning as a Key ­Concept in Big History’, in L. Grinin, D. Baker, E. Quaedackers and A. Korotayev (ed.) Teaching and Researching Big History: Exploring a New Scholarly Field, Volgograd, Uchitel Publishing House: 41–64. Baker, D.(2016) ‘Collective Learning: A Potential Unifying Theme in Big History’, Journal of World History, 26, 1: 77–104. Big History: Connecting Knowledge, Macquarie University, viewed 5 May 2017, https:// www.coursera.org/learn/big-history 359

Tracy Sullivan

Big History Institute (2016)Big History School, viewed 10 December 2017, https://www. mq.edu.au/bighistory/our-courses/big-history-school Big History Project LLC, Big History Project 2016, viewed 15 August 2017, https://school. bighistoryproject.com/bhplive Big History Project LLC (2012), Big History Project FAQ. Big History Project LLC (2015), Summary of Big History Project Research 2014/15 School Year. Christian, D. (2008) Big History: The Big Bang, Life of Earth and the Rise of Humanity, ­Chantilly, The Teaching Company. Christian, D. (2011) The History of our World in 18 minutes, TED, viewed August 10 2017, https://www.ted.com/talks/david_christian_big_history Christian, D. (2016, October)Interviewed by author, Sydney. Christian, D. (2005)Maps of Time: An Introduction to Big History, California, University of California Press. Christian, D., Stokes Brown, C. and Benjamin, C. (2014) Big History: Between Nothing and Everything, New York, McGraw Hill. David, C. (2010) ‘The Return to Universal History’ History and Theory, 49, 4: 6–27. Commonwealth of Australia (2009) The Shape of the Australian Curriculum, viewed 10 July 2017, https://acaraweb.blob.core.windows.net/resources/The_Shape_of_the_­Australian Curriculum_May_2009_file.pdf Dewey, J. (1916) ‘Democracy and Education’, in Boydston, J. A. (1976) ed. John Dewey: The Middle Works, 1899–1924, Carbondale, Southern Illinois University Press, 82–83. Godinho, S. and Shrimpton, B. (2008) Interdisciplinary Curriculum: A Sustainable Future or an Unattainable Vision in a Changing Educational Climate?, AARE International Education Research Conference Papers, Brisbane. Johnston, A. and Sampson, K. (2013) ‘Big History’ Connections’, in Connections: Connecting the Somerville House Community, vol. 11, 1, 12–15. Kitchenham, A. (2008) ‘The Evolution of John Mezirow’s Transformative Learning ­Theory’, Journal of Transformative Education, vol. 6: 104–123. Lane, B. (2012) ‘Thinking Big’, The Weekend Australian, 11 August, viewed 10 August 2017. Lovat, T. and Smith. J., (1990), Curriculum Action on Reflection, Wentworth Falls, Social Science Press. Mezirow, J. (2003) ‘Transformative Learning as Discourse’, Journal of Transformative Education, vol. 1, 1: 58–63. Mezirow, J. (1997) ‘Transformative Learning: Theory to Practice’ New Directions for Adult and Continuing Education, no. 74, Summer: 5–12. Australian Curriculum, Assessment and Reporting Act (2008), viewed 10 August 2016, https:// www.legislation.gov.au/Details/C2008A00136. Shaw, A. (2009) ‘Education in the 21st Century’, Ethos vol. 17, 1: 11–17. Sullivan, T. (2013) ‘Big History, Interdisciplinarity and 21st Century Curriculum: The Biggest of Pictures’, Teaching History, vol. 47, 1: 56–60. Sullivan, T. (2014) ‘Big History and the Secondary Classroom’, in L. Grinin, D. Baker, E.Quaedackers and A. Korotayev (eds.) Teaching and Researching Big History: Exploring a New Scholarly Field, Volgograd, Uchitel Publishing House: 317–327. Taylor, E. W. (2000) ‘Analysing Research of Transformative Learning Theory’, in J. ­Mezirow and Assoc. (eds.) Learning as Transformation: Critical Perspectives on a Theory in Progress, San Francisco, Jossey-Bass: 285–328. Thompson Klein, J. (2010), ‘A Taxonomy of Interdisciplinarity’, in R. Frodeman, J. ­Thompson Klein and C. Mitcham (eds.) Oxford Handbook of Interdisciplinarity, ­Croydon, Oxford University Press: 15–29.

360

16 BIG HISTORY TEACHING IN KOREA Seohyung Kim

Introduction Since the nineteenth century, many historians have focused on the absolute cause of historical phenomenon within a single scale of time and space according to the ­specialization of academic fields. Historical studies of some nations in Western E ­ urope and Asia have offered a base to develop modern nations by emphasizing national identity or the importance of a unified nation to define people as the single community. In this process, discourses in which the superiority of Western Europe was overemphasized or hegemony appeared, and these are recognized today as ­Eurocentric. As a result, some nations of Western Europe have justified their historical experiences as the universal history of all human beings.1 In this historical discourse, many other regions except Western Europe were excluded, and Europe was identified as the center of world history. However, in today’s global society, in which all the regions have close interconnections because of rapid progress of technologies, Eurocentrism or fragmented and ­specialized historical studies have revealed their limitations. Single and absolute centrism was the result of an incorrect understanding of world history, so it could not be world history in the true sense of the term. As a consequence, historians who study modern world history today understand that “the origin of modernity is not ­European, but global.”2 In this atmosphere, new trends appeared in an attempt to analyze the whole history of the Earth from a more balanced perspective. Most historians have considered that only human beings have a history. In this sense, big history is a new perspective because it tries to examine origin stories and the histories of human beings, life and the Universe together by tracing back to Big Bang, 13.8 billion years ago. Of course, lots of origin stories about the world, life and human beings were handed down all over the planet, but big history is different from these stories. The foundations of stories and explanations in big history are scientific knowledge and evidence based on technological developments, which big history uses to analyze many changes and historical meanings. In this way, big history has emerged as a bridge that connects natural sciences and humanities within the largest context of time and space. 361

Seohyung Kim

Bill Gates, the former CEO of Microsoft, and David Christian, the founder of big history, began to extend big history education to teenagers in the United States and Australia from 2012.They launched the Big History Project, which provides videos about several important parts of big history and lots of materials about the Universe, life and human history. South Korea also has begun to teach big history in middle schools and high schools since 2012.The purpose of big history education in Korea is to provide the big picture, which connects everything in the world and encourage students to understand interconnections between different subjects and to develop new levels of creativity. So, in this article, I will examine the history of big history education in Korea and especially consider teachers’ workshop as a mechanism to spread big history education. With this analysis, we can understand the reason why we have to teach big history in the twenty-first century and its importance and historical meaning in the global society.

Big history education in a Korean University The first example of big history education in Korea, which expands its scale of analysis from the Big Bang to the present and the future and tries to discover inter-­ relationships between the natural sciences and humanities, began at Ewha Womans University in 2009. The Institute of World and Global History at Ewha Womans University invited David Christian to teach big history to undergraduate students and to work together with Ewha for five years, as part of a special program sponsored by the Korean Government. The goal of this program was to invite influential world scholars for the development of education and research in Korean universities, so its name was ‘World Class University’ (WCU). As the result of the WCU program, ­David Christian taught big history at Summer school in English, and I taught the first Korean big history curriculum for five years. The title of big history course was “A History of Everything after the Big Bang,” a three-credit class for 16 weeks per semester that explained stories about the Universe, the Earth, life and human beings on the foundation of scientific knowledge and information, and also many historical documents. The purpose of this curriculum was to introduce big history and to encourage students to comprehend interconnections between many disciplines in order to achieve a better understanding of global society, which is connected complicatedly. Also, we tried to inspire attention and interest about different disciplines for students as part of convergence e­ ducation. Convergence education is a broader concept of “Science, Technology, Education, ­Mathematics,” so called as STEM, because it tries to expand its boundaries of education not only to scientific fields, but also to the liberal arts. Many experts and scholars in ­Korea have created the new concept of      “STEAM,” which means    “­Science, ­Technology, ­Education, Arts, and Mathematics.” But this concept has limited the boundary of Arts as the actions of painting and making or other activities, so many scholars suggested much broader idea to include all the humanities, such as history, geography, politics, economics and philosophy. In this sense, convergence education is a new attempt to make more balanced perspectives, which deals with every discipline beyond their academic boundaries. David Christian taught 64 hours of the big history course for 4 weeks every summer vacation for five years, and I taught this course during one semester, 16 weeks. 362

Big history teaching in Korea

At first there was no book or article in Korean that I could use as the reference, so I had to find out lots of material to prepare my course. Because my field is history, I gave more attention to explaining human history within a larger context, but I also began to read and explain scientific events, such as creation of the Universe, and the many changes to the Earth and the evolution of life and human beings. Within this big history course, I could collect and accumulate lots of information and data from modern sciences and begin to recognize common features between the natural sciences and humanities. One of these common features is that both scientists and historians make use of the most reliable data for writing and explanation. For example, in big history, many cosmologists have suggested that beginning of the Universe was 13.7 billion years ago, but this date is changed as 13.8 billion years ago with more reliable scientific evidence. The same is true for archeologists or historians, because we also use the most credible relics and documents of the past. Big history education is the essence of convergence education in that it not only combines all the disciplines, but tries to discover interconnections between them. Moreover, big history has different features from other fusion educational models, because it has the narrative, which connects the Universe, the Sun, the Earth, life and human beings according to a type of chronicle. So, big history is the new trend in analyzing origins and changes within much larger contexts of time and space based on scientific narratives. Big history education is important because it recounts and explains not only the history of human beings, but also the history of life, the Earth and the Universe, which surrounds and interacts with humans. It is literally the history of everything in the world. Many historians think big history can be the foundation for writing histories of the whole of humanity on the Earth, like universal history. In the nineteenth century, Lord Acton suggested that “universal history is different from unified history of all nations… Universal history can be mentioned in the higher level according to degree of devotion to whole common future.”3 Leopold von Ranke, widely regarded as the father of modern history, also claimed that “all periods belong to the whole, the universal history. Knowledge about human history must be properties of all human beings.”4 Influential modern world historian, William McNeill, also emphasized universal history, which could provide not only the complexities of many peoples and cultures, but also the perspective of human beings as the whole.5 Since the nineteenth century, most historical narratives and explanations had been based on the nation state, but big history, which expands its frame to different levels of environments, has enabled historians to write the history of human beings as the whole. Big history is thus important for two reasons. The first reason is the accumulation of so much information. Rapid development of modern science after the twentieth century has close relevance with not only the natural sciences, but the humanities, such as history, archeology or anthropology. Lots of scientific discoveries, knowledge and information could provide necessary and objective data for historians, archeologists and anthropologists to study the past of human beings. It is now possible to describe the creation of the Universe, which happened 13.8 billion years ago, the formation and stabilization of the Earth, and the appearance of many species of humans on the Earth because of the development of scientific technologies and interconnections between the natural sciences and humanities. 363

Seohyung Kim

The second reason is that the task of historians to write a history of all human beings becomes more realistic by emphasizing the importance of the various ‘knowledge communities’ of the Earth. It is very important for us to place humans in the many different ethnic communities and knowledge communities they have constructed. This kind of history is the history we need in today’s global society, and it enables us to explain concepts and the importance of human beings as the whole. Of course, it is very difficult to maintain objective historical narrative and perspectives about all human beings. After all, any narrative of human history cannot but be diverse and overlapped according to different point of views. In this sense, big history, which tries to write to history of humans as the whole, can be the bridge to overcome any form of centrism, which confines history to the limited historical experiences of some people in the world; rather, big history offers a multi-centric perspective. With this in mind, big history education at Ewha Womans University was the first attempt to implement a new history course by expanding scales of time and space from human beings to other organisms, the Earth and even the Universe. We were sure many students could find out “something that they could not see before” and analyze existing phenomena with the new perspective of big history. Big history education brings the history of the Universe, the Earth, life and human beings together and investigates diversities and universalities in history. So, it is important to consider both a deeper understanding about the past, and prospects the future, in order to help solve the many serious problems confronting global society today. From this first course at the university level, big history education in Korea spread to the level of middle school and high school. More and more people began to sympathize with messages of convergence and interconnections between several disciplines through big history.

Big history as an extra-curriculum course The first big history education as an extra-curriculum middle school course began in 2011.The course was composed of five classes which were assisted by the Institute of World and Global History at Ewha Womans University and the Korean Foundation for the Advancement of Science and Creativity. The first purpose of this extra-­ curricular course was to help students understand today’s global society through the convergence of many different scientific perspectives.6 Since the nineteenth century, the natural sciences and the humanities have been separated and regarded as totally different disciplines. However, as we have been arguing, it is necessary to combine and link different disciplines together to understand the nature of global society. The second purpose was to cultivate a sense of the unity and the importance of common sense or virtue of modern global society.7 There are so many global problems, such as global warming, nuclear war or poverty, and it is very difficult for us to consider solutions to these problems because of conflicts of complicated interests between nations. Big history, which tries to emphasize human beings as a whole, is helpful in understanding the possibility and necessity of a single global community on the Earth. There were five parts for the first big history Middle School extra-curricular course. The first part was “Big Bang and the origin of the Universe,” within which I explained the process of the creation of the Universe and the importance of humans 364

Big history teaching in Korea

in the universal scale of time and space. Especially in this part, I emphasized the reason why we needed a scientific foundation to understand the nature of human beings and characteristics of human society. The second part was “Creation and development of life on the Earth,” and my goal was to comprehend the evolution of organisms and the historical meaning of biological diversities. The third part dealt with “Era of foragers,” with the appearance of many different species of hominids and proto-humans on the Earth. In this part, I explained the importance of the natural environment of Africa and its importance in human history. Also, I discussed human life ways during this period, in which most people gained energy for their survival from the environment by hunting and gathering. These were the basic foundation of subsequent complexities and developments within human societies. In the fourth part, I described the domestication of plants and animals and rapid changes within human societies that followed, such as the appearance of the city, nation, class and power. Agriculture was the driving force of civilization for humans, and human societies could become more and more complex compared with other organisms in the world and even in the Universe. The final part of the course was “Global Era,” which traced back to the formation and development of several networks and how the Earth came to be connected into a single global network. Also, I pointed out the limitations of traditional world history, which had emphasized only some historical experiences of Western European nations. Moreover, I discussed the influences of human beings after the Industrialization in the nineteenth century and the role of human beings and global community of human beings as the whole. Every student who participated in this course was a talented student with excellent ability in sciences and mathematics. So most of them already had an interest in the natural sciences including cosmology, biology, physics or engineering. However, while taking the big history course, they became more interested in the humanities, such as geography or archeology, and even history. Most students in Korea think that history is a boring subject, and few students have an interest in the history of Korea or the world. In this sense, the big history extra-cuticular course could be the catalyst to create new interest and excitement amongst these scientifically talented students. Besides, even though, many students did not know big history, they had lots of knowledge and information about the Universe, the Earth and life. So, big history could help explain facts with different and new perspectives, as it attempts to combine and interconnect the natural sciences and the humanities within the largest scale of time and space in the world. With the success of big history course, the Institute of   World and Global History and Korean Foundation for the Advancement of Science and Creativity (KOFAC) again began to create a new extra-curriculum approach. The goals of extra-­curriculum in Korea are to expand the chances for cultivating upright character and public e­ ducation, to change education from a supply and demand consumerist approach and to increase learning choices of students because they can choose various extra-­curricular courses. Of course, there are also many advantages for schools. Schools can develop the quality of their classes and make available different levels of courses d­ epending on the ability and standard of their students. Most of all, students can manage self-­directed learning through many activities because it is very difficult for s­tudents to experience a range of activities within the formal education system in Korea. 365

Seohyung Kim

As a result, we created new big history extra-curricular courses, composed of two hours per week. Teachers explained key concepts and terms in big history for the first hour, and students could do other activities for the second hour. This new curriculum provided the opportunity for students to develop the ability of convergence and develop an interest in many different subjects at school. For teachers, the big history curriculum could be the opportunity to increase their responsibilities and improve the quality of their classes. For parents, big history courses offered them the opportunity to reduce private education, because public education schools were now in charge of this new education program. It was very complicated and difficult to introduce a new course into Korean schools, and this was the main we chose big history as the extra-curriculum. In a way that was different from the United States or Australia, where big history courses were being introduced for teenagers, only the Ministry of Education can be involved in public education in Korea. In this situation, if the principal of each school permits the incorporation of big history course as the extra-curriculum, then students can learn big history. In order to have the big history course be accepted as the extra-curriculum, we had to develop new syllabi for classes and conduct teachers’ workshops to help teachers understand big history and prepare themselves to teach the course. In accordance with the academic calendar of Korean schools, we decided to make 10 categories within the big history extra-curricular course. In this course, teachers taught Big Bang, Stars and Element, Solar system and the Earth, birth of organisms, evolution of life, appearance of humans and global migration, agriculture, the formation of cities and nations, global networks and Anthropocene. The most important difference between the former subjects and big history course is to emphasize the co-existence and interconnections of many subjects. Students could understand larger frame of the Universe, the Earth and human society and develop the ability to link the contents of curricula together. The Institute of World and Global History at Ewha Womans University conducted the first two big history teachers’ workshops in 2012 and 2013. The first teachers’ workshop was held from July to September 2012, and eight teachers participated in this workshop. Five of them were science teachers, two were history teachers and one was geography teacher. David Christian gave a presentation in the first class of the workshop and explained the approach of big history, the Big History Project, and key concepts and thresholds in big history. There were teachers from both middle school and high school, and collectively we decided which level was appropriate for big history education in Korea. Also, there were two different types of school, the public school and private school. During the teachers’ workshop, every teacher was very eager to understand the largest patterns and scales of big history, and important thresholds beyond their subject. This was the first experience for Korean teachers to teach other subjects based on the largest possible narrative from the beginning of the Universe to human history. In order to create more interest among students, we also created activities for each class. For example, in the first class to explain big history and its scales, teachers made timelines to experience the whole scale of time from the beginning of the Universe to the present. Some of the activities were adopted from those used by the Dominican University of California during its Summer Institute, but others were invented 366

Big history teaching in Korea

with Korean teachers. Activities from Dominican University included experiencing skulls from different hominid species, or making a tunnel book; and our own activities included making a story about the formation of a constellation, or building a city or nation under certain conditions. By doing these activities, teachers could understand big history more easily and realized the necessity of developing new teaching methods as an escape from traditional education, which had emphasized only knowledge and the answer to questions. During the teachers’ workshop, we also focused on the differences between traditional education and big history. Big history has the potential to be a very important driving force for students to do their own thinking and to discuss several points with other students by using possible evidences or proofs. In this sense, big history as the extra-curriculum was ‘real’ education. As a result of the workshops, teachers developed a detailed understanding of the interconnections between cosmology, physics, chemistry and history in the part of Big Bang, and they also came to realize the difference between the ways of understanding of the world from the past to the present. In considering the evolution of the first organisms, they could comprehend the Goldilocks conditions that were necessary for life, and also the historical meaning and importance of bacteria. Of course, they also considered the evolution of life and came to understand the “Tree of Life.” Science textbooks in Korea do teach evolution, but there was no explanation about “Tree of Life,” so most students and even teachers misunderstood how evolution worked. It was the same in section focused on the evolution of human beings. Most textbooks stressed how our ancestor, Homo sapiens was smart to survive, but there was no explanation about the co-existence of other species with Homo sapiens. In the part on global networks, every world history textbook emphasized the historical experiences of Western Europeans. In this way, the big history teachers’ workshop was a watershed in the recognition of problems and errors in Korean textbooks and education system. The second teachers’ workshop was held from January to February 2013, and 20 teachers participated. Some of them came from other Korean cities, far from Seoul. Teachers who had participated in the first big history teachers’ workshop were used as lecturers for the second teachers’ workshop. There was one Korean literature teacher, and she had no chance to learn or study science subjects after being a teacher. However, she had considerable interest in the Universe and human history and was eager to learn more at our big history workshop in Seoul. Her experience demonstrated that more teachers could understand big history and sympathize with big history as convergence education, which connects the natural sciences and the humanities based on the scientific evidences and stories.This was a new and powerful education model in Korea which helped develop creativity.

Free semester and big history course The teachers who participated in big history teachers’ workshops opened big history at their school as the extra-curriculum course, and one high school introduced a regular big history course for the first time. Hana Academy Seoul is the one of the best private high schools in Korea, and two teachers, a science teacher and a geography teacher, were successful in creating this regular course. It is an Independent Private 367

Seohyung Kim

High school, so teachers have free choice to create new courses, without being subject to the demands of the Ministry of Education. Of course, it was not so easy to make a big history course, because the goal of most Korean high schools is to educate students to be successful in the entrance examination. However, teachers at Hana Academy Seoul are very eager to spread big history to their students and almost 40 students take the big history course each semester. These were the first teachers to teach big history at high school, and their students were the first to learn big history in Korea. After the sudden death of late Professor Ji-Hyung Cho at Ewha Womans University, seven scholars established the “Big History Academy” to help spread big history. Most of the seven are famous and influential scientists in Korea. There are currently 30 people in Big History Academy today, and we are working to make big history a regular course for middle school students. There is a special semester in Korean middle schools, called the “Free Semester.” Free Semester is the semester without evaluation or examination which allows students to explore courses for the future. The former administration established this new educational policy to provide more chances and opportunities for students to discover their dreams and aptitudes. This educational system is similar to the “Transition Year” of Ireland, which has been liberating students from examinations and introducing them to a wide range of learning experiences since 1974. Some Irish schools leave examinations out altogether or introduce activity programs for future careers with the help of private companies or communities. Every middle school in Korea has had to introduce a “Free Semester” since 2016. There are currently 3,204 middle schools in Korea today. The Ministry of Education initially selected 42 middle schools in 2013 to introduce the “Free Semester,” and all middle schools in Jeju Special Self-Governing Province began this new educational system in 2015. When David Christian and I visited Jeju, we met the superintendent of Jeju Special Self-Governing Provincial Office of Education, and several other people in charge, and discussed ways to teach big history during “Free Semester.” However, because Jeju is an island in the ­Southern Sea of Korea, it was hard to manage a program for teachers in Jeju. Some teachers have participated in Big History Academy and we discussed possibilities of big history education in Korean schools. Finally we arrived at a conclusion that the best way to spread big history education in Korea is through the “Free Semester.” Middle school students have learned basic sciences and history, so it is natural to combine the natural sciences and the humanities within big history course. Besides, there is no evaluation or examination during “Free Semester,” so this is the best time to experience several activities and unfold the largest patterns or structures to help develop student creativity and their ability to prepare for their near future. However, this development had ramifications for the big history teachers’ workshop. The former workshops were for the extra-curricular model, so there were not so much burden for teachers. But if we want to make big history as regular course, we have to provide more detailed content to teachers. First, most teachers in Korea have to take job training as teachers, and when they complete this training course the Ministry of Education gives teachers credit. It is very strict to be a part of the institution that gives credit to teachers, so the Big History Academy worked collaboratively with Seodaemun Museum of Natural History, which was the first Natural 368

Big history teaching in Korea

History Museum established by a local government in Korea in 2003. The Director of museum is eager to spread big history and he has plans to make this museum the first big history museum in the world. As result, the Big History Academy and ­Seodaemun Museum of Natural History submitted a proposal to teach big history to Seodaemun County. The type of this workshop is job training for teachers, which gives necessary credit to teachers when they complete the big history education program, which is entitled “Big History Teachers’ Workshop for Free Semester.” This is the first big history teacher’s workshop for big history education as the regular course, so we decided to limit applications to 20 teachers in Seoul. The goal of this workshop is to teach big history, which systematizes information and knowledge from other subjects based on convergence education, and to pursuit interconnections between different subjects within the narrative of big history. From the Big Bang – the beginning of the Universe – to the present and to the future, big history can link the Universe, life and human beings at all scales of time and space through big questions, and develop creativity amongst students by seeking diverse answers. This new version of the big history teachers’ workshop is composed of 15 parts, and each part is two hours long.  The first part is Big Bang, beginning of the Universe and we will examine changes in the Universe after the Big Bang and understand the process of beginning of the world with scientific evidence. In the part focused on Stars, teachers can comprehend phenomenon in the Universe such as the birth and death of stars, and common feature between stars and human beings. In the part ­focused on various elements, we can search for elements, which were made in stars, and way everything in the world was made with the combination of elements. It is also important to grasp the birth of the Sun and interconnections between the Sun and other planets in the part of Solar system. Teachers will examine the process of creation of the Earth and common or different features between the Earth and other planets and they can discuss the Goldilocks conditions necessary for the emergence of life on the Earth, and the “Tree of Life.” Also, the course will explain the creation and extinction of species in the history of the Earth, and teachers will exchange opinions about the historical meaning of the Great Extinctions in the history both the Earth and human beings. In the part of the workshop focused on the first humans and evolution, teachers can understand various conditions surrounding the appearance of human beings and the evolution of several species. Also, in the section that considers the Paleolithic era, we can comprehend developments and changes in human history and determine the conditions necessary for agriculture and the rapid changes in human societies that followed. In the section on the city and nation, we discuss common and different features of several cities or nations; and when we consider empires, teachers will analyze several empires in world history, and how the development and collapse of empire has a­ ffected human history. Of course, we will also try to identify interconnections between regions and nations in the part on global networks and how Western Europe came to dominate other worlds, such as Africa, Asia and America; and also the historical meaning of Industrialization in modern times. And in the last part, teachers have the opportunity to think about the influence of human beings in terms of the ­Anthropocene, and the possibilities for co-existence between other species and humans. 369

Seohyung Kim

Big history as the regular course is for 14-year-old students, which means we can better explain some key concepts in each part, and tell stories that let students link these things together. Of course, there will be many activities for teenagers to experience knowledge and information from class by doing things. In this sense, the big history teachers’ workshop is composed of 30 hours, with one hour of explanation and one hour of activity for each part. Teachers who participated in former teachers’ workshops and professors from the Big History Academy will be lecturers for the workshop. From the first day of August, this workshop will be held for 7 days, with 4 hours per day. So, teachers will take 30 hours of big history course to help prepare them to teach the big history regular course; and after completion they will teach the big history course for the “Free Semester.” For this workshop, teachers, who had already attended previous big history teachers’ workshops, wrote the textbook. The temporary title of this textbook is “Big History from Teachers,” and it is composed of two sections. The first section is the science part, which explains how to see the world through the lens of big history, beginning of the Universe, dust from stars, elements and the family of the Solar ­system, the living Earth, the miracle of life, and the Great Extinctions. The second part is focused on human history, including Homo sapiens, the Paleolithic era, civilization and agriculture, the appearance of large human communities, empires, connected worlds, Industrialization and the rise of the Western Europe, the ­Anthropocene and future. The Director of Seodaemun Museum of Natural H ­ istory, the Director of the Seoul Science Center, and I supervised all the contents and stories in this textbook. And now we are preparing a big history workbook, which is full of many stories about the Universe, the Sun and the Earth, life and human society and various activities for students. This new version of the teachers’ workshop is very important, because it is the first trial for big history as part of the public education curriculum. With the big history teachers’ workshop, teachers can understand the importance and necessity of big history education in global society and help develop a more suitable curriculum for Korean educational system. For the twenty-first century, big history is the essence of convergence education in both Korean society and global society.

Conclusion In the twenty-first century, we desperately need a new creative education model, which can overcome the limitations of traditional teaching methods. It is particularly crucial for Korea, which is developing into a leading country in the world, to develop creative talented people and convergence between the natural sciences and the humanities is the most powerful driving force for growth of next generations. In this sense, big history, the epitome of convergence knowledge between the natural sciences and the humanities, is the common sense and foundation for leaders in global society.Through big history, students can understand the general structures and frames of all disciplines with storytelling, and they are able to interpret events in the Universe, in the Earth and in human society from many different perspectives. Thus, big history is the answer to enhanced communication, co-existence and consilience between disciplines. 370

Big history teaching in Korea

In this regard, big history education is one of the most important education models in Korea today. It is very important to develop curriculums for teachers to overcome limitations of subjects through an understanding of big history. For the previous big history teachers’ workshops but big history was part of the extra-curriculum at that time. With this new trend to provide more opportunities to think about students’ future, big history may become the regular course in Korean public education. To achieve this goal, the Big History Academy will hold big history teachers’ workshops to prepare them to teach the regular course, and teachers can develop creative education model, which accommodate the particular situations and levels of learners. There have been so many criticisms of the Korean education system, because the entrance examination has been seen as the most important goal. This has meant that students had no opportunity to seriously consider their future, and which subjects or disciplines were the best fit for them. Today’s global society confines individuals to one field. But young generations need to cultivate convergence knowledge and prepare for unpredictable situations in global society. Only big history education can provide understanding at different levels, and insights that connect the past, the present and the future. For this purpose, big history education and teachers’ workshop have the potential to be watersheds in both the Korean and global societies of the twenty-first century.

Notes 1 Ji-Hyung Cho, “Beyond Eurocentrism to Global History,” Ji-Hyung Cho, Yongwoo Kim eds., Challenge of Global History: How Overcome Eurocentrism (Paju: Seohaemunjip, 2010), 16. 2 David Christian, “New Imaginary Community: From Ethnic History to History of Humanity,” Ewha Sahak Yeongu 40 (2010), 6. 3 Ernest Breisach, Historiography: Ancient, Medieval, and Modern (Chicago, IL: University of Chicago Press, 2007), 321. 4 Fritg Sten ed., The Varieties of History: From Voltaire to the Present (New York: William Collins, 1956), 61–62. 5 William McNeill, “Mythistory, or Truth, History, and Historians,” American Historical Review 91:1 (February, 1986), 1–10. 6 Seohyung Kim, “Big History and Convergence Education in South Korea,” Ricard B. Simon, Mojgan Behmand and Thomas Burse eds., Teaching Big History (Oakland, CA: University of California Press, 2014), 69. 7 Seohyung Kim, “Big History Education and Convergence Education Program for the Talented Students,” Ewha Sahak Yeongu 44 (2012), 272.

371

17 CROSSING THRESHOLDS Using big history to meet challenges in teaching and learning in the United States Robert B. Bain

The Big History Project (BHP) has been more than a curriculum in big history. It has been a design experiment in teaching and learning.1 In addition to providing a course in big history for thousands of schools and hundreds of thousands of students around the world, the BHP team2 used research and theory to engineer, study, and revise forms of learning to help meet some enduring challenges that history teachers and students face in “doing” the teaching and learning of history. While embracing BHP’s central goal of teaching effectively a modern origin story that, as David Christian explained, students will “understand the challenges that face us and the opportunities that face us…[so] big history will become a vital intellectual tool”, the project team recognized the chance to do even more.3 In creating a syllabus to teach Christian’s elegant narrative of increasing complexity,4 we also seized the opportunity to engineer a syllabus that could increase the complexity by which students could engage in disciplinary thinking.5 More specifically, the BHP design team used theory and research on thinking and learning to improve the way students could read historical and scientific sources, write evidence-based arguments, reason about causation, and use concepts drawn from multiple disciplines while learning the big history story and its concepts. We resisted assuming that if students learned about Christian’s Thresholds of Increasing Complexity and the evidence that makes that story so compelling, they would also increase their capacity for critical thinking. It is a tempting assumption. The big history narrative asks vital questions that invite investigation, the weighing of evidence, and assessing others’ claims. Critical thinking was inherent in creating big history. It was tempting to believe critical thinking also inheres in learning big history. However, three experiences disturbed any chance that the BHP design team would hold such a belief. First, the pilot teachers6 quickly pointed out that the videos and course texts that scholars made for the BHP not only posed big questions but also answered the questions in an engaging manner. In many ways, the course materials did the thinking for the students, mitigating the need for further inquiry, except to clarify complicated scientific concepts (such as Cosmic Background Radiation, accretion, or natural 372

Crossing thresholds

selection). Though the materials modeled brilliantly scientific or historical thinking, the teachers worried that there was little in the course narrative to help students do such thinking on their own. Second, the public and educators in the United States expect courses both to teach content and to advance students’ literacy and thinking skills. Administrators and teachers reviewing the course could see how learning a big history story would further students’ understanding of scientific and historical content. However, they could not see how the narrative would enhance students’ literacy skills or even their ability to apply the new concepts they learned. Often this hindered course adoption. Sold on big history and the content it offered students, educators wanted evidence that adding the BHP course would help further students’ capacity to read, write, and think. Finally, at least 100 years of research on students’ cognition challenged the belief that learners are passive recipients of what a curriculum, even an exciting curriculum, teaches.7 Cognitive research has shown how students actively use their existing, pre-instructional practices, beliefs, and knowledge to construct new understandings. What students know or believe affects how they interpret new content, sometimes supporting what they learn while at other times hindering it. Effective curricular design rejects the view of students as vessels into which interesting content flows. Rather, research and theory on learning holds that curriculum and instruction must anticipate and plan for how students’ previous knowledge, experiences, or understandings might affect what they learn. For curriculum designers, knowledge of learners is as important as is the knowledge of the “stuff ” taught. From the outset, therefore, the BHP design team in the United States faced questions raised by teachers who wanted activities to engage their students in disciplinary thinking, by administrators who sought ways to meet the literacy demands on their schools, and by the research and theory urging learner-centered as well as ­knowledge-centered environments. Fortunately, Christian’s big history narrative gave the project team a coherent, elegant structure on which we could use research and theory to build out and test lesson sequences to enhance students’ capacity to “do” big historical reading, writing, and thinking. In this chapter, after briefly introducing the BHP, I offer an example of a challenge students faced in engaging in critical thinking about causes and the lesson sequence, grounded in research and theory not originally part of a big history narrative, the project team designed to meet that challenge. Then, the chapter discusses how the project helped teachers learn unfamiliar content and instructional practices needed to teach BHP course, and to learn all this while teaching big history. I hope to show how the BHP, like big history itself, drew on theory and research from multiple disciplines to construct a pathway of increasing complexity for students and teachers, extending Christian’s central concept to help teachers and students to cross thresholds of increasingly complex learning.

The Big History Project The BHP is a comprehensive program offering a course in big history for secondary students. It delivers to teachers and students all its resources and materials online. 373

Robert B. Bain

Bill Gates’ interest in Christian’s course in big history launched BHP in 2011. Gates had viewed Christian’s video course “Big History: The Big Bang, Life on Earth, and the Rise of Humanity” and fell in love with the concept of integrating “all” knowledge into one coherent story.8 Wondering why there were few big history courses for secondary students, Gates contacted Christian to create an online way to do just that.9 Working with scholars of history and history education at the University of ­Michigan, a group of designers from Intentional Futures, and a group of teachers in the US and Australia, Gates and Christian piloted the first syllabus in 2011 with five schools in the United States and three in Australia. The number of pilot schools expanded to 55 schools in the US and 9 in Australia in 2012. The project team opened the course to the public in 2013 and it has remained free, open, and online ever since. As the project closes its ninth year, there are more than 1,600 schools offering the BHP course in the US alone and hundreds more internationally. Christian constructed the first syllabus, spreading the eight Thresholds of Increasing Complexity across 20 units. His big history origin story remains the overarching and organizing structure for BHP’s course in big history.The initial goal for BHP was to offer Christian’s big narrative to schools around the world, trusting that teachers would add what was needed to make the story fit their context. The design team created a set of video lectures from scholars such as Christian, Henry Louis Gates, and Janna Levin, biographies of important people by Cynthia Stokes Brown, and infographics to populate the 20 units in the original syllabus as well as assessments to evaluate the effect on students and teachers. However, using data and teacher feedback from the first pilot schools, the BHP design team saw a need and the opportunity to do more. In addition to a more streamlined unit structure in the 2012 school year, the team added new content and more rigorous pedagogical approaches, such as disciplinary literacy, d­ ocumentary-based question-like investigations, and instructional routines to help teachers and students become critical consumers and users of big history.10 For example, in the second pilot the teachers had the chance to use ten text-based investigations with their students. In the investigations, students worked with libraries of primary and secondary sources drawn from history and the sciences to develop evidence-based ­arguments in response to big research questions (i.e., When and why have scientist change their minds? To what degree, has the modern age been a positive or negative force?). Teachers assigned one investigation in the first week or so of the course, another at mid-term, and another at the end-of-course, sending their students’ papers to external evaluators.11 Evaluated along four variables – structure of the argument, use of evidence, application of disciplinary concepts, and writing mechanics – these assessments helped the project team to determine growth in how students analyzed texts and the reasoned with them to construct arguments. Even though the students showed growth, the data enabled teachers and the team to refine instruction each year. This continuous improvement kept the course fresh and effective, critical as the course scaled to more and more teachers and students. Slowly, BHP situated the big history course within a set of slightly bigger goals such as deepening and improving students’ thinking using multiple scales, reading primary and secondary texts, use of evidence and disciplinary concepts, crafting well-written arguments, and making causal claims. Each year the project added to and refined the 374

Crossing thresholds

instructional materials and activities to align to the expanding disciplinary goals. We also increased our contact with teachers to help them navigate content in which they were not familiar and, in many cases, enhance their instructional practices in teaching writing, reading, and disciplinary thinking, as well as the big history concepts. Beyond teaching big history, BHP’s course became an experiment in “big” teaching and “big” learning, as I hope the two examples that follow demonstrate.

Causal thinking: a threshold of learning “The study of history,” E.H. Carr claimed “is a study of causes,” a claim that should resonate with those who study big history.12 Big history is hyper-focused on change, big changes over vast expanses of time and space, and the causes of those changes. A critical outcome of studying BHP big history, then, should find students developing a deeper understanding of causation and an enhanced ability to make sophisticated causal claims. Unfortunately, the BHP team did not find that was the case and so we designed and added lessons to deepen the ways students reasoned about causation. It is important to note that BHP students could explain causes of big history’s major threshold changes. We were pleased by how well students used the concept of Goldilocks’ Conditions to recount the situations under which the eight new forms of complexity emerged. However, we had little evidence that students could transfer this causal understanding to think about or analyze the causes of other changes in history or in their lives. Re-narrating the complex and sophisticated causal explanations of experts for the appearance of heavier elements, the formation of the Earth, or the development of agriculture did not seem to enable students to use such expert thinking themselves when considering other changes in history.We discovered that it is more difficult to equip students to do complex causal thinking on their own than it was to teach them multi-faceted causes for eight specific changes in the past.There seemed to be a gap between the type of causal thinking disciplinary experts did and the causal thinking of our students. Illuminating likely gaps between expert thinking and those of novices has been a staple of cognitive research and theory since at least John Dewey’s 1902 essay, The Child and the Curriculum.13 Over the past 50 years, scholars of historical thinking have identified the differences in novice-expert thinking about reading, writing, historical empathy, scale switching, and the nature of the discipline itself.14 For curriculum designers interested in developing students’ thinking as well as their knowledge, such research has been instrumental in operationalizing the learning goals (expert thinking) and the likely cognitive distance a learner needs to travel to reach those goals. While not predictive for every classroom or student, such research does sharpen educators’ eyes to possible and even probable context within which learning will occur.This was particularly true in how a line of research assisted BHP team’s efforts to improve the causal thinking of our students. The research on expert historians and scientists thinking about causation shows them using dynamic processes to employ a type of thinking that has a long history.15 Humans have long wondered how and why things change, why they don’t change, and what has been or could be the impact of change. Indeed, causes appear to shape almost all origins stories, so it is not surprising as people developed systematic ways to 375

Robert B. Bain

study the world around them, analyzing and making claims about causes became central to their disciplinary efforts. And while disciplines have distinctive concepts and different ways to launch, conduct, and represent their research, there are some common characteristics that define their causal thinking. Most historians and scientists work to avoid monocausal explanations, identifying dynamic relationships among events, structures, context, environment, and other possible agents to explain change. They frequently develop a hierarchy of causes that vary in duration, temporal and spatial proximity, importance or significance. Scholars typically consider causes that are unintentional, often happening far away in time and space or are a consequence of long trends or patterns.While recognizing the importance of   human agency in explaining a change, most scholars of history and science rarely consider human action independent of context. Nor do they necessarily privilege events closest in time or space as the most significant in explaining change. They have a rich vocabulary and visual tools to describe, represent, evaluate, and make causal claims. Research on learners suggests most students see change differently.16 Scholarship has pointed to novices’ tendency to seek a single cause, or to line up single causes into a linear chain, like dominoes. It is not unusual for learners to think that the most important cause is the one closest in time and space to the event or is the byproduct of intentional human actions. Several studies point to students’ tendency to gravitate toward the “personal” or “near”, ignoring entirely large-scale structural, contextual, distant, unintentional, or accidental factors. For example, in a noteworthy study, Carretero and his colleagues asked sixth, eighth, tenth, and twelfth graders as well as graduate psychology and history majors to explain the causes of pivotal events in history.17 The researchers asked the students to put in rank order five types of historical explanations: personal, political, economic, ideological, or global. The researchers gave the students a short description of causes (i.e., “The personal motives of Columbus and the Spanish Queen and King” to explain the “Discovery of the New World”) but did not categorize the explanation as a causal type (i.e., as personal, political, etc.). Analyzing the ways different ages or levels of students ranked the causes, the researchers found similarities among the sixth, eighth, tenth, and twelfth graders as well as Psychology graduate students. All the students, but the History grad students, typically ranked the intentions and desires of historical actors as the most compelling explanation for the event, ranking the events most distant in time or space as the least compelling. Only the history graduate students situated personal actions within larger structural contexts or varied their ranking depending upon the historical event. Most surprising was how little change in causal reasoning occurred between six grade and, in some cases, graduate school. Causal thinking, it seems, might not grow automatically with age or with advanced teaching. Specific, explicit instruction in doing causal thinking, as grad students in history would likely experience, appears to make a difference. Or so the BHP design team concluded. If teachers want students to improve their ability to “do” dynamic causal analysis, then the curriculum would have to do more than teach about the Goldilocks Conditions and the specific conditions that fostered the major changes in the big history narrative. Learning that experts developed multi-faceted causal explanations for the thresholds did not seem to deepen students’ capacity to do such thinking on their own.To increase what students learned, we theorized that the curriculum would have to (1) create opportunities 376

Crossing thresholds

for students to “see” the features of complex causal thinking in experts’ explanations, not just learn the products of such thinking; and (2) equip students with the tools to engage in and practice causal thinking for themselves. Over the past three years, the BHP team created, trialed, and modified a sequence of lessons placed judiciously through the course to assist students in advancing their skills in doing causal analysis and making causal claims. A few examples of the learning progression should give a glimpse of the work. In the first new lesson on causation, we made the processes of cause and consequence explicit. After viewing a video early in the course where Christian explains why stars light up, students begin their work on causal thinking with this short paragraph and a causal mapping activity: Historical events rarely have a single cause. Some causes happen right before the event; some long before; others, a really long time before. Some play a central role; some merely contribute to the event; others trigger the event. Some causes are essential while others are less significant. And, the significance of each cause may change depending on the timeframe within which you choose to examine a historical event. Or, the significance may change depending on the questions you ask about that event.18 The text address head-on what research suggests is likely to be students’ pre-­ instructional view of cause as singular, proximate, and undifferentiated. It points out that experts rarely see change as happening from a single cause, that scholars look for cause of different durations, and that play different roles in generating change, such as triggering. The activity made the case to students that doing causal thinking helps develop a relationship “between events over time, which gives us the opportunity to connect historical events to our own lives.”19 In this first lesson on causation, the students get a map that illustrates the processes of star formation as described in Christian’s video How Stars Form. The map does not identify the triggering event, leaving that work to the students as well as identifying the other events as causes, consequences, or both (Figure 17.1). Here the curriculum makes visible, in a simple way, the relationships embedded in Christian’s (or anyone’s) verbal explanation of change. It begins to unpack for students the complexity of thinking that they too often encounter in a linear fashion. In addition, it introduces the idea of  “triggering” causes, a concept not mentioned in the video, and has students consider that an event or process could be both a cause and a consequence. By abstracting a picture of the process from the specific explanation Christian offered, we sought to help students make the first step toward a meta-­ cognitive understanding of causation and a causal argument. After studying Threshold 3, New Chemical Elements, students revisit and complicate their causal map by adding the causes and consequences about the death of stars and the emergence of new chemical elements, as Christian and others explain in the course materials. This second activity asks students to do more than just identify a triggering event or causes and consequence. By not providing the causes, it requires students to identify, analyze, and construct the relationships among the events embedded in experts’ explanations.20 Students have the option to add more circles to indicate more causes or consequences (Figure 17.2). 377

Figure 17.1  Cause & Consequence Activity 1.

Figure 17.2  Cause & Consequence Activity 2.

Crossing thresholds

Of course, such mapping also helps students learn the experts’ explanations for how new chemical elements form. Concept mapping thus serves an important knowledge acquisition function. However, this activity expands the students’ development of causal thinking by showing how, at times, causes and consequences are dynamic, multi-faceted and interconnected. It also helps students see that causal reasoning involves making claims about processes. The lessons on causation, then, spiral by increasing slowly the cognitive load that students take on. Each successive activity in the causal learning progression adds a degree of complexity to the students’ work while reinforcing their prior learning. For example, in the next unit, students work in groups to list all the reasons – well over 20 – experts use in the videos and course material to explain the development of the Earth’s atmosphere. Students then place each reason in a “cause” column and match it to a consequence that they place in the appropriate column.21 As before, the consequences can become causes as students use most processes or events twice.This reconstruction of an explanation requires students to work actively, thinking about and then shifting the relationships between situational causes and consequences.The lesson closes with having students label and categorizes which of the scientific ­disciplines – ­Astrophysics, Biology, Chemistry, or Geology – contributed to our collective understanding about the formation of new chemical elements. Again, the course materials did not overtly identify causes or consequences, nor which discipline contributed to explaining how the atmosphere developed. The activity required students to analyze and then extend the expert’s explanation, reinforcing the dynamic relationships between cause and consequence while introducing the role that different disciplines play in constructing explanations. These activities increasingly add cognitive demands on students who must take another’s explanation and while learning it also analyze it for its structural features. In Unit 6, the course uses a delightful and powerful activity we adapted from the work of Arthur Chapman and James Woodcock.22 The students analyze the story of Alphonse the Camel to determine if it really was the straw that broke poor, old ­Alphonse’s back. The story, a fanciful tale created by Chapman and modified slightly by Woodcock, has many interconnected causal explanations for why Alphonse died. For example, Alphonse had a congenital birth defect, an oppressive camel driver, horrible working conditions, worked with other camels that could not create a union, lived within an exploitative economic structure in a mountainous region. All these and others, formed the potential features of a Goldilocks Conditions that might explain why when Frank the Camel Driver added just one more piece of straw to Alphonse’s load, the poor oppressed camel collapsed and died. The students’ job was to determine if it really was the straw that broke the camel’s back.The story has a host of causes of different grain size, duration, and type: some causes are from long ago while others are more near-term; some causes triggered other causes and others were catalysts or underlying causes; some people might consider some of these major and others minor; some were economic, others political, others still geographic, structural, or even personal. The story is a cornucopia of causes for students to identify and then, for the first time in the course, to make and warrant their own causal claims. After students make their initial analysis, we introduce students to new maps, new categories, and new language to analyze and explain change. The curriculum 379

Robert B. Bain

provides students with charts to locate a cause temporally, giving students a way to see long, medium, or short-term causes. The activity introduces students to institutional categories, such as political or economic, for students to use in categorizing the causes of Alphonse’ death. Then students consider the role they think different causes played by using language such as contributing, triggering, and underlying causes. And finally, the students weigh the various causes by degrees of significance, determining which causes they thought were necessary, which were sufficient, and which most significant. With so many causes and no expert explanation offered, students must go ­beyond just identifying causes. They must now do the type of thinking experts do in figuring out the relationships among events, specifying the relative role each played in Alphonse’s death. The maps, categories, and language help them develop and make causal claims, without determining the causal claim they make. That is, some students will argue that it was Alphonse biological back problem was the most significant reason for his death, while others will dispute that claim by arguing since other camels had died that might only be a contributing factor. In short, the story is causal tinker toy set, enabling students to build a causal model that they can and must defend. And, the causal models they produce tend to contextualize human agency or proximate events within a more dynamic, less myopic frame. Students, then, must develop their own explanation and causal argument. The tools support that work. After working with new maps, new language, new causal categories and after building their own models of causal change in the Alphonse story, students apply their new tools to other explanations in the course (i.e., the development of farming or the modern age) and then practice using their maps and language on events outside the big history curriculum, including their research on their little BHPs. These lessons, charts, maps, and new language assist students in establishing conceptual c­ larity and nuance in making claims about causes in the natural and human realms. From the first activity, the lessons explicitly encourage students to look beyond the human and the proximate in considering long- and short-term causes and the different roles conditions, events, and people play in creating change conditions. These additional lessons do not alter the big history story that drives the BHP course forward. Rather, the lessons help students see how David Christian and others structured their causal arguments. The activities reveal the tools scholars use to build explanation as students learn those explanations. By adding such goals and instruction, students learn more than the big history content since they also learn to use the tools, skills, and reasoning needed to assess and make causal claims. The teaching and learning, thus, grew a bit bigger. However, using these lesson sequences takes additional time, adds increasing complexity to what teachers and students do, and exists within a theory of causal change in students’ thinking. Adding such lessons also makes BHP’s work of preparing teaches more challenging. The BHP team had to think about ways to help teachers see the big picture of the instructional processes, to understand the various ways these lesson spiral to support changes in students’ thinking, and how they might use, modify, or ignore lessons or part of lessons. I turn now to the work of helping teachers learn to teach BHP. 380

Crossing thresholds

Supporting teachers: another threshold of increasing complexity BHP assumed the job of preparing teachers to teach big history, no small task considering the speed at which the course grew from eight schools to thousands of schools globally. While trusting in the teachers’ professional skills, we worried about the justice and wisdom of leaving teachers to their own devices or local support in learning to teach this new course.Taking seriously the idea of collective learning, BHP sought to meet the instructional challenges teachers faced in learning to teach the BHP course, such as mastering novel and complex content and pedagogical approaches. And each year, the BHP team encountered the need to provide support for more teachers and students, working in very different situations and contexts. In the project’s first year, teacher education was incidental to building out the online platform and getting quality resources in the hands of teachers and students. However, teacher education soon became central to the project’s goals. Our work with teachers has helped us identify five challenges or thresholds teachers would, to varying degrees, cross to teach the course effectively. New and veteran teachers must learn to teach with (1) unfamiliar content, (2) familiar content in unfamiliar contexts, (3) little familiarity with how students learn BHP content, (4) ­ambitious inquiry and literacy practices of the disciplines, and (5) BHP’s new online resources and course-management tools. A brief explanation of each illuminates the complexity of these challenges for teachers and for the project to mediate. BHP’s big history course draws on concepts, texts, and literacy practices taken from the disciplines of history, cosmology, physics, chemistry, geology, biology, paleontology, anthropology, economics, theology, and sociology. Given the range of concepts drawn from so many disciplines and the course’s temporal and spatial scope, teachers inevitably work with new and unfamiliar content. For most teachers, BHP pushes them outside their areas of expertise and experience. It is not much of an exaggeration to say that no secondary teacher has had the preparation to handle the challenge of teaching so much material outside their areas of certification and experience. BHP teachers must also find ways to meaningfully place familiar content into big history’s historical narrative. For example, history teachers with experience teaching China’s early dynasties or science teachers teaching fusion discover that there is far less space in the big history narrative and far less time in the BHP course calendar to treat these topics as they might in a typical world history or chemistry course. Making hard choices, then, about what concepts and facts to minimize, collapse, or drop, and how much depth of understanding to sacrifice has proven to be as much a challenge as learning to teach unfamiliar concepts and processes. In addition to teaching new content or familiar content in new contexts, most BHP teachers have little experience with how students learn big history’s content or practices. History teachers, for example, have generally not taught students about Big Bang cosmology, plate tectonics, or evolution. They have had little need in their history courses to attend to students’ background knowledge, misconceptions, or emerging ideas about those scientific concepts. Likewise, science teachers have little familiarity with ways students see political or economic revolutions, patterns of historical continuity and change, or ways presentism influences students’ thinking 381

Robert B. Bain

about government, inequality, or personal history. In addition, most history teachers have little experience intervening to help students develop more sophisticated ways of thinking about such scientific concepts, nor science teachers with those historical concepts. Teachers, then, not only had to learn new content, but they needed to develop a new understanding of how their students might approach such content. As discussed above, BHP offered lessons to develop students’ critical thinking and advance their literacy practices in addition to teaching a big history story, its key concepts, or facts. Adding disciplinary thinking and literacy goals, such as evaluating cause or writing evidence-based arguments, required teachers to use more ambitious, less didactic pedagogy.23 Unfortunately, learning to think, read, and write critically does not come bundled automatically with learning a historical narrative, no matter how compelling. The project team recognized that if the course was going to develop more complex thinking, then students needed to participate actively in the inquiry, literacy, and discourse practices of the disciplines. Students needed to work on authentic problems, join in substantive, historical and scientific conversations, and engage in disciplinary thinking. To accomplish this, the BHP team and our teachers created and tested lessons that centered on key intellectual problems, ideas, and sources and required attention to students’ emerging ideas and practices. Teachers had to teach using disciplinary-specific models of inquiry, while pressing students for evidence-based narratives, arguments, or explanations.24 Didactic instruction, while useful to convey and help students learn information, was insufficient for the other course goals. In the United States, such ambitious pedagogy has long been the goal of educational reform.25 The BHP has made such teaching a key “selling” point to school districts and administrators. By the project’s second year, it shaped most of the curricular changes while adding disciplinary literacy and thinking demands on teachers. Finally, BHP resources live in a specially designed, online repository with different levels of accessibility for teachers and students. Teachers in the BHP work without a hard-bound textbook or student workbook, the most common instructional resources for teaching history, social studies, and the sciences. In addition to housing all the resources, the BHP course site embeds a specially designed course-management tool that teachers must learn to use to provide students access to the course, administer assessments, and track progress. This technological aspect of teaching the BHP course is another challenge that teachers and their students most learn to navigate. And with differing levels of technology accessibility and savvy, BHP also had to figure out ways to help teachers learn to use the technology well. Learning how to teach big history well while teaching it is a complicated endeavor for both the teachers and for the BHP team. Equally complicated were the imperatives that defined the context within which BHP worked. First, the team needed to provide professional development at no cost to teachers or districts. Since Bill Gates publicly promised that the BHP course would be free forever, the project team tacitly extended that promise to providing professional development to BHP teachers. Second, professional development activities needed to provide support for both beginning and experienced BHP teachers. With each new year, the BHP worked with more first-time teachers and a growing group of “experienced” BHP teachers. Third, the project wanted to provide community support for teachers teaching in 382

Crossing thresholds

a wide-range of contexts, including teachers teaching in under-resourced areas or those without much or any local support or connections.26 BHP’s professional development challenges were daunting and remain so, complicated by research on professional development that improves student learning.27 What type of professional development is most likely to affect how students learn? Research suggests that quality professional development provides teachers structured opportunities to try what they are learning in their classrooms, get and give feedback, and analyze the results of their teaching. To engage in such active learning, then, the professional development activities must be sustained through a semester or a school year and not compartmentalized in a summer workshop or a one-off presentation. The BHP team saw the need to do more than deliver quality curriculum but also wanted to invest in the teachers, professionals who quickly became our partners in co-creating the curriculum and the professional development provided to their peers. While there are many facets to the professional development activities of the BHP, two features stand out as most innovative and effective in helping both the teachers and project cross the learning thresholds needed for effective instruction: the educative curriculum and the online teacher community.

Educative curriculum For almost 60 years, curricular scholars and teacher educators have recognized the potential for a K-12 curriculum to develop teacher learning as well as expanding student learning.28 Ball and Cohen coined the term educative curriculum to describe curricula that would teach teachers as well as students.29 Such a curriculum narrows the gap between the curricular design and enactment by helping teachers acquire the knowledge, understanding, and practices needed to teach the content and skills effectively to all their students.30 Educative materials help teachers deepen their understanding of the content by discussing “alternative representations of the ideas and connections among them,” moving teachers’ understanding of the content beyond what they expect students to learn.31 Further, the curriculum would make teachers aware of “students’ probable ideas about the content at hand, and about the trajectories of their learning that content” and the range of ways students might respond or make meaning of specific curricular material.32 Finally, Ball and Cohen argue that the curriculum should “expand teachers vision about where lessons fit in larger learning progression and how units relate to each other during a course, a year, or larger.”33 From the beginning the BHP team realized the student-facing resources would also play a role in educating teachers about key scientific and historical concepts. The project team also realized the role teacher-facing materials could play in helping teachers use the inquiry and literacy practices of the disciplines and situate pedagogical moves in relationship to changes in students understanding and thinking. Slowly, the team has added new material to help teachers navigate the more ambitious features of the BHP instructional goals. For example, the teacher-facing pages in the Investigations (discussed above) did more than provide teachers with the instructional steps in teaching each Investigation. It explained to teachers some of the reasons why the specific inquiry problem has been of interest and concern to the scholarly community, thus helping teachers 383

Robert B. Bain

locate the work within the world outside the course and outside schools.The teacher instructions also suggested some research-based speculations about the challenge students might face in doing the work and offered tools to help students manage those challenges (i.e., reading scaffolds; pre-writing activities). Beyond a procedure for teaching Investigations, the material presented rationales and perspective to help teachers locate the work in what came before and what will come after. Teachers report that this has helped them increase their understanding of the content that students are studying, how that content fits into the larger course structure, and possible, if not probable, student reactions.34 In another example, the BHP team created a series of lessons to help teachers use the BHP writing rubric and improve students’ capacity to write well-crafted arguments. Judiciously placed into the course at key points, the new lessons focus teachers’ attention on a probable learning progression to “grow” students’ skill in creating a well-structured, well-supported and carefully analyzed essay. In addition to providing activities, the BHP team provides information for teachers to help them see how each activity is situated within a temporal frame and this is making visible the pedagogical theory of learning. Such was also true for how the curriculum helped teachers use BHP’s newly designed lessons to deepen student thinking about causation. In attending to teachers’ learning as well as the students’ learning, the new lessons had short discussions of possible challenges students might face in making causal claims. Without overwhelming teachers, these short, research-grounded discussions alert teachers to possible misconceptions or obstacles to learning. In addition, the teacher materials explain the trajectory of activities ranging from helping students see causal relationships to mapping causation to making their own complex, causal claims using disciplinary language. Just as David Christian’s 18-minute TED talk gives students (and teachers) a big picture they might use to organize the subsequent pieces of the puzzle they study, so the teacher-facing materials offer a big picture of the learning sequence. And like the content big pictures, these help teachers connect individual lessons to a bigger instructional narrative or the students’ learning story. To varying degrees, these examples fit the research discussions of educative curriculum in that they explicitly seek to help teachers develop and apply a deeper understanding of the content or skills at hand, students’ likely initial ideas and learning trajectories about that content, and ways specific lessons within a learning trajectory and the larger temporal frame. Such intentional educative curriculum has helped teachers mediate the challenges that seem to come with teaching unfamiliar content, familiar content in unfamiliar contexts, little familiarity with how students learn unfamiliar science or history content, and the ambitious inquiry, disciplinary thinking and literacy practices that define big history and the BHP.

Online professional development activities In the projects third year, the BHP team launched an online professional community. This community has new and veteran teachers find their footing, allowed for the exchange ideas and resources, and enabled the project team to expand its reach to teachers.35 While there are many facets to the online community, three stand out 384

Crossing thresholds

for helping teachers manage the course and advance student learning: (1) embedded messaging, (2) the Yammer community, and (3) public blog posts. First, BHP embedded a message system into the course site for teachers. BHP team can deliver messages, reminders, or directions to all teachers as they log on to the course site. While essentially informational, the team can point teachers to ­under-used or utilized lessons or activities, or cue teachers to important instructional information. Every time a teacher logs onto the course site, they see most recent updates or messages from the project team. The messaging system is two-way since any teacher can message the team using a dedicated connection. Teachers use this to get the help they need when they need it, particularly about navigating the course site or materials. Second, the team created and maintains a dynamic online community, open only to BHP teachers, using the Yammer platform.There are over 40 special interest groups in which teachers can post messages and from which they can receive messages.36 The Yammer community provides teachers with the chance to post specific questions, lesson plans, lesson modifications, or links to new resources and then have exchanges around their questions or posts. Teachers regularly use Yammer to figuratively open the door to their classrooms. They ask questions of the community regarding an upcoming activity, post about a lesson’s success or shortcoming, explain how they modified a lesson to fit their teaching context, or upload new materials, videos, or links for others to use. The discussions are free-flowing and unstructured, though member of the project team seed discussions with questions or new resources. A group of teacher leaders maintain an active presence, frequently posting, answering questions, or making visible how they are teaching the course. Their activity is great help to “lurkers”, the teachers who are active followers of the Yammer conversations though rarely post.  To keep the conversations and knowledge discussed fresh and growing, BHP has monthly “expert” conversations with scholars and experienced teachers, engaging in online discussions with the community. Yammer also is tied to an online induction program for brand new teachers to BHP, who introduce themselves to the community immediately and get veteran teachers for online mentoring should they seek it. While Yammer merits more substantive analysis, it comes close to providing participating teachers with “just-in-time” contact with more experienced and informed teachers. Through Yammer, BHP teachers get access to specific ideas, advice, and materials about teaching specific big history content when they need it. The asynchronous aspect of this online community enables teachers to dig into a conversation when most relevant to them. In my very brief analysis of some of these conversations, I found examples where a teacher posted an activity and then six months later another teacher tried it, offered praise, and modification to which the originator responded even later. In many ways, this community is a living example of collective learning, a lab for the testing, and the dissemination of ideas about big history and teaching big history. The BHP Teacher Blog is a more formal extension of collective learning since it entails longer, more polished publications by BHP teachers, the project team, or scholars than lives on Yammer. The BHP team solicits, edits, and publishes the blog posts related to teaching big history.37 Blog topics range from teachers’ reflections on teaching the course to using data images in class to some of the ways they are 385

Robert B. Bain

documenting learning to extensions they’ve added to the curriculum. Occasionally a research scholar will share a new study on learning for teachers, replicating the journal clubs that define so much professional learning in medicine and law. In short, BHP has surrounded BHP teachers with a community of support through the educative curriculum or the online community. Utilizing the virtue of the online environment that enables the team to be responsive to teachers quickly and the value of teacher-to-teacher exchanges, BHP has made teacher collective learning an essential feature of its course in big history. Both the educative curriculum and all the facets of the online community provide teachers with access to information, guidance, ideas, and feedback when they need it, when they ask for it or want it. It has enabled the team to support the vast and growing community of BHP teachers.

Conclusion The two examples I have discussed are but a modest sample of how the BHP team added ambitious goals and used research to design and test lessons to help teachers and student reach those goals. I could have discussed in more detail the reading and writing goals, the lesson sequence the team design to achieve those learning objectives, or the data to indicate the growth in students their ability to make coherent arguments, use evidence, and apply disciplinary concepts to think about new problems. Or I might have discussed teacher learning beginning with the 90% of the BHP teachers who report that they have improved their skills by teaching the course.38 Rather, I want to close pointing out (again) that the BHP design team merely seized the opportunity offered by Christian’s compelling, well-researched, and ­coherent modern origin story. Unlike most history or science courses without an over-arching story or through line, the big history narrative itself went a long way in helping students manage the fragmentation that plagues most courses and schooling. The story of the Thresholds of Increasing Complexity solved the “knowledge-­inpieces” problem beautifully, enabling the project team to turn to ratcheting up the learning expectations of what a course in big history might offer students and teachers. Big history gave us, all of us, the space in which we could explore what might be educationally possible. It turns out, there is a lot that is possible. Over these past few years, I have come to understand how in many ways we re-purposed the concepts central to the big history narrative– Thresholds of Increasing Complexity, Goldilocks Conditions, Collective Learning – to think about ways to assist so many students and teachers to expand not only their understanding of the Universe and change over largest scales of time and space, but also to expand their capacity to engage in disciplinary thinking and disciplinary literacy. We realized that big history’s Thresholds of Increasing Complexity helped us see the thresholds of increasingly complex learning, the intellectual progressions we sought as student outcomes and added to the course. It required scale shifting on our part as we identified and connected the different scales at which instruction and learning occurred from the micro interactions of individual students with individual texts to a series of lessons building upon each other over a semester or a year. The concept of collective learning helped us to see what was happening in the online community and encouraged us to develop new networks to increase the interactions. 386

Crossing thresholds

Big history, then, enabled us to expand what we could accomplish by using big history to frame an approach to “big” teaching and learning, to support, stimulate, and enable teachers and students to transcend the limits of their own classroom, own community, or their own experience.

Notes 1 Paul Cobb, et al., “Design Experiments in Educational Research,” Educational Researcher 32, no. 1–2 (2003): 9–13. 2 The project’s design teams have been quite small. During the pilot years (2011–2014), the team included David Christian and Tracy Sullivan of Macquarie University, Bob Bain and Tamara Shreiner of the University of Michigan, Greg Amrofell and Mike Dix of Intentional Futures, and Andy Cook of bgC3. Also, the first seven pilot teachers in the United States were instrumental in the design of the course. Beginning in 2013, Bob Regan joined Andy Cook at bgC3 to take the lead on the BHP course. Bain stayed on as a consultant to the project though Amrofell, Dix, and Shreiner ended their active work in 2014. 3 David Christian, “The History of the World in 18 Minutes,” TED Conference (Long Beach, California, 2011). https://www.ted.com/talks/david_christian_big_history, accessed June 2011. 4 David Christian, “The Case for Big History,” Journal of World History 2, no. 2 (1991): 223–238; David Christian, Maps of Time: An Introduction to Big History, California World History Library (Berkeley: University of California Press, 2005); David Christian, Origin Story: A Big History of Everything (New York: Little Brown and Company, 2018). 5 The BHP can be accessed at www.bighistoryproject.com. 6 BHP began with seven pilot teachers in the United States and three in Australia in 2011–2012. These ten teachers gave regular and active feedback to the design team, working quite closely to co-create the course. In the 2012–2013 school year, the pilot teacher group increased to over 60 teachers in the US and Australia. 7 See, for example, John Dewey, “The Child and the Curriculum,” in John Dewey on Education: Selected Writings., ed. Reginald D. Archambault (New York: Modern ­Library, 1902/1964), 339–358; Jerome S. Bruner, The Process of Education (Cambridge: ­Harvard University Press, 1960); John D Bransford, et al., How People Learn: Brain, Mind, ­Experience, and School (Washington, DC: National Academy Press, 2000); John Bransford and Suzanne Donovan, eds., How Students Learn History, Mathematics, and Science in the Classroom (Washington, DC: National Academy of Sciences, 2005). 8 David Christian, Big History: The Big Bang, Life on Earth, and the Rise of Humanity, The Great Courses (2008) at https://www.thegreatcourses.com/courses/big-history-thebig-bang-life-on-earth-and-the-rise-of-humanity.html. 9 Andrew Ross Sorkin, “So Bill Gates Has This Idea for a History Class,” New York Times Magazine: Education Edition, 7 September 2014. 10 A strong indication of the how data, and feedback from teachers and students helped create the course appeared in the first year when the pilot teachers criticized the original 20-unit structure as being too fragmented and compartmentalized. The US pilot teachers worked closely with the design team to figure out a way to merge units, reducing the number to 10, where it has remained since 2012. 11 In the first few years, the University of Michigan worked with a group of teachers and graduate students to evaluate the papers. More recently, a team of graduate students at Arizona State University evaluated the papers. 12 E.H. Carr, What is History? (London, New York: Macmillan; St. Martin’s Press, 1961), 81–82. 387

Robert B. Bain

1 3 See, for example, John Dewey, “The Child and the Curriculum,” in John Dewey on Education: Selected Writings, ed. Reginald D. Archambault (New York: Modern Library, 1902/1964), 339–358; John D. Bransford, et al., How People Learn: Brain, Mind, Experience, and School (Washington, DC: National Academy Press, 2000); John Bransford and Suzanne Donovan, eds., How Students Learn History, Mathematics, and Science in the Classroom (Washington, DC: National Academy of Sciences, 2005). 14 See, for example, James F. Voss and Mario Carretero, “Cognitive and Instructional Processes in History and the Social Sciences” (Hillsdale, NJ: Lawrence Erlbaum ­A ssociates, 1994), 1–14; Gaea Leinhardt, Teaching and Learning in History, eds. G. Leinhardt, I.L. Beck, and C. Stainton (Hillsdale, NJ: Lawrence Erlbaum Associates, 1994), 209–255; Peter N. Stearns, Peter. Seixas, and Samuel Wineburg, Knowing, Teaching & Learning History: National and International Perspectives (New York: New York University Press, 2000); Sam Wineburg, Historical Thinking and Other Unnatural Acts: Charting the Future of Teaching the Past (Philadelphia, PA: Temple University Press, 2001); M. Gail Jones, et al., “Experienced and Novice Teachers Concepts of Spatial Scale,” International Journal of Science Education 30, no. 3 (2007): 409–429; Robert B. Bain, “Challenges of Teaching and Learning World History,” in A Companion to World History, ed. Douglas Northrop (Cambridge: John Wiley & Sons Ltd., 2012), 111–127. 15 For example, Frederick J. Teggart, “Causation in Historical Events,” Journal of the History of Ideas 3, no. 1 (1942): 3–11; Edward Hallett Carr, What Is History? (London, New York: Macmillan; St. Martin’s Press, 1961); Richard J. Evans, In Defense of History (New York: W.W. Norton, 1999); Louis O. Mink, Historical Understanding (Ithaca, NY: Cornell University Press, 1987); Michael Stanford, The Nature of Historical Knowledge (­Oxford, New York, NY: Blackwell, 1990); Charles. Tilly, Why? (Princeton, NJ: Princeton University Press, 2006); Tina A. Grotzer, Understanding of Consequences: Learning About Causality in a Complex World (Lanham, MD: Rowman and Littlefield, 2012) 16 Mario Carretero, Asunción López-Manjón, and Liliana Jacott, “Explaining Historical Events,” International Journal of Educational Research 27, no. 3 (1997): 245–253; Bjorn Andersson, “The Experiential Gestalt of Causation: A Common Core to ­Pupils ­Preconceptions in Science,” European Journal of Science Education 8, no. 2 (1986): 155–171; Olla Hallden, “On the Paradox of Understanding History in an Educational Setting,” in Teaching and Learning in History, eds. Gaea Leinhardt, Isabel L. Beck, and ­Catherine ­Stainton (Hillsdale, NJ: Lawrence Erlbaum Associates, 1994), 27–46; Peter Seixas, “Conceptualizing the Growth of Historical Understanding,” in Handbook of Education and Human Development: New Models of Learning, Teaching and Schooling, eds. David R. Olson and Nancy Torrance (Oxford, UK: Blackwell, 1996), 765–783; Ola Hallden, “Conceptual Change and Contextualization,” in New Perspectives on Conceptual Change, eds. Wolfgang Schnotz, et al. (Oxford New York: Pergamon, 1999), 53–65; Arthur Chapman, “Camels, Diamonds and Counterfacutals: A Model for Teaching Causal Reasoning,” Teaching History 112 (2003): 46–53; Robert B. Bain, “Challenges of Teaching and Learning World History,” in A Companion to World History, ed. Douglas Northrop (Cambridge: John Wiley & Sons, Ltd, 2012), 111–127; Michelene T. H. Chi, et al., “Misconceived Causal Explanations for Emergent Processes,” Cognitive Science 36, no. 1 (2012): 1–61; Tina A. Grotzer and S. Lynneth Solis, “Action at an Attentional Distance: A Study of Children’s Reasoning About Causes and Effects Involving Spatial and Attentional Discontinuity,” Journal of Research in Science Teaching 52, no. 7 (2015): 1003–1030. 17 Mario Carretero, Asunción López-Manjón, and Liliana Jacott, “Explaining Historical Events,” International Journal of Educational Research 27, no. 3 (1997): 245–253. The events that the researchers used were the fall of the Soviet Union, French Revolution, Spain sails to the New World, and the Second World War. 388

Crossing thresholds

18 The Big History Project, “Understanding Cause and Consequences, Part 1”, Unit 3, Activity 3.0. 19 Ibid. 20 The Big History Project, “Understanding Cause and Consequences, Part 1”, Unit 3, Activity 3.1. 21 The Big History Project, “Categorizing Causes”, Unit 4, Activity 4.0. 22 Arthur Chapman, “Camels, Diamonds and Counterfacutals: A Model for Teaching Causal Reasoning,” Teaching History 112 (2003). 23 While calls for active learning of disciplines and ambitious pedagogy has a long history, recent descriptions of ambitious teaching can be found in Fred M. Newmann, Authentic Achievement: Restructuring Schools for Intellectual Quality (San Francisco: J­ossey-Bass Publishers, 1996); John D. Bransford, Ann L. Brown, and Rodney R. National Research Council Cocking, eds., How People Learn: Brain, Mind, Experience, and School (­Washington, DC: National Academy Press, 2000); Robert B. Bain, “‘They Thought the World Was Flat?: Applying the Principles of How People Learn in Teaching High School History,” in How Students Learn History in the Classroom, eds. John Bransford and Suzanne Donovan (Washington: National Academy of Sciences, 2005), 179–214; ­Robert B. Bain, “Rounding up Unusual Suspects: Facing the Authority Hidden in ­H istory Textbooks and Teachers,” Teachers College Record 108 (2006): 2080–2110; ­Elizabeth Birr Moje, “Developing Socially Just Subject-Matter Instruction: A Review of the Literature on Disciplinary Literacy,” in Review of Research in Education, ed. ­Lawrence Parker (Washington, DC: American Educational Research Association, 2007), 1–44; Magdalene Lampert, Timothy A. Boerst, and Filippo Graziani, “Organizational Resources in the Service of School-Wide Ambitious Teaching Practice,” Teachers College Record 113, no. 7 (2011): 1361–1400; Mark Windschitl, Jessica Thompson, and ­Melissa Braaten, “Ambitious Pedagogy by Novice Teachers: Who Benefits from Tool-­ Supported Collaborative Inquiry Into Practice and Why,” Teachers College Record 113, no. 7 (2011): 1311–1360; Marcy Singer-Gabella, et al., “Learning to Leverage Student Thinking: What Novice Approximations Teach Us About Ambitious Practice,” The Elementary School Journal The Elementary School Journal 116, no. 3 (2016): 411–436. 24 Examples of such instruction can be found in Fred M. Newmann, Authentic Achievement: Restructuring Schools for Intellectual Quality (San Francisco: Jossey-Bass Publishers, 1996); John D Bransford, et al., How People Learn: Brain, Mind, Experience, and School (Washington, DC: National Academy Press, 2000); Robert B. Bain, “‘They Thought the World Was Flat?: Applying the Principles of How People Learn in Teaching High School History,” in How Students Learn History in the Classroom, eds. John Bransford and Suzanne Donovan (Washington, DC: National Academy of Sciences, 2005), 179–214; Robert B. Bain, “Rounding up Unusual Suspects: Facing the Authority ­H idden in History Textbooks and Teachers,” Teachers College Record 108, no. 10 (2006): 2080–2110; Elizabeth Birr Moje, “Developing Socially Just Subject-Matter Instruction: A  Review of the Literature on Disciplinary Literacy,” in Review of Research in Education, ed. Lawrence Parker (Washington, DC: American Educational Research Association, 2007), 1–44; Magdalene Lampert, Timothy A Boerst, and Filippo Graziani, “Organizational Resources in the Service of School-Wide Ambitious Teaching Practice,” Teachers College Record 113, no. 7 (2011): 1361–1400; Mark Windschitl, Jessica Thompson, and Melissa Braaten, “Ambitious Pedagogy by Novice Teachers: Who Benefits from Tool-Supported Collaborative Inquiry Into Practice and Why,” Teachers College Record 113, no. 7 (2011): 1311–1360; Marcy Singer-Gabella, et al., “Learning to Leverage Student Thinking: What Novice Approximations Teach Us About Ambitious Practice,” The Elementary School Journal The Elementary School Journal 116, no. 3 (2016): 411–436. 389

Robert B. Bain

25 Reform urging schools use disciplinary inquiry to teach stretches back at least to 1895 and the Committee of Ten with more recent and relevant examples being the Common Core Literacy Standards, College and Career Readiness Standards, Next Generation Science Standards, and the C3 Framework. 26 For research that has stressed the value of such collective participation see Michael S. Garet, et al., “What Makes Professional Development Effective? Results from a ­National Sample of Teachers,” American Educational Research Journal (2001): 915–945; Laura M. Desimone, et al., “Effects of Professional Development on Teachers’ Instruction: Results from a Three-Year Longitudinal Study,” Educational Evaluation and Policy Analysis, no. 2 (2002): 81–112; Laura M. Desimone, “Improving Impact Studies of Teachers’ ­Professional Development: Toward Better Conceptualizations and Measures,” Educational Researcher 38, no. 3 (2009): 199, 181. 27 Thomas R. Guskey and Michael Huberman, eds., Professional Development in Education: New Paradigms and Practices (New York: Teachers College Press, 1995); Michael S. ­Garet, et al., “What Makes Professional Development Effective? Results from a National ­Sample of Teachers,” American Educational Research Journal 38, no. 4 (2001): 915–945; Laura M. Desimone, et al., “Effects of Professional Development on Teachers’ Instruction: Results from a Three-Year Longitudinal Study,” Educational Evaluation and Policy Analysis, no. 2 (2002): 81–112; Laura M Desimone and Michael S Garet, “Best Practices in Teachers’ Professional Development in the United States,” Psychology, Society and Education 7, no. 3 (2015): 252–263. 28 Jerome S. Bruner, The Process of Education (Cambridge: Harvard University Press, 1960); Jerome S. Bruner, Toward a Theory of Instruction (Cambridge, MA: Belknap Press of ­Harvard University, 1966); Deborah Loewenberg Ball and David K. Cohen, “Reform by the Book: What Is: Or Might Be: The Role of Curriculum Materials in Teacher Learning and Instructional Reform?” Educational Researcher 25, no. 9 (1996): 6–14; ­Elizabeth A Davis and Joseph S Krajcik, “Designing Educative Curriculum Materials to Promote Teacher Learning,” Educational Researcher 34, no. 3 (2005): 3–14; Gabriel J. Stylianides and Andreas J. Stylianides, “The Role of Instructional Engineering in Reducing the Uncertainties of Ambitious Teaching,” Cognition and Instruction Cognition and Instruction 32, no. 4 (2014): 374–415; Corey Drake, Tonia J. Land, and Andrew M. Tyminski, “Using Educative Curriculum Materials to Support the Development of Prospective Teachers’ Knowledge,” Educational Researcher Educational Researcher 43, no. 3 (2014): 154–162; ­Elizabeth Davis et al., “Designing Educative Curriculum Materials: A Theoretically and Empirically Driven Process,” Harvard Educational Review 84, no. 1 (2014): 24–52. 29 Deborah Loewenberg Ball and David K. Cohen, “Reform by the Book: What Is: Or Might Be: The Role of Curriculum Materials in Teacher Learning and Instructional Reform?” Educational Researcher 25, no. 9 (1996): 6–14. 30 Over last 20 years, others have elaborated upon and extended the concept of educative curriculum. Davis and Krajcik, for example, provided a useful set of design heuristics, Davis, et al offered an example of design-based research to construct such educative curriculum, and Bain and Shreiner analyzed examples of how teachers used a less thoughtfully designed curriculum that attempted to be educative. However, for this paper, I think the three features Ball and Cohen offer – attention to student understanding, teacher understanding, and larger learning progressions over time remain the most useful. See Davis and Krajcik, op. cit.; Davis, et al., op. cit.; Robert B. Bain and Tamara Shreiner, The Power and Potential of World History for Us All, Unpublished evaluation study for World History for Us All (2014). 31 Ball and Cohen, op. cit., 7.

390

Crossing thresholds

3 2 Ibid. 33 Ibid. 34 They were necessarily modest in part because the educative curricular rationale has not always been understood by BHP team members and at times, there have been attempts to eliminate or collapse these to reduce the footprint or the amount of information for teachers with which to work. This tension has been under-explored in the literature as well as the BHP work. Indeed, as I’ll note in last section, BHP has little formal data on the impact of professional development or curriculum on teachers’ knowledge, teachers’ practice. 35 While the online community was part of the original design of BHP, its fruition has been the work of Bob Regan. There is not under-estimating both its value and its importance for the success and future of the BHP. 36 Among the many groups are groups for each of the BHP’s 10 units, the IB Program, BHP Science, Little Big Histories, Mentoring and Collaboration, Standards Alignment, Big History Book Club, and then a few groups connected to the Professional Learning Experience for brand new teachers. 37 The blogs can be accessed at https://blog.bighistoryproject.com/. 38 The BHP site publishes its research study each year. I encourage anyone interested in more data about how students are doing in the course and the perceptions of both students and teachers to visit the site to review those reports.

Bibliography Alvarez, Walter. A Most Improbable Journey: A Big History of Our Planet and Ourselves. New York: Norton, 2016. Bain, Robert B. “Rounding up Unusual Suspects: Facing the Authority Hidden in History Textbooks and Teachers.”Teachers College Record108 (2006):2080–2110. ———. “They Thought the World Was Flat? Applying the Principles of How People Learn in Teaching High School History.” In How Students Learn History in the Classroom, edited by John Bransford and Suzanne Donovan,179–214. Washington: National Academy of Sciences, 2005. Bain, Robert B., and Tamara Shreiner. The Power and Potential of World History for Us All. Unpublished evaluation study for World History for Us All, 2014. Ball, Deborah Loewenberg, and David K. Cohen. “Reform by the Book: What Is: Or Might Be: The Role of Curriculum Materials in Teacher Learning and Instructional Reform?”Educational Researcher25, 9 (1996):6–14. Bransford, John D., Ann L. Brown, and Rodney R. National Research Council Cocking, eds. How People Learn: Brain, Mind, Experience, and School. Washington, DC: National Academy Press, 2000. Brown, Cynthia Stokes. Big History: From the Big Bang to the Present. New York: New Press, 2007. Bruner, Jerome S. The Process of Education. Cambridge: Harvard University Press, 1960. Bryson, Bill. A Short History of Nearly Everything. New York: Broadway Books, 2003. Carretero, Mario, Asunción López-Manjón, and Liliana Jacott. 1997. “Explaining Historical Events.”International Journal of Educational Research27, no.3:245–253. Chaisson, Eric J. Cosmic Evolution: The Rise of Complexity in Nature. Cambridge, MA: ­Harvard University Press, 2001. Christian, David. Big History: The Big Bang, Life on Earth, and the Rise of Humanity. ­Chantilly VA, The Great Courses, Teaching Company Production, 2008. ———. “The Case for ‘Big History’.”The Journal of World History 2, no.2 (1991):223–238.

391

Robert B. Bain

———. Maps of Time an Introduction to Big History. Berkeley: University of California Press, 2004. ———. “The Return of Universal History,” History and Theory 49, no.4 (2010):6–27. Davis, Elizabeth A., and Joseph S. Krajcik. “Designing Educative Curriculum Materials to Promote Teacher Learning.”Educational Researcher34, no.3 (2005):3–14. Davis, Elizabeth, Annemarie Sullivan Palincsar, Anna Maria Arias, Amber Schultz ­Bismarck, Loren Marulis, and Stefanie Iwashyna. “Designing Educative Curriculum Materials: A Theoretically and Empirically Driven Process.”Harvard Educational Review84, no.1 (2014):24–52. Desimone, Laura M. “Improving Impact Studies of Teachers’ Professional Development: Toward Better Conceptualizations and Measures.”Educational Researcher 38, no.3 (2009):199, 181. Desimone, Laura M., and Michael S. Garet. “Best Practices in Teachers’ Professional Development in the United States.”Psychology, Society and Education 7, no.3 (2015):252–263. Desimone, Laura M., Andrew C. Porter, Michael S. Garet, Kwang Suk Yoon, and B ­ eatrice F. Birman. “Effects of Professional Development on Teachers’ Instruction: Results from a Three-Year Longitudinal Study.”Educational Evaluation and Policy Analysis, no.2 (2002):81–112. Dewey, John. “The Child and the Curriculum.” In John Dewey on Education; Selected Writings, edited by Reginald D. Archambault. New York: Modern Library, 1902/1964: 339–358 Drake, Corey, Tonia J. Land, and Andrew M. Tyminski. “Using Educative Curriculum Materials to Support the Development of Prospective Teachers’ Knowledge.”Educational Researcher Educational Researcher 43, no.3 (2014):154–162. Garet, Michael S., Andrew C. Porter, Laura Desimone, Beatrice F. Birman, and Kwang Suk Yoon. “What Makes Professional Development Effective? Results from a National Sample of Teachers.”American Educational Research Journal, 2001: 915–945. Guskey, Thomas R., and Michael Huberman, eds. Professional Development in Education: New Paradigms and Practices. New York: Teachers College Press, 1995. Hesketh, Ian. “The Story of Big History,”History of the Present 4, no.2 (2014):171–202. Lampert, Magdalene, Timothy A. Boerst, and Filippo Graziani. “Organizational Resources in the Service of School-Wide Ambitious Teaching Practice.”Teachers College Record113, 7 (2011):1361–1400. Megill, Allan. “‘Big History’: Old and New: Presuppositions, Limits, Alternatives,”Journal of the Philosophy of History 9, 2 (2015):306–326. Moje, Elizabeth Birr. “Developing Socially Just Subject-Matter Instruction: A Review of the Literature on Disciplinary Literacy.” In Review of Research in Education, edited by Lawrence Parker,1–44. Washington, DC: American Educational Research Association, 2007. Moje, Elizabeth Birr, and Jennifer Speyer. “The Reality of Challenging Texts in High School Science and Social Studies: How Teachers Can Mediate Comprehension.” In Best Practices in Adolescent Literacy Instruction, edited by Kathleen A. Hinchman and Heather K. Sheridan-Thomas, 185–211. New York: The Guilford Press, 2008. Newmann, Fred M. Authentic Achievement: Restructuring Schools for Intellectual Quality. San Francisco: Jossey-Bass Publishers, 1996. Singer-Gabella, Marcy, Barbara Stengel, Emily Shahan, and Min-Joung Kim. “Learning to Leverage Student Thinking: What Novice Approximations Teach Us about Ambitious Practice.”The Elementary School Journal, 116, no.3 (2016):411–436. Sorkin, Andrew Ross. “So, Bill Gates Has This Idea for a History Class.”New York Times Magazine, The Education Issue, 6 September 2014, http://www.nytimes.com/2014/09/07/ magazine/so-bill-gates-hasthis-idea-for-a-history-class.html Spier, Fred. The Structure of Big History from the Big Bang until Today. Amsterdam: A ­ msterdam University Press, 1996. 392

Crossing thresholds

———. “Big History: The Emergence of a Novel Interdisciplinary Approach,”Interdisciplinary Science Reviews 33, no. 2 (2008): 141–152. ———. Big History and the Future of Humanity. Malden, MA: Wiley-Blackwell, 2010. Stylianides, Gabriel J., and Andreas J. Stylianides. “The Role of Instructional Engineering in Reducing the Uncertainties of Ambitious Teaching.”Cognition and Instruction 32, no.4 (2014):374–415. Windschitl, Mark, Jessica Thompson, and Melissa Braaten. “Ambitious Pedagogy by ­Novice Teachers: Who Benefits from Tool-Supported Collaborative Inquiry into Practice and Why.”Teachers College Record113, no.7 (2011):1311–1360.

393

PART V

Big history and the future

18 BIG HISTORY AND THE FUTURE OF TECHNOLOGY Leonid E. Grinin and Anton L. Grinin

Introduction: between human and post-human revolutions This contribution is devoted to the history of technological development in terms of production (or technological) revolutions. The analysis of its current state and forecasts is made with respect to big history. Technologies have been playing a significant role in the history of humankind from the very origin of Homo sapiens. Numerous facts show that already after 50,000 BP technologies were developed in various fields: from hunting and cooking to primitive painting. Agriculture, building, transportation, and many other human achievements could not have happened without certain technologies. Thus, one can argue that technologies play a very important role in Big History. They played a special role in the collective learning which is defined as the sixth threshold of increasing complexity. This Homo Sapiens’ achievement which happened at the beginning of the Upper Paleolithic was probably one of the most important events in the whole human history, and sometimes is termed as the Human revolution (e.g., Shea 2006).1 Today we are at the threshold of another important transition which is often called ‘post-human revolution’, which could bring quite radical changes to society and even transform the human biological nature.

Technological dimension of big history Three production revolutions – three big history thresholds The whole historical progress can be divided into four big periods which we denote as four production principles2: (1) (2) (3) (4)

Hunter-Gatherer3; Craft-Agrarian; Trade-Industrial; Scientific-Cybernetic. 397

Leonid E. Grinin and Anton L. Grinin

Each production principle starts with great technological breakthrough which we denote as production revolution.There were three such revolutions: (1) the Agrarian or Neolithic Revolution (12,000/10,000–5,500/3,000 BP); (2) the Industrial Revolution (the last third of the fifteenth to the first third of the nineteenth centuries); (3) the newest Cybernetic one (1950 to the 2060/2070s). In respect of big history, these revolutions are tightly related to the main big history thresholds (the Agricultural Threshold; Modern Revolution Threshold; and the Cybernetic Revolution are related to the Future Ninth Threshold4). Production revolutions are technological breakthroughs which change the whole structure of society and the way of life. Each production revolution has its own cycle consisting of three phases: two innovative phases and between them – a modernization phase (see Tables 18.1, 18.2; Figure 18.1). At the initial innovative phase a new productive sector emerges. The modernization phase is a long period of distribution and development of innovations. It is a period of progressive innovations when the conditions gradually emerge for the final innovative breakthrough. At the final innovative phase new innovations dramatically spread and improve for the new production principle, which, at this time, attains full strength. The Agrarian Revolution was a great breakthrough from hunter-gatherer production principle to farming. Its initial phase was a transition from hunting and gathering to primitive hoe agriculture and animal husbandry (that took place around 12,000–9,000 BP). The final phase was a transition to intensive agriculture (with large-scale irrigation and plowing) which started around 5,500 years ago (for more details, see Grinin 2007a; Grinin A. and Grinin L. 2015; Grinin L. and Grinin A. 2016). These changes are also presented in Table 18.1. The Industrial Revolution was a great breakthrough from craft-agrarian production principle to machine industry, marked by intentional search for and use of scientific and technological innovations in the production process. Table 18.1  The phases of the Agrarian Revolution Phases

Type

Name

Dates

Changes

Initial

Innovative

Manual agriculture

12,000–9,000 BP

Middle

Modernization

Diffusion of agriculture

9,000–5,500 BP

Final

Innovative

Irrigated and plow agriculture

5,500–3,500 BP

Transition to primitive manual (hoe) agriculture and cattle-breeding Emergence of new domesticated plants and animals, development of complex agriculture, emergence of a complete set of agricultural instruments Transition to irrigative or plow agriculture without irrigation

398

Big history and the future of technology

Table 18.2  The phases of the Industrial Revolution Phases

Type

Initial

Innovative

Name of the phase

Innovative

Changes

Manufacturing the last third of the fifteenth to sixteenth centuries

Middle Modernization Diffusion of industrial enterprises Final

Dates

Machinery

Development of shipping, technology and mechanization on the basis of water engine, development of manufacture based on the division of labor and mechanization Seventeenth to Formation of complex industrial early eighteenth sector and capitalist economy, centuries increase in mechanization, and division of labor 1730–1830s Formation of sectors with the machine cycle of production using steam energy

Its initial phase started in the last third of the fifteenth and sixteenth centuries with the development of shipping, technology, and mechanization based on the watermill as well as with a ‘more organic’ division of labor.The final phase was the well-known breakthrough of the eighteenth and nineteenth centuries with the introduction of various machines and steam energy (for more details about Industrial Revolution, see Grinin 2007b; Grinin A. and Grinin L. 2015; Grinin L. and Grinin A. 2016; Grinin and Korotayev 2015a). These changes are presented in Table 18.2. The Cybernetic Revolution is a great breakthrough from industrial production to the production and services based on self-regulating systems. Its initial phase dates back to the 1950–1990s. The breakthroughs occurred in the spheres of automation, energy production, synthetic materials production, space technologies, exploration of space and sea, agriculture, and especially in the development of electronic control facilities, communication and information. We assume that the final phase will begin in the nearest decades, that is in the 2030s or a bit later, and will last until the 2070s. We denote the initial phase of the Cybernetic Revolution as a scientific-­ information one, and the final – as a phase of self-regulating systems. Today we are in its modernization phase which will probably last until the 2030s. This intermediate phase is a period of rapid distribution and improvement of the innovations made at the previous phase (e.g., computers, Internet, cell phone, etc.). The technological and social conditions are also prepared for the future breakthrough. We suppose that the final phase of the Cybernetic Revolution will lead to the emergence of various self-regulating systems (see below). The scheme of the Cybernetic Revolution is presented in Figure18.1. Each phase of big history is accompanied by the emergence of new evolutionary mechanisms. In particular, certain preconditions and preadaptations can be already detected within its previous phase. The same refers to the development of productive forces. Within previous production principle, some prerequisites of technologies 399

Leonid E. Grinin and Anton L. Grinin

Figure 18.1  The phases of the Cybernetic Revolution.

appear, which then flourish during the next production principle. For example, many mechanisms, engines and machines emerged within the Craft-Agrarian production ­principle, especially during its last centuries (the twelfth to fourteenth centuries). But for production revolution to start technological changes should occur. Thus, the Industrial production revolution began at the end of the fifteenth century and lasted until 1830.

Cybernetic revolution, self-regulation, and artificial intelligence in terms of big history The theory of production revolutions proceeds from the assumption that the essence of these revolutions can be clearly observed only during their final phase. The most important thing about the final phase of Cybernetic Revolution will be a wide use of self-regulation in different technological and bio-socio-technological systems. The analysis of such systems can be based on cybernetics which is a transdisciplinary approach for exploring complex regulatory systems via the processes of receiving, transformation and transfer of information (see, e.g., Wiener 1948; Beer 1959;Von Foerster and Zopf 1962; Umpleby and Dent 1999). The most important characteristics and trends of Cybernetic Revolution are the following: (1) (2) (3) (4)

The increasing amount of information and growing complexity; Consistent development of the system’s abilities to the regulation and self-regulation; Mass use of artificial materials with new properties; Application and control of systems and processes of various nature including living material and new levels of organization of matter (including different nanoparticles as building blocks); 400

Big history and the future of technology

(5) Miniaturization and microtization as a trend of the constantly decreasing mechanisms, electronic devices, implants, etc.; (6) Ubiquitous resource and energy saving; (7) Individualization as one of the most important technological trends; (8) Implementation of smart technologies and a trend towards humanization of their functions (use of the human language, voice, movements, etc.); (9) Control over human behavior and activity to eliminate the negative influence of the so-called human factor. Some of these trends coincide wholly or partially with the perceptions of artificial intelligence (AI) and its future development (though this concept is quite vague and difficult to define). But other trends cannot be included into the concept of AI (for more details see below). Self-regulation can be defined as a system’s ability to preserve stability and basic parameters within changing environment. Self-regulation as a broad concept incorporates various aspects of maintaining stable state of a system at all phases of big history and especially at the biological and social ones. Self-regulation is of great importance for big history since it is one of the most developed levels of growing complexity (see Grinin 2016). Self-regulation has already revealed at the early phases of big history, in fact, with the emergence of the first systems (e.g., the first stars). The emergence of life is tightly connected with self-regulating systems. In the course of chemical evolution, chemical substances gradually became more complex until some of them got the ability for self-regulation. For example, lipids, which are able to change their form when interacting with water, while retaining its chemical structure. One of the most important features of living organisms is an existence of a code molecule. RNA is considered as the first self-replicating molecule. The further formation of complex systems, such as DNA, proteins, enzymes, etc., required the creation of a complex system of regulation. The more complicated the system became, the more complicated was its regulation. In order to overcome the entropy, systems sought to isolate themselves from direct contacts with the environment, forming protective (insulating) shells. Presumably that is how the first coacervates were formed, and later – the cells. A cell became the main self-­regulating living system due to which many organisms were formed in the process of evolution. Biological systems demonstrate the complexity growing up to the level of self-regulation within evolution. Due to collective learning human society has also developed into a complex self-regulating system. During the next decades the technological complexity is supposed to rapidly increase thus promoting the ability for self-regulation. At present there are already many self-regulating systems around us, for example, self-driving cars, the artificial Earth satellites, pilotless planes, navigators laying the route for a driver. Another good example is life-supporting systems (such as medical ventilation apparatus or artificial heart). They can regulate a number of parameters, choose the most suitable mode and detect critical situations. It is also worth mentioning the genetic engineering which is used for the creation or changing biological and physiological self-regulating systems. 401

Leonid E. Grinin and Anton L. Grinin

We suppose that during the final phase of the Cybernetic Revolution different developmental trends should produce a cluster of technological innovations. The medical sphere has unique opportunities to combine the abovementioned technologies into a single complex. In our opinion, the general driver of this cluster will be medicine, which can connect additive technology, nanotechnology, biotechnology, robotics, information and cognitive technologies.We denote this technological cluster as a MANBRIC-complex (an acronym for the included technologies).5 Figure 18.2 shows the citation frequency of MANBRIC-technologies in scientific publications and relations between the technologies forming the complex. The thickness of line demonstrates the intensity of interactions while the direction of arrows shows the sphere of application of technologies. As one can see from Figure 18.3, medicine and biotechnologies are most closely related. There is also distinguished a separate direction of biomedicine (Pankhurst et al. 2003; Gupta A. and Gupta M. 2005). The development of MANBRIC-complex can be examined, for example, via the analysis of patent applications in medicine, pharmaceuticals, and biotechnologies which also demonstrate converging growth rates (Grinin, Grinin, and Korotayev 2016).6 An important question is in what sphere will the final phase of the Cybernetic Revolution start? First of all, one should remember that the ‘breakthrough’ sphere is usually quite narrow as it happened during the Industrial Revolution (when the breakthrough occurred in a specific field – cotton industry). In a similar way, given

Figure 18.2  T  he relationship between citation frequency in scientific publications and the technologies forming MANBRIC, according to the Web of Science, 2010–2015. 402

Big history and the future of technology

Figure 18.3  D  ynamics of the global combined share of four technologies with the highest share of patent applications in 1985 (electrical machinery, measurement, machine tools, and other special machines) in comparison with the dynamics of the global combined share of patent applications in four top categories (medical, pharmaceutical, computer, and biotechnologies), 1985–2014. Source: WIPO IP Statistics Data Center 2016.

the general vector of scientific achievements and taking into account that a future breakthrough area should be commercially attractive, we think that the final phase (the one of self-regulating systems) of the Cybernetic Revolution will begin in one of the newest branches of medicine. It probably has already formed (such as biomedicine or nanomedicine) or it can form as a result of the uptake of other innovative technologies into medicine. It is important that in the nearest decades not only the developed but also developing countries will face the problems of population ageing, shortage of labor resources and the necessity to support a growing number of elderly people. The progress in medicine can contribute to the extension of working age (as well as to the general increase of the average life expectancy) of elderly people and to more active involvement of disabled people into labor activities. Thus, elderly people and people with disabilities could more and more subsist for themselves. At present medicine is closely related to biotechnologies (see Figure18.2) through pharmaceuticals, gene technologies, new materials, etc. The distinctive feature of modern medical science is its ‘bio-related trends’ – a wide use of approaches based on the methods of molecular and cell biology. At present medicine is highly computerized especially in the field of diagnostics, various automatic control systems have been developed; for example, for the control 403

Leonid E. Grinin and Anton L. Grinin

of breathing, nutrient supply to specific organs, blood pressure, control over the functioning of some internal organs, etc. Medicine (supported by both government and private funding) has been a major influence on gross domestic product (GDP). Taking into consideration the anticipated faster growth rates of GDP in the developing countries and a rapid formation of the middle class there, one can suppose that in general, spending on health care will increase significantly. For example, in Germany a number of health care personnel constitute 22% of the total number of employed people while the share of automobile industry is only 2.3% (Nefiodow L. and Nefiodow S. 2014). We have no opportunity to describe the whole range of MANBRIC-­technologies with the equal attention. So in this paper, we will focus on the most important spheres.

Future technologies Big history, technologies, and rules of evolution When considering the issue of future technologies in terms of big history, one should emphasize that the growing technological complexity is connected with some other aspects of big history. Elsewhere we formulated a number of evolutionary rules, which can be applied for the analysis of different big history directions (e.g., Grinin, Markov, and Korotayev 2008, 2009; Grinin 2016; Grinin 2014; Grinin, ­Korotayev, and Markov 2017). Among these evolutionary rules, we single out three rules which are of particular importance for the development of technologies.

Rule 1.  Evolution occurs only in a small part of a system According to data obtained from Planck observatory, the Universe is composed of 5% of ordinary (baryonic) matter, 24% of dark matter, and 70% of dark energy. Thus, the most bulk of our Universe is occupied by dark matter and energy which can hardly evolve. In living organisms, for example, an estimated percentage of the non-coding DNA reaches 98%. In social evolution, according to some sources, the number of innovators in a society is about 3–5%. The same is in evolution of technology, for example, only a small number of startup projects appear successful.

Rule 2.  Evolutionary block assemblage In evolution, there emerge some basic and more complex components which assemble in various combinations. In this sense, evolution is similar to construction, where ready-made units are used to build new creations. Thus, in cosmic evolution atoms are universal components for the formation of molecules while chemical evolution in space started with the emergence of a sufficient variety of atoms. On Earth the atoms and non-organic molecules launched the geological development, and later – the emergence of organic molecules, and eventually life. In biological evolution block 404

Big history and the future of technology

assemblage principle of formation can be observed, for example, at the level of cells, tissues, and organs. Many molecules, e.g., of DNA, also consist of peculiar semantic units. Similarly in social evolution, religion or legal systems are often borrowed by other countries. All technologies are made according to this principle. For instance, a modern vehicle is a result of numerous technological achievements: from wheel, alloys, mechanical systems, plastic, fabrics to the onboard computer.

Rule 3.  The increasing diversity Variety is as a universal evolutionary trend. Thus, within cosmic evolution there appeared a growing number of chemical elements and molecules, as well as stars and planets. In biological evolution the number of species has been continually increasing for a long time. However, the growth cannot keep going constantly since evolution always balances around the optimum.Thus, it is not surprising that there are periods of reverse development and reduction of the diversity (e.g., during mass extinctions). In social evolution, there is a growing diversity of political forms, cultures, and religions. In the evolution of technologies the growing diversity is very impressive. From 1980 to 2014 the number of patent applications has increased ten-folds.7 The rules given above are used as examples to demonstrate that technological development we observe nowadays is not unique. Under similar conditions and preconditions, evolution in different systems may proceed the following similar patterns. Thus, based on the evolutionary rules, the theory of production principles and other aspects we give some forecasts of upcoming technological revolution.

On some future medical technologies Constant health monitoring as a self-regulating supersystem. During the final phase of the Cybernetic Revolution a very important direction of self-regulation can develop as different health monitoring systems for early diagnosis and preventing diseases. The key compounds of such devices are biosensors and similar tiny devices. One can easily imagine that in the future they will be able to become an integral part of a human life, providing a constant scanner of an organism or a certain organ and transmitting the information to a medical center in case of potential or real threats. On the whole, medicine will develop towards increasing individualization and personification through the selection for individual therapy while the use of mass drugs and standard therapeutic technologies will be reduced. Economy, optimization of resource consumption, and miniaturization. The achievements in medicine will make a significant contribution to the optimization of resource consumption, for example, due to the targeted drug delivery and minimization of interference with the organism. Hospital treatment will be less used as the operations will be more targeted, and the rehabilitation period will be minimal. More people will be treated at home since the development of remote treatment is rather probable when doctors control the indices of a patient online and can make the necessary prescriptions remotely. It could sharply decrease a cost of medical treatment which now is exorbitant one for a great number of people. Saving money (as well as resources) is one of the most important directions for the economy. 405

Leonid E. Grinin and Anton L. Grinin

Medicine develops in the direction of growing miniaturization. There is a trend of constantly decreasing size of instruments to micro- and nanoscale. For example, repairing heart tissue destroyed by a heart attack usually requires invasive open-heart surgery. But now researchers have developed a technique that allows using a small needle to inject a repair patch, without opening up the chest cavity (Montgomery et al. 2017). The perspective direction in medicine is slowing down the ageing process. It is highly probable that human medicine will significantly increase life expectancy. Already nowadays in some countries the average life expectancy is more than 80 years. We suppose that increase of life expectancy will occur as a result of a breakthrough in medical technologies in the 2030s–2050s. In the 2050s the average life expectancy will increase by 15 years or even more. It is quite possible that genetic methods will significantly increase life expectancy. In this respect, the study of telomeres, which were found to play an important role in cell division, seems to be promising (Slagboom, Droog, and Boomsma 1994).8 Transplantation. Another important branch of medicine is regeneration and transplantation of organs and tissues of a human body. At present, medicine achieved great results in organ transplantation, (e.g., heart, lungs, liver, pancreas, and kidneys). However, human donor organs are scarce, and people who donate donor organs without special agreement are brought to criminal responsibility all over the world. Medicine and biotechnology will provide an opportunity to design different artificial organs, such as skin, retina, trachea, vessels, heart, ear, eye, limbs, liver, the lungs, pancreas, bladder, ovaries. Many of them are already designed today. Even new organs or combinations are possible.There is already an opportunity of tissue engineering. In laboratories, scientists cultivate new cells to replace injured bone or cartilage. For example, recently, the soft artificial heart was created from silicone using a 3D-printing, lost-wax casting technique; it weighs 390 grams and has a volume of 679 cm3.This artificial heart has a right and a left ventricle, just like a real human heart, though they are not separated by a septum but by an additional chamber. This chamber is inflated and deflated by pressurized air and is required to pump fluid from the blood chambers, thus replacing the muscle contraction of the human heart (Cohrs, Petrou, Loepfe et al. 2017). This technology has the potential to develop cell therapy and methods of tissue regeneration. One can expect that opportunity to ‘deceive’ the immune suppression will be one of the main breakthroughs in the field of regeneration and transplantation of organs and tissues. Changing human reproductive capabilities is an especially important field of medicine. The number of incurable diseases causing infertility decreases. Nevertheless, the only opportunity for some patients is to use in vitro fertilization. Besides, due to the development of medicine there increases a number of women who want to have children after their reproductive age is over (e.g., at present, it is possible to grow an embryo outside the woman’s body).

The future of biotechnology One can suppose that at the very first stage of Cybernetic Revolution biotechnology, as an independent direction, will play a less important role than medicine. It will be 406

Big history and the future of technology

rather an important component of medical technologies, providing breakthroughs in treatment of diseases or monitoring of organism functions. Genetic engineering will play important role in different spheres of biotechnologies (see below). Gene modification. On the basis of the genetic data the most appropriate treatment will be adapted for individual patients, and if it is necessary the defective genes will be corrected. Presumably, first gene therapy will manifest itself in sport medicine as enormous investments are made in it and the best minds are engaged in this field. When choosing the appearance of a future child (color of eyes, skin, etc.), gene therapy can be used. In future, it might be possible that babies will be born almost by order, these will be ‘the perfect babies’ (Fukuyama 2002).9 In other words, parents will choose desirable features of a child before his/her birth. Achievements in self-regulation. The level of controllability will increase considerably within a number of important systems connected with biotechnologies. Thus, probably, while transforming an organism, scientists will insert not a separate useful gene (Simon, Priefer, and Pühler 1983), but a whole set of necessary genes which will operate depending on environmental conditions. Such characteristics will be extremely important in the case of climate changes which are quite probable. It will become possible to choose the most optimal varieties of seeds for a unique combination of weather conditions and territory. Consequently, huge databases of such plant varieties and variations will be created. It is quite possible that in the future the whole process of getting a transgenic plant will take place without human participation, thus, it will become self-regulating. It is possible to assume that by the end of the final phase of the Cybernetic Revolution the agricultural biotechnologies will be already developed to a degree that the modified products will be able to response even to the smallest fluctuations of local conditions. In other words, it will be possible for farmers to select individual fodder and drugs by means of programs and to order them via the Internet. Even an individual will be able to invent a houseplant hybrid suitable for the interior and to order its production and delivery. The same refers to domestic animals: it will be possible to breed animals with peculiar characteristics within separate breeds of animals (or even by the individual order). It is probable that the selection of animals on the basis of genetic engineering will also develop in the direction of decreasing human participation Creation of new materials. In the 1940–1970s, one of the main directions was the development of industrial production of already known substances (e.g., vitamins) or their analogues; however, during the same period there appeared the elements which did not exist in natural environment (e.g., Humalog, which is a widely applied synthetic analogue of human insulin) (Woollett 2012). This sequence reminds the history of development of chemistry: at first, people learned to produce the known substances, and then the artificial materials. Due to biotechnologies, many new materials are produced, for example, bioplastics. The main advantage of this material is that unlike ordinary plastic it can biodegrade. Thus, the main goal of bioplastics production is preserving environment, reducing the production of goods from non-renewable resources and cutting the discharging of carbon dioxide into the atmosphere. This is an important step to the 407

Leonid E. Grinin and Anton L. Grinin

creation of self-cleaning ecological systems in the future and also to the preservation of the environment. The increase and cheapening of food production is a global challenge for the humankind taking into account that the population number will continue to increase for several more decades (first of all in the poor and poorest countries, especially in Africa), perhaps, reaching nine or more billion people (see UN Population … 2012). Biotechnologies can make a huge contribution to the solution of the problem. The solution of urban and some environmental problems. Biotechnologies are successfully used for cleaning up oil spills, in wastewater treatment, etc. According to the Organization for Economic Cooperation and Development (OECD), the potential market for bioremediation, that is, the use of living microorganisms to degrade the environmental contaminants (including plants for soil purification), amounts tens of billions dollars. So important changes will certainly take place with respect to the use of biotechnologies for the solution of environmental problems. Here it is possible to assume that biotechnologies will be intruded first of all into the urban ecology. It is necessary to consider that in the coming decades the urban population will increase by 40–50% (see, e.g., NIC 2012). Among the problems which can be potentially solved by means of the development of biotechnologies, there are those related to water cleaning, recycling of waste, liquidation of stray animals (it will be promoted by introducing genes for sterility or something of that nature). Already today the ­micro-organisms for water cleaning are applied; with their help we also get bio-gas from waste recovery. But in the future, these and similar problems will be solved by the development of self-regulating systems that will make it possible to solve a number of technical and scientific problems. But the problem of ecological self-regulating systems, naturally, is not limited by the cities; it has to be extended to the cleaning of reservoirs and other ecosystems. The creation of ecological self-regulating systems will considerably reduce expenses and free huge territories occupied by waste deposits, as well as allow breeding fish in self-cleaning reservoirs. One can assume that an important direction will be the creation of self-regulating environmental systems in resort and recreational territories which will provide the best conditions for rest and business.

The breakthrough in the sphere of resource saving Resource and energy saving is one of the main tasks and outcomes of introduction of biotechnology. The basic opportunities with respect to resources saving are connected with an opportunity to influence the genetic organization of living beings which at present serves the basis for the agricultural (‘green’) biotechnology which has already become a part of the initial phase of the Cybernetic Revolution. The breakthrough in this area is connected with totipotency, that is an ability of plants to form a full-fledged organism from a single cell. With the necessary gene transfer, one can make, for example, a variety of potato resistant to the Colorado beetle, or reduce the susceptibility to drought, cold, and other stresses (Grinin et al. 2010). New agricultural technologies are of great importance for the developing countries. For example, genetically modified and pest-resistant varieties of cotton plant and corn demand 408

Big history and the future of technology

much smaller usage of insecticides which is more cost-effective and eco-friendly.The individualization is also noticeable in the animal genetic engineering which develops more slowly, but even now and in prospect it has an enormous value for agriculture and medicine (by means of genetic engineering, it is possible to increase milk production, to improve quality of wool, etc.). Biotechnology can help to solve many global issues, for example, to cheapen the production of medicines and foodstuffs including producing and making them in ecologically sound ways that can also keep or make the environment pristine, thereby considerably expanding their production.The solution to the food problem will proceed in different ways, in particular due to the mass production of food protein whose shortage is sharply perceived in many societies (at present the feed protein for animals is generally produced in this way). Even now, there are results based on the production of food proteins or, for example, imitation meat. But so far such a production is too expensive. A gram of laboratorial meat costs US$ 1,000 dollars (Zagorski 2012) but this is part of the usual process from the laboratory to mass cheap production.

AI, robots, nanotechnologies, additive and cognitive technologies Self-regulating systems and AI. Within big history the period after a new threshold is sometimes anticipated as a period of rapid development and even predominance of AI. We agree that the future investigations in the fields of big history and evolution are closely associated with the development of AI. Thus, it turns important to define similarities and differences between self-regulating systems and AI. On the one hand, the notion of ‘self-regulating systems’ correlates rather closely with AI which has become a subject of intensive research in the recent decades (see, e.g., Poole, Mackworth, and Goebel 1998; Russell et al. 2003; Hutter 2005; Luger 2005; Neapolitan and Jiang 2012; Keller and Heiko 2014; Hengstler, Enkel, and Duelli 2016). ‘Intelligent’ machine is often defined as the one that takes actions that maximize its chance of success at some goal (e.g., Russell et al. 2003). Of course, such a machine can be also considered as a self-regulating system. The notion of AI is usually connected with machines, IT-technologies, robots, and sometimes equated with technical intelligence (Zhang et al. 2016). On the other hand, the notion of ‘self-regulating systems’ is wider than the notion of ‘artificial intelligence’ (AI) since the former includes various self-regulating systems that can function independently, but can hardly be regarded as AI. For example, biotechnological systems designed to neutralize pollution, or the ones connected with human physiology (e.g., artificial immunity on the basis of artificial antibodies, or systems based on the use of various other proteins or viruses, or genetic engineering technologies that are able to control certain physiological processes and so on). In addition, we expect the emergence of self-regulating systems of mixed ­nature – for example, biochemiotechnical. One should also note that they can function within more complex systems, like a human organism. As examples of such self-regulating systems, one may mention artificial organs grown in laboratories and incorporating a number of biosensors and other technical elements. Thus, any AI can be regarded as a self-regulating system, but not all self-regulating systems can be associated with AI. 409

Leonid E. Grinin and Anton L. Grinin

Robots and self-driving cars in the future The opportunities of using robots are undoubtedly vast. In particular, only these devices can help to solve the problem of care of growing numbers of elderly people and to some extent the associated problem of labor shortage. In general, there is no doubt that robots will play a significant role in the transition to self-regulating systems. We assume that in the 2020s, certain although not revolutionary achievements in this area will occur; in the 2030–2040s, we will witness a much more significant rise in robotics, but an explosive development of robots will happen a bit later in the 2050–2060s. By this time, it is also possible to expect the creation of really ‘smart’ robots. We believe that in the next two decades robotics will develop rapidly in the service sector. At present there are many publications on how robots may replace humans in many fields. We agree that the changes in this sphere will be enormous yet, they will take several decades to occur. In future, robotic servants may replace household chores as well as perform some complicated tasks, for example, they will be more and more involved into investigation of space bodies and other tasks that can be dangerous for humans (military, rescue, space activities, etc.). Hardly all of them will be anthropomorphous; their design will be most likely defined by functions. However, universal robots are also likely to emerge. Robots will play a very important role in medicine, for example, in surgery and in the sphere of social nursing care. The number and variety of surgical robots grow every year. According to some forecasts, surgical robotics market will grow up to US$ 28.8 billion by 2020.10 Surgical robotic systems are a combination of equipment, accessories, software, and services, which help doctors to perform minimally invasive surgeries including gynecological, cardiac, neurological, and orthopedic. Robotic systems allow surgeons to automate the surgical procedure, improve efficacy and precision during the procedure, and minimizing post-surgical complications (see about robotics and additive technologies in footnote 9). Robotic systems will continue to be used extensively in transportation, in particular they will also be used in the development of self-driving vehicles. The latter might be especially important. Taking into account the above-described ‘meaning’ of the Cybernetic Revolution (as a revolution of self-regulating systems), the breakthrough will most probably occur in the direction of autonomous transport.Vehicles and other transport systems will become self-driving and will use the electric vehicle technologies. Even today, there are attempts of realizing this opportunity. A vivid example here is Tesla’s self-driving cars. But other groups of companies also announced their self-driving cars. For example,‘Mercedes-Benz’ has presented the concept of driverless car (della Cava 2015). Google works to create such a car by 2020 (see Muoio 2015), but it already tests the Toyota self-driving car in California (and arranges joint projects with Ford). Just as in 1997 the computer defeated the world chess champion; recently self-driving car has beaten the racing driver at speeds over 200 kilometers per hour. Some researchers even work to understand how to make self-driving cars become capable of making moral and ethical decisions just like humans do. Any decision that involves risk of harm to a human or even an animal is considered to be an ethical decision. It also includes quite rare situations when a collision is unavoidable, but a decision can be made as to which obstacle to collide with. Researchers believe that by 410

Big history and the future of technology

algorithms it is possible to make self-driving car decide whether to use a sophisticated algorithm or a simple rule such as ‘always stay in the lane’ (Sütfeld, Gast et al. 2017). However, the development of such systems as self-regulating systems is an important forerunner of the forthcoming start of the final phase of the Cybernetic Revolution (in the 2030s). The self-driving electric vehicles with a new accumulator together with roads allowing free recharge can become a powerful source of technological development during the final phase of the Cybernetic Revolution.

The future of nanotechnology Nanotechnologies have become one of the most popular technologies of modern times. Despite the fact that some indices show the slowing down development, nanotechnologies have many prospects.11 First of all, different nanocoatings will rapidly develop. Nanocoatings are used in different fields: industry, aircraft building, and electronics. The components of nanoelectronics, photonics, neuroelectronic interfaces, and nanoelectromechanical systems are also promising. They will allow further micronization of devices.We believe that self-regulating technologies, (e.g., self-cleaning coatings which regulate the temperature) will gain a widespread use. Similar technologies have already been created. For example, recently in the University of Central Florida a flexible anti-reflection film on smartphones and tablets was made. It makes the screen bright and sharp as well as scratch resistant and self-cleaning. The film contains tiny uniform dimples, each about 100 nanometers in diameter (about one one-thousandth of the width of a human hair) (Guanjun Tan et al. 2017). In future, the nanotechnologies will provide excellent opportunities for the self-assemblage of nanoelements and nanodevices. It will become possible to make a transition to controlled self-assemblage of nanosystems, creation of three-dimensional networks, nanorobots, etc. One may also speak about the use of molecular devices, atomic design. There are rather attractive prospects in the development of nanomechanics, nanomachinery, and nanorobotics. Long ago there started to develop the idea of creation of computers that process and store information through not special condition of environment (magnetic, electric, and optical) but through nanotechnologies, for example, via silicone (the main material in the production of semi-conductor devices) chips replaced by carbon nanotubes. In this case a bit of information can be written in the form of a cluster, for example, of 100 atoms. This would reduce their size several fold and at the same time increase quick response. Quantum technologies (e.g., the creation of quantum computer) will be one of the most important technological breakthroughs in this context. It is very difficult to predict which path the development of information technologies will follow, but one can assume that as a result of the completion of the Cybernetic Revolution, the information capacity will increase by an order of magnitude.

The future of 3D-printers The opportunities provided by 3D-printers are great: from building to cooking, from a house workshop to museums, from medicine to children toys, from training models 411

Leonid E. Grinin and Anton L. Grinin

to design. At present, 3D-printing is used in aircraft construction and rocket engineering to produce individual elements, for example, support stand for an aircraft engine (see, e.g.,Turichin 2015). And just because they are used in such high-­technology spheres, their development needs considerable investments. In fact, 3D-printers constitute a universal house workshop or a universal production or construction factory. And in the future they will acquire new functions and incorporate new subsystems. Additive 3D-printing (i.e., merging (fusing) of materials and creation of certain objects) is a very promising direction. Thus, in future 3D-printers will help to produce any material needed, even the biological one. Great opportunities are especially associated with the opportunities to grow human organs and tissues, including through the usage of patient’s own tissues. Soon it will suffice to have a sketch and to make (to ‘print’, ‘fuse’) any detail at home or in a 3D-printing center. It will also be possible to organize a small single-piece production. Engineers could also develop simple 3D-food printers which can print, for example, candies or pizza. Undoubtedly, the development of additive technologies will be connected with other directions of MANBRIC-complex, for example, with robotics (additive technologies will be used to create robots, and at the same time the robots themselves will use additive technologies in their activities).12

Cognitive technologies: neural interfaces or brain–computer interfaces (BCI) A brain–computer interface (BCI) is a direct communication pathway between brain and an external device.This technology implements the interaction between brain and computer systems that can be realized via electrode contact with head skin or via electrodes implanted into brain. The implementation of neural interfaces is already widespread, for example, in artificial visual systems or bionics.The most notable device is the cochlear implant, which has been implanted in more than 220,000 people worldwide. In three or four decades, disabled people will get another chance in life. BCIs may improve rehabilitation for people with strokes, head trauma, and other disorders. At present there already exist devices which allow paralyzed people to speak, write, and even work at the computer as, for example, in the case of the famous astrophysicist, Stephen Hawking. Those who can pay and want to increase their abilities will be able to replace their body parts by bionic ones. Also in three or four decades, small scalp electrodes will make remote brain control possible. So people will be able to turn TV on only by thinking about it. In the future, neural interfaces can be applied not only in medicine, but also in daily pursuits, for example, to control condition of a driver’s or an operator’s brain and in case of falling asleep to awake him automatically. In general the achievements in cognitive science are already in use and their application will increase even more in the areas which move towards self-regulating systems – from medicine to robotics, from cybernetics to problems of AI, and, of course, for the military purposes. However, serious technical and social difficulties can hamper the development of this direction (see below). After surpassing these constraints, the development of neural interfaces will promptly reach a new level. 412

Big history and the future of technology

Some ideas on other future technologies: smart devices Everyday technologies become more self-regulating, complicated, and more intelligent. It is even clear from their names. The word ‘smart’ is used as a prefix for many devices.Today smartphones have become ubiquitous, while smartwatches are becoming popular, people watch smart TV, and in schools they use smart boards. Here are just a few examples of smart things: smart kettle, smart swimsuit, smart stroller, smart cup, smart rope, smart T-shirt which tracks your posture, smart cane with GPS for elderly people, smart bottle which automatically tracks hydration and temperature, smart highway with nanocoating which changes its color according to the weather and warns drivers of potential risk. There is also developed a concept of a smart city with smart traffic signs and traffic light signals, as well as smart cars. This will also allow time and resource saving. Exoskeleton will allow people to perform hard work with fewer efforts. We assume that this trend will continue and thus, in three or four decades the majority of everyday devices will be smart. An absolute majority of them will be connected to the smartphone and Internet. One can predict that we will live in smart homes with smart kitchens, while a smart climate control system will maintain the required temperature 24 hours a day.

Mobile phone as an integrating device The key feature of the future technologies is that most of them will be integrated with mobile phone or similar devices. A mobile phone will be a universal control panel and analytical center. It will collect all data from smart technological devices, for example: how many meters one walks, how many calories one consumes, how many hours one sleeps, how much money one spends, how many hours one plays basketball, and how many points one scores. The mobile phone will become a powerful means of control not only over an individual, but over pets and children. For example, a smart bracelet that monitors a child’s clean hands and signals if the child takes, for example, unwashed fruit. At any time parents can check the purity of their child’s hands using a special application in smartphone. Thus, the important future technological trend is the development of virtual reality through different devices especially mobile phone or in another form of such integrative device (as it is known, there are different ideas on this future forms, e.g., glasses). It is quite probable that such devices will be able to adopt the functions and become a new type of sensory organ and source of information for people.Thus, special glasses will allow connecting vision and hearing with the high resolution virtual reality devices. In the future, virtual reality may be not only seen but also felt. A small band with a device on the arm is already designed which will enable users touch the object in virtual reality.

Conclusion: will the development of the Cybernetic Revolution proceed in the direction of cyborgization? There is no doubt that future development within big history and evolutionary paradigms is connected with the development of intelligence and transformation of intellectual creatures. As to the direction and speed of this transformation, there are 413

Leonid E. Grinin and Anton L. Grinin

many points of view, including those (which we do not share) that AI will be able to unite billions of people’s minds into a new system (Kurzweil 2000) or that humans will soon become immortal (see below). On the other hand, the development of medicine and self-regulating systems, which will constitute the nucleus of changes in the final phase of the Cybernetic Revolution, will undoubtedly lead to the increasing interference in human body. In this context, we would like to conclude the paper by the reflections about the ways this interference in human body can change the human biological nature and transform a human into a cyborg. A very popular word ‘cyborg’ (short for ‘cybernetic organism’) derives from the word ‘cybernetic’. Cyborg is defined as a theoretical or fictional being with both organic and b­ iomechatronic parts.13 The term ‘cyborg’ often applied to an organism that has restored function or enhanced abilities due to the integration of some artificial component or technology that relies on some sort of feedback. It is obviously that many achievements in medicine will impel our civilization to the state in which more and more humans can become partial cyborgs. Thus, we are following the path of development of self-­regulating systems of a new type which will be constituted by the elements of different origin: biological and artificial. All that we have written about artificial organs and tissues will contribute to the breakthrough in the field of both production of absolutely new materials which will expand the implementation of non-biological elements in the human body. Thus, the Cybernetic Revolution is closely connected with the process that can be designated as cyborgization. We should be aware of the fact that this actually means not only the formation of a new direction in medicine, but also the moving towards the cyborgization of a human being. Of course, this can cause a certain and quite reasonable anxiety. On the other hand, expanding the opportunities for not just a long but also an active life is hardly possible without significant support for the sensory organs and other parts of the body which weaken as a result of ageing and other reasons. Finally, glasses or contact lenses, artificial teeth, tooth fillings, bones, aerophones, artificial blood vessels, mitral valves, etc. allow hundreds of millions of people to live and work, and these people still remain humans. The same is true with respect to more complex systems and functions. Thus, people with disabilities can benefit from the development of medicine and cyborgization as they will be able to significantly compensate their drawbacks. However, we suppose that the idea that someday the human body will be fully replaced by non-biological material and only the brain or the organs which support the senses will remain is just fantasy; this will never come true. People who propose such solutions, for example, to replace supposedly less lasting and comfortable biological material by the technological inventions (such as replacement of haematocytes by billions of nanorobots, etc.) in their forecasts try to use the outdated logic that was widespread several decades ago in science fiction or scary stories: the replacement of biological organisms with technical ones. The modern logic of scientific and technological progress including the latest achievements in bioengineering shows the shift towards the synthesis of biological forms and technical solutions into a unified system. Still there are numerous obstacles here. Let us take, for example, the above-mentioned possible opportunities for brain control which may be hampered by the immune rejection in the first turn. Second, many nanostructures, for example, nanopipes, which had been predicted a bright future appeared very toxic for 414

Big history and the future of technology

human body (Kotov et al. 2009). Third, the implantation of external devices leads to traumatization of the whole organism despite all serious attempts to reduce this impact. Another problem is the different electric conductance of biological material and of a technical device, though there is certain progress in the solution of this problem (Abidian and Martin 2009). But even if we solve these problems we will still need some powerful software capable to handle brain signals. Technical achievements can hardly replace the biological mechanisms which have been selected for many millions of years. On the contrary, we should follow the path of ‘repair’, improvement, the development of self-regulation, and support of biological mechanisms via some technical solutions. The human brain is very tightly connected with the body and sensory organs, most of its functions are based on the control of the body that does not imply its full-fledged work outside its biological foundation. The opportunities of science and medicine to replace worn organs will increase but the biological foundations of a human will always exist and must prevail. If one can help the human body by different means including methods of activization of immune system, opportunities of genetics, the methods of blocking, or decelerating the process of ageing, etc., it is much more reasonable to preserve the human biological foundation. In any case, in the nearest decades in the process of cyborgization quite radical breakthroughs are possible, but nevertheless the process of cyborgization will not go too far. Thus, we believe that in the next 100 years the human lifestyle and biological nature will experience crucial changes which can become a turning point in the transition to the post-human society. However, these changes, no matter how profound they are, will be very far from the images drawn by modern wishful-thinking scientists.

Notes 1 Sometimes we denote it as the Upper Paleolithic Revolution. 2 See Grinin (2006a, 2006b, 2007a, 2007b, 2012); Grinin L. and Grinin A. (2013, 2015); Grinin A. and Grinin L. (2015). 3 It lasted till 12th mil. BP. 4 A number of big history researchers connect this threshold with so-called singularity. 5 Namely: Medicine, Additive, Nano, Bio, Robotic, Information, and Cognitive technologies. For the convenience of pronunciation the technologies are listed not in order of priority. 6 See also Appendix to Ch. 9 in Grinin L. and Grinin A. 2015 at https://www.socio nauki.ru/book/files/ot_rubil_do_nano/online_version/9_chapter_appendix/266p.php 7 For the dynamics of patent application see Appendix to Grinin L. and Grinin A. 2015, 274ff. URL: https://www.socionauki.ru/book/files/ot_rubil_do_nano/online_­ version/9_chapter_appendix/274p.php. 8 In 2009, Elizabeth H. Blackburn, Carol W. Greider and Jack Szostak were awarded the Nobel Prize for the discovery of the way chromosomes are protected by telomeres and the enzyme telomerase from terminal underreplication. 9 It is difficult to say how ‘perfect’ they will be and what kind of problems will appear as a result of these technologies. E.g., the possibility to predict the baby’s gender resulted in gender imbalance in China. As a result, there are a disproportionate number of boys. Thus, we agree with Francis Fukuyama, who believes that the future achievements of the ‘biotechnology revolution’ should be accepted with great prudence (Fukuyama 2002). 415

Leonid E. Grinin and Anton L. Grinin

10 URL: https://www.alliedmarketresearch.com/surgical-robotics-market?surgicalrobotics-market. 11 See Appendix to Grinin L. and Grinin A. (2015, p. 288ff ).https://www.socionauki.ru/ book/files/ot_rubil_do_nano/online_version/10_chapter_appendix/288p.php. 12 E.g., the Stormram 4, as the robot is named, is made from 3D-printed plastic and is driven by air pressure. This robot can be used in an MRI scanner. Carrying out a biopsy (removing a piece of tissue) during a breast cancer scan in an MRI significantly increases accuracy. The Stormram 4 is a stimulus for the entire diagnostic phase of breast cancer; the accurate needle control, effectively real-time MRI scanning and a single, thin-needle biopsy enable quicker and more accurate diagnoses to be made (University of Twente 2017). 13 The term was coined in 1960 by Manfred Clynes and Nathan S. Kline (Halacy 1965).

References Abidian M. R., and Martin D. C. 2009. Multifunctional Nanobiomaterials for Neural Interfaces. Advanced Functional Materials 19(4): 573–585. doi:10.1002/adfm.200801473. Beer S. 1959.Cybernetics and Management. London: English Universities Press.doi:10.2307/ 3007308. della Cava M. 2015. Captain Kirk, Your Mercedes F 015 Awaits. URL: http://www.usatoday. com/story/tech/2015/01/06/mercedes-f-015-luxury-in-motion-debuts-at-consumerelectronics-show/21359239. von Foerster H., and Zopf G. W. 1962. Principles of Self-Organization. Self-Organization. Ed. by M. C. Yovits, and S. Cameron, pp. 31–50. New York: Pergamon Press. Cohrs N. H., Petrou A., Loepfe M., et al. 2017. A Soft Total Artificial Heart-First Concept Evaluation on a Hybrid Mock Circulation. Artificial Organs.doi:10.1111/aor.12956. Fukuyama F. 2002.Our Post-human Future: Consequences of the Biotechnology Revolution. New York: Farrar, Straus, and Giroux. Grinin A. L. 2016. Self-Regulation as a Global Evolutionary Mega-Trend. Evolution: Evolution and Big History: Dimensions, Trends, and Forecasts. Ed. by L. E. Grinin and A. V. Korotayev, pp. 139–164. Volgograd: Uchitel. Grinin A. L., and Grinin L. E. 2015. The Cybernetic Revolution and Historical Process. Social Evolution & History1: 125–184. Grinin A. L., Kholodova V. P., and Kuznetsov V. V. 2010. A Comparative Analysis of Physiological Mechanisms of Salt-Endurance of Different Varieties of Mustard. Vestnik ­Rossiyskogo Universiteta Druzhby Narodov 1: 27–38. In Russian (Гри-нинА. Л., ХолодоваВ. П., and КузнецовВ. В. Сравнительныйанализфизиологическихмехан измовсолеустойчивстиразличныхсортовгорчицы.  ВестникРоссийскогоуниверси тетадружбынародовx 1: 27–38). Grinin L. E. 2006a. Periodization of History: A Theoretic-Mathematical Analysis. History and Mathematics: Analyzing and Modeling Global Development. Ed. byL. E. Grinin, V. de Munck, and A. Korotayev, pp. 10–38. Moscow: KomKniga. Grinin L. E. 2006b.The Productive Forces and Historical Process. 3rd edition. Moscow: ­KomKniga. In Russian (ГрининЛ. Е. Производительныесилыиисторическийпроцесс. 3-еизд. М.: КомКнига/URSS). Grinin L. E. 2007a. Production Revolutions and the Periodization of History. Herald of the Russian Academy of Sciences 77(2): 150–156. Grinin L. E. 2007b. Production Revolutions and Periodization of History: A Comparative and Theoretical-Mathematical Approach. Social Evolution & History 6(2): 75–120. Grinin L. E. 2012. Kondratieff Waves, Technological Modes and Theory of Production Revolutions. Kondratieff Waves: Aspects and Prospects: Yearbook. Ed. by A. А. Akaev, 416

Big history and the future of technology

R. S. Greenberg, L. E. Grinin, A. V. Korotayev, and S. Yu. Malkov, pp. 222–262. Volgograd: Uchitel. In Russian (ГрининЛ. Е. 2012. Кондратьевскиеволны, технол огическиеукладыитеорияпроизводственныхреволюций.  Кондратьевскиеволны: аспектыиперспективы: ежегодник. Отв. ред. А. А. Акаев, Р. С. Гринберг, Л. Е. Гринин, А. В. Коротаев, С. Ю. Малков, с. 222–262. Волгоград: Учитель). Grinin L. E. 2014. The Star-Galaxy Era of Big History in the Light of Universal Evolutionary Principles. Teaching & Researching Big History: Exploring a New Scholarly Field. Ed. by Leonid E. Grinin, David Baker, Esther Quaedackers, and Andrey V. Korotayev, pp. 163–187. Volgograd: ‘Uchitel’ Publishing House. Grinin L. E., and Grinin A. L. 2013. Macroevolution of Technology. Evolution: Development within Big History, Evolutionary and World-System Paradigms. Ed. by L. E. Grinin, and A. V. Korotayev, pp. 143–178. Volgograd: ‘Uchitel’ Publishing House. Grinin L. E., and Grinin A. L. 2015.From Bifaces to Nanorobots. The World on the Way to the Epoch of Self-Regulating Systems (History of Technologies and Description of Their Future Development). Moscow: Moscow branch of Uchitel Publishing House. In Russian (ГрининЛ. Е., ГрининА. Л. Отрубилдонанороботов. Мирнапутикэпохесамоуп равляемыхсистем (историятехнологийиописаниеихбудущего). М.: Московска яредакцияизд-ва«Учитель»). Grinin L., and Grinin A. 2016.The Cybernetic Revolution and the Forthcoming Epoch of Self-­ Regulating Systems. Moscow: Moscow branch of Uchitel Publishing House. Grinin L. E., and Korotayev A. V. 2015a.Great Divergence and Great Convergence: A Global Perspective. New York: Springer. Grinin L. E., Grinin A. L., and Korotayev A. 2016. Forthcoming Kondratieff Wave, Cybernetic Revolution, and Global Ageing.Technological Forecasting and Social Change. doi:10.1016/j.techfore.2016.09.017. Grinin L. E., Markov A. V., and Korotayev A. V. 2008. Macroevolution in Animated Nature. Moscow: LKI/URSS. In Russian (ГрининЛ.Е., МарковА.В., КоротаевА.В.Мак роэволюциявживойприродеиобществе.М.: ЛКИ/URSS). Grinin L. E., Markov A. V., and Korotayev A. V. 2009. Aromorphoses in Biological and Social Evolution: Some General Rules for Biological and Social Forms of Macroevolution. Social Evolution & History 8(2): 6–50. Grinin L. E., Korotayev A. V., and Markov A. V. 2017. Biological and Social Phases of Big History: Evolutionary Principles and Mechanisms. From Big Bang to Galactic Civilizations. A Big History Anthology Volume III. The Ways that Big History Works: Cosmos, Life, Society and Our Future. Ed. by Rodrigue B., L. Grinin, and Korotayev A. New Delhi: Primus Books. Gupta A. K., and Gupta M. 2005. Synthesis and Surface Engineering of Iron ­Oxide Nanoparticles for Biomedical Applications. Biomaterials 26(18): 3995–4021. doi:10.1016/j. biomaterials.2004.10.012. Halacy D. S. Jr. 1965.Cyborg: Evolution of the Superman. New York: Harper. Hengstler M., Enkel E., and Duelli S. 2016.Applied Artificial Intelligence and Trust –The Case of Autonomous Vehicles and Medical Assistance Devices.Technological Forecasting and Social Change. Elsevier 105: 105–120. Hutter M. 2005.Universal Artificial Intelligence, Machine Learning.doi:10.1145/1358628.1358961. Keller J., and Heiko A. 2014. The Influence of Information and Communication Technology (ICT) on Future Foresight Processes – Results from a Delphi Survey. Technological Forecasting and Social Change. Elsevier 85: 81–92. Kotov N. A., Winter J. O., Clements I. P. et al. 2009. Nanomaterials for Neural Interfaces. Advanced Materials 21(40): 3970–4004.doi:10.1002/adma.200801984. Kurzweil R. 2000. The Age of Spiritual Machines: When Computers Exceed Human Intelligence. New York: Penguin Groups. 417

Leonid E. Grinin and Anton L. Grinin

Luger G. F. 2005.Artificial Intelligence: Structures and Strategies for Complex Problem Solving. Pearson education. Montgomery M., Ahadian S., Locke Huyer D. et al. 2017.Flexible Shape-memory Scaffold for Minimally Invasive Delivery of Functional Tissues. Nature Materials. doi:10.1038/nmat4956. Muoio, D. 2015. 10 companies making a bold bet that they’ll have self-driving cars on the road by 2020. Business Insider, October 12. URL: http://www.businessinsider.com/ google-apple-tesla-race-to-develop-self-driving-cars-by-2020-2015-10 Neapolitan R. E., and Jiang X. 2012.Contemporary Artificial Intelligence. Boca Raton, Florida: CRC Press. Nefiodow L., and Nefiodow S. 2014. The Sixth Kondratieff. The New Long Wave of the World Economy. Rhein-Sieg-Verlag: Sankt Augustin. NIC – National Intelligence Council 2012.Global Trends 2030: Alternative Worlds. URL: www.dni.gov/nic/globaltrends. Pankhurst Q. A. et al. 2003.Applications of Magnetic Nanoparticles in Biomedicine. Journal of Physics D: Applied Physics 36(13): R167. URL: http://stacks.iop.org/0022-3727/36/ i=13/a=201. Poole D. L., Mackworth A., and Goebel R. G. 1998.Computational Intelligence and Knowledge. Computational Intelligence: A Logical Approach (Ci): 1–22. URL: https:// www.cs.ubc.ca/~poole/ci.html. Russell S. J. et al. 2003.Artificial Intelligence: A Modern Approach. Pearson Education Limited. New York: Pearson. Shea J. J. 2006.The Human Revolution Rethought, Evolutionary Anthropology: Issues, News, and Reviews. Edinburgh: Edinburgh University Press. doi:10.1002/evan.20085. Simon R., Priefer U., and Pühler A. 1983. A Broad Host Range Mobilization System for in Vivo Genetic Engineering: Transposon Mutagenesis in Gram Negative Bacteria. Nature Biotechnology 1(9): 784–791.doi:10.1038/nbt1183–784. Slagboom P. Е, Droog S., and Boomsma D. I. 1994. Genetic Determination of Telomere Size in Humans: A Twin Study of Three Age Groups. American Journal of Human Genetics 55: 876–882. Guanjun Tan, Jiun-Haw Lee, Yi-Hsin Lan, Mao-Kuo Wei, Lung-Han Peng, I-Chun Cheng, and Shin-Tson Wu. 2017.Broadband Antireflection Film with Moth-eye-like Structure for Flexible Display Applications. Optica 4(7): 678.doi:10.1364/OPTICA.4.000678. Sütfeld L. R., Gast R., König P., and Pipa G. 2017.Using Virtual Reality to Assess ­Ethical Decisions in Road Traffic Scenarios: Applicability of Value-of-Life-Based Models and Influences of Time Pressure. Frontiers in Behavioral Neuroscience 11.doi:10.3389/ fnbeh.2017.00122. Turichin G. A. 2015.Additive Technologies in Modern Production. Report at the 2nd International Seminar ‘Basic Technologies of the First half of the 20th Century (structural and cyclical analysis)’. Saint-Petersburg, 1–2 October. In Russian (ТуричинГ. А. Аддитивныетехнологиивсовременномпроизводстве.Докладнавтороммеждун ародномсеминаре«Базисныетехнологиипервойполовины XIX в. (структурноциклическийанализ)». Санкт-Петербург, 1–2 октября. Umpleby S. A., and Dent E. B. 1999. The Origins and Purposes of Several Traditions in Systems Theory and Cybernetics. Cybernetics and Systems 30(2): 79–103.doi:10.1080/ 019697299125299. UN Population Division of the Department of Economic and Social Affairs of the United Nations Secretariat 2012.World Population Prospects: The 2010 Revision. URL: http://esa. un.org/unpd/wpp/index.htm. University of Twente 2017 3D-Printed Robot Aims to Fight Cancer. ScienceDaily.URL: www.sciencedaily.com/releases/2017/07/170703121134.htm. Accessed February 5, 2018. 418

Big history and the future of technology

Wiener N. 1948.Cybernetics, or Control and Communication in the Animal and the Machine. Scientific American. JSTOR 179(5): 14–19. WIPO = World Intellectual Property Organization. 2016. World Intellectual Property ­O rganization database. URL: http://www.wipo.int/portal/en/index.html. Woollett R. 2012. Innovation in Biotechnology: Current and Future States. Clinical ­Pharmacology and Therapeutics 91(1): 17–20. doi:10.1038/clpt.2011.219. Zagorski I. 2012. Not by Meat Alone: They Promise to Create Leather Jackets in Laboratories. Vesti.ru. September 20. URL: http://www.vesti.ru/doc.html?id=9120 84& cid=2161. In Russian (ЗагорскийИ. 2012. Немясомединым: кожаныекурткибудутвы ращиватьвлаборатории. Вести.ру, 20 сентября). Zhang Y. et al. 2016. Technology Roadmapping for Competitive Technical Intelligence.Technological Forecasting and Social Change 110: 175–186. doi:10.1016/j.techfore.2015.11.029.

419

19 BIG HISTORY AND THE SINGULARITY Akop P. Nazaretyan

The first to use a specific mathematical term Singularity in his futurist reasoning was John von Neumann. In 1958, he said to his colleague Stanislaw Ulam: “The ever-accelerating progress of technology and changes in mode of human life…gives an appearance of approaching some essential Singularity in the history of the race beyond which human affairs, as we know them, could not continue” (quoted from [Eden et al. 2012: 4]). Later on, the exotic concept appeared in some Russian history books in a similar context, although with contrasting interpretations [Porshnev 1966; Dyakonov 1994]. A series of subsequent discoveries has called growing interest in this mathematical extrapolation referring to the near future among scientists and philosophers. In 2008, Singularity University was founded in collaboration with NASA. This paper shows how the concept of Singularity is related to big history and complexity theory worldviews, which give it universal grounds and help in tracing the palliative global developments in the twenty-first century.

The constructs of world, global, and universal (big; mega-) history Three competing patterns were kept on the agenda in historical discussions throughout the nineteenth and twentieth centuries. One was a Eurocentric, linear, and ­teleological view of history as a consistent progress “from the worse to the better” aimed at the perfect future condition. Another (unintentionally fortified by thermodynamics) was a descent from the deific past to an atheistic chaos. The third pattern was that there had never been a “human history” but the cycles of ascent, flourishing, and descent of regional civilizations without causal successions or universally valid events. Meanwhile, multiple discoveries in twentieth century sciences made it possible to single out no less than seven crucial landmarks in panhuman history and prehistory (like the Neolithic Revolution or the Axial Revolt, etc.) and a distinct succession in humankind’s development in spite of never-ending cycles in regional stories. More than that, the prevailing vectors in social evolution continue those observed in the evolution of biosphere and, after all, the cosmo-physical evolution of the “Metagalaxy”. 420

Big history and Singularity

From there, we can discriminate between the subjects of world history, global history and Universal (Big or Mega-) History, which together give us an integrated view of the past so far as it is traceable today and a context for careful anticipations. In continuation, adapted fragments from the monograph [Nazaretyan 2015] are exposed with minimal references. For more detailed arguments, examples and bibliography see the author’s publications available in English [Nazaretyan 2005a, 2005b; 2009; 2010a, 2010b; 2014a, 2014b]. The world history paradigm was formulated in the nineteenth century, together with the national histories, under the influence of the ideas of humanism and progress. It is based on the evolutionist methodology, and nowadays involves all of social and cultural events since the Paleolithic up to modern times. The global history concept is a product of the first half of the twentieth century, as the close mutual influence of geological, biotic, and social processes was discovered. It studies successive births and transformations of the planet’s spheres in which first biota and thereupon culture became the leading agents. The global history founders, Pierre Teilhard de Chardin and especially Vladimir Vernadsky, like most of their contemporaries, believed that Earth and Solar system were the maximum domain of evolution, for the universe was infinite in space and time, invariant and therefore deprived of history. Later on, cosmology expelled the stationary model and so the integral image of the past enlarged up to the evolving Metagalaxy.The final crystallization of ­Mega-History subject is due to the discovery of one more crucial fact: we can distinctly trace back the common vectors for the successive transformations in the cosmic Universe, Earth’s crust, biosphere, society, and intelligence. For all that, although no direct contradictions with the physical irreversibility laws are found, the orientation of the vectors discord with the classical natural science paradigm. Namely, the Metagalaxy has been successively evolving from the more probable random states (or “natural” ones, from the “entropy” point of view) to the less probable (“unnatural”) ones, so that the histories of biosphere and anthroposphere are the localized phases of the single universal process. The growing complexity mega-trend so apparently contradicts the suggestions inferring from the classical natural history (time as growing entropy; heat death theory) and so reliably corroborated by the empirical data of modern sciences and humanities that the astrophysicists have to distinguish between the thermodynamic arrow of time and the cosmological arrow of time and look for their causal relations [Chaisson 2006]. A question at the heart of all this is why evolution has gone in such a strange ­direction; we find various answers up to the obviously teleological and theological ones in relevant literature. An effective background for cross-disciplinary patterns free from the divine and/or telic assumptions is provided by modern complexity theories (their equivalents are called synergetics in Germany and Russia, non-linear thermodynamics in Belgium and France or dynamic chaos theory in USA). Such theories allow seeing the perfection of negentropy mechanisms not as the aim but as a means for non-­equilibrium systems’ (nature and society) resistance in the conditions of decreased sustainability. Thus, in the self-organization pattern, “human history is the story of one… system, which exists on the scale of a million or so years” [Christian 1991: 238], and has to evolve to sustain itself. 421

Akop P. Nazaretyan

However, is it true that the vectorial model correctly describes the empirical data of social history? David Christian has noted that the heated discussions around these problems are mainly due to the opponents’ unwillingness to alternate the distances, the exposures and the optic instruments in order to vary the pictures. Many details are perceived through the microscope, whereas perspectives and trajectories vanish. The wide-angle lens shows how civilizations, tribes and families grow, flourish and degrade, and how all the lines break, branch out and often curve down. At the same time, in this case the researcher finds no correlation between the parameters of social transformation in different local objects and comes to the conclusion that history is multiline or cyclic. He notices separate trees, bushes, branches and leaves, each of which is mainly original, but a wide-angle lens captures no long-term trends or regularities. Not to miss the forest for the trees, a telephoto lens is required, which opens the smallest scale and thus very large time and space blocks. This makes it possible to compare the states of society for highly remote time sectors. In this case, we may observe a set of reliable correlations and also reveal that neither tribes or states, nor “civilizations” but humanity in its broadest sense, and even the whole hominidae family, has been the subject of evolution. Similarly, to discover the global biological evolution one should de-emphasize separate populations, species or even ecosystems and compare the conditions of the biosphere at different levels of the geological timetable; this way the successive growth of morphological and behavioral diversity and “intellectual” qualities, and the increasing influence of the biota on the geological processes are evident. Since hominids have once and for all turned to tool making, in spite of countless divergences, migrations and isolations, culture as a super-natural reality has been a single and common planetary phenomenon, which is proved by multiple particular observations. As to the explosive growth of local varieties since the Middle ­Paleolithic, it was a typical process of the evolving system’s inner diversification. Turning to the telescopic retrospection, it makes obvious the fact of directional social transformations in the sweep of time, as well as the conjugation of the vectors, which can hardly be shown on the scale of separate societies. We have singled out six conjunct social evolution vectors: growth of world population, technological power, organizational complexity and information capacity of the intelligence, perfection of mechanisms for cultural regulation and growing specific weight of virtual realities. The first three vectors are deduced as “empirical generalizations” and may be easily supplied with mathematical figures. The other three have been traced back by special methods and arguments. However, all of them keep within the integral grotesque formula “moving away from the natural state”, i.e. the integral society-nature system has been successively withdrawing from the “natural” (“wild”) condition, assuming more and more anthropomorphic and culture-centered qualities; the degree of tool (including sign) mediation in society-nature and intra-social relationships and individual psychic reflection has been increasing. So the kernel of global causalities was successively shifting towards the mental phenomena, especially after the first Neolithic agrocenoses marked the initial regeneration of the wild biosphere into the anthroposphere or Noosphere – an integral nature-culture system. Leadership in the many-thousand-years marathon has intermittently shifted from one region or continent to another including Australia (the first cave pictures, stone 422

Big history and Singularity

tools with polished handle and blade, and the first boats were invented there); Europe and then North America were preeminent for the latest centuries. The most paradoxical fact is discovered as the anthropogenic crises, especially the global ones, are analyzed. In spite of some modern ecologists’ slogans (“Back to the Nature!” etc.), each aggravation in society-nature relations has been radically overcome not by convergence between society and nature but vice versa, by a next spire of “denaturalization” of the society together with its native habitat. We may easily reveal this fact by comparing the hunter-gatherer economy with the food production, or the industrial civilization with the agricultural one, or the information society with the industrial one. Each leap was foregone by a complex crisis of the former activities and accompanied by transformations of all the conjunct parameters. As a result, humans’ ecological niche broadened and deepened, the population increased, along with technologies, needs and ambitions, and… the path towards the next crisis continued.

The pattern of techno-humanitarian balance During World War II, the German philosopher and sociologist Norbert Elias, a Jew who had lost his relatives in the Holocaust, demonstrated with figures that the “civilizing process” had been reducing the percentage of violent deaths [Elias 1939/2000]. Later on, this suggestion was confirmed by the comparative calculations made by ­British, American [e.g. Pinker 2011] and Russian scholars. Thus, we used a cross-­ cultural index – Bloodshed Ratio (BR), or the ratio of the average number of k­ illings (K) per unit of time to a population size (P ) during a given period (Δt). The number of killings included wars, political repression and everyday violence:

BR =

K ( ∆t ) P ( ∆t ) 

(1)

A more specified formula is applied to consider the BR per centuries. In a whole, specific estimates have demonstrated that over the course of millennia the violent death rate has been non-linearly but successively decreasing while both the destructive potential and population densities have had a distinctly upward trend.1 It can hardly mean humans’ “lowering aggressiveness”: inversely, the psychological experiences show that population densities beyond the natural ecological niche make humans like the rest of animals increasingly aggressive. To explain the contrasting combination of the long-term trends, we should assume a more likely factor, which has compensated for the growth in tool potential. A hypothesis to explain its essence arises from different empirical data; in fact, our calculations are conducted to check a corollary of the hypothesis. Summing up diverse information from cultural anthropology, history and historical psychology concerning anthropogenic catastrophes, we find a regular relation between three variables: technological potential, quality of cultural regulation and social sustainability. The law of techno-humanitarian balance states that the higher the power of production and war technologies, the more advanced behavior-restraint is required to enable self-preservation of the society. 423

Akop P. Nazaretyan

What we refer to as the law is inferred from empirical observations. Respective hypotheses claim that this has been a selective mechanism of viable social systems over the time of human history and prehistory. The circumstances of the early hominids’ existence were of the kind that only an essential development of tool intelligence gave them a chance to survive. Meanwhile, having begun tool making, they dramatically interfered with the ethological balance between the force of wild animals’ natural weapons and the instinctive inhibition of intra-species killing. The power of artificial weapons rapidly exceeded the power of instinctive aggression-inhibition (the Homo habilis in the Olduvai Gorge used to crush one another’s skulls with their choppers), and the proportion of mortal conflicts grew to the extent incompatible with the original tool-makers’ further existence. This can be the main reason for the fact demonstrated in archeology: many groups seem to have been on the borderline between animals and proto-humans, yet very few could have crossed it. Since the individuals with normal animal motivation were doomed to mutual destruction in the new unnatural conditions, certain psychasthenic and hysterical individuals got selective privileges. Their survival required artificial (beyond biological instincts) collective regulation, which was paradoxically provided by pathological changes in the psycho-nervous system, abnormal mental liability, suggestibility, and phobias.Thus, the origins of animism and irrational fear of the dead and posthumous revenge is supposed to strongly restrain in-group aggression and stimulate care for the handicapped: archeology gives us evidence of such biologically senseless facts in the Early Paleolithic. The assumption of a “herd of neurotics” as our remote ancestors has been thoroughly argued by neurologists, cultural anthropologists and psychologists. Here, the relevant point is that the initial forms of proto-culture and proto-morals emerged as an outcome of the first existential crisis in human prehistory. From the Habilis on, hominids’ unnatural intra-species killing facility seems to have been a key problem of pre-human and human history: the ways of solving this existential problem influenced essentially the forms of social organization, cultural and spiritual processes. So far as further life of the hominidae family (­including our own species, the Neoanthropes) has not had a natural background any longer, it was to a great extent enabled by the adequacy of cultural regulation with technological power. As the tool-makers were increasing their power and aggressiveness, culture developed more and more intricate means of ­aggression-sublimation to adjust to the growing destructive facilities; the mechanism of techno-­humanitarian balance was discarding social organisms that could not adapt to their tools’ power. The pattern resolves the paradox of decreasing physical violence versus growing destructive resources. Besides, it helps explain causally both the sudden collapses of flourishing societies and the breakthroughs of humanity into new historical epochs (which often look still more mysterious). For an initial and rough guide, a formal apparatus distinguishes between internal and external sustainability. The former, Si, expresses the social system’s capability to keep away from endogenous catastrophes. The latter, Se, is capability to withstand fluctuations in the natural and geopolitical habitat. 424

Big history and Singularity

If we refer to the quality of cultural regulation as R, and technological potential as T, a simple equation represents the pattern:

Si =

f1 (R ) f 2 (T ) 

(2)

It stands to reason that T > 0, for in case of no technology at all we are dealing with a herd (not a society) where biological causalities are effective. If technological potential is very low, primitive regulation means is sufficient, as in case of the Paleolithic tribes (like the regular infanticide to prevent demographic overflow). A system is highly sustainable, up to stagnation, as cultural regulation quality exceeds the technological might (Medieval China is a textbook example). Finally, the denominator growth raises the probability of anthropogenic crises if it is not compensated by the numerator growth. The aggravating misbalance usually provokes the psychological effects, which entail a crisis-causing behavior. In essence, once the new technologies exceed the former cultural restrictions, public attitudes and sentiments get peculiar features. A sense of omnipotence and permissiveness is intensified together with the increasing needs and ambitions. Success euphoria produces an impatient expectation of new successes and an irrational thirst for “small victorious wars” – a mass complex of catastrophophilia, in terms of Peter Sloterdijk [1983].The subjugation process and a search for new moderately resisting enemies are getting self-valuable, while as we know from the specific experiments [Petrenko 2010], strong emotions flatten the worldview (reduce the semantic space dimensionality). A more primitive worldview entails impulsive decision-making, and the numerator index in equation (2), instead of increasing in proportion to the denominator’s growth, is falling.Thus the cultural imbalance lowers the society’s sustainability. Abstracting here from more psychological details, suffice it to note that the unbalance is fraught with ruinous effects either in case of war or production technologies. For instance, Toynbee [1987] cited various examples to illustrate the inverse relationship between “military and social progress” and was puzzled by the fact that this was true about production tools as well as weapons. William McNeill [1992: 148] wrote: “It certainly seems as though… every heightening of efficiency in production were matched by a new vulnerability to breakdown”. Numerous facts gathered in relevant papers testify to the distressing destiny of societies that could not anticipate the delayed effects of their economic activities. In spite of all peculiarities, a common script was simple: increasing intervention into the ecosystem → landscape destruction → social catastrophe. In contrast, particular studies of wars and natural hazards have demonstrated that the external sustainability is the technological potential’s positive function:

Se = g (T ...)

(3)

Hereby, growing technological potential makes a social system less vulnerable to external fluctuations and more vulnerable to the internal ones, i.e. mass mental states, failed decisions of influential leaders or other destructive individual activities (less “fool-proof”). One more conclusion is that the specific weight of anthropogenic crises versus the ones caused by outside factors (spontaneous climate fluctuations, geological or 425

Akop P. Nazaretyan

cosmic cataclysms, aggressive nomads, and so on) has been historically increasing. Accordingly, time intervals between the global man-made crises in history have been successively shortening. What is still more remarkable, this trend keeps on the biospheric trend of evolution-acceleration.

The Singularity puzzle In any visual representation the cosmological arrow appears rectilinear; yet, the cumulative changes have not been uniform. During the first billions of years after the Big Bang, evolution was slowing down until heavy elements were synthesized in the depths of the first generation stars and ejected into the cosmic space by supernova explosions. This initiated an additional self-organization mechanism with competition for free energy (the heavy elements unlike the light ones need energy feed from outside). Thus about 10 billion years ago, as evolution went its way towards organic molecules and living matter, the slowdown changed into acceleration [Panov 2005b]. The solar system emerged around 4.6 billion years ago, and the first signs of living organisms on Earth are recorded since about 4 billion years ago2; thus our planet was one of (probably various) points on which further cosmic evolution was localized. Although the fact of its consecutive acceleration is obvious for any global analyst, an additional and wonderful discovery was made in recent decades.The Australian economist Graeme Snooks [1996], the Russian physicist Alexander Panov [2005a, 2005b] and the American mathematician Raymond Kurzweil [2005] independently on different sources and with different mathematical apparatus compared the successive time intervals between the phase transitions in biospheric, pre-social and social evolution. The calculations demonstrate that the intervals have been shortening in accordance with a rigorous decreasing progression, and thus the evolution on Earth has been accelerating under the logarithmic law (see Figure 19.1).

Figure 19.1  Scaling law in the phase transitions. From [Panov 2005b]. 426

Big history and Singularity

Like all of the fundamental discoveries, the scale is highly counter-intuitive, i.e. it strongly conflicts with the intuitive suggestions. Traditionally, the researchers tended to explain the global catastrophes (like the pangolins’ extinction on the boundary of the Mesozoic and the Cenozoic or the megafauna extinction on the boundary of the Pleistocene and the Holocene) by appealing to some outside challenges: large meteorites, powerful volcanoes, climate changes, etc. Those versions are extremely vulnerable in each particular case, but the table of hyperbolic acceleration debunks this approach for good and all. Continents have been drifting, meteorites falling down, volcanoes erupting and climate changing during the 4 billion years; later on, the wayward Homo sapiens intervened with their free will and never-ending extravagances, and near ten thousand years ago (the Neolithic) the Noosphere started to arise. Nevertheless, the global transitions, which were foregone each time by crises and catastrophes, followed as if there were a schedule. This paradoxical fact turns us to the synergetic pattern, which appeals to accrual entropy accumulation and progressive perfection of anti-entropy mechanisms enabled by the growth of complexity. Particular analysis of the crucial episodes – or transitory singularities – shows that the events could have developed otherwise in each case: the evolution of the biosphere and then the anthroposphere could have suspended (in compliance with the Lotka–Volterra oscillation circuit in ecology) or the sustainable non-equilibrium system could have collapsed in a global catastrophe. In synergetic terms, we call simple attractor the scenarios related to system’s intensive degradation and simplification after the polyfurcation phase. Those related to suspension (interim stabilization on the achieved level of non-equilibrium without complication, which is fraught with system’s gradual degradation in a long-term perspective) refer to the horizontal strange attractor. Yet, you and I live on this planet and enjoy the fruits (and experience the troubles) of postindustrial civilization thanks to the fact that evolution has gone towards the vertical strange attractors in all the turning points, i.e. global sustainability was each time reestablished on a higher level of non-equilibrium and complexity. One more consideration is originated in the system theory and its implementation principle: all of the possible events do happen. From there, we must assume that there are multiple hearths of evolution in the Universe in which all possible scenarios are realized. Very few of them achieve a level comparable to the one we find on Earth while the others implement all of the dead-end scenarios. Finally, having extrapolated the curve into the future, the researchers came to a unanimous and still more striking result: around the mid-twenty-first century the hyperbole comes to the final Singularity point. It turns into a vertical, i.e. the speed of the evolutionary processes tends to infinity and the time intervals between new phase transitions tend to zero. How can we interpret this mysterious mathematical result? Obviously, the evolution on Earth cannot continue the algorithm it has followed for the latest four billion years and a conclusive phase transition comparable to the emergence of life is to occur over the twenty-first century. In other words, the planetary history intrigue is expected to be resolved in this or that way during the next few decades! The most elementary suggestion is that the anthroposphere is approaching the top of possible complexity after which evolution passes into its “descending brunch”: the 427

Akop P. Nazaretyan

anthroposphere will degrade to an unmanned biosphere with further degradation to the sphere of equilibrium. Thus the kernel of the simple attractor is that with a lapse of time Earth will become a “normal” cosmic body like Moon or Mars free from res cogitans and living matter at all. We can trace various scenarios in the network of the same attractor and the duration of the degradation process depends on how exactly the events will go on. It is always more problematic to trace the strange attractors beforehand or even to clear up whether or not they do exist beyond this singularity. The horizontal one might be seen as a kind of Hegelian-like “End of history”. Although the details of a long-term stabilization on the peak of complexity are now hardly imaginable, we must assume its compromise status: sooner or later, the known natural mechanisms will bring the anthroposphere to collapse. Still more difficult is to imagine a vertical strange attractor. In this context, we pay attention at the remarkable turn in modern cosmological thinking. In the t­wentieth century, only some of the Soviet astrophysicists (or the descendants from the USSR) influenced by the “Russian Cosmism” dared to assume human’s potential intervention in the cosmic-scale processes and perspectives. In contrast, serious ­Western scholars shared the belief that life, society, culture, and mind were nothing but epiphenomena (side effects) of spontaneously evolving material structures without any mutual influence on the cosmic processes and doomed to traceless vanish with time. The Nobel Prize winner Steven Weinberg [1993] expressed this common belief by noting that only the awareness of the unavoidable end imparts a tint of a “high tragedy” to the “farce” of human existence. Meanwhile, those “naturalist” scenarios lost their popularity by the beginning of the twenty-first century: following recent publications, we can note a radical change of mind. Assertions about consciousness as a “cosmologically fundamental fact”, the conclusive influence of the developing knowledge on subsequent evolution of the Metagalaxy and the perspectives of  “living cosmos” are widespread among physicists up to an exotic idea of deliberate creation of new universes with preset parameters for posterior emergence of life, etc. (see [Deutsch 1997; Rees 2003; Davies 2004; Smolin 2006; Kaku 2010] and others). We also appeal to the studies in gestalt-psychology and heuristics, which have demonstrated that any boundaries imposed on engineering by physical laws, are surmountable by a change of the cognitive meta-system. Specifically, those parameters of the problem that are uncontrollable constants inside one model become manageable variables within a more complex meta-model; this implies that the facilities of intellectual control may be potentially unlimited. From there, the implementation principle suggests one more conclusion. If the intelligence originated on Earth destroys itself before it realizes those potential universal developments, the role will be fulfilled by another, “presumably some extraterrestrial intelligence” [Deutsch 1997: 353]. Earlier, by extrapolating some arguments derived from the evolution of creative intelligence and its growing intervention into the mass-energy processes on Earth, we supposed that humanity is now unwittingly participating in a universal natural selection of planetary civilizations [Nazaretyan 1991]. As we have assumed that very few of the local hotbeds of evolution achieve the level comparable to the one we find on our planet, this implies the following suggestion. Only those of technologically 428

Big history and Singularity

advanced civilizations (perhaps a single one), which succeed in progressive adjusting their ­aggression-regulation to unlimitedly growing power, can break out to the cosmic stage of evolution. The rest remains universal evolution’s active storage as well as the planetary bio- and Noospheres, which interrupt their evolution at earlier stages. Thus the mechanism described in the pattern of techno-humanitarian balance might remain the determinant at the conclusive stage of civilizations’ planetary histories to enable their selective cosmic relevance. Here, it goes without saying that human intelligence has potentially unlimited capacity to perfect its self-control in compliance with the growing technological power; yet, this belief is not indisputable for a psychologist. It may turn out that some intrinsic attributes (like the innate gestalts) restrain the flexibility of minds, and thus the range of self-control is narrower than the range of technological ingenuity. For instance, the analysis of the historical episodes makes us suspect that both human and perhaps “post-human” (symbiotic; man-machine) intelligence needs an image of enemy for effective group solidarity (“them–us” archetype) and a strategic meaning formation is hampered by prolonged lack of competing agents. Emotional ambivalence programmed in the limbic structures of our brain intermittently induces an unconscious search for the “negative” experiences like fear and hatred and provokes corresponding activities. Although since the most ancient times culture has been developing measures like rituals, art, sports, TV programs or computer games to relieve those functional drives, sooner or later people feel bored with the sublimation measures and the longing for the “not for fun” passions is activated. Grotesquely, it looks as if there were a kind of natural self-destruction program embedded in the mind’s base plate to prevent a cosmic outburst of intelligence. If no measures to effectively overcome those irrational fluctuations are possible, we must suggest that the evolution of complexity on any planet has an extreme boundary and no planetary Noosphere can escape conclusive self-destruction; thus the “Silence of Cosmos” gets a most trivial and pessimistic explanation. This would mean that in spite of our intuitive belief, the mental realities are more rigid than the physical ones. In other words, the intellectual agent has potentially more power over the mass-energy world than over his own mental conditions and what is feasible from the physical point of view is excluded by the immanent laws of psychology and cultural anthropology. If it is so, this unexpected circumstance can play a fatal role in civilizations’ destiny: just because of it, life and intelligence are indeed no more than epiphenomenal effects and the future cosmic developments are exhaustively described in the naturalist scenarios. In case we still accept that mind’s self-regulating capacity is potentially commensurate to its unlimited technological evolution, we get back to the hypothesis of universal natural selection. So, the nodal question shifts to another realm: whether or not the Earthly intelligence will succeed in upgrading its self-regulation to balance the accelerating breakthrough in technologies prior to their destructive effects become irreversible. As the latest biophysical and paleontological researches have shown (see in particular endnote 2), a spontaneous emergence of a living cell is too highly improbable to happen repeatedly on various planets: once appeared, the biota have most probably “infected” all of the available points in the cosmic space. In all likelihood, if the formation of a Cosmic intelligence is possible, it must be just as unique in its degree and might occur only once at certain stage of universal evolution. 429

Akop P. Nazaretyan

How high are the chances of Earth civilization to implement this unique opportunity? More than ten years ago, the famous English astronomer Sir Martin Rees [2003] estimated its chances to survive the twenty-first century (and obtain cosmic relevance) as 50/50. This corresponded to our own estimation at that time, but now it looks too optimistic. Humanity established a historical record of nonviolence in the first decade of this century: by the UN and the WHO data, from 2000 to 2010, the total sum of violent deaths in the world (armed conflicts, political repressions and everyday violence) numbered about 500,000 a year, while the total population was verging on 7 billion [Krug et al. 2002; Global… 2011]. Although the number of killings looks terrible, this Bloodshed Ratio is unprecedentedly low (lower than the yearly number of suicides in the same period). Some regions show indexes of one and less killings a year for 100,000 habitants. The encouraging facts gave the analysts a timid hope that the trend of virtualization (violence prevalent in the media news, films and computer games) would continue.We expected something like the advanced computer programs for the user’s multisensory involvement in virtual battles to undergo intensive emotional experiences and thus relieve the psychological tensions by means of substitute activity and so on. Perhaps, we underestimated the dynamism of the irrational mood fluctuations among both political leaders and the mass. Unfortunately, since 2011, the situation has taken a turn for the worse. The euphoria and catastrophophilia symptoms first manifested in the late 1990s in the USA (as a result of the victory in the “Cold War”) and in some Muslim regions. Lately, the nostalgia for “small victorious wars” has infected other regions and become a relevant motivation. The political leaders’ intellectual qualities and readiness to estimate the delayed consequences are decreasing (compared to their forerunners in the 1970–1980s), international law is being abandoned and the global geopolitical system is losing its sustainability. Earth civilization successfully completed the twentieth century for it had managed to solve the global menaces of those times. Actually, we have anyhow learned to deal with the population growth and ecological contaminations and psychologically adjusted to the nuclear weapon, but are facing the new global problems. In Bill Joy’s [2000] words, the century of weapon of mass destruction was changed by the century of knowledge-enabled destruction. The boundaries between the states of peace and war as well as between war, production and everyday technologies are diffusing (so it was in the Paleolithic), while spreading access to education makes the destructive means every year cheaper and more easily accessible. So the “sophisticated” weapons are slipping out of governments’ control and falling in hands of irresponsible groups and individuals free from the habits of long-term and system anticipation. Another aggravating crisis is still more paradoxically related to the greatest successes in humanist culture. In the early nineteenth century, 1/3 of English children outlived the age of five years, while current children’s mortality in the post-industrial regions is less than 1%. The integral longevities have increased four times during the 200 years and the pay-off for the unprecedentedly high value of individual lives in modern societies is genetic load exponential accumulation. The humans’ biological

430

Big history and Singularity

wellbeing depends more and more on life comfort, perfecting medical care and other artificial conditions. A linear extrapolation shows that the trend of biological degeneration can irreversibly affect human brains around the mid-twenty-first century if effective contra-measures are not undertaken. Thus, without genetic engineering and other technological interventions into the most intimate foundations of humans’ existence, our species is doomed to peter out, whereas the newly developing technologies carry new menaces of both destructive errors and abuse. Researching the global dangers, we find one that may be the pivotal problem in the next decades; it is related to meaning formation. Over millennia, humans have been seeking their meanings of life mainly in the context of religious or quasi-­ religious ideologies, which are always built in the matrix of friend-or-foe discrimination. Tribes and states, confessions, nations and classes have been designing their inner solidarity (in-group aggression-sublimation) by means of shared aversion to the “strangers”. Service to the macro-group sacred idols and expected reward for the confrontation against the alien (“hostile”) ones has been the background for both the group and individual life meanings. As soon as an ideological content with similar group identity involved across a vast geographical and cultural region, discriminations inevitably followed (by religious sects and movements, nations and sub-nations, testament or class confrontations) to abate in-group aggression by venting it outside. This anti-entropy mechanism has worked effectively throughout history. Meanwhile, the synergetic law of delayed dysfunction claims that the productive mechanisms at the previous stages of the system’s development turn destructive (fraught with a catastrophic entropy growth) at a following stage.Thus, while the task of the humanitarian culture was putting in order and transferring social violence (to escape as much as possible its chaotic forms), ideological worldviews served for social sustainability. Since the new historical stage has set the task of removing physical violence as a condition for global survival, most of the outdated sustaining procedures are counter-productive. Therefore, turning back to the pattern of techno-humanitarian balance, the key question for the Earth or any other planetary civilization’s destiny behind the ­Singularity is whether or not the strategic life meanings can be designed above ideological worldviews and macro-group discriminations for a non-confrontational solidarity. In another drafting, the same question might sound as follows: How far can the development in morals and concomitant aggression-restrictors go? To what extent of conscience can our mind and even our brain elevate without losing its motivations and the will towards activity? Theoretically, modern cross-disciplinary worldviews accumulated in Mega-History, unlike the classical naturalism, might warrant new universal meanings and motivations free from ideologies; yet, how real are the chances to massively assimilate it in the next few decades? Accelerating technological development and spreading education are unprecedentedly raising the global role of the individual activities and mentalities. In view of the approaching Singularity, the crossroads of the current historical phase look extremely dramatic: perhaps, our earthly wives are now giving birth to either the potential gods with access to some forms of immortality and cosmic supremacy or the generation of suicides who will finally crumble the Noosphere…

431

Akop P. Nazaretyan

Notes 1 Thus, the twentieth century does not look as incomparably sanguinary as we usually see it proceeding from the habitual Eurocentric position. In fact, Europe had lived relatively unworried (compared to other regions) during 266 years between the ­Westphalian Peace Treaty (1648) and the World War I (1914), till the outside world remained a large reservoir for the aggression-overshoot. As we consider globally, the nineteenth century is not inferior to the twentieth century even in the absolute figures of war, genocide, and everyday violence (the Chinese historians indicate that from 60 to 100 million people perished in sum of the Opium Wars and the Taiping Insurrection [Wang Yumin 1993; Cao Shuji 2001]) and exceeds it several times in relation to the population sizes. As we compare remote historical époques (even coexisting in time), the difference achieves orders of magnitude [Keeley 1996]. 2 Recent discoveries in paleontology, biophysics, and cosmology have reinforced the hypothesis of the cosmic origin of life: the first organisms supposedly emerged somewhere in the Galaxy, were carried by meteorites and nestled all of suitable planets during 215 million years (one Galactic year). In particular, their first signs on Earth precede the appearance of the oceans [Rozanov 2009].

Bibliography Cao Shuji. A history of the Chinese population: The Qing dynasty. Shanghai: Fudan Univ. Press, 2001 (in Chinese). Chaisson E.J. Epic of evolution. Seven ages of the cosmos. New York: Colombia Univ. Press, 2006. Christian D. The Case for ‘Big History’. Journal of World History 2 (2), 1991: 223–238. Davies P. The cosmic Blueprint. New discoveries in nature’s creative ability to order the Universe. Philadelphia & London: Templeton Press, 2004. Deutsch D. The Fabric of Reality. London, New York: Allen Lane, The Penguin Press, 1997. Dyakonov I.M. The ways of history. From the ancient humans to modern times. Moscow: “Oriental Literature” RAS, 1994 (in Russian). Eden A.H., Moor J.H., Søraker J.H. and Steinhart E. (Eds.). Singularity hypotheses. A scientific and philosophical assessment. Berlin Heidelberg: Springer-Verlag, 2012. Elias N. The civilizing process: Sociogenetic and psychogenetic investigations. Rev. ed. Cambridge, MA: Blackwell, 1939/2000. Global study of homicide. Trends, contexts, data. UNODC, 2011. Joy B. Why the future doesn’t need us?. Wired, 2000, April: 238–262. Kaku M. Physics of the Future: How Science Will Shape Human Destiny and Our Daily Lives by the Year 2100. New York, etc.: Doubleday, 2010. Keeley L.H. War before civilization. The myth of the peaceful savage. New York: Oxford Univ. Press, 1996. Krug E.G., Dahlberg L.L., Mercy J.A., Zwi A.B. and Lozano R. Eds. World report on ­v iolence and health. Geneva: World Health Organization, 2002. Kurzweil R. The singularity is near: When humans transcend biology. New York: PG, 2005. McNeill W.H. Control and catastrophe in human affairs. The Global Condition: Conquerors, Catastrophes and Community. Princeton, NJ: Princeton Univ. Press, 1992: 133–149. Nazaretyan A.P. Book review. Minds & Machines, 27, 2014a: 245–248.

432

Big history and Singularity

Nazaretyan A.P. Evolution of non-violence: Studies in big history, self-organization and historical psychology. Saarbrucken: LAP, 2010a. Nazaretyan A.P. Fear of the dead as a factor in social self-organization. Journal for the ­Theory of Social Behaviour, 35 (2), 2005a: 155–169. Nazaretyan A.P. Global crises and the meaning of life. World Affairs – The Journal of ­International Issues, 18(3), 2014b: 10–34. Nazaretyan A.P. Intelligence in the Universe: Sources, Formation, and Prospects]. Moscow: Nedra, 1991 (In Russian). Nazaretyan A.P. Nonlinear Futures: Mega-History, complexity theory, anthropology & psychology for global forecasting. 3rd ed. Moscow: Argamak-Media, 2015 (in Russian). Nazaretyan A.P. Technology, psychology and catastrophes. Social Evolution & History. Studies in the Evolution of Human Societies, 8 (2), 2009: 102–132. Nazaretyan A.P. Virtualization of social violence: A sign of our époque?. Societal and ­Political Psychology International Review, 1(2), 2010b: 23–36. Nazaretyan A.P. Western and Russian traditions in Big History: A philosophical insight. Journal for General Philosophy of Science, 36, 2005b: 63–80. Panov A.D. Scaling law of the biological evolution and the hypothesis of the self-consistent Galaxy origin of life. Advances in Space Research, 36, 2005a: 220–225. Panov A.D. The singular point of history. Social Sciences Today, (1), 2005b: 122–137 (in Russian). Petrenko V.F. The multidimensional mind: A psycho-semantic paradigm. Moscow: New Chronograph, 2010. Pinker S. The better angels of our nature. The decline of violence in history and its causes. New York: Viking Penguin, 2011. Porshnev B.Ph. Social psychology & history. Moscow: Nauka, 1996 (in Russian). Rees M.J. Our final century: Will the human race survive the twenty-first century? New York: Basic Books, 2003. Rozanov A.Yu. Life conditions on early Earth after 4.0 bil. years ago. Problems of the ­emergence of life. Moscow: RAS, 2009: 185–201 (in Russian). Sloterdijk P. Kritik der zynischen Vernunft. 1 und 2. Bnd. Frankfurt am Main: Edition Suhrkamp, 1983. Smolin Lee. The trouble with physics. Houghton Mifflin and Penguin (UK), 2006. Snooks G.D. The dynamic society. Exploring the sources of global change. London and New York: Routledge, 1996. Toynbee A.J. A Study of History: Abridgement of Volumes I–VI. New York, Oxford: ­ Oxford Univ. Press, 1987. Wang Yumin. Debating the so-called ‘death toll exceeding one hundred million’ during the Taiping Revolution period. Academic Monthly, (6), 1993: 41–50 (in Chinese). Weinberg S. The First Three Minutes: A Modern View of the Origin of the Universe. New York: Basic Books, 1993.

433

20 UNDERGROUND METRO SYSTEMS A durable geological proxy of rapid urban population growth and energy consumption during the Anthropocene Mark Williams, Matt Edgeworth, Jan Zalasiewicz, Colin N. Waters, Will Steffen, Alexander P. Wolfe, Nicholas J. Minter, Alejandro Cearreta, Agnieszka Gałuszka, Peter Haff, John McNeill, Andrew Revkin, Daniel deB. Richter, Simon Price, and Colin Summerhayes Introduction Global human population grew slowly between 10000 BCE and 1000 CE from ~ 4 million to around 265 million (Ortiz-Ospina & Rosen, 2016). It rose more rapidly to circa 900 million by 1800, and thereafter to 7 billion people by 2011. As human population has grown, the urban population has expanded even more rapidly, exceeding more than 50% of the total population for the first time probably during 2007 (UNDESA, 2012). Urban populations have had a significant impact on energy flow in the biosphere and concurrently in the technosphere (Haff, 2014), the globally emergent system that includes technology, humans, their domesticated animals and plants, and the structures and networks that connect and support these components: current urban energy consumption is estimated at 371 EJ (1 EJ = 1018 J) (including energy from fossil fuels, nuclear reactors, and renewable sources; see below, and Table 20.1). A direct measure of urbanisation is the growing development of mechanised transport systems since the early nineteenth century that allowed for the rapid transit of goods and people (Figure 20.1).These reflect new styles of human behaviour, involving long-distance commuting between work and home, and the massing of people in urban centres.These new patterns of movement correspond closely with accelerating rates of flow in the distribution of material resources such as fossil fuels – the main source of energy for powering the growing industrial economy. Part of this development of rapid transit systems is the construction of underground metro systems to service an increasing urban population. The world’s first underground metro was 434

Table 20.1  Urban primary energy use: 1850–2010 Year Total global population

1850 1860 1870 1880 1890 1900 1910 1920 1930 1940 1950 1960 1970 1980 1990 2000 2010

Urban population

Total Urban primary population energy use

Urban primary energy use

Urban primary energy use

Per capita Number of urban operating energy use metro systems

(PT , billions) (PU , billions) (% of PT)

(ET , 1018 J) (EU , 1018 J) (% of ET)

(PCEU , 109 J)

1.28 1.34 1.34 1.42 1.54 1.66 1.78 1.92 2.10 2.32 2.53 3.02 3.69 4.44 5.29 6.12 6.91

27.8 29.8 32.9 36.0 42.1 49.4 61.7 69.2 78.9 89.9 112.3 160.0 230.7 313.3 379.7 434.4 547.0a

39 40 44 45 48 51 59 60 61 62 69 80 92 101 100 97 106

0.13 0.14 0.15 0.18 0.22 0.27 0.32 0.38 0.47 0.57 0.74 1.00 1.34 1.75 2.29 2.86 3.51

10 11 11 13 14 16 18 20 22 25 29 33 36 39 43 47 51

5.1 5.7 6.8 8.1 10.4 13.7 18.8 22.9 28.7 35.7 50.7 79.4 123.0 177.0 229.0 277.0 371.0

18 19 21 22 25 28 30 33 36 40 45 50 53 56 60 64 68

0 0 1 1 1 5 11 14 17 20 21 29 37 61 87 112 138

a Extrapolated from 2008 value. Data sources: ET: Grubler et al. (2012); PT and PU: History Database of the Global Environment (2013). Note that whilst urban population became half of the global population between 2000 and 2010, urban energy use became half of global energy use around 1960.

Figure 20.1  L  inear correlation between the number of operational metros (1859–2010) and global urban population (data from Table 20.1). A nearly identical correlation (R 2=0.99; t=3.15; p=0.006) is obtained between metro numbers and total primary urban energy use.

Mark Williams et al.

the Metropolitan Line, opened in London, UK, on January 10th, 1863. London and its underground system are part of a phenomenon that has been unfolding for over 150 years (Figure 20.2), with the expansion of metro systems, especially since the 1950s, producing a global pattern on all continents with the exception of Antarctica (Metrobits.org). Metro systems form complex structures extending beneath hundreds of km2 of cities (Figures 20.2 and 20.3). Their tunnels are cut through pre-existing geology, but they also have surface expressions grounded in the evolving urban stratigraphy where they are located. Though clearly contained within a historical narrative with firm dates on construction, when viewed from a distant future geological perspective, the global spread of metros will appear essentially isochronous (becoming global in circa 15 decades), depending on the scale of analysis and degree of stratigraphic resolution applied. In this paper, we show how the geological signal of underground metro systems constitutes a proxy for changes in urban population and energy consumption in the nineteenth to twentieth centuries that reflects a substantial modification of energy flow through the biosphere, and which also signals the growing importance of the technosphere (Haff, 2014), particularly its urban component (Zalasiewicz et al., 2016a).Viewed in this sense, metros are also a component of anthroturbation (human

Figure 20.2  D  evelopment of  the London Underground System from 1863 to present, showing a broad reduction in age of construction and transfer from subsurface to surface lines and stations towards the network periphery. Above ground network information supplied by ‘Transport for London [https://tfl.gov.uk]’ and age of construction determined from ‘Evolution of the London Underground – YouTube’ downloaded June 4th, 2016 [https://www.youtube.com/watch?v=U9Tldw1c0K0].

436

Underground metro systems

Figure 20.3  T  he relationships of metros with surface anthroturbation, the archaeosphere, and deep geology. Metros are part of the set of human trace fossils (including mines and quarries) that, uniquely for the biosphere, penetrate deep into the subsurface geology. Note that the above ground city, the archaeosphere, and the metro tunnels are not drawn to scale.

bioturbation, as defined in Zalasiewicz et al., 2014b) with their signature commencing in the mid-nineteenth century (1860s). Whilst this inception predates the most geographically widespread stratigraphic signals of human impact on Earth by some nine decades (Waters et al., 2016), metros also form large, durable features linked with and related to surface strata formed during the proposed Anthropocene Epoch of geological time (Crutzen and Stoermer, 2000; Crutzen, 2002; Zalasiewicz et al., 2008). We define metros here as the broad equivalent of subways, meaning rail-driven urban and suburban mass transport systems that run mainly underground and differ from light rapid transit and commuter trains by having dedicated trackways that operate independently of other trains and other forms of traffic (such as automotive, pedestrian, and cycle).

Energy flow in the biosphere from an urban perspective Metro systems developed in conjunction with urban population growth and increasing energy consumption, and they are a physical stratigraphical proxy of the change of Homo sapiens from a rural- to dominantly urban species, and of the highly concentrated energy consumption by humans in urban areas. In 1800, less than 5% of the world’s population was urban. This rose to circa 13% in 1900, and 30% in 1950. By 2007, 50% of the world’s population lived in urbanised areas (including cities, suburbs, and smaller urban centres), and by 2014, this had risen to 54% (United Nations, 2014, using a variety of definitions of ‘urban’, with the most basic unit having 2500 inhabitants). Forward projections suggest 60% of the 437

Mark Williams et al.

human population will be urban by 2050 (United Nations, 2014). Thus, after nearly three million years of being rural, the genus Homo became dominantly urban in the twenty-first century. This change, most of which occurred in the twentieth century, is driven by a variety of factors that include the concentration of major investment and economic opportunities in urban areas, with 97% of the world’s GDP based on industry and services in and around urban centres (Satterthwaite, 2007). Efficient infrastructure, such as metro systems provide, helps support the dominant role of cities in human productivity (Bettencourt et al., 2007) by enhancing connectivity between urban residents. World human total energy consumption in 2013 was 567 EJ. This corresponds to the energy release of 13,5411 million tonnes of oil equivalent (‘Mtoe’; Key World Energy Statistics, 2015), a figure that includes hydrocarbon, nuclear, hydroelectric and renewable sources. Urban energy consumption forms a major c­ omponent of this. A recent study of the Great Acceleration (Steffen et al., 2015), defined as the interlinked late twentieth century changes simultaneously sweeping across the ­socio-economic and biophysical spheres of the Earth System, illustrates the ­extremely rapid increase since about 1950 in global population, urban population, and world primary energy use. Urban primary energy use can be calculated from these three other datasets if the per capita energy use of urban and non-urban dwellers is known, and if the relationship between urban and non-urban is assumed to remain constant. The International Energy Agency (IEA) estimated that about 67% of world primary energy use in 2006 was in urban areas (Grubler and Schulz, 2013). They noted, however, that there is much variation amongst countries, regions, and urban areas in the ratio of per capita energy use in urban and non-urban areas, so their global assessment of urban energy use is only a rough estimate. In 2006, the urban population was very close to 50% of the total population, so the IEA estimate implies that urban dwellers, on average, used twice-as-much energy per capita as their non-urban counterparts. For this first-order estimate of urban energy use, we assume that this 2:1 relationship holds for the 1850–2010 period. Given the low global energy use before the mid-twentieth century, our estimate for the 1850–1950 period would not change much if this 2:1 scaling did not hold during that period. Total primary energy use is given by:

ET =

(PCE )(P ) U

U

+

(PCE )(P ) NU

NU

(20.1)

where: ET = total primary energy use PCEU= per capita primary energy use in urban areas PCENU= per capita primary energy use in non-urban areas PU= population of urban areas PNU= population of non-urban areas Based on the IEA estimates, we then assume that:

(

)

PCE NU = ( 0.5) PCE U  438

(20.2)

Underground metro systems

This allows us to simplify (20.1) to:

(

)( )

( )

E T = PCE U  PU

+ 0.5 PNU  

(20.3)

Or, solving for the per capita energy use in urban areas:

( )

PCE U = E T /  PU

( )

+ 0.5 PNU 

(20.4)

The urban primary energy use is then estimated simply by multiplying the outcome of equation (20.4) by the urban population:

EU =

(PCE )(P ) U

U

(20.5)

Based on this approach, estimates of urban energy use from 1850 to 2010 at 10-year intervals are given in Table 20.1. Such energy consumption figures bear comparison with the available energy in the terrestrial biosphere, not least because if the urban population continues to grow in the twenty-first century, energy consumption in urban areas may reach 730 EJ by 2050 (Creutzig et al., 2014). In the terrestrial biosphere the total net primary production amounts to approximately 118 billion tonnes of plant by dry matter, equating to a gross calorific value of 2190 EJ/yr (the above ground harvestable component of this is 67 billion tonnes or 1241 EJ/yr; see GEA, 2012, chapter 7, section 7.7.1): humans currently harvest or destroy about 30% of the above ground Net Primary Production (NPP), equating to 373 EJ/yr (see also Haberl et al., 2007, 2013). Together with their use of hydrocarbon, nuclear, and renewable sources, this represents a significant modification to patterns of energy flow in the biosphere by one species (Homo sapiens) and is one measure of human impact on the biosphere. The geological legacy of these demographic and energetic transformations is considerable. Urban structures contain many novel materials that have a high geological preservation potential, including bricks, concrete, glass, and various metals. These contribute to a trace- and technofossil record of the development of urban centres, reflecting the globalization of building materials and architectural styles in the twentieth century, and the commensurate rapid growth of the urban fabric. They are therefore a physical record of the increasing urban population of the twentieth and twenty-first centuries, of increasing urban energy use (Table 20.1), and of the growth of the urban technosphere (Zalasiewicz et al., 2016a). One component of this urban stratigraphical record is the trace fossil signature of underground metro systems, which, because of their deep burrows (tunnels) into pre-existing urban structures and geology (incorporating the novel materials noted above into the tunnel architecture), are likely to provide a long-lived geological record.

Metros as geological traces of human activity Here we deal with metros as geological phenomena, though note that they can also be analyzed in terms of the archaeological, historical, and cultural development of humans. As such, we use terminology developed in geology to examine these 439

Mark Williams et al.

structures, being cognisant that they can also be explained as part of the artifactual evidence of humans, and in terms of archaeological stratigraphy. Trace fossils are biogenic sedimentary structures that result from the behavioural interactions between organisms and substrates; and include trackways and trails, burrows, borings, bite marks, and coprolites (Bertling et al., 2006; Minter et al., 2007). The trace fossil record of organism activity is evident more than 541 million years ago in the latter part of the Proterozoic Eon of geological time (Jensen, 2003; Matz et al., 2008). This record becomes diverse in marine sedimentary rocks from the beginning of the Phanerozoic Eon, the last ~541 million years of Earth history, reflecting the rapid evolution of complex animal-based marine ecosystems early in the Cambrian Period. Indeed, the beginning of the Phanerozoic is defined by the appearance of the trace fossil Treptichnus pedum as a biostratigraphic marker (Landing et al., 2013). The trace fossil record subsequently extended to the land during the Ordovician and Silurian periods and provides a wealth of information on organism behaviour through time, and of responses to major evolutionary transitions (­Plotnick, 2012; Mángano & Buatois, 2016a, 2016b), latterly including that of humans. Urban underground metros can be seen as one of the latest components of this record, recording significant changes in the behaviour of humans (for example, massing in huge cities), and therefore of the biosphere, in the nineteenth to ­twenty-first centuries. Metros can also be seen as a distinctive signature of the evolving technosphere (Haff, 2014). Trace fossils are diagnosed on the basis of their morphology, which is a function of three factors: the organism or organisms responsible, their behaviour, and the substrate. Bertling et al. (2006) considered signs of human biology to be within the realms of trace fossils, whereas signs of human technology are not. A fine line of distinction may be drawn here, with plastic in and of itself not being a trace fossil (but rather a technofossil); however, a plastic beaker would be a trace fossil with the plastic being the substrate that has been modified by the behaviour of the producer into a beaker. The beaker would, in the sense of Zalasiewicz et al. (2013), also be a technofossil. To be a trace fossil, a structure must be produced by an individual organism or similar (homotypic) organisms; show evidence of how its morphology is influenced by behaviour; and also that it involves modification of a substrate, which may also influence its morphology. It is demonstrable that metros fulfill all three of these criteria. More contentious amongst trace fossil taxonomists will be whether recent structures should be considered as trace fossils and whether they should receive trace fossil names. Trace fossils may be produced by intrusion, compression, backfilling, or excavation (Bromley, 1996). Metros fall into the latter category of production.Whilst undeniably produced by humans, in comparison to other excavated traces they are biomechanically unique in that they are mainly produced by machines, as extensions of human tunnelling capacity, rather than by organism ‘muscle power’ alone (though early metro trenches were excavated by labourers using hand-held tools such as spades, picks, and barrows, with horse-drawn carts and river barges used for transportation of spoil). In addition to the method of trace production, organisms may rework sediment through a variety of means, including regeneration, whereby material is actively brought to 440

Underground metro systems

the surface from below and may then be subject to other processes (François et al., 1997). The materials extracted from metros are often moved considerable distances elsewhere and are used to construct other human trace fossils, such as embankments, in or beyond the original urban fabric. As such, through excavating metros, humans act as large-scale sediment-regenerators. In bringing large quantities of material up from depth and redistributing it on or near the surface, such human activity also constitutes a form of geological uplift. Morphologically, metros are on a scale of kilometres to hundreds of kilometres, with individual structures (e.g. tunnels) typically metres to tens of metres in diameter: many metro systems are between 5 and 100 km in track length, some being truly colossal, including Shanghai (588 km), Beijing (554 km), London (402 km), Paris (383 km), New York (370 km), Moscow (334 km), Seoul (327 km), Tokyo (305 km), ­Mexico (227 km), and Delhi (215 km); (figures from Metrobits.org). They form complexes of interrelated locomotion (tunnels), resting (stations, platforms) and energy-flow traces (e.g. energy cables, used here as an analogy to feeding traces). Some metro systems, as in London, have evolved over many decades (Figure 20.2), whilst others, like Kunming (operational from 2012), are recent. All metros are dynamic structures, with long-lived systems having lines or sections that have ‘died’ as in ­London’s Acton Town branch of the District Line, or which have ‘hibernated’, such as those in B ­ erlin that reopened after re-unification (Metrobits.org). The simplest ‘Line’ forms are expressed in cities such as Liverpool, Bilbao, Mumbai and Lima. Many underground metros show major morphological variation between different systems that reflect their functions. Overall morphologies include circular (e.g. Glasgow, ­Moscow), looping and radial (London) or spiral (Tokyo) components (Metrobits. org). Comparisons may be drawn with graphoglyptids, a group of complex patterned, meandering, spiral and networked trace fossils (Seilacher, 1977). Stations, platforms, and tunnel shapes also differ in their morphologies. Whilst metro systems express a range of different topologies at the gross scale, these do not appear to cluster as regional types: thus circle-radial types can be found as far apart as London and Tokyo (­Metrobits.org). Other factors, such as local geology or geomorphology, and surface urban structure and cost, control design (e.g. in London, see Levinson, 2007). The dynamics underlying metro shapes is suggested by studies of organismal systems such as slime mold networks, which can replicate the patterns of the Tokyo rail system in response to the distribution of resources in their environment (Tero et al., 2010); and ant nests (Tschinkel, 2003; Minter et al., 2012) and army ant swarm raids (Franks et al., 1991), whose morphologies are also influenced by environmental factors. The morphologies of trace fossils may also be influenced by substrate. In the case of the London metro system, construction was only made possible by the properties of London Clay which, despite its local vertical spatial variability in geotechnical properties, formed an ideal tunnelling medium (Paul, 2009). Geotechnical properties, including undrained shear strength of the London Clay Formation, have been shown to depend on lithology, state of weathering, presence of discontinuities, and stress history (Cripps & Taylor, 1986; Chandler & Apted, 1988; Hight et al., 2003). The undrained shear strength of the London Clay Formation typically ranges between 100–175 kN/m2 in its weathered state and 100–400 kN/m2 in its unweathered state (Cripps & Taylor, 1986). This substrate also influenced metro morphology in more 441

Mark Williams et al.

specific ways. For example, some of the excavated clay was used to make bricks to line the tunnels. The world’s first railway tunnel, Fritchley Tunnel, Derbyshire was constructed in 1793, and lined with sandstone blocks. Brunel’s Thames Tunnel, which started construction in 1825 and opened in 1843, was the first tunnel to be successfully constructed beneath a river and to use a tunnelling shield apparatus, with the tunnel lined by bricks to support the tunnel excavation. This technique was subsequently utilised during the development of the London metro.A major advance in tunnel construction came with the development of sprayed concrete lining (shotcrete/gunite), a process patented in 1911, but which became commonly used in the 1970s, allowing much faster lining of tunnels compared with brick lining. This was followed by the use of enhanced tunnel boring machines (TBMs) that excavated tunnels and supported the walls whilst pre-cast concrete cylinder linings were inserted behind the TBM. This is the state-of-the-art construction technique currently being used in the ‘Crossrail’ development under London. Parallels may be drawn here with the boxwork burrower trace fossil genera Ophiomorpha and Thalassinoides. Ostensibly they have the same gross morphology, but differ in that Ophiomorpha has a pellet lining whilst Thalassinoides is unlined (Bertling et al., 2006). These morphological differences are manifest as behavioural responses to substrate conditions. Ophiomorpha is constructed in soft sediment and therefore requires reinforcement through a pellet lining, whereas Thalassinoides is excavated within firm sediment and so does not need to be reinforced. As trace fossils, underground metro systems are uniquely human in their function, scale and complexity. Thus metros have significantly greater complexity and scale as trace fossils than any examples from earlier periods of Earth history. In function they track the development of large urban areas, beginning in London in the nineteenth century. In 1801, London had a population of about 1.1 million (Emsley et al., 2016) and midway through the nineteenth century with circa 3 million people had acquired seven major railway termini delivering commuters to the city. The above ground human congestion this produced was mitigated by the construction of the Metropolitan underground railway line (the line later lent its name to the Paris ‘Metro’, and is the name most synonymous with underground mass transit systems worldwide) between 1861 and 1863.This 6 km long line connected Paddington, St Pancras, Euston, King’s Cross, and Farringdon (Figure 20.2). However, it is important to note that this was not a completed and closed subterranean feature. It continued to be extended through subsequent construction of this and other lines into a much larger and more complex stratigraphic entity, and in fact is still being extended today (Figure 20.2), notably with the construction of Crossrail, currently Europe’s largest construction project, adding 42 km to the network and removing 3 million tonnes of excavated material (http://www.crossrail.co.uk/news/crossrail-in-numbers).

The stratigraphical context of metros Metros are part of the evolving body-fossil, technofossil, trace fossil, and chemical record of the genus Homo that extends back some 2.8 million years in Africa (Villmoare et al., 2015), and is underpinned by a technofossil record of stone tools that is shared with hominins other than Homo (Harmand et al., 2015). This fossil record 442

Underground metro systems

extends throughout Eurasia during the Pleistocene (Garcia et al., 2010), whilst the earliest records of anatomically modern humans is in Morocco at circa 300,000 years ago (­Richter et al., 2017), and Ethiopia 195,000 years ago (McDougall et al., 2005). The early human biostratigraphic signal is sporadic, but widespread extinctions of other large terrestrial animals (‘megafauna’) beginning from circa 50,000 years ago in ­Australia (Roberts et al., 2001), and extending into the Americas from between 13,500 and 11,500 years ago (Koch & Barnosky, 2006; Barnosky, 2008), coupled with a technofossil record of stone tools, provides a clearer though still scattered signal of humans during the latest Pleistocene. The biostratigraphic record of humans becomes more widespread during the Holocene, in fossil evidence for the domestication of animals and plants, the spread of agriculture and animal husbandry (Martin & Sauerborn, 2013 and references therein), and the emergence of early urban centres. Associated with these developments is the gradual accumulation in places of anthropogenically modified ground, often rich in artefact inclusions, and typically coalescing over time into larger formations (Zalasiewicz et al., 2014b; Edgeworth et al., 2015; see ­Figure 20.3). This process accelerated during the Industrial Revolution with a rapid increase in the growth of anthropogenic ground (sometimes referred to as humanly modified or artificial ground), due to the introduction of steam-­powered and later petrol-powered earth-moving machinery, and with associated artefacts and chemical signatures. The size of some artificial features such as quarries was also subject to a vast increase in scale for the same reason: formerly dug by hand using hand-held tools, the new machinery facilitated extraction of materials at unprecedented speeds and volumes, whilst new forms of transport made possible the rapid distribution of materials quarried. A profound change in the stratigraphic record of humans occurred during the twentieth century (Waters et al., 2016), with globally discernible human influences on the biosphere, atmosphere, biogeochemical cycles (e.g. of carbon, nitrogen, and phosphorus), and patterns of terrestrial and marine sedimentation, accompanied by unique new technofossil markers such as plastics (Zalasiewicz et al., 2013, 2016b). This record includes radiogenic isotopes produced by military nuclear detonations from 1945 onwards, which have left a durable (geologically long-lived) signal of human activity in both marine and terrestrial successions, and which has been proposed as a stratigraphic marker for a potential Anthropocene Epoch of geological time (Waters et al., 2015, 2016). Metros are part of this unfolding geological record of humans, with large-scale (tens of metres to kilometres) tunnelling being a human activity extending back into classical times and prehistory, as evidenced, for example, by the Neolithic flint mines at Grimes Graves in Norfolk, England, which have vertical shafts with horizontal galleries extending out laterally (Russell, 2000), by the Siloam tunnel under Jerusalem, thought to date from the eighth century BCE, and by the subterranean aqueducts of Rome and the Roman Empire. Extensive systems of qanats for channelling water throughout the Middle East made use of vertical access shafts combined with horizontal or gently sloping tunnels, some of which extend for tens of kilometres (Döring, 2007). In the last millennium evolving sewer systems beneath great cities, like that of Paris originating from the fourteenth century, presage the development of mechanisms to transport people, and not just their physical waste.Victor Hugo, in his 1862 novel Les Miserables, had already described the Paris system as the ‘intestine of the Leviathan’, describing its function in terms of ‘its arteries, and its circulation, which is of mire and minus the human form’. 443

Mark Williams et al.

Metros are characterised by their connection to the surface at the extant urban stratum and therefore coalesce with and cut through many other surface human traces (Figure 20.3). They are part of the ‘archaeosphere’ – defined by Edgeworth et al. (2015) as the totality of humanly modified ground when considered on a global scale. In common with mines and other artificial subterranean structures that entail the digging of vertical shafts, they penetrate deep below the existing surface, unconformably cross-cut pre-existing urban and natural geological deposits, some being consolidated units many millions of years in age, whilst their components at ground level are approximately conformable with surface anthropogenic strata. In terms of ichnofabrics (the overall texture and fabric of a sediment that arises from bioturbation and bioerosion; Taylor et al., 2003), metros may be considered as deep tier traces, penetrating the homogenised mixed layer of thorough bioturbation into the heterogenous transition layer of partial bioturbation, but above the historical layer of no active bioturbation. Deep tier traces are particularly prevalent in the fossil record. In some cases they are also excavated within lithified hardgrounds (typically carbonate or phosphate cemented surfaces developed at or near the seafloor). Despite being modern traces, they therefore have a high preservation potential. This negates some of the arguments of whether recent structures may be considered as trace fossils because metros have already potentially passed the fossilization barrier (Bertling et al., 2006). There is debate regarding whether modern traces should be given trace fossil names. Trace fossil taxa based upon fossil examples may be applied to recent traces (Bertling et al., 2006) because the essence of the system is for communication amongst researchers; however, we do not advocate the need to name metros formally with ichnogeneric and ichnospecies names. Instead they already possess names. The construction of metro systems has more recent fore-runners in the first long tunnels used to facilitate transport of goods and people, the earliest being the Malpas Tunnel (1679) of the Canal du Midi in southern France. The earliest occurrence of an underground metro system per se is 1855, with the test tunnel built in northeast ­England at Kibblesworth,Tyne, and Wear in 1855. Although this is not associated with a major conurbation, it could be seen as a precursor to the later Tyne and Wear Metro that opened in 1980. Therefore, clear correlation to Anthropocene (and particularly urban) sedimentary deposits forming at that time is not easily constrained.As already mentioned, the first metro tunnel directly associated with a major city and its accumulating stratigraphy is the Metropolitan Line in London, beginning construction in 1861, opening in 1863 and associated with several early stations. It was followed shortly afterwards by the District Line. The initial construction used the ‘cut and cover’ method (Figure 20.4). Deep tunnelling only commenced with the construction of the Northern Line, inaugurated in 1890 (Levinson, 2007) as the system continued to grow throughout its more than 150-year history. Electrification of the London ­Underground in 1890 sparked the spread of metro systems throughout ­Europe and eastern North America (Figures 20.5 and 20.6), though using various propulsion systems. In the 1890s–1900s there ­followed Budapest (1896), Glasgow (1896), Boston (1897), ­Chicago (1892, though much of this is above ground), Paris (1900), Berlin (1902), New York (1904), and Philadelphia (1907) with the first system in South America opening in B ­ uenos Aires in 1913. In the interwar era underground metro systems spread to East Asia (Tokyo 1927, Osaka 1933), and thereafter to the  Indian  subcontinent (Calcutta 1984),  and  Africa  (Cairo  1987). 444

Underground metro systems

Figure 20.4  C  utaway of London’s five levels of traffic at Charing Cross Station (now ­Embankment Station), taken from the Popular Science Magazine, January 1921, pp.  44–45, drawing by S.W. Clatworthy. The District Railway (just b­elow ground) represents an example of a cut and cover construction, whilst the ­Hampstead and Bakerloo railways are deep tunnels. In this way, metro systems like that in London demonstrate the evolution of behaviour within a single trace fossil complex, evolving from ‘at surface’ burrows to deep burrows over time. P ­ icture available through the Popular Science Archive at: http://www.popsci.com/ archive-viewer?id=YSoDAAAAMBAJandpg=44andquery=charing%20cross.

In  the  early twenty-first century metro construction focussed heavily on regions of major economic growth in China and India (Figure 20.6). A rapid geographical spread of underground metro systems occurred in the postWorld War II period (Figures 20.5 and 20.6), coinciding with the interval associated with the Great Acceleration (Steffen et al., 2015), which immediately post-dates the mid-twentieth century (post-1945) radiogenic signature from nuclear arms testing (Waters et al., 2015; Zalasiewicz et al., 2015). Metro systems and lines that were constructed in the period when detonations are likely to have left the strongest radiogenic signals include Stockholm 1950, Toronto 1954, Rome and St Petersburg 1955, Nagoya 1957, Lisbon 1959, and Kiev 1960 (Metrobits.org). The rapid geographical spread of metro systems, their high preservation potential, and their association with a multitude of technofossils (sensu Zalasiewicz et al., 2013) of urban origin suggests that they will form a geologically long-lived signature of changes to urban environments, human population, and energy flow that are part of the evolution of an Earth state that is dominated by humans in the twentieth and twenty-first centuries (Crutzen & Stoermer, 2000; Crutzen, 2002;Williams et al., 2016).Whilst metros are by and large not used for provision of bulk resources, or for outflow of waste 445

Figure 20.5  (A,  B) Urban population growth and energy use 1850–2010 (based on figures in Table 20.1); (C) number of metro systems plotted against time (see also Gonzalez-Navarro & Turner, 2016); (D) urban population as a percentage ­ of ­total population plotted against number of metro systems. The increase in number of metro systems in the post 1950 period is evident and approximates ­economic changes associated with the Great Acceleration (Steffen et al., 2015).

Figure 20.6  G  lobal spread of major metro systems, 1863 to present, with inset maps for Europe and China (data in Metrobits.org). A–D represents the spread of metro systems at various stages since 1863.

Mark Williams et al.

products in the bulk metabolism of a city, they clearly play a vital role in facilitating the daily transport of humans who organise these processes. In fact, the metro system, as it brings in fresh workers in the morning and discharges them in the afternoon, resembles the action of a pair of lungs when a biological organism imports, uses, and then exports a critical resource. Whether it be by reference to lungs, intestines, or other biological processes, analogies between large social organizations and organisms are often deployed to illustrate a point, as when Thomas Hobbes (1651) sketched out his picture of the State as an ‘Artificial Man’. For present purposes, however, the connection between city and organism is much deeper than that of a simple analogy. Recent quantitative studies of urban metabolic processes (e.g., West, 2017) have uncovered a close relationship between metabolic patterns in cities and those in animals, arising from the fact that both kinds of systems are bound by similar physical constraints on resource distribution. Thus urban ichno- and technofossils are not just analogous to biological ichno- and body-fossils, but, at least to the extent they record the past existence of transport mechanisms, as in the case of metros, they share (or shared) the same dynamical relationship with the system that created them as do many fossils of biological origin. Metro systems do more than just passively track the growth of urban areas and their energy use: they actively enhance that growth too. Humans are archetypal ecosystem engineers, modifying their environment in ways that affect themselves and other organisms (Jones et al., 1994). Over geological time, the phenomenon of ecosystem engineering is recognised as having macroevolutionary consequences (­Erwin, 2008). Metros are certainly influencing the community ecology of humans; for example, Golders Green in north London was barely a village when the Northern Line reached it in 1907, but within seven years over 500 new houses had been built (Wolmar, 2004). Thus, they provide an important proxy for the increasing importance of urban centres as regards energy consumption and human population during the twentieth and twenty-first century, particularly so from the 1960s. Whether or not they lead to longer temporalscale macroevolutionary changes will be revealed by the fullness of geological time. Metro systems do not provide a precise biostratigraphic definition for the proposed Anthropocene Epoch of geological time (Waters et al., 2016), as the structures of metro systems themselves do not provide the annual resolution that is important for defining this geological boundary. Speleothems, tree-rings, ice cores, lake, river, and marine sediment successions that cover the interval of the past few hundred years provide this degree of stratigraphical resolution (Zalasiewicz et al., 2014a). Nevertheless, metros do contain artefacts (technofossils), architectural features and materials, even written information (media technofossils) that are often dateable from a historical perspective and broadly correlatable with the ever-accumulating urban strata at the ground surface of most urban centres. Such technofossils include, for example, lighting systems that are identical to those used in above ground buildings and products sold in vending machines such as canned drinks. Both of these technofossil types contain components that might potentially fossilize (that is, they contain recalcitrant materials that might be preserved as geological phenomena) both within metro systems as well as in strata accumulating at ground level, although a portion of this evidence will be diachronous (time-transgressive) on short timescales. Metro systems also preserve distinctive chemical and mineralogical signatures, for example, from airborne particulate matter. Over long time frames these might be flushed out of the tunnel system via groundwater flow, producing a halo effect in 448

Underground metro systems

adjacent deposits. Air in metros shows 2–14 times higher concentrations of particulate matter than found in the ambient air at street level (Cartenì et al., 2015). Metros have their own internal source of aerosols, derived from steel dust resulting from mechanical wear of brake pads, abrasion of rails and wheels. Nanoparticles and their aggregates are produced by condensation of gaseous iron or melted iron generated during electric sparking between catenaries and pantographs (overhead power cables, and the structure mounted on the roof of a train/tram to collect power from them; Jung et al., 2010; Kam et al., 2011; Eom et al., 2013). Studies on the chemical composition of particle samples from metros in different countries have shown that iron is the most abundant element (Nieuwenhuijsen et al., 2007). Of the other elements, the most often reported in aerosol and dust samples are: Ba, Cr, Cu, Mn, Mo, Ni, Pb, Sb, Sn, Sr, and Zn (Chillrud et al., 2004; Guo et al., 2014; Cusack et al., 2015; Qiao et al., 2015). Probable sources of elevated levels of barium and antimony are through wear on brake pads containing barium sulfate and antimony pentasulfide as friction materials (Cusack et al., 2015). Other elements, such as Cr, Mn, Ni are constituents of steel alloys (Chillrud et al., 2004). When viewed from a far-future geological perspective, metro trace fossils may have better preservation potential than many of the more precise stratigraphical records, including those of radiogenic isotopes. From this perspective, metros should be viewed as an important part of the stratigraphic narrative expressed during the Anthropocene, as a proxy that simultaneously tracks fundamental changes in human behaviour, energy use, and the development of the urban technosphere.

Evolutionary significance of metros It can be argued from a range of different parameters that humans have fundamentally changed the biosphere (Ellis, 2015; Williams et al., 2015a, 2016; Zalasiewicz et al., 2016a; Steffen et al., 2016 and references therein).These changes include directed evolution of other organisms, the global translocation of organisms (the ‘neobiota’), extinction (Barnosky et al., 2011, 2012), the appropriation of a major component of terrestrial net primary production (NPP) (e.g. Krausmann et al., 2013) supplemented with fossil NPP and other energy sources, and the increasing interaction of organisms with the technosphere (sensu Haff, 2014).These changes are unique when viewed from the perspective of circa four billion years of evolution of the biosphere. Metro systems are apt stratigraphic markers of this most recent suite of planetary changes, given they developed precisely during the interval when human appropriation of energy resources on Earth became dominant over natural processes (Table 20.1). These changes are also associated with the Great Acceleration (Steffen et al., 2015), and with its global technofossil signal of mass-produced goods (Zalasiewicz et al., 2013, 2016a). The spread of metro systems is coincident with these developments where complex and large-scale (multi-kilometric-scale) structures with the same function (in this case underground mass transit) became essentially global over the period of two consecutive human life times (see Figures 20.1, 20.5, 20.6). Metro systems are therefore a geological marker of the expansion of the urban technosphere (Zalasiewicz et al., 2016a). The overall geographical distribution of metro systems also reflects patterns of human population concentrations in Europe, the eastern seaboard of the Americas and East Asia (Figure 20.6), though some systems are geographically remote from 449

Mark Williams et al.

others, such as Novosibirsk (founded 1986) and Sydney (current extension due for completion in 2019). The latitudinal spread of metros between 60°N (Helsinki, ­northernmost) and 34°S (Buenos Aires, southernmost) is consistent with a geographical pattern that reflects the expansion of the human population during an interglacial phase of the Cenozoic icehouse planetary state, with no metro systems (or major population centres) at polar latitudes. Though this latter point may seem obvious at present, from the perspective of a far-future geological analysis, it would be critical evidence for understanding the prevailing climatic context of human-induced changes to the Earth System in the early twenty-first century, and that this had occurred in an interglacial, rather than glacial stage of the Cenozoic Icehouse. Many underground metro systems are also causally associated with rivers and estuaries along which major cities developed, their constructions extending into deposits well beyond their active zone of sedimentation of the river, thus making them unique trace fossils associated with river and estuarine basins (Williams et al., 2015b). Those rivers and tidal channels are commonly conduits of energy flow via which manufactured goods, food, and energy supplies travel. Whilst metro systems are only a localised component of the global transportation system, their association with rivers and estuaries signals this connectivity. They are also likely to have much greater potential for preservation than road, surface rail, shipping, and air networks, and therefore will signal the increasing population and economic activity of urban areas from the nineteenth to twenty-first centuries. The scale, complexity, and rapid expansion of metro systems and their association with a period of accelerating change to the biosphere may be compared with past fundamental changes to life in deep time. One comparison might be with the traces of organisms used as the primary definition of the Proterozoic–Phanerozoic boundary at circa 541 million years ago, which is founded on the appearance of a suite of trace fossils, most notably Treptichnus pedum, in a succession of marine sedimentary rocks on the Avalon Peninsula in Newfoundland, Canada (Landing et al., 2013). Treptichnus pedum may represent the burrow of a priapulid worm (Vannier et al., 2010) and is a marker for the rapid (from a geological perspective) evolution of Phanerozoic-style marine ecosystems, with their complex inter-relationships between animals. The traces of metro systems also signal a profound reorganisation of the biosphere, one in which humans have become conspicuous consumers, supplemented by their use of energy sources from fossil, nuclear, and renewable sources (Barnosky, 2015; Williams et al., 2016), and one in which urban areas are the most significant consumers of energy (Table 20.1). These urban changes are characteristic of an Earth state where humans have become veritable agents of geological change (Crutzen & Stoermer, 2000; Crutzen, 2002). Whilst the structures and tunnels of metro systems do not provide a stratigraphical resolution on a sub-annual or annual basis that would be necessary to define the boundary of an Anthropocene Series within an historical context (for which see Waters et al., 2016), they do provide a remarkable geological signal, developed in just 15 decades, that may bracket the eventual choice of such a boundary.

Conclusions Metros are trace fossil systems of human activity that extend over hundreds of km2, are of a scale and complexity greater than any biologically mediated tunnelling or 450

Underground metro systems

burrowing structures in pre-human nature, and include uniquely human materials such as concrete, steel, and plastics. Metros have high preservation potential into the geological future, as they penetrate unconformably (“bioturbate”) earlier geological formations. They have spread rapidly across the planet, from their inception in 1861, to become a widely distributed phenomenon of the mid-latitudes of all ice-free continents. Therefore, from a far-future geological perspective their spread will appear effectively instantaneous and will form a significant geological marker of the transition of Homo sapiens from a dominantly rural to urban species in the early twenty-first century, of the immense energy consumption that humans use to sustain urban structures, and of the development of the technosphere, especially its urban component. All of these features reflect the accelerating scale of the human enterprise on urban landscapes, and of the associated energy flows in both the biosphere and technosphere. The tunnels, structures, and artefacts within metro systems originate at, and contribute to, the level of accumulating urban strata of the nineteenth to twenty-first centuries. Therefore, metros also provide an important signal of sedimentary deposits that may be associated with a potential Anthropocene Epoch of geological time.

Acknowledgments We are very grateful to Martin Head (Brock University) for his extensive and perceptive comments on an earlier draft of this manuscript.We also thank three reviewers for their useful comments.We also wish to thank Marco Gonzalez-Navarro (University of Toronto) and Matthew Turner (Brown University) for sharing data pertaining to the temporal evolution of global metro networks.

Note 1 This figure is steadily rising, and by 2014 was 13,699 Mtoe, Key World Energy Statistics (2016).

References Barnosky, A.D., 2008. Megafauna Biomass Trade Off as a Driver of Quaternary and Future Extinctions. PNAS 105, 11543–11548. Barnosky, A.D., 2015. Transforming the Global Energy System is Required to Avoid the Sixth Mass Extinction. MRS Energy and Sustainability: A Review Journal 2. doi:10.1557/ mre.2015.11. Barnosky, A.D., et al., 2011. Has the Earth’s Sixth Mass Extinction Already Arrived? Nature 471, 51–57. Barnosky, A.D., et al., 2012. Approaching a State-Shift in the Biosphere. Nature 486, 52–56. Bertling, M., Braddy, S.J., Bromley, R.G., Demathieu, G.R., Genise, J., Mikulas, R., Nielsen, J.K., Nielsen, K.S.S., Rindsberg, A.K., Schirf, M., Uchman, A., 2006. Names for Trace Fossils: A Uniform Approach. Lethaia 39, 265–286. Bettencourt, L.M.A., Lobo, J., Helbing, D., Kühnert, C., West, G.B., 2007. Growth, Innovation, Scaling, and the Pace of Life in Cities. PNAS 104, 7301–7306. Bromley, R.G., 1996. Trace Fossils: Biology, Taphonomy and Applications. Chapman and Hall. Cartenì, A., Cascetta, F., Campana, S., 2015. Underground and Ground-Level Particulate Matter Concentrations in an Italian Metro System. Atmospheric Environment 101, 328–337. 451

Mark Williams et al.

Chandler, R.J., Apted, J.P. 1988. The Effect of Weathering on the Strength of London Clay. Quarterly Journal of Engineering Geology 21, 59–68. Chillrud, S.N., Epstein, D., Ross, J.M., Sax, S.N., Pederson, D., Spengler, J.D., Kinney, P.L., 2004. Elevated Airborne Exposures of Teenagers to Manganese, Chromium, and Iron from Steel Dust and New York City’s Subway System. Environmental Science and Technology 38, 732–737. Creutzig, F., Baiocchi, G., Bierkandt, R., Pichler, P.-P., Seto, K.C., 2014. Global Typology of  Urban Energy Use and Potentials for an Urbanization Mitigation Wedge. PNAS 112, 6283–6288. Cripps, J.C., Taylor, R.K. 1986. Engineering Characteristics of British Over-Consolidated Clays and Mudrocks. I. Tertiary Deposits. Engineering Geology 22, 349–376. Crutzen, P.J., 2002. Geology of Mankind – the Anthropocene. Nature 415, 23. Crutzen, P.J., Stoermer, E.F., 2000. The Anthropocene. Global Change Newsletter 41, 17–18. Cusack, M., Talbot, N., Ondráček, J., Minguillón, M.C., Martins, V., Klouda, K., Schwarz, J., Ždímal, V., 2015. Variability of Aerosols and Chemical Composition of PM10, PM 2.5 and PM 1 on a Platform of the Prague Underground Metro. Atmospheric Environment 118, 176–183. Döring, M., 2007. Wasser für Gadara. 94 km langer Tunnel antiker Tunnel im Norden Jordaniens entdeckt. Querschnitt 21, 24–35. Edgeworth, M., Richter, D.deB., Waters, C., Haff, P., Neal, C., Price S.J., 2015. Diachronous Beginnings of the Anthropocene: The Lower Bounding Surface of Anthropogenic Deposits. The Anthropocene Review 2, 33–58. Ellis, E.C., 2015. Ecology in an Anthropogenic Biosphere. Ecological  Monographs 85, 287–331. Emsley, C., Hitchcock, T., Shoemaker, R., 2016. London History – A Population History of London, Old Bailey Proceedings Online (www.oldbaileyonline.org, version 7.0, 27th April 2016) Erwin, D.H., 2008. Macroevolution of Ecosystem Engineering, Niche Construction and Diversity. Trends in Ecology and Evolution 23, 304–310. Eom, H.J., Jung, H.J., Sobanska, S., Chung, S.G., Son, Y.S., Kim, J.C., Sunwoo, Y., Ro, C.U. 2013. Iron Speciation of Airborne Subway Particles by the Combined Use of ­Energy Dispersive Electron Probe X-ray Microanalysis and Raman Microspectrometry. Analytical Chemistry 85, 10424–10431. François, F., Poggiale, J.-C., Durbec, J.-P., Stora, G., 1997. A New Approach for the Modeling of Sediment Reworking Induced by a Macrobenthic Community. Acta Biotheoretica 45, 295–319. Franks, N.R., Gomez, N., Goss, S., Deneubourg, J.-L., 1991. The Blind Leading the Blind in Army Ant Raid Patterns: Testing a Model of Self-organization (Hymenoptera: Formicidae). Journal of Insect Behavior 4, 583–607. Garcia, T., Féraud, G., Falguères, C., de Lumley, H., Perrenoud, C., Lordipanizde, D., 2010. Earliest Human Remains in Eurasia: New 40Ar/39Ar Dating of the Dmanisi Hominid-Bearing Levels, Georgia. Quaternary Geochronology 5, 443–451. GEA, 2012. Global Energy Assessment – Toward a Sustainable Future. Cambridge University Press, Cambridge, UK and New York, NY, USA and the International Institute for Applied Systems Analysis, Laxenburg, Austria. http://www.iiasa.ac.at/web/home/research/Flagship-Projects/Global-Energy-Assessment/Chapters_Home.en.html Gonzalez-Navarro, M., Turner, M.A., 2016. Subways and Urban Growth: Evidence from Earth. SERC Discussion Paper 195. Grubler, A., Johansson, T.B., Mundaca, L., Nakicenovic, N., Pachauri, S., Riahl, K., Rogner, H.-H., Strupelt, L., 2012. Chapter 1-Energy Primer. In: Global Energy Assessment (GEA)  – Toward a Sustainable Future. Cambridge: Cambridge University Press and Laxenburg: 452

Underground metro systems

International Institute for Applied Systems Analysis. 99–150. http://pure.iiasa.ac.at/id/ eprint/10068/1/GEA%20Chapter%201%20Energy%20Primer.pdf Grubler, A., Schulz, N.B., 2013. Urban Energy Use. In: Energizing Sustainable Cities, edited by Grubler, A. and Fisk, D. Routledge, Abingdon, Oxon, UK, pp. 57–70. Guo, L., Hu, Y., Hu, Q., Lin, J., Li, C., Chen, J., Li, L., Fu, H., 2014. Characteristics and Chemical Compositions of Particulate Matter Collected at the Selected Metro Stations of Shanghai, China. Science of the Total Environment 496, 443–452. Haberl, H., Erb, K.-H. Krausmann, F., Gaube, V., Bondeau, A., Plutzar, C., Gingrich, S., Lucht, W., Fischer-Kowalski, M., 2007. Quantifying and Mapping the Human A ­ ppropriation of Net Primary Production in Earth’s Terrestrial Ecosystems. PNAS 104, 12942–12947. Haberl, H., Erb, K.-H., Krausmann, F., 2013. Global Human Appropriation of Net ­Primary Production (HANPP). The Encyclopedia of Earth http://editors.eol.org/eoearth/ wiki/Global_human_appropriation_of_net_primary_production_(HANPP) Haff, P.K., 2014. Technology as a Geological Phenomenon: Implications for Human Well-Being, In:A Stratigraphical Basis for the Anthropocene, edited by Waters, C. N., ­Zalasiewicz, J. A.,Williams, M., Ellis, M.,Snelling, A., vol. 395, pp. 301–309, Geological Society, London, Special Publications. Harmand, S.et al. 2015. 3.3-Million-Year-Old Stone Tools from Lomekwi 3, West ­Turkana, Kenya. Nature 521, 310–315. Hight, D.W., McMillan, F., Powell, J.J.M., Jardine, R.J., Allenou, C.P. 2003. Some C ­ haracteristics of London Clay. In: Characterisation and Engineering Properties of Natural Soils, edited by Tan, T.S., Phoon, K.K., Hight, D.W., Leroueil, S. Rotterdam: Balkema, pp. 851–907. History Database of the Global Environment (2013) Netherlands Environmental Assessment Agency. HYDE. Available at: http://themasites.pbl.nl/tridion/en/themasites/ hyde/basicdrivingfactors/population/index-2.html (accessed 15 February 2013). Hobbes, T., 1651. Leviathan, Available at Project Gutenberg, http://www.gutenberg.org/ files/3207/3207-h/3207-h.htm#link2H_4_0001. Hugo, V., 1862. Les Miserables, Part V, Book II. Translated by Isabel F. Hapgood 1887. Available at Project Gutenberg, https://www.gutenberg.org/files/135/135-h/135-h.htm. Jensen, S., 2003. The Proterozoic and Earliest Cambrian Trace Fossil Record; Patterns, Problems and Perspectives. Integrative and Comparative Biology 43, 219–228. Jones, C.G., Lawton, J.H., Shachak, M., 1994. Organisms as Ecosystem Engineers. Oikos 69, 373–386. Jung, H.J., Kim, B., Ryu, J., Maskey, S., Kim, J.C., Sohn, J., Ro, C.U., 2010. Source Identification of Particulate Matter Collected at Underground Subway Stations in Seoul, K ­ orea Using Quantitative Single-Particle Analysis. Atmospheric Environment 44, 2287–2293. Kam, W., Ning, Z., Shafer, M.M., Schauer, J.J., Sioutas, C., 2011. Chemical Characterization and Redox Potential of Coarse and Fine Particulate Matter (PM) in Underground and Ground-Level Rail Systems of the Los Angeles Metro. Environmental Science and Technology 45, 6769–6776. Key World Energy Statistics, 2015. International Energy Agency. https://www.oecd-ilibrary. org/energy/key-world-energy-statistics-2015_key_energ_stat-2015-en [online resource retrieved in 2016]. Koch, P.L., Barnosky, A.D., 2006. Late Quaternary Extinctions: State of the Debate. ­Annual Review of Ecology Evolution and Systematics 37, 215–250. Krausmann, F., Erb, K.-H., Gingrich, S., Haberl, H., Bondeau, A., Gaube, V., Lauk, C., Plutzar, C., Searchinger, T.D., 2013. Global Human Appropriation of Net Primary Production Doubled in the 20th Century. PNAS 110, 10324–10329. Landing, E., Geyer, G., Brasier, M.D., Bowring, S.A., 2013. Cambrian Evolutionary ­Radiation: Context, Correlations, and Chronostratigraphy—Overcoming Deficiencies of the First Appearance Datum (FAD) Concept. Earth Science Reviews 123, 133–177. 453

Mark Williams et al.

Levinson, D., 2007. Density and Dispersion: The Co-development of Land Use and Rail in London. Journal of Economic Geography 8, 55–57. McDougall, I., Brown, F.H., Fleagle, J.G., 2005. Stratigraphic Placement and Age of ­Modern Humans from Kibish, Ethiopia. Nature 433, 733–736. Mángano, M.G., Buatois, L.A. (eds)., 2016a. The Trace-Fossil Record of Major Evolutionary Events. Volume 1: Precambrian and Paleozoic. Topics in Geobiology 39. Springer. Mángano, M.G., Buatois, L.A. (eds)., 2016b. The Trace-Fossil Record of Major Evolutionary Events. Volume 2: Mesozoic and Cenozoic. Topics in Geobiology, 40. Springer. Martin, K., Sauerborn, J., 2013. Chapter 2, Origin and Development of Agriculture. 9–48 In: Agroecology, edited by Martin, K., Sauerborn, J. Springer Verlag, Netherlands. Matz, M.V., Frank, T.M., Marshall, J.N., Widder, E.A., Johnsen, S., 2008. Giant Deep-Sea Protist Produces Bilaterian-Like Traces. Current Biology 18, 1849–1854. Metrobits.org at: http://Metrobits.org Minter, N.J., Braddy, S.J., Davis, R.B., 2007. Between a Rock and a Hard Place: ­A rthropod Trackways and Ichnotaxonomy. Lethaia 40, 365–375. Minter, N.J., Franks, N.R., Robson-Brown, K.A., 2012. Morphogenesis of an Extended Phenotype: Four-Dimensional and Nest Architecture. Journal of the Royal Society Interface 9, 586–595. Nieuwenhuijsen, M.J., Gomez-Perales, J.E., Colvile, R.N., 2007. Levels of Particulate Air Pollution, Its Elemental Composition, Determinants and Health Effects in Metro Systems. Atmospheric Environment 41, 7995–8006. Ortiz-Ospina, E., Roser, M., 2016. World Population Growth. Published online at OurWorldInData.org. Available at: http://ourworldindata.org/data/population-growth-­ vital-statistics/world-population-growth/ [Online resource accessed 29th July, 2016]. Paul, J.D. 2009. Geology and the London Underground. Geology Today 25, 12–17. Plotnick, R.E., 2012. Behavioral Biology of Trace Fossils. Palaeobiology 38, 459–473. Qiao, T., Xiu, G., Zheng, Y., Yang, J., Wang, L., Yang, J., Huang, Z., 2015. Preliminary Investigation of PM 1, PM 2.5, PM 10 and Its Metal Elemental Composition in Tunnels at a Subway Station in Shanghai, China. Transportation Research Part D: Transport and Environment 41, 136–146. Richter, D., Grün, R., Joannes-Boyau, R., Steele, T.E., Amani, F., Rué, M., Fernandes, P., Raynal, J.-P., Geraads, D., Ben-Ncer, A., Hublin, J.-J., McPherron, S.P. 2017. The Age of the Hominin Fossils from Jbel Irhoud, Morocco, and the Origins of the Middle Stone Age. Nature 546, 293–296. Roberts, R.G., Flannery, T.F., Ayliffe, L.K., Yoshida, H., Olley, J.M., Prideaux, G.J., Laslett, G.M., Baynes, A., Smith, M.A., Jones, R., Smith, B.L., 2001. New Ages for the Last Australian Megafauna: Continent-Wide Extinction about 46,000 Years Ago. Science 292, 1888–1892. Russell, M., 2000. Neolithic Flint Mines in Britain. Tempus: London. Satterthwaite, D., 2007. The Transition to a Predominantly Urban World and Its Underpinnings. Human Settlements Discussion Paper Series. Theme Urban Change 4. Accessed at: https://pubs.iied.org/pdfs/10550IIED.pdf Seilacher, A., 1977. Evolution of Trace Fossil Communities. In: Patterns of Evolution as Illustrated by the Fossil Record (Hallam, A. ed). Elsevier. 359–376. Steffen, W., Broadgate, W., Deutsch, L., Gaffney, O., Ludwig, L., 2015. The Trajectory of the Anthropocene: The Great Acceleration. The Anthropocene Review 2, 81–98. Steffen, W., et al., 2016. Stratigraphic and Earth System Approaches to Defining the ­A nthropocene. Earth’s Future 4, 324–345. Taylor, A., Goldring, R., Gowland, S., 2003. Analysis and Application of Ichnofabrics. Earth-Science Reviews 60, 227–259. 454

Underground metro systems

Tero, A., Takagi, S., Saigusa, T., Ito, K., Bebber, D.P., Fricker, M.D., Yumiki, K., ­Kobayashi, R., Nakagaki, T., 2010. Rules for Biologically Inspired Adaptive Network Design. Science 327, 439–442. Tschinkel, W.R., 2003. Subterranean Ant Nests: Trace Fossils Past and Future? Palaeogeography, Palaeoclimatology, Palaeoecology 192, 321–333. UNDESA, 2012. World Urbanisation Prospects, 2011. Revisions. United Nations Department of Economic and Social Affairs, New York. http://hdrstats.undp.org/en/tables/display ByRow.cfm. United Nations, Department of Economic and Social Affairs, Population Division, 2014. World Urbanization Prospects: The 2014 Revision, Highlights (ST/ESA/SER.A/352). https://esa.un.org/unpd/wup/publications/files/wup2014-highlights.pdf. Vannier, J., Calandra, I., Gaillard, C., Żylińska, A., 2010. Priapulid Worms: Pioneer ­Horizontal Burrowers at the Precambrian-Cambrian Boundary. Geology 38, 711–714. Villmoare, B., Kimbel, W.H., Seyoum, C., Campisano, C.J., DiMaggio, E.N., Rowan, J., Braun, D.R., Arrowsmith, J.R., Reed, K.E., 2015. Early Homo at 2.8 Ma from ­Ledi-Geraru, Afar, Ethiopia. Science 347, 1352–1355. Waters, C.N., Syvitski, J.P.M., Gałuszka, A. Hancock, G.J., Zalasiewicz, J., Cearreta,A., Grinevald, J., Jeandel, C., McNeill, J.R., Summerhayes, C., Barnosky, A., 2015. Can Nuclear Weapons Fallout Mark the Beginning of the Anthropocene Epoch? Bulletin of the Atomic Scientists 71, 46–57. Waters, C.N., et al., 2016. The Anthropocene is Functionally and Stratigraphically Distinct from the Holocene. Science 351, 137. West, G., 2017. Scale, The Universal Laws of Growth, Innovation, Sustainability, and the Pace of Life in Organisms, Cities, Economies, and Companies. Penguin Press, New York. Williams, M., Zalasiewicz, J.A., Haff, P.K., Schwägerl, C., Barnosky, A.D., Ellis, E.C., 2015a. The Anthropocene biosphere, The Anthropocene Review 2, 196–219. Williams, M., Zalasiewicz, J., Davies, N., Mazzini, I., Goiran, J.-P., Kane, S., 2015b. ­Humans as the Third Evolutionary Stage of Biosphere Engineering of Rivers. Anthropocene 7, 57–63. Williams, M., et al., 2016. The Anthropocene: A Conspicuous Stratigraphical Signal of Anthropogenic Changes in Production and Consumption Across the Biosphere. Earth’s Future 4, 34–53. Wolmar, C., 2004. The Subterranean Railway: How the London Underground Was Built and How It Changed the City Forever. Atlantic Books, London. Zalasiewicz, J., et al., 2008. Are We Now Living in the Anthropocene? GSA Today 18, 4–8. Zalasiewicz, J., Williams, M., Waters, C.N., Haff, P.K., Barnosky, A.D., 2013. The Technofossil Record of Humans. Anthropocene Review 1, 34–43. Zalasiewicz, J, Williams, M., Waters, C N., 2014a. Can an Anthropocene Series Be D ­ efined and Recognized? In: A Stratigraphical Basis for the Anthropocene, edited by Waters, C.N., Zalasiewicz, J.A., Williams, M, Ellis, M.A.,Snelling, A.M. Geological Society, London, Special Publication 395, 39–54. Zalasiewicz, J., Waters, C.N., Williams, M., 2014b. Human Bioturbation, and the ­Subterranean Landscape of the Anthropocene. Anthropocene 6, 3–9. Zalasiewicz, J., et al., 2015. When Did the Anthropocene Begin? A Mid-twentieth Century Boundary Level is Stratigraphically Optimal. Quaternary International 383, 196–203. Zalasiewicz, J., et al., 2016a. Scale and Diversity of the Physical Technosphere: A Geological Perspective. The Anthropocene Review 4, 9–22. Zalasiewicz, J., et al., 2016b. The Geological Cycle of Plastics and Their Use as a S­ tratigraphic Indicator of the Anthropocene. Anthropocene 13, 4–17.

455

21 THE COMING ENERGY TRANSITION What comes after fossil-fueled civilization?1 Joseph Voros Introduction The modern scientifically based understanding of how humankind came to be here—called, among other things, Cosmic Evolution, The Epic of Evolution, Universal History, as well as, more recently, big history—is an intellectually exciting and very powerful conceptual model for making sense of the entire past, leading from the Big Bang nearly 14 billion years ago to our present planet-wide i­nformation-based technological civilisation (e.g., Brown 2008a, 2017; Chaisson 2001, 2007, 2008; Christian 2004, 2008, 2018; Christian, Brown & Benjamin 2013; Delsemme 1998; Jantsch 1980; Spier 1996, 2010, 2015). It represents a remarkable synthesis of diverse knowledge domains and scholarly disciplines brought together into a unified account of many different dynamical processes arising since the beginning of the universe. It also allows us to identify some of the major forces and drivers of change in human history operating over a number of different spatial and temporal scales, providing insights into how the globalised world we know today has come to be the way it is. In short, in the words of the International Big History Association (2016), big history “seeks to understand the integrated history of the Cosmos, Earth, Life, and Humanity, using the best available empirical evidence and scholarly methods”. While there have been many examples of earlier attempts to synthesise the sum total of human knowledge in this way (see, e.g., Spier 2010, ch. 1, 2015, ch. 1), Erich Jantsch (1980) wrote perhaps the first account of cosmic evolution/big history based on the modern understanding of non-equilibrium thermodynamics, drawing strongly upon the work of, amongst many others, Nobel Laureate Ilya Prigogine, to whom he dedicated his book. It remains a stunning work of synthetic scholarship and remarkably prescient insight even after nearly four decades. Two of the key concepts (among several Jantsch considered) that are found in the big history approaches of Eric Chaisson (e.g., 2001, 2004), Fred Spier (e.g., 1996, 2005, 2011, 2015), David Christian (e.g., Christian 2004, 2008, 2018), and Frank Niele (2005), are energy (or more precisely, energy flows), and complexity.2 Accordingly, our attention here will be primarily focussed upon the issue of the energy available for use by human society, 456

The coming energy transition

the degree of social complexity that can be supported by it, and how this relationship may be viewed within the context of the ‘eight thresholds’ approach to big history developed by David Christian as applied to thinking about the future of human civilisation on a big history timescale. In other words, we shall use the ‘eight thresholds’ view of big history to explore what the ‘next’ such Threshold might look like—what we might as a result call ‘Threshold 9’.To aid us in this exploration, we will also make use of James Dator’s (e.g., 1998) four generic ‘archetypal’ ‘images of the future’ to ‘contour’ the overall ‘shape’ of the coming global future at this macro scale. As a physicist-turned-futurist, I am interested in using scale-appropriate frameworks of understanding to generate ideas for further exploration into the dynamics that are shaping our present world, and which are likely to be involved in shaping the future (Voros 2003, 2005). By choosing frameworks of appropriate scope, we may look for insights about potential futures at a ‘deeper’ level than merely extrapolating ‘surface’ trends, and thereby undertake profoundly ‘deeper’ futures thinking than that engendered by merely ‘reading’ these trends (Voros 2006a, 2017a). The grandest model currently available for use in this way would seem to be the all-­encompassing scenario of Cosmic Evolution itself, which can be viewed as a broad nomothetic process that includes the specific idiographic case of how that evolution has played out in this corner of the universe here on Planet Earth, namely, (our) big history. One can readily imagine that there could also be other civilizations, or at least inhabited worlds, which may also have their own unique analogous versions of big history. Because big history takes such a vast ‘big picture’ view of human history, it is very well suited and ideally conceived as a framing perspective for looking at global-scale changes and very long-term processes. Consequently, this chapter will use big ­history as a framework for foresight at the global-civilisational scale. In essence, we will here be using big history as a ‘scaffold’ or framework for undertaking long-term broadly outlined global-scale foresight.3 The discussion here should be regarded as merely an initial and very preliminary exploratory sketch of a few of many possible ideas, being done as much to show the process of undertaking such foresight-focussed exploratory work based on the big history framework as much as for any insights that might be generated by it. I hope that it can contribute in some way to a wide-ranging continuing ­conversation—among big historians, sociologists, futurists, and any other similarly interested ­scholars—around the issue of the energy basis supporting our common global future. I also fondly hope that the activity of generating such a sketch—as well as demonstrating the thinking process that underpins it—will also in some way help contribute to the successful navigation by humanity of some of the major issues we need to confront at the civilizational, planetary, and even species level as we steer our way into the rapidly emerging and increasingly dangerous and uncertain future which lies ahead…

Contemplating ‘Epoch 8’, profiling ‘Threshold 9’ In Eric Chaisson’s approach to Cosmic Evolution, he considers seven major ‘epochs’ of increasing material-energetic complexity in the unfolding of the evolution of the cosmos: particulate; galactic; stellar; planetary; chemical; biological; and cultural (Chaisson 2007). An interactive website based on his body of work over decades also 457

Joseph Voros

adds an 8th Epoch, ‘future evolution’ where “the cosmic-evolutionary scenario is extended in time” (Chaisson 2008). It forms a broad palette for anticipations of how aspects of the process of cosmic evolution may further unfold in our instance of that process. In David Christian’s approach to big history, he considers eight major ‘thresholds’ of this increasing material-energetic complexity: the origin of the universe; the first stars and galaxies; the formation of chemical elements; the formation of the Earth and solar system; the arising of life; the arising of humanity; the transition to a­ griculture; and the “modern revolution” (Christian 2004, 2008, 2018; Christian, Brown & ­Benjamin 2013). Clearly, these eight thresholds can be seen to fit within the seven epochs described above, albeit with an obvious emphasis on the cultural epoch, as his telling of the account is framed from the perspective of an historian not an astronomer. The question of the future is also considered in Christian’s work (2004, ch. 15, 2018, ch. 12), and, indeed, in the work of other Big Historians as well (e.g., Brown 2008a, ch. 13, 2017, ch. 11; Niele 2005, ch. 7; Spier 2010, ch. 8, 2015, ch. 8). If our interest in the coming future is focussed on humanity and human civilisation at the global scale, then our contemplations of  ‘Epoch 8’ naturally find expression in the activity of attempting to characterise the unfolding of global dynamics over the next few decades and centuries.4 Given that our present civilisation has arisen over the last few centuries following the emergence of ‘Threshold 8’—based upon the ever-increasing use of non-renewable fossil fuel energy—as a futurist I naturally find myself thinking about what the next threshold of big history might be—when these fuels are either much less, or perhaps even no longer, easily available for our use, or if they are in the process of being replaced by other primary sources of energy. This will no doubt be a major energy transition, one of only a few in world history (Niele 2005; Smil 1994, 2017). Vaclav Smil (2010b, p. viii) has observed that such energy transitions are “inherently protracted”, and “usually … take decades to accomplish”. This will surely have profound implications for our present civilisation, so it would seem wise to undertake some serious foresight-based anticipatory thinking in order to begin to prepare for the consequences arising from such a change in the global energy system. This, then, is the key logical starting point of our current futures exploration based on big history: to recognise that there will be a time in the future when a new big history threshold has been crossed—one where fossil fuels are no longer the primary source of energy powering human societies; what we might therefore call ‘­Threshold 9’. One wonders then what forms human society might take, and what the effect will be on social complexity that is based on different sources of energy than these. Will social complexity continue to increase in new and emergent ways owing to the discovery of newer more energy-dense sources of energy? Or, will it perhaps be reduced to relatively simpler lifeways due to the availability of only lessdense energy sources? It is prudent to consider deeply what may happen to human civilisation when easy access to these finite sources of energy inevitably begins to tighten in the not-too-distant future.Thus, in our present contemplations of Epoch 8, we find ourselves focussing upon and ‘profiling’ Threshold 9. We will find that the most likely projected future global-civilisational trajectory currently unfolding—in the absence of a major catastrophic shock, technological 458

The coming energy transition

energy breakthrough, or similar low-probability ‘wildcard’ event—is a slowly unfolding ‘descent’ over a timescale of decades to centuries towards a global society constrained by ever-declining access to easily available sources of high-density fossil fuel-based energy. Such an energy-constrained future clearly has major implications for the level of complexity possible for human civilisation. This observation in turn suggests undertaking an increasingly urgent programme of continuing anticipatory research and exploration into both the underlying nature, as well as the potential emergent characteristics of, the coming transition to ‘Threshold 9’, in order to prepare for, and perhaps to mitigate, to the degree possible, its more unwelcome aspects. In what follows, in order to set the context, the main ‘human’ thresholds—6, 7, and 8—are very briefly considered from the perspective of energy use. Since post-­ Threshold 8 modern industrial civilisation is so overwhelmingly based on easy access to cheap abundant fossil fuel-based energy, confronting the uncomfortable question of what comes after such easy access would be aided considerably by some sort of organising framework to guide our thinking. To this end, Dator’s four ‘generic’ archetypal futures are examined to see how they can be used to ‘contour’ our thinking about the implications for the longer-term human future that arise from considering this far-reaching question. We then briefly examine past dynamics in an earlier Threshold to look, by analogy, for potential insights into the upcoming transition period, and end by reflecting on how we may need to prepare ourselves for the coming transition to a post-fossil fuel-based civilisation.

Reviewing the ‘human’ thresholds In David Christian’s ‘thresholds’ conception of  big history, the three thresholds which pertain to humanity and so are of most direct interest to us here are: Threshold 6, some 200–300 millennia ago (in the Palaeolithic Era) when our species Homo sapiens emerged as distinct from other closely related hominines; Threshold 7, some 10–11 millennia ago (the transition into the Agrarian Era) when humans began changing their main approach to making a living from foraging to farming; and Threshold 8, some two to three centuries ago (the transition into the Modern Era) when humans began to utilise more extensively the energy stored in highly energy-dense fossil fuels, in our many social, economic and other activities. At Threshold 6, it appears that gaining access to the (chemical) energy stored in foodstuffs is simply a ‘given’ as the primary goal of biological survival. In this sense, from the point of view of energy use, it would seem to differ little from Threshold  5, the emergence of life on Earth.5 Following the human migrations that ultimately extended to all major land masses except Antarctica, human freedom to range widely into hitherto unutilised territory eventually began to become more constrained, leading to an increasing ‘intensification’ of the use of existing lands, rather than simply extending our presence into newer lands (‘extensification’). This transition of ­techno-economic base from foraging to agriculture—Threshold 7—seems to have been a fairly gradual—and possibly initially unwelcome—process (Brown 2008a, ch.  5; Christian 2004, ch. 8). Here the utilisation of environmental resources and energy also intensified, as humans domesticated plants and animals, and subsequently began to deliberately harness wind and water energy. 459

Joseph Voros

By Threshold 8, human energy usage had begun to unlock the stored solar energy encapsulated in long-dead organisms—the ‘fossil fuels’ made up of peats, coals, crude oils, and some natural gases.The much higher energy densities (energy per unit mass) of these fuels compared to previously utilised biomass fuels and dispersed forms of renewable energy made them very attractive, as did their relative abundance (Smil 1994, pp. 153 & 219), as well as their relative ease of accumulation or extraction. ­Efficiencies of energy use have also improved markedly over the last several centuries, from less than 5% in open wood fires to 94–97% for modern gas-fired space heaters (Smil 2010b, pp. 7–8). Today almost every aspect of the modern industrialised world has become utterly dependent upon fossil fuels. Smil (1994, 2017) has called this dependence “fossil-fuelled civilization”, while Niele (2005) has even characterised modern fossil fuel-using and fuel-dependent human beings as a distinct sub-species: Homo sapiens carbonius.

What comes after ‘fossil-fuelled civilisation’? However, these luxuriantly energy-dense non-renewable fossil fuels obviously cannot last forever.6 At some stage, access to fossil fuel energy will inevitably face a severe bottleneck of availability and then an overall decline. The energy infrastructure that powers contemporary industrial civilisation, and which is hugely interdependent with and upon the social systems and institutions that are themselves powered by it, will then undergo a crisis of stability which will have a flow-on effect to industrial civilisation itself. A large number of contemporary writers have examined this emerging civilisational crisis, generally framed around climate change, energy decline, or economic instability (e.g., Ahmed 2010; Brown 2008b, 2011; Greer 2008; ­Heinberg 2010; Heinberg & Lerch 2010; Holmgren 2009; Kunstler 2005; Lynas 2008; Roberts 2005; Slaughter 2010). The investigations undertaken by Nafeez Ahmed (2010, 2017) are particularly notable for their attempt to look beyond disciplinary boundaries and specialisations to examine the many mutually reinforcing interactions between the various crises that different experts focus upon, as well as the identification of 11 systemic-level ‘structural’ issues that will need to be urgently addressed if we are to transition smoothly to what has been called a “post-­ carbon civilisation” (e.g., Ahmed 2010; Heinberg & Lerch 2010). There are many, many recent and contemporary commentators around the ‘peak oil’ (e.g., Hall & ­Klitgaard 2012, ch. 15), ‘peak energy’, or even ‘peak everything’ (Heinberg 2010) debate, with many arguing positions both pro and con; far too many to list here. In sketch, though, the ‘con’ position usually tends to counter to the ‘pro’ position (which argues that easy access to energy is rapidly running out), by claiming that there are vast reserves still left in the ground which will last many decades or centuries yet, or that some new technological innovation in the future will surely occur to mitigate the problem. However, on a big history timescale, the next few decades or centuries are only a momentary ‘blip’ in the overall trajectory of the human species and of planet Earth. On this scale, the availability of highly concentrated energy-dense fossil fuels is but a “brief anomaly” (Floyd 2012) in the long history of the Earth, so a properly 460

The coming energy transition

diachronic view of human and Earth history needs to look well beyond the present ‘peak’ debates to the longer term when fossil fuels are no (inevitably) longer so readily available. Thus the question of what effect the running-down of fossil-fuelled energy systems will have on the structure and complexity of human civilisation is a genuinely serious one which deserves somewhat better than to be willfully ignored, deferred to future generations, or wished-away through a fairy-tale-like reliance upon a hoped for miraculous technological salvation. That is why the futures-thinking approach based on big history taken here is to imagine a time in the future when fossil fuels have either effectively run out or, at the very least, that our fossil-fuelled civilisation has run out of easy access to these fuels, and is increasingly based on other primary sources of energy. It is this that I am calling ‘Threshold 9’—a time in the future, when the predominant sources of energy powering human societies are no longer fossil fuels (note that this view is agnostic about what those sources may be; it merely notes that fossil fuels will not be those sources). This is a way to bypass the sometimes rather heated and often unproductive current argumentation about energy scarcity, locked as it is in a present-moment ­perspective, to simply acknowledge that fossil fuels are indeed finite—something which no one could seriously argue against—and use that incontestable fact as our foundational starting point to seek to generate insights about the coming global longer-term future. This has the important effect of enabling us to avoid getting ‘stuck’ in the present debates about the imminent coming, or not, of  ‘peak oil’ or ‘peak energy’ and to instead simply take up a stance in the farther future when these debates will simply be over because the facts are beyond dispute. This is an example of a ‘discontinuous’ method of thinking about the future: we ‘escape’ from the tyranny of the present, and of limited imagination extrapolations based on our constrained synchronic view of the present, by purposefully ‘jumping’ to a point in the future (Voros 2006a). From such a future perspective, we are then able to ‘look back with different eyes’ (as it were) to see, in a different way, what sort of future-history trajectories might emerge from our current situation.

Scoping future dynamics Four generic ‘images’ of the future So, where to from here? How can we begin to seriously examine alternative futures for our present global civilisation on a big history timescale? James Dator has studied the ways that different cultural groups and societies think about the future. According to him, all ideas about—or ‘images’ of—the future, can be grouped into four broad generic archetypal classes, as follows (e.g., Dator 1998, 2002):7 • •

Continuation—the current historical trajectory continues, most usually conceived of as continued economic growth; Collapse—a breakdown of the social order due to one or more of a number of possible causes, such as economic instability, environmental overload, resource depletion, moral degeneration, military conflict such as an external attack or internal civil war, a meteor/comet impact, etc; 461

Joseph Voros

• •

Disciplined or Constrained Society—a society organised around some set of over-arching constraining values, whether ancient, traditional, ideological, natural, environmental, God-given, etc; and Transformational Society—which sees the end of current forms of behaviour, beliefs, norms, or organisation, and the emergence of new forms (rather than a return to older or traditional ones, as above), possibly even including intelligent life-forms. The two main sub-variants are ‘high-tech’ (technological) and ‘highspirit’ (spiritual/consciousness) transformation.

These four archetypal futures provide a useful structure for thinking about the possible futures of societies in general. But let us now choose ‘energy availability’ as the central organising principle or focus of examination through this framework to see what potential insights we might be able to extract from this particular analysis about the possible future dynamics of the world system.

Continuation Much of what is known as the ‘technoliberal optimist’ literature (Wagar 1991) assumes that things in general will simply continue improving as they have been for the last two or three centuries or so, guided by the further spread of democratic ideals, unfettered free enterprise, and unbounded technological progress (e.g., Diamandis & Kotler 2012; Schwartz, Leyden & Hyatt 2000). And it makes perfect sense from a straightforward (if somewhat naïvely unsophisticated) application of ‘extrapolative evolution’ (Voros 2006a) of the world system’s dynamics and technological changes over the past few centuries. The Continuation archetype characterises precisely this viewpoint. It takes as its baseline the historical trajectory of industrial civilisation since it emerged, and simply extends this trend line into the future as the single uncritically projected future. Therefore, in this view, we can expect ever more economic growth, and ever-more numbers of humanity to be lifted out of poverty as ‘progress’ continues to improve the lot of humankind. This admittedly quite attractive view of the future of human history is fairly ­prevalent—and not without some basis (e.g., Millennium Project 2012)—not only in the OECD countries, who stand to remain in their present comfortable lifestyles, but also in many industrialising nations, who sense that there is much to be gained as they seek to attain the living conditions and lifestyles of the richer industrialised countries. Unfortunately, such an uncritical extrapolation of past trends into the future is based upon the (usually unchallenged) assumption that the deeper underlying system dynamics that have made this possible will also continue into the future.These system dynamics have, to a very large extent, been based upon the energy sources fuelling industrial civilisation—dynamics which, as we have noted above, are quite literally running out of fuel. Thus, regrettably, the Continuation scenario for ever-more economic growth and well-being is almost certainly a phantom based on the (most likely) delusional assumption that easy access to cheap, abundant energy will certainly continue without abatement. Barring a technological breakthrough—which puts us into the hightech transformation sub-class of the Transformational Society archetype, to be described 462

The coming energy transition

below—such an anodyne view of the future is not sustained by the currently available evidence. Indeed, as mentioned above already, the preponderance of emerging evidence seems to point to the inevitable decline of easy access to abundant, concentrated sources of energy (e.g., Niele 2005; Smil 1994, 2010a, 2010b, 2017; Turner 2008, 2014, and many other contemporary works, including those cited earlier), with the result that Continuation cannot be sustained beyond a fairly brief time into the future—if at all—and certainly not on the big history timescale we are utilising as our basis for anticipatory prospection. My point here is not to argue nor necessarily to disagree with the optimists who assume that we will find a positive way forward, but simply to note that this cannot be assumed as a natural or inevitable continuation of the past. Rather, it will require some kind of innovation or breakthrough for this to occur—as Richard Heinberg (2009) puts it, we are “searching for a miracle”—and that is the key point that needs to be borne in mind. It is akin to basing a retirement plan on the assumption that one will simply win the lottery (again, in this case, since the discovery of fossil fuels already constitutes such a lottery win for our species!) when it becomes necessary to do so in order to be able to keep on making a living. It would of course be quite nice and very welcome! However, it is by no means certain, and is probably not the wisest strategy to pursue for one’s long-term future. It is perhaps for the best to at least have a credible backup ‘Plan B’, ‘just in case’.

Collapse In contrast to the ‘business as usual’ projected endless-growth mindset, there exists a considerable literature dealing with what may happen when energy sources, usually abbreviated to just ‘oil’, begin to run dry (e.g., Brown 2008b; Kunstler 2005; R ­ oberts 2005) or when the biosphere can no longer tolerate and absorb the stresses that human civilisation is placing upon it by burning them (e.g., Brown 2011; D ­ iamond 2005; Farnish 2009; Lynas 2008; Meadows, Randers & Meadows 2004).This is precisely the Collapse archetype, although variants of the other sub-classes of this class also exist, such as asteroid or cometary impact (e.g., Chapman 2004; Collins, ­Melosh & Marcus 2005; Schweickart et al. 2008). Of course, from the perspective of  big ­history, this latter possibility is not as farfetched and improbable a ‘wildcard’ (Petersen 1997, 1999) as some people may think, not only since we owe the ascent of mammals over large reptiles in significant part to just such an event 65 million years ago (e.g., Alvarez 1997), but also because there has actually been an impact in the very recent past that could have had devastating effects had it hit densely populated areas—the Tunguska Impact of 1908 (Di Martino, Farinella & Longo 1998; Gasperini et al. 2007). Had this occurred during the Cold War, one wonders what the consequences might have been, or indeed might still be if any similar future impact blast during a time of high international tension is mistaken for a nuclear detonation (cf. Sagan 1980, p. 76). The Chelyabinsk bolide in early 2013 was a minor example of this; and a rather lucky escape, too. The Collapse archetype came to wide popular attention with the publication of Jared Diamond’s (2005) eponymous best-seller, which continues a tradition of scholarly study of the collapse of complex societies in history that includes the important 463

Joseph Voros

earlier work of Joseph Tainter (1990). However, the concept of ‘collapse’ itself is a somewhat imprecise one, and carries a certain connotation of rapidity that may not be entirely useful for our current purposes. As some scholars have noted, from the present we tend to see the historical ‘collapses’ of the distant past through a greatly foreshortened perspective, which can make even very drawn-out processes appear to seem somewhat abrupt from this vantage point. For example, the ‘collapse’ of the Western Roman Empire is generally thought to have taken some three centuries or so to occur (e.g., Tainter 1990). To those living through it, however, it would hardly have been noticeable over a lifetime, let alone experienced as the kind of rapid decline which the term ‘collapse’ connotes. From the perspective of the longer-term human future on a big history timescale, the end of fossil-fuelled industrial civilisation may very well come to be viewed similarly as a ‘collapse’, due to the same foreshortening of timescales that we ourselves experience when looking to the past. But, on our own timeframe—the timeframe of our individual lives and those of our immediate descendants and subsequent ­generations—we will almost certainly not experience this as a rapid ‘collapse’ in the sense that the term is commonly used.8 Rather, there will almost certainly be a kind of  ‘envelope’ of declining fossil fuel energy availability, which will shape the contours of the degree of complexity that is possible for human civilisation. Thus while this might eventually be considered a ‘collapse’ by historians from the farther future, if any there be, it will almost certainly not take place on the same timeframe as other more rapid events which might more fittingly merit the term ‘collapse’, such as a nuclear war or an asteroid/cometary impact. For this reason, some commentators conceive of the end of fossil-fuelled industrial civilisation due to energy scarcity not in the ‘rapid’ terms of a ‘collapse’, but more along the lines of a drawn-out ‘decline’ or ‘descent’ to an eventual new form of societal organisation with a techno-economic base founded upon renewable forms of energy—notable examples being  John Michael Greer (2008, 2009), David Holmgren (2009) and Richard Heinberg (2010). Of course, other forms of rapid societal collapse may also occur—acute environmental disasters, sudden economic recessions or depressions, unexpected social upheavals and unrest—and they should of course also be borne in mind. But our present futures assessment is based around a focus on energy availability, and the implications for social complexity that such availability allows. This therefore suggests not a sudden rapid ‘apocalyptic’-type end to fossil fuel energy sources, but rather a more gradual decline contoured by a narrowing energy availability ‘envelope’ as we inevitably move down the descending side of the empirically derived bell-shaped Hubbert Curve which was initially developed to describe oil production (e.g., Hall & Klitgaard 2012).There are even some commentators who, perhaps wryly, look upon the coming decline as an opportunity (e.g., Homer-Dixon 2006; Orlov 2008).

Disciplined/constrained society The notion of constraints on what human society can do being forced upon it by relative energy scarcity can be considered an unusual twist on the Disciplined/ ­Constrained Society archetype. In essence, at the end of the energy descent process, 464

The coming energy transition

human civilisation is in this case constrained not by the social values held by the dominant elite of the populace (that is, by an ‘endogenous’ constraint), which is the more usual form of this archetype, but by the fact that cheap abundant energy is no longer easily available for use by the society—an exogenous constraint. The fact that most forms of this archetype have generally been of the ‘internal values’ kind may be indicative of how pervasive the underlying (and unexamined) assumptions of abundant energy, continued progress, and lack of constraints on recent human societies and ambition have been. The negative reactions to the observations and ­policy-response explorations made, for example, by the authors of The Limits to Growth (Meadows et al. 1972; Meadows, Randers & Meadows 2004) is a telling case in point, and the almost gleeful manner in which this work was systematically and deliberately misrepresented and widely ridiculed makes for blood-boiling and blood-chilling reading for anyone wanting to influence global public policy (Bardi 2008, 2011). This is all the more ironic since subsequent follow-up work tracking the actual trajectory of ­human civilisation—as it has compared to the ‘baseline’ ‘business-as-usual’ ‘do-­ nothing-differently’ ‘standard run’ of The Limits to Growth—shows that global civilisation is right on track for the resulting baseline ‘overshoot and collapse’ scenario to unfold, essentially right on schedule (Turner 2008, 2014). We seem to have blithely wasted the very narrow window of opportunity we would have had to avert this undesirable future which heeding and acting on the message of The Limits to Growth might have afforded us (Randers 2012). Perhaps the vindictively gleeful ridicule was somewhat premature, and it would now appear that the voices who so delightedly and loudly did so are rather less vocal these days. A related ‘constraint’-type work from around the same time which also garnered negative reactions was The Population Bomb (­Ehrlich 1971), and it too might also prove rather unwelcomely prescient.9 Greer (2009) outlines a series of stages through which he argues our civilisation’s “long descent” (2008) will likely pass—from our current ‘abundance economy’ based on still relatively freely available high-density fossil fuel energy; to a ‘scarcity industrialism’ wherein the constraints on society are becoming increasingly prevalent; to an age of ‘salvage’ where earlier infrastructure is dismantled and re-used due to an increasing difficulty of manufacturing new materials; and ultimately the move to a new “ecotechnic” age, some one to three centuries or so hence, based on sustainable forms of techno-economy and organic agriculture. In all these stages there is a high likelihood of considerable economic and human turmoil. Make no mistake, this is no ‘sweetness-and-light’ transition to a Utopian paradise of blissful co-existence with Nature (what is sometimes known in the literature as ‘ecotopia’), nor is it a return to Marshall Sahlins’ (1972) somewhat idealised “original affluent society” of the ­Palaeolithic. Rather, it is a process of de-industrialisation—with all of the problematic consequences that the winding-back of many of the accomplishments of the last few centuries implies. This trajectory of what Greer (2008, passim) calls “catabolic collapse” is compared with similar cases from history, most especially the Mayan collapse, and his analysis is very aware of a big history energy perspective, even if he does not explicitly name it as such. These types of energy-constrained societies are a very important class of futures to be aware of, as I suspect they will become increasingly vital in guiding our collective thinking about the decline of readily available dense-energy sources 465

Joseph Voros

over the next decades and centuries. Greer is one of very few authors I am aware of writing about the next stages in human history which are expected to emerge on a timeframe of generations—as opposed to the more common timeframe of a few decades—as well as dealing with a disciplined/constrained society that has not totally and utterly ‘collapsed’, which latter devastated post-Apocalyptic image is a common theme also. Another notable author is Warren Wagar (1991, 1999), whose future society is admittedly disciplined mainly by socialist values rather than by energy availability per se, as well as by the physical aftermath of a global nuclear war (i.e., a prior very abrupt ‘collapse’), and is itself merely a transitional stage on the way to an eventual ‘transformational’ world civilisation (see below). With respect to the coming ‘energy discipline’, Smil’s work provides an important ‘envelope function’ of realistic energy system possibilities over the longer term which could be used to help guide our collective thinking—and, hopefully, collective learning (Christian 2010)—about the next transition in the configuration of our global energy system (Smil 2010a, 2010b). And finally, drawing on big history itself, perhaps we can make use of our understanding of what we might call the ‘anabolic’ ascent or rise of ­modern fossil-fuelled industrial civilisation over the past few centuries as a framework to generate some ideas and potential insights into the prospective ‘unwinding’ processes of decline that may well lie ahead during our civilisation’s all-too-­plausible ‘catabolic descent’.

Transformational society Of course, there could always be some sort of breakthrough or transformation that radically changes the nature and form of human society. These, Dator suggests, are usually conceived of as being either technological or spiritual/consciousness-based in nature. The case of   The Singularity could be considered an intriguing ‘hybrid’ form of these idealised types wherein consciousness transfers itself onto a technological substrate (e.g., Broderick 1997; Eden et al. 2012; Kurzweil 1999, 2006; Smart 2003).10 Let us consider the technology sub-type first. It is certainly true that technology has in the past radically altered humans’ relationship with nature, and especially our ability to utilise environmental energy. Thus, in this view, therefore, it will simply be just another such technological advance that will mitigate the impending energy problems we currently face in prospect.There are several very well-known possibilities frequently mentioned in contemporary commentaries, so it will suffice here to simply mention a few of them very briefly. Almost all renewable forms of energy—such as photovoltaic, wind, running water, and wind-generated wave energy—are ultimately based on the energy output of the Sun (Smil 2010a, 2010b), which we must therefore fervently hope will continue to have a nice stable lifetime on the main sequence for a good while yet! The other main forms of renewable energy not derived wholly from the Sun are geothermal, to be discussed later, and tidal. The trouble with these otherwise attractive clean sources is that they are nowhere near as energy-dense as fossil fuels. Electricity can be generated from a variety of sources, including renewable and, while therefore an attractive form of potentially cleaner energy, it is not necessarily suitable for all tasks—aircraft transport is a case in point. 466

The coming energy transition

Hydrogen, which is often mentioned in discussions of energy transitions, is only useful as a means for transporting energy, not as a source, although one can envisage, for example, a solar-powered conversion plant producing hydrogen gas from electrolysing seawater, with a view to its application as a potential transport fuel. It has the distinct advantage of producing only gaseous water (i.e., steam) as a by-product of combustion—a greenhouse gas, certainly, but somewhat less troublesome than carbon dioxide or methane, two of the main offenders. A transition to a ­hydrogen-based transport fuel system would require re-working of existing fuel infrastructures, and—were it to even be feasible—there does not as yet appear to be the political will or commercial appetite to undertake such an extensive programme of concerted action. Fusion energy is also often assumed to be a potential clean energy goldmine, but it has consistently not made the progress that was being hoped for it in the mid-latter part of the twentieth century CE.Yet, despite this, it persists in the public consciousness, and—if we are being fair-minded—it just might turn out to be more substantial on the longer big history timeframe we are considering. So if we are to remain alert to the existence of  ‘wildcard’ breakthroughs, we need to keep it somewhere in mind. There is also some hope held out for more attractive forms of nuclear fission energy, which, if the promise of, for example, thorium-based nuclear power is fulfilled—as opposed to the increasingly unpalatable uranium and its by-products—might be a more socially and politically acceptable way to act as a bridge to the fully renewable system we will almost certainly need to eventually create. The reader is referred to Smil (2010a, 2010b) for a comprehensive analysis of many of the conventional forms of energy noted here. Then there are the possibilities of some astonishing technological breakthroughs about which we are at present wholly ignorant and unable to even speculate. An ability to concentrate the more-diffuse renewable forms of energy into higher densities might be just such a development. Or perhaps some entirely new ‘miraculous’ source of energy is found, such as tapping the quantum vacuum energy of space-time, to give but one extreme example (cf. Clarke 1999). I have suggested elsewhere that there is merit in entertaining ‘preposterous’ futures ideas (Voros 2012, 2015), so perhaps ‘cold fusion’ can also be mentioned here, since Clarke dedicated his short story—albeit perhaps with tongue-in-cheek—to the two scientists who first announced it, but for which there has as yet not been any widely accepted experimental confirmation. The other major variant of the Transformational Society is one of consciousness or spiritual transformation—some new form or aspect of  human consciousness emerges and re-defines our value systems, such that we become focussed on ‘higher’ goals than the mostly materialistic ones we currently pursue. A number of futurists have considered this from the point of view of either a contemporary transition to a new ‘expanded’ worldview (e.g., Harman 1998) or from a sequence of changes over the next few centuries, which is precisely the time-frame we are using for considering Threshold 9. Thus, Wagar (1991, 1999), mentioned above, considers ‘three futures’, which pass through the four major classes we have been discussing here, culminating in a more spiritually informed consciousness-based planetary civilisation, while Duane Elgin also sees a transition to planetary civilisation over the ensuing centuries, based upon an expanded awareness of our place in the universe (2001, 2009).11 467

Joseph Voros

It may be that it is just such a new sense of ourselves and of our place in the Cosmos that gives rise to the mindset required to live within our energy means, and might indeed be one of the main prerequisites for us to successfully make the transition to what Elgin (1994) has called “initial maturity” as a “sustainable species-­ civilization”.  Thus, it might be argued that cosmic evolution, big history and other related cognitive frameworks may themselves provide the foundation for a new more integrated worldview, onto which an almost spiritual dimension could be read.There are several authors who are, to varying degrees, pursuing something like this line of thought (e.g., Abrams & Primack 2011; Christopher 2013; Genet et al. 2009; Primack & Abrams 2006; Swimme & Tucker 2011). However, Carl Sagan—who could rightly be considered one of the early pioneers of the modern scientifically based worldview we now know as cosmic evolution or big history (e.g., Sagan 1973, 1980)—­maintained that any meaning or significance to our existence was to be found not out there in the Cosmos, but rather here within us (Sagan 1995, p. 57): The significance of our lives and our fragile planet is then determined only by our wisdom and courage. We are the custodians of life’s meaning. … If we crave some cosmic purpose, then let us find ourselves a worthy goal. Perhaps one such worthy goal could be for humankind to dedicate itself wholeheartedly to building towards a fully sustainable and wholly equitable global civilisation… Finally, it is perhaps fitting to end this discussion of a more cosmically aware integrative worldview by returning full circle to the cosmic-evolutionary pioneer Erich Jantsch, whose decades-old work is still able to strengthen and deepen our understanding of the many material-energetic, informational and complexity-related processes occurring in cosmic evolution and big history ( Jantsch 1980). His thoughts on what he called ‘the evolutionary vision’—which can help us not only understand our past history and present place in the universe, but also to confront through our disciplined anticipations our coming future—were published posthumously after his untimely death, as literally among his last words in print ( Jantsch 1981, p. 213): The evolutionary vision is itself a manifestation of evolution. The reward for its elaboration will not only be a new (or partly revived) natural philosophy or an improved academic understanding of how we are interconnected with evolutionary dynamics at all levels, but also an immensely practical philosophy to guide us in a time of creative instability and major restructuration of the human world …. With such an orientation, science will also become more realistic and meaningful for the concerns of human life. It will be not merely an end product of human creativity, but a key to its further unfolding in all domains.

Insights from earlier thresholds Another approach to thinking about the future arises from using as an interpretive framework an adapted form of ‘macrohistory’, the study of how social systems change over time, in search of patterns or even ‘laws’ of social change (Galtung & Inayatullah 1997; Inayatullah 1998; Voros 2006b). In this approach, one looks for regularities in 468

The coming energy transition

the key dynamics of historical change and uses these as a way to seek insights into the situation being studied. In the context of our use here of the ‘thresholds’ view of big history, this would imply utilising earlier thresholds to try to generate insights about the next one. I call this use of earlier dynamics as a trigger for seeking insights ‘reiterative analogy’. There is naturally no assumption that the dynamics will repeat, so this approach should never be confused with a predictive stance; rather, we merely make use of some aspect or aspects of those dynamics as a cognitive trigger to try to generate new ideas and potential insights. Of course, the most recent threshold, ­T­hreshold 8—which was the transition to wide-scale use of fossil fuels—was clearly the initial stimulus for our thinking here about the next threshold,Threshold 9, which I have thereby defined as a corresponding transition away from fossil fuels. Let us now, by way of further exploration, go back one further threshold to Threshold 7, and seek to draw some insights from considering the characteristics of that one. Threshold 7 was characterised, in essence, by the transition from a techno-­ economic base of foraging to one of farming. This was a way to obtain food more reliably and predictably than relying upon simply finding it in the environment, although ­Palaeolithic food-gathering does seem likely to have been undertaken fairly methodically from existing knowledge of growth and seasonal cycles (Christian 2004), rather than being as haphazard as this brief characterisation might suggest. Clearly, there is an analogy being suggested here between ‘food’ and ‘energy’, which is, one could argue, not entirely unreasonable or unfounded. If we consider our approach to energy today, we can see that for several centuries we have effectively been ‘foraging’ for fossil fuel energy by searching the environment to see where it may be located ‘just lying around’, as it were. We have then ‘gathered’ it by mining or other forms of extraction and then moved on to look for other new deposits when the ready supply has become exhausted or no longer able to yield commercially useful quantities. In this way, one can see a clear resonant parallel between the extensification of human foraging range during the Palaeolithic Era and the increasingly extensive exploration of the Earth’s near-surface for energy reserves during the Modern Era. In the late Palaeolithic, we eventually ran out of new ranges to enter—extensification ‘ran down’, as it were—and we were forced to settle down and intensify the production and harvesting of in situ food-energy sources in order to continue to make a living, which was precisely the transition to agriculture and food farming.This was probably not, as noted earlier, an easy—or even welcome—process, and not necessarily everyone automatically took to it enthusiastically. Similarly, in the late Modern Era we again find ourselves running out of readily exploitable energy reserves—i.e., our energy-foraging extensification is starting to ‘run down’—and so we now once again find ourselves beginning to be forced to seek ways to intensify production and harvesting, this time of in situ ‘ambient’ environmental (as opposed to fossil fuel-based) energy. We are being forced to become, as it were, ‘energy farmers’; and, as before, this does not appear to be an easy, or indeed welcome, development! There are further ideas that can be drawn out from this analogy. One intriguing thought is that, as certain geographical areas were found to be conducive to agriculture and farming food with the result that human populations were increasingly drawn to those areas in the transition to Threshold 7, so might there perhaps be certain new and different areas which are found to be conducive to farming energy in 469

Joseph Voros

the transition to Threshold 9, so that human populations may increasingly be drawn to those regions. In effect, a demographic-geographic shift might take place to new centres of energy farming and the economic activities supported by it, analogously to the shift which took place in the early Agrarian Era for food farming. This will be an interesting potential dynamic to watch for. One final idea to briefly consider here is whether there might be any potential analogy between conventional fossil fuel energy—given how it revolutionised human civilisation and gave rise to Threshold 8—and any other form of energy that may have been analogously ‘fossilised’ in some way. What comes quickly to mind, of course, is geothermal energy, which has, in an admittedly fairly loose sense of the term, been ‘fossilised’ from the time of the formation of the Earth, and is due to a combination of the remnant heat of formation itself, as well as to the subsequent decay over time of radioactive elements present in the initial accretion disc that gave rise to the solar system. This ‘fossil heat’ is, in effect, driving the (so to speak) ‘internal convection engine’ which powers the tectonic movements that have had such an important role in Earth’s long-term geological and geographical history, not to mention in the distribution of valuable minerals in the Earth’s outer layers.This is an intriguing resonance with the role that the internal combustion engine has had in the Modern Era. In the context of this futures assessment, my interest is in whether this non-solar form of renewable energy could perhaps become an important source of power not only in the energetic sense, but also in the geopolitical sense, given that it is not subject to the problems of intermittency that other renewables have, like solar or wind.12 Thus, in the same way that some countries have had considerable advantages conferred upon them from the geographical distribution of fossil fuel reserves—one thinks of OPEC, for example, the Organisation of Petroleum Exporting Countries—it is interesting to wonder about whether, in an energy-constrained disciplined-society future, relatively easier access to the non-intermittent renewable energy from geothermal sources might also become a source of advantage for another group of countries. One might imagine a geothermal analogue to OPEC; perhaps an ‘OGAC’—an Organisation of Geothermal-Accessing Countries—who use this geographical good fortune to their techno-economic advantage. If this access to geothermal energy were used to generate hydrogen, for example, then there could emerge a system of fuel distribution based on hydrogen that is analogous to the present system of oil trading and distribution, with perhaps the attendant geostrategic implications that would flow from that emergence. This, too, will be an interesting potential dynamic to watch for.

Concluding remarks These have been some preliminary exploratory ideas based upon the awareness that, as the fossil fuel-based energy sources which have powered industrial civilisation— since what one well-known approach to big history calls ‘Threshold 8’—begin to become scarcer, there is an increasingly urgent need to confront and make sense of the wider implications this fact has for our present civilisation. This prospective new threshold in history—a time when fossil fuels are no longer the primary source of energy used to power human society—has been referred to here as ‘Threshold 9’. 470

The coming energy transition

Why is this approach useful? It is often difficult to clearly see the present l­onger-term dynamics of the world system while we are still so completely immersed within them. So, as a futures-thinking device to provide an entirely different perspective that is not, so to speak, ‘lost in the present’, we instead look to a future time when these confusingly complex dynamics may have largely played themselves out, and use that position as our vantage point from which to seek some clarity. In the current exploration, we were interested in the longer-term big history view of the global future engendered by considering what an imagined prospective ‘Threshold 9’ might look like. We then examined the shapes of some resulting conjectured potential futures to see what insights they might give us into what some of the important aspects of the present are that we will need to consider carefully as we move into the actual unfolding future. It is an oft-quoted aphorism that ‘hindsight is always “20-20”’. In the approach taken here, we are, in essence, seeking to generate deeper insight into our present situation by taking a long-term foresight view based on a vantage point in the more distant future, and then, so to speak, ‘looking back to the present’ from that future-based perspective in order to generate, as it were, ‘artificial hindsight’. This is obviously not a perfect process, but given the absence of practical time travel and the logical impossibility of future-revealing technology, it is currently among the most powerful approaches we have at our disposal. Conducting the same thought experiment using different analytical frameworks would of course generate different ideas, so that many different perspectives and approaches could thereby be tapped and integrated in this way. The most probable global trajectory emerging in prospect—barring a major rapid Collapse episode such as nuclear war, asteroid impact or similar event, or a miraculous Transformational Society brought about by an astonishing technological ­breakthrough— appears to be an energy system decline/descent over many human lifetimes to an eventual Disciplined/Constrained Society where the discipline is imposed upon us by much more limited access to energy than we have so greatly enjoyed for the past few centuries. The main point here is that our civilisation’s current e­ nergy-intensive lifeways can only continue, it seems, by way of some quasi-­miraculous technological breakthrough. Many people already know this, of course, and fully expect it to occur, given so many earlier technological innovations in recent history. However, the crucial difference here in our time is that, whereas the remarkable technological innovations of the past several centuries have essentially relied upon easy access to sources of readily available energy, this time it is access to energy itself which is the major bottleneck and which requires the breakthrough. And that cannot be regarded or treated in the same way as earlier technical innovations, nor can any technical circumvention of this issue be simply assumed or unquestioningly relied upon to occur.Therefore, it is as well for us to remember to be cautiously sceptical about present-day uncritical technological optimism. While other possibilities remain open, and should of course always remain in consideration, a wise course of action would appear to be conducting detailed multi-­ perspectival multi-disciplinary anticipatory research into both the underlying nature and emergent characteristics of a big history threshold of this post-fossil fuel form. Indeed, the very concept of ‘profiling Threshold 9’—to get a general sense of the broad longer-term future trajectory of our planetary civilisation—could be 471

Joseph Voros

considered a useful orienting direction for an entire joint research programme in both big history as well as futures studies scholarship, not only as a very interesting academic pursuit—which it most certainly would be— but also as an enormously prudent practical step towards preparing for the coming of Threshold 9. This programme would involve re-imagining all facets of human social organisation from the perspective of utilising primary energy sources that, while they may by that time be renewable, are nonetheless almost certainly likely to be much more diffuse and much less energy-dense than those we currently have access to. It will be no small task to reconceptualise the entirety of human civilisation in this way, as the example of the recent fundamental historical transition to modernity via the use of fossil fuels shows very clearly as a case-in-point. These many facets would range from, among other things, agriculture, transport, domestic and industrial energy use (and probably, by then, large-scale in situ harvesting or production), climate adaptation, forms of work and organisational design, and so on, to perhaps the very nature of the human relationship with the Earth itself, as well as potential new worldviews and forms of consciousness, founded upon a re-connection with the natural world and the Cosmos at large. As Primack and Abrams (2006) have put it, to “think cosmically, act globally”. Yet if we do face the future squarely and prepare ourselves properly, we might just be able to guide this transition with some relative agency and dignified freedom to act, rather than find ourselves being unwillingly dragged kicking and screaming into a future where we may be forced to abandon much of what we have accomplished in our history so far. As Bertrand de Jouvenel observed long ago (1967, p. 276): The proof of improvidence lies in falling under the empire of necessity. The means of avoiding this lies in acquainting oneself with emerging situations while they can still be molded, before they have become imperatively compelling. In other words, without [foresight] there is effectively no freedom of decision. Whether our species’ transition to Threshold 9 is dignified or not, skillful or not, orderly or not, or commended or not, will depend very much on the seriousness of our anticipatory preparations and our commitment to the necessary actions. Let us hope that our eventual descendants, wherever they may be, will look favourably upon and approve of how we will ultimately choose to respond to the ever-looming energy transition crisis that our fossil-fuelled civilisation is now facing. Moreover, let us make it our number one and over-riding civilisational priority, subsuming and subordinating all others, to do everything we can, in order to ensure that they can. For, in the words of the late Warren Wagar (Marien 2005): We are the link between the traditional civilizations of a well-remembered past and the emergent world civilization. We stand between. If we break under the strain there will be no future. All posterity is in our keeping. Let us therefore deliberate carefully and act with as much wisdom and foresight as we can possibly muster. ***** 472

The coming energy transition

Notes 1 This chapter is a lightly edited, revised, and updated version of earlier work which first appeared in Voros (2013), as well as later in a slightly revised form as part of Voros (2017b). 2 Jantsch also wrote about information as a key concept, one which is increasingly becoming recognised in studies of complexity science (e.g., Lineweaver, Davies & Ruse 2013), as well as about consciousness. Sadly, space does not permit us to explore these any further here, although we will have occasion to briefly mention these two key ideas again later. 3 Interestingly, although we cannot pursue it any further now, Jantsch himself also noted that ‘anticipation’ arises as an emergent property of increasing complexity in the realm of cognition and consciousness (see his Figure 40 and the accompanying discussion, Jantsch 1980, p. 208ff ). 4 Terms frequently used for foresight work include: prediction, projection, forecasting, prognostication, prophecy, conjecture, prognosis, inference, speculation, and a wide variety of other terms. None of these is entirely satisfactory; they all carry certain connotations that may not be helpful, such as the connotation of certainty that the word ‘prediction’ carries, or the connotation of ungrounded guesswork that the word ‘speculation’ has. What is needed is a neutral term that is as free of connotations as possible so as to be merely indicative of the activity of thinking about the future. For nearly two decades now I have used the term ‘prospection’ to denote this activity. As described in, for example, Voros (2003), this is formed from: ‘pro’ = ‘forward’, ‘spect’ = ‘look’, and ‘-tion = the noun form of the action. Thus, the word ‘prospection’, where the stress falls on the second syllable, has the meaning ‘the activity of purposefully thinking about the future to create “forward views” and ideas about, or “images” of, the future.’ In the present context, I will instead use the somewhat more familiar term ‘profiling’, following Clarke (2000), as it carries the appropriate sense of trying to discern broad contours in the complex dynamics of our world (Voros 2009); we will not undertake the entirely futile task of attempting to pinpoint specifics. 5 It is possible to argue, however, that perhaps the defining aspect of  Threshold 6 is the emergence with physiologically modern humans of what appears to be a greatly-­ expanded ability to utilise information, brought about through increases in brain size and perhaps also through changes in the brain’s structural organisation, as well as other physical changes, such as the development of more refined vocal capabilities. This ability to process (through cognition), store (via memory) and transmit information (via speech and/or symbols), from individual to individual, would seem to be the major disjunction that marks this as a Threshold. Christian has argued at length in many places that it is the capacity for what he has called collective learning that is the defining characteristic of our species (e.g., Christian 2010). That capacity is very likely founded upon this intensification of the capacity for information processing, something noted in several places by Jantsch, and manifested in the interaction of the individual and social domains of human groups as an enhanced ability for symbolic informational exchange, which Christian has identified as the capacity for collective learning. 6 Of course, on the timescale of big history, strictly speaking these fuels are renewable, but only on timeframes beyond any practical utility for humans, these being ‘on the order of ’ tens to hundreds of millions of years. Thus, for all practical human purposes, they are effectively non-renewable. 7 The four archetypal futures can be considered a more nuanced and extended expression of a simpler two-class approach, comprising (i) extrapolative evolution, where the system dynamics are assumed to continue relatively smoothly; and (ii) disjunctive 473

Joseph Voros

revolution, where the dynamics are assumed to deviate sharply from smooth continuity (Voros 2006a). 8 The question of how to meaningfully and rigorously use such ‘collapses’, ‘shocks’, ‘wild cards’ or similar ‘discontinuities’ in futures thinking is and remains a difficult one (­H iltunen 2006; van Notten, Sleegers & van Asselt 2005). 9 In this regard, we observe that there has been a stabilisation of the so-named historical ‘Malthusian’ population collapse cycles which occurred prior to the modern revolution (Christian, Brown & Benjamin 2013, p. 190), as those appear to have been driven primarily by inabilities for agricultural productivity to meet the consumption needs of human societies. This has not been so large a problem since the “energy bonanza” of Threshold 8—pandemic disease events and warfare notwithstanding—although with energy scarcity and resource constraints again looming in the future, it may be that such Malthusian cycles might once again make an unwelcome mirror-image return to the long view of human history. It is certainly something that The Limits to Growth studies portend, as well. 10 One might wonder aloud, perhaps facetiously, about who will run the servers or computers, and how they will be powered… 11 The question of the future of human consciousness itself is also an interesting focus of study, although much of this work is still emerging (see, e.g., Ghose 2003; Wilber 2017, 2018). 12 As before, strictly speaking, on the long view of Big History geothermal energy is of course a finite resource, as is solar energy. But both can be expected to last for a very long time, with geothermal energy depletion only likely to become a problem on the same order of timescale as the post-main-sequence red-giant death of the Sun (i.e., some billions of years). One hopes humanity has found a way to move on before this becomes a pressing issue! In the meantime, we can treat both forms of energy as effectively indefinitely renewable; or at least, to a very good approximation!

References Abrams, NE and Primack, JR 2011, The new universe and the human future: How a shared cosmology could transform the world, Kindle edn, Yale University Press, New Haven, CT. Ahmed, NM 2010, A user’s guide to the crisis of civilisation: And how to save it, Pluto Press, London. ——— 2017, Failing states, collapsing systems: BioPhysical triggers of political violence, Springer Briefs in Energy: Energy Analysis, Springer International, Switzerland. doi:10.1007/978-3-319–47816-6 Alvarez, W 1997, T. rex and the crater of doom, Princeton University Press, Princeton. Bardi, U 2008, Cassandra’s curse: How “The Limits to Growth” was demonized, The Oil Drum: Europe [weblog], 9 March 2008, viewed 29 September 2015, http://europe.theoildrum. com/node/3551. ——— 2011, The Limits to Growth revisited, SpringerBriefs in Energy, Springer, New York. doi:10.1007/978-1-4419–9416-5 Broderick, D 1997, The spike: How our lives are being transformed by rapidly advancing technologies, Reed International Books, Kew, VIC, Australia. Brown, CS 2008a, Big history: From the big bang to the present, New Press, New York. Originally published 2007. ——— 2017, Big History, small world: From the big bang to you, Berkshire Publishing Group, Great Barrington, MA, USA. Brown, LR 2008b, Plan B 3.0: Mobilizing to save civilization, W.W. Norton & Co., New York. ——— 2011, World on the edge: How to prevent environmental and economic collapse, W. W. Norton & Company, New York. 474

The coming energy transition

Chaisson, EJ 2001, Cosmic evolution: The rise of complexity in nature, Harvard University Press. ——— 2004, ‘Complexity: An energetics agenda’, Complexity, vol. 9, no. 3, Mar, pp. ­14–21. doi:10.1002/cplx.20009 ——— 2007, Epic of evolution: Seven ages of the cosmos, new paperback edn, Columbia University Press, New York. Originally published Dec 2005. ——— 2008, ‘Cosmic evolution: From the big bang to humankind’, Interactive web pages, Harvard University, https://www.cfa.harvard.edu/~ejchaisson/cosmic_evolution/. Chapman, CR 2004, ‘The hazard of near-Earth asteroid impacts on Earth’, Earth and Planetary Science Letters, vol. 222, no. 1, May 15, pp. 1–15. doi:10.1016/j.epsl.2004.03.004 Christian, D 2004, Maps of time: An introduction to big history, Kindle edn, University of California Press, Berkeley. ——— 2008, Big history: The big bang, life on earth, and the rise of humanity [audiovisual material], 48 lectures, 30 mins/lecture, The Great Courses, course no. 8050, The Teaching Company, Chantilly, VA. ——— 2010, Reinventing intelligence: Why collective learning makes humans so different [Conference Presentation], 31m59s, Techonomy Conference, Lake Tahoe, California. ——— 2018, Origin story: A Big History of everything, Little, Brown and Co., New York. Christian, D, Brown, CS and Benjamin, C 2013, Big history: Between nothing and everything, McGraw-Hill Education, Boston, MA. Christopher, D 2013, The holy universe: A new story of creation for the heart, soul and spirit, New Story Press, Santa Rosa, CA. Clarke, AC 1999, ‘Improving the neighbourhood’, Nature, vol. 402, no. 6757, Nov 4, p. 19. doi:10.1038/46906 ——— 2000, Profiles of the future: An inquiry into the limits of the possible, Millennium edn, Orion Books, London. Originally published 1962. Collins, GS, Melosh, HJ and Marcus, RA 2005, ‘Earth impact effects program: A webbased computer program for calculating the regional environmental consequences of a meteoroid impact on Earth’, Meteoritics and Planetary Science, vol. 40, no. 6, Jun, pp. ­817–840. doi:10.1111/j.1945–5100.2005.tb00157.x Dator, JA 1998, ‘The future lies behind! Thirty years of teaching future studies’, American Behavioral Scientist, vol. 42, no. 3, Nov, pp. 298–319. Issue: ‘Special issue: Futures studies in higher education’, JA Dator (ed.). ——— (ed.) 2002, Advancing futures: Futures studies in higher education, Praeger studies on the 21st century, Praeger, Westport, CT. de Jouvenel, B 1967, The art of conjecture, trans. N Lary, Weidenfeld and Nicholson, London. Originally published 1964. Delsemme, A 1998, Our cosmic origins: From the Big Bang to the emergence of life and intelligence, Cambridge University Press, Cambridge. Di Martino, M, Farinella, P and Longo, G 1998, ‘Foreword of the Tunguska issue’, Planetary and Space Science, vol. 46, no. 2–3, Feb-Mar, p. 125. Issue: ‘International Workshop Tunguska 96’. doi:10.1016/S0032–0633(97)00164-5 Diamandis, PH and Kotler, S 2012, Abundance: The future is better than you think, Free Press, New York. Diamond, J 2005, Collapse: How societies choose to fail or survive, Allen Lane (Penguin), London. Eden, AH, Moor, JH, Søraker, JH and Steinhart, E (eds) 2012, Singularity hypotheses: A scientific and philosophical assessment, The Frontiers Collection, Springer, Berlin. doi:10.1007/978-3-642–32560-1 Ehrlich, PR 1971, The population bomb, Ballantine; Friends of the Earth; Pan, London. Originally published 1968. 475

Joseph Voros

Elgin, D 1994, ‘Building a sustainable species-civilization: A challenge of culture and consciousness’, Futures, vol. 26, no. 2, Mar, pp. 234–245. Issue: ‘Visions of sustainability’. doi:10.1016/0016–3287(94)90112-0 ——— 2001, Promise ahead: A vision of hope and action for humanity’s future, ebook edn, HarperCollins, New York. ——— 2009, The living universe: Where are we? Who are we? Where are we going?, Berret-­ Koehler, San Francisco. Foreword by Deepak Chopra. Farnish, K 2009, Time’s up!: An uncivilized solution to a global crisis, Green Books, London. Floyd, J 2012, ‘Beyond this brief anomaly: An inquiry into energy and society’, weblog, viewed March 21, 2012 http://beyondthisbriefanomaly.org. Galtung, J and Inayatullah, S (eds) 1997, Macrohistory and macrohistorians: Perspectives on individual, social, and civilizational change, Praeger Publishers, Westport, CT. Based on a seminar given Spring 1989 at the Dept. of Sociology, University of Hawaii, Honolulu. Gasperini, L, Alvisi, F, Biasini, G, Bonatti, E, Longo, G, Pipan, M, et al. 2007, ‘A possible impact crater for the 1908 Tunguska Event’, Terra Nova, vol. 19, no. 4, Aug, pp. 245–251. doi:10.1111/j.1365–3121.2007.00742.x Genet, C, Genet, R, Swimme, B, Palmer, L and Gibler, L (eds) 2009, The evolutionary epic: Science’s story and humanity’s response, Collins Foundation Press, Santa Margarita, CA, USA. Foreword by David Christian. Ghose, A 2003, The future evolution of man: The divine life upon Earth, trans. PB Saint-Hilaire, Lotus Press, Twin Lakes, WI, USA. Originally published 1963. Greer, JM 2008, The long descent: A user’s guide to the end of the industrial age, New Society Publishers, Gabriola Island, BC. ——— 2009, The ecotechnic future: Envisioning a post-peak world, New Society Publishers, Gabriola Island, BC. Hall, CAS and Klitgaard, KA 2012, Energy and the wealth of nations: Understanding the biophysical economy, Springer, New York. doi:10.1007/978-1-4419-9398-4 Harman, WW 1998, Global mind change: The promise of the twenty-first century, 2nd rev. and expanded edn, Berrett-Kohler, San Francisco, CA. Originally published 1987. Heinberg, R 2009, Searching for a miracle: “Net energy” limits and the fate of industrial society, False Solutions Series, no. 4, Post-Carbon Institute and International Forum on ­Globalization, San Francisco. ——— 2010, Peak everything: Waking up to the century of declines, New Society Publishers, Gabriola Island, BC. Heinberg, R and Lerch, D (eds) 2010, The post carbon reader: Managing the 21st century’s sustainability crises, Contemporary Issues Series, Watershed Media, Healdsburg, CA. Hiltunen, E 2006, ‘Was it a wild card or just our blindness to gradual change?’, Journal of Futures Studies, vol. 11, no. 2, Nov, pp. 61–74. Holmgren, D 2009, Future scenarios: How communities can adapt to peak oil and climate change, Chelsea Green Publishing, White River Junction, VT. Homer-Dixon, TF 2006, The upside of down: Catastrophe, creativity, and the renewal of civilization, Island Press, Washington, DC. Inayatullah, S 1998, ‘Macrohistory and futures studies’, Futures, vol. 30, no. 5, Jun, pp. ­381–394. doi:10.1016/S0016–3287(98)00043-3 International Big History Association 2016, ‘What is Big History?’, web page, https:// bighistory.org/whatisbighistory/ Jantsch, E 1980, The self-organizing universe: Scientific and human implications of the emerging paradigm of evolution, Pergamon Press, New York. ——— (ed.) 1981, The evolutionary vision: Toward a unifying paradigm of physical, biological and sociocultural evolution, American Association for the Advancement of Science selected symposium, vol. 61, Westview Press (for AAAS), Boulder, CO. 476

The coming energy transition

Kunstler, JH 2005, The long emergency: Surviving the end of oil, climate change, and other converging catastrophes of the twenty-first century, Grove Press, New York. Originally published 2004. With a new Afterword by the author. Kurzweil, R 1999, The age of spiritual machines: When computers exceed human intelligence, ­A llen & Unwin, Sydney. Originally published 1999 (Viking Penguin), New York. ——— 2006, The singularity is near: When humans transcend biology, Penguin Books. Originally published 2005 (Viking Adult). Lineweaver, CH, Davies, PCW and Ruse, M (eds) 2013, Complexity and the arrow of time, Cambridge University Press, Cambridge. Lynas, M 2008, Six degrees: Our future on a hotter planet, National Geographic Society, Washington, DC. Marien, M 2005, ‘Celebrating the life of futurist W. Warren Wagar’ [Visionaries], The Futurist, vol. 39, no. 2, Mar–Apr, p. 68. Meadows, DH, Meadows, DL, Randers, J and Behrens, WW 1972, The limits to growth: A report for the Club of Rome’s project on the predicament of mankind, Universe Books, New York. Meadows, DH, Randers, J and Meadows, DL 2004, Limits to growth: The 30-year update, Chelsea Green Publ., White River Junction, VT. Millennium Project 2012, ‘State of the future index’, web page, The Millennium Project, viewed Oct 13, 2012 http://www.millennium-project.org/millennium/SOFI.html. Niele, F 2005, Energy: Engine of evolution, Elsevier, Amsterdam. Orlov, D 2008, Reinventing collapse: The Soviet example and American prospects, New Society Publishers, Gabriola Island, BC. Petersen, JL 1997, ‘The wild cards in our future: Preparing for the improbable’, The Futurist, vol. 31, no. 4, Jul–Aug, pp. 43–47. ——— 1999, Out of the blue: How to anticipate big future surprises, 2nd edn, Madison Books, Lanham, MA. Originally published 1997. Primack, JR and Abrams, NE 2006, The view from the center of the universe: Discovering our extraordinary place in the cosmos, Riverhead (Penguin), New York. Randers, J 2012, ‘The real message of The Limits to Growth: A plea for forward-looking global policy’, Gaia, vol. 21, no. 2, pp. 102–105. Roberts, P 2005, The end of oil: On the edge of a perilous new world, Houghton Mifflin ­Harcourt, Boston. Originally published 2004. Sagan, C 1973, The cosmic connection: An extraterrestrial perspective, Doubleday, New York. ——— 1980, Cosmos, Illustrated (First) edn, Random House, New York. ——— 1995, Pale blue dot: A vision of the human future in space, Headline, London. Sahlins, MD 1972, ‘The original affluent society’, in Stone age economics, Aldine de Gruyter, New York, pp. 1–40. Schwartz, P, Leyden, P and Hyatt, J 2000, The long boom: A vision for the coming age of prosperity, Orion Publishing, London. Schweickart, RL, Jones, TD, von der Dunk, F and Camacho-Lara, S 2008, Asteroid threats: A call for a global response, trans. AIPoAT Mitigation, Association of Space Explorers, Houston, TX. Tok, Jessica, (ed.). Slaughter, RA 2010, The biggest wake up call in history, Foresight International, Indooroopilly (Brisbane), Australia. Smart, J 2003, ‘Considering the singularity: A coming world of autonomous intelligence (A.I.)’, in HF Didsbury, Jr. (ed.) 21st century opportunities and challenges: An age of destruction or an age of transformation, World Future Society, Bethesda, MD, pp. 256–262. Smil, V 1994, Energy in world history, Essays in world history, Westview Press, Boulder, CO. ——— 2010a, Energy myths and realities: Bringing science to the energy policy debate, American Enterprise Institute for Public Policy Research (distr. by Rowman & Littlefield, Lanham, MD), Washington DC. 477

Joseph Voros

——— 2010b, Energy transitions: History, requirements, prospects, Greenwood Publishing Group, Santa Barbara, CA. ——— 2017, Energy and civilization: A history, MIT Press, Cambridge, MA. Spier, F 1996, The structure of big history: From the big bang until today, Amsterdam University Press, Amsterdam. ——— 2005, ‘How big history works: Energy flows and the rise and demise of complexity’, Social Evolution and History, vol. 4, no. 1, Mar, pp. 87–135. Issue: ‘Exploring the horizons of big history’. ——— 2010, Big history and the future of humanity, Wiley-Blackwell, Chichester. ——— 2011, ‘Complexity in big history’, Cliodynamics, vol. 2, no. 1, pp. 146–166. ——— 2015, Big history and the future of humanity, 2nd edn, Wiley-Blackwell, Chichester. Swimme, BT and Tucker, ME 2011, Journey of the universe, Kindle edn, Yale University Press, New Haven, CT. Tainter, JA 1990, The collapse of complex societies, New studies in archaeology, Cambridge University Press, Cambridge. Originally published 1988. Turner, GM 2008, ‘A comparison of The Limits to Growth with 30 years of reality’, Global Environmental Change, vol. 18, no. 3, Aug, pp. 397–411. doi:10.1016/j.gloenvcha.2008.05.001 ——— 2014, Is global collapse imminent? An updated comparison of The Limits to Growth with historical data, MSSI Research Paper No. 4, Melbourne Sustainable Society Institute, The University of Melbourne. http://sustainable.unimelb.edu.au/research-papers. van Notten, PWF, Sleegers, AM and van Asselt, MBA 2005, ‘The future shocks: On discontinuity and scenario development’, Technological Forecasting and Social Change, vol. 72, no. 2, Feb, pp. 175–194. doi:10.1016/j.techfore.2003.12.003 Voros, J 2003, ‘A generic foresight process framework’, Foresight, vol. 5, no. 3, pp. 10–21. Won a Highly Commended Award in the Emerald Literati Club Awards for Excellence 2004: see Foresight, vol.6, no.5, 2004. doi:10.1108/14636680310698379 ——— 2005, ‘A generalised “layered methodology” framework’, Foresight, vol. 7, no. 2, pp. 28–40. doi:10.1108/14636680510700094 ——— 2006a, ‘Introducing a classification framework for prospective methods’, Foresight, vol. 8, no. 2, pp. 43–56. Won a Highly Commended award in the Emerald Literati Network Awards for Excellence 2007. doi:10.1108/14636680610656174 ——— 2006b, ‘Nesting social-analytical perspectives: An approach to macro-social analysis’, Journal of Futures Studies, vol. 11, no. 1, Aug, pp. 1–21. ——— 2009, ‘Morphological prospection: Profiling the shapes of things to come’, Foresight, vol. 11, no. 6, pp. 4–20. Won the Outstanding Paper Award in the Emerald Literati Awards for Excellence 2010. doi:10.1108/14636680911004939 ——— 2012, ‘Macro-prospection: Thinking about the future using macro- and Big History’. Paper presented at Global Future 2045 International Congress, Radisson-Slavyanskaya Hotel, Moscow, Feb 17–20. http://hdl.handle.net/1959.3/216156. ——— 2013, ‘Profiling “Threshold 9”: Using Big History as a framework for thinking about the contours of the coming global future’, in LE Grinin & AV Korotayev (eds), Evolution: Development within Big History, evolutionary and world-system paradigms, Evolution Almanac, vol. 3, Uchitel Publishing House, Volgograd, Chap. 6, pp. 119–142. ——— 2015, ‘The Anthropocene, “Threshold 9” and the long-term future of humankind’. Paper presented at The Big History Anthropocene Conference: A Transdisciplinary Exploration, Big History Institute, Macquarie University, Sydney, 9–11 Dec, Dec 11. https://youtu.be/Z7gU3ZsSWr0. ——— 2017a, ‘Big Futures: Macrohistorical perspectives on the future of humankind’, in B Rodrigue, L Grinin & A Korotayev (eds), The way that Big History works: Cosmos, life,

478

The coming energy transition

society and our future, From Big Bang to galactic civilizations: A Big History anthology, vol. III, Primus Books, Delhi, Chap. 22, pp. 403–436. ——— 2017b, ‘Big History and anticipation: Using Big History as a framework for global foresight’, in R Poli (ed.) Handbook of anticipation: Theoretical and applied aspects of the use of future in decision making, Springer International, Cham, Chap. 95. doi:10.1007/978-3-319–31737-3_95-1 Wagar, WW 1991, The next three futures: Paradigms of things to come, Contributions to sociology, no. 98, Greenwood Press, New York. ——— 1999, A short history of the future, 3rd edn, University of Chicago Press. Originally published 1989, 1992. Afterword by Immanuel Wallerstein. Wilber, K 2017, The religion of tomorrow: A vision for the future of the Great Traditions—more inclusive, more comprehensive, more complete, Shambhala, Boulder. Also available as Kindle and Audible editions. ——— 2018, Integral Buddhism and the future of spirituality, Shambhala, Boulder. Also available as Kindle and Audible editions.

479

INDEX

3D network molecules 285 Abel, Wilhelm 208–15, 219 abundance of elements 284 accumulation of parasites, bacteria and predators 289 aerial photography 303 Age of Discovery 186 AHA (American Historical Association) 181 AI (Artificia Intelligence) 126, 401, 409, 412 aja’ib 112 Alphonso the camel 379 Alpine Orogeny 306 Alvarez, Walter 57 animals brick building 288–90 animism 260 Annales 202, 210, 212, 215 Anthropocene 437, 443, 450 Anthropocene Epoch 30 anthropography 109,110 anthropology 112, 113, 115–25 anthroturbation 436 ANZAC 322, 323 archaeosphere 444 Armstrong, Stuart 61 Arnold, Matthew 28 asabiyyah 205, 216 Atemporality 41 attractor 427 Augustine 4 Australian Imperial Forces 320 Axial Age 110, 265

Benner, Steven 64 Berman, Marshall 29 Berrill, Norman 38 Big Chill 198 Big History Academy, Korea 368, 371 Big History and the Future of Humanity 8 Big History Institute 21 The Big History of Civilization (Great Courses) 8 Big History Project 8, 175, 339–41, 343–9, 351, 353, 355, 357, 359, 362, 372–3 Big History School 9 Big History: From Nothing to Everything 9 Big History: From the Big Bang to the Present 9 big learning 375 big teaching 375 biosphere 434–6, 439 Biotechnology 402, 406, 408, 409 Bishop Orosius 4 Black Day of the German Army 331 bloodshed ratio 423 Bodin, Jean 16 Bois, Guy 209, 212–14, 216 Bostrom, Nick 62 Braudel, Fernand 202, 203, 210, 220 Brenner, Robert 211–14, 219 brick 282 Brin, David 60, 62 Brunhes, Bernard 48 building with certain elements 284–8 building with gases, liquids or solids 284–5 Calder, Nigel 38 Cambrai 324 481

Index

Cambrian explosion 308 Cannadine, David 20 Cantabrian Mountains 306 carbon 300 carbon cycle 307 Carboniferous forests 309 Carr, E.H. 17 Carter, Howard 156 Chaisson, Eric 21, 26, 38, 58, 456–8 cheap building 285 chemistry of politics 196 Childe,Vere Gordon 157, 163 chimpanzee 182, 184, 193 Cho, Ji-Hyung 368 Christian, David 39, 362, 456–9, 466, 469 Chronometric Revolution 5 ChronoZoom 170 claim testers 353, 357, 359 Clark, Grahame 158 Cloud, Preston 23, 38 Co2 sink 309 coal deposits 301, 309 coal mining 313 Collapse 461, 463–6, 471, 474 Collective learning 27, 202, 204, 209, 219, 220, 221, 223, 224, 466 Collingwood, R.G. 19 Communist Manifesto 29 commuting 289 competition 254 complementarity 283 complexity 177,190, 191, 193, 197, 198, 202, 204, 214, 220, 223, 224, 397, 400, 401, 404 concentration of carbon on Earth 318 connections 178 constant universe 172 constraining and refining behaviour 291 Continuation 462, 463 Cosmic Background Radiation (CBR) 42 Cosmos 173 Crab Nebula 41 Crash Course Big History 8 Crick, Francis 38 crisis 423–4 critical thinking 372 Crosby, Alfred W. 8 Cultural intelligence hypothesis 91 cultural regulation 423, 429 Currie, Madame 47 Cybernetic Revolution 405–8, 410–16, 418

Darwin, Charles 206, 219, 220 Dator, James 457, 459, 461, 466 deep tier traces 444 denaturalization 423 design experiment 372 Diodorus Siculus 4 directing energy flows 293 Disciplined/constrained society 464, 465 diseases 313 divine kingship 264 DNA 193, 195 Dobb, Maurice 209 Dominican University California 10 Dover Beach 28, 29 Drake Equation 58 duration of building use 289 Durkheim, Emile 18, 25, 28 educative curriculum 383–4 Einstein, Albert 38 energy consumption 434–5, 437–8 energy flows 293 energy-rate density 127 Epic of Gilgamesh 109 eternal life 267 ethnology 118 eusocial 290 Ewha Womans University 362 Exoplanets 59 external social sustainability 424 Fermi Paradox 58 Fernandez-Armesto, Felipe 22, 23 final innovative phase 398 foresight 457, 458, 471 formation of carbon in stars 316–18 Fraser, T.J. 37, 39 Fusion energy 467 future of religion and morality 271–4 Gaia Space Telescope 66 Garrod, Dorothy 158 gender 180, 181, 183, 185, 190, 197, 198 Gene modification 407 General intelligence hypothesis 91 genotemporality 50 geographical distribution of metro systems 449–50 geological evolution 306 Gliese 445 66 Gliese 710 66 global civilization 113, 114, 120 482

Index

globalization 113, 181, 185–7, 190, 192, 198 goldilocks circumstances 293 Goldstone, Jack 215, 219, 224 Gordon, Richard 63 Gouldsblom, Johan 21 grand narratives 293 Grand Valley State University 11 Grantham 214 Great Acceleration 438, 445, 449 Great Filter 60–2

International Society for the Study of Time (ISST) 37 investigations 374, 383 isochronous 436

Haber, Fritz 327 Hallpike, Christopher 75 Hana Academy, Seoul 367 Hansen, Robin 60 Hawking, Stephen 5 Hegel, George 4 Herto, Ethiopia 161 Hess, Harry 49 Hewlett-Packard walk-through-time exhibit 38 high status knowledge 341 Hindenburg Line 332 historical record of nonviolence 430 History Channel 8 history of little big histories 281 honest signal 291 Hughes-Warrington, Marnie 7 human brick use 290–2 human footprint 303 human language ability 257 Hutton, James 45, 46

Kant, Immanuel 4 Kondratiev, Nikolai 208, 216 Korean Foundation for the Advancement of Science and Creativity (KOFAC) 365

Ibn Khaldun 4, 205 ichnofabrics 444 implementation principle 427 Influenza Pandemic of 1918–1919 334 initial innovative phase 398 inquiry and literacy practices 381–2 Institute of World and Global History, Ewha Womans University 362, 364–6 intelligence amplification 126 intensive care 257 interdisciplinarity 115, 124, 340, 342, 343, 351, 359 interdisciplinary 177, 281 Intergovernmental Panel on Climate Change 309 internal social sustainability 423 International Big History Association (IBHA) 11 International Geophysical Year 49

James Webb Space Telescope (JWST) 59 Jeju Special Self-Governing Province, Korea 368 Journal of Big History 16 Journal of World History 21

Ladurie, Emmanuel La Roy 203, 204, 210–14, 219, 220, 224 Lamarckian mechanism 81, 82 language 4, 5, 15, 23 Laniakea Supercluster 41 law of delayed dysfunction 431 law of techno-humanitarian balance 423 layered structure of little big histories 283–4 Leakey, Louis 157, 158 Leakey, Mary 157 learning progression 384 lenses 174 levels of complexity 279 LIDAR 164 limestones 308, 315 linguistics 118, 130 little big histories 177, 279 little big history assignment 282 longue durée 219, 220, 222, 224 Lord Acton 363 Lord Carnarvon 156 Lyotard, Jean-Francoise 20 McNeill, John 30 McNeill, William H. 24, 32, 363 Macquarie University 9, 21 magnetotemporality 48 major phases in big history 283 Malthus, Thomas 205–16, 219 mammal morality 255 Manifest Destiny 181 Marx, Karl 4, 19, 207, 209–15, 219 Massive Open Online Courses (MOOCS) 21 materialism 251 materialist 119 Mazlish, Bruce 7 483

Index

meaning formation 431 Mears, John 21 Messines 325, 326 Metagalaxy 421, 428 metro 434–6 miniaturization 401, 405, 406 modernization phase 398 monocausal explanations 376 Moore’s Law 63 moral codes 266–7 morality and religion in the modern state 268 A Most Improbable Journey. A Big History of Our Planet and Ourselves 57 multi-faceted causal explanations 376 Muqaddimah 4 Myos Hormos 165 Nalón basin archeological record 312–13 Nalón basin faunal and floral record 310 Nalón River 300 nanomaterials 316 nanotechnology 402, 411 nationalism 181 natural history 171–2 natural philosophy 172 natural, human-based morality 268 naturalism 251, 270 Nazaretyan, Akop 123, 126, 130 net primary production 437, 449 neutral building 292 new questions 282, 288, 290, 292 nitrogen 321 Noosphere 422, 427 online repository 381–2 Oort Cloud 66 The Origin of Continents and Oceans 49 origin story 260 Oumuamua (“messenger from afar arriving first”) 65 pair bonding 257 paleolithic/tribal morality and religion 258 Panama Canal 322 pedagogical framework 340, 349, 352–4, 356–9 perspective 174 petrotemporality 42–4 Piggott, Stuart 156 Pilbara, Western Australia 47 political identity 180, 181, 185, 186

post-carbon civilisation 460 post-human revolution 397 Postan, Michael 210, 212–14 potassium argon dating 160 pre-human morality 254 pre-instructional practices, beliefs and knowledge 373, 377 predisposition to moral behavior 256 preservation potential 439, 444–5, 449 Principles of Social Evolution (1986) 75 professionalization 118 Proxima Centauri 65 race 180–3, 185, 186, 190, 197, 198 Radical Enlightenment 269 radiometric dating 23 religion, in chiefdoms 262–3 religion, definition 252 religion, in early agrarian societies 263–5 religion, in expanding agrarian civilisations 265–8 religion, origins 259 religions, portable universal 266 religious ecology 273 religious naturalists 273 Renfrew, Colin 23 Ricardo, David 207 Riddley, Matt 51 rock cavities 310 Royal Air Force 329 Sagan, Carl 38 salvation 267 scaling law in phase transitions 426 Schrodinger, Erwin 22 Science, Technology, Education, Arts and Mathematics (STEAM) 362 Search for Extra-Terrestrial Intelligence (SETI) 59 The Second Coming 29 self-discipline 262 Seodaemun Museum of Natural History, Korea 368, 370 separation of church and state 268 shaman 261 Sharov, Alexei 63 silicate mineral glue 286 Sima Qian 4 singularity 420, 427 size 175 Snow, C.P 18, 39 social evolution vectors 422 484

Index

Spier, Fred 6, 7, 39 standardized building practices 288–90 state religions 267 Steno, Nicholas 43 stratigraphical resolution 448 student learning process 381–2 submarines 323 substrate 440–1 Sung Dynasty (960–1279) 81 supporting teachers 381 Swaziland Supergroup 47 Synthetic Theory of Evolution 82 technofossil 440, 449 technological potential 423 technosphere 434–6, 439, 449 TED talk Big History 8 telescopes 316 theoretical framework 175 theory of mind 260 This Fleeting World 8 Threshold 457–9, 467, 469, 471, 472 thresholds of increasing complexity 372, 386 thresholds of increasingly complex learning 373 Toynbee, Arnold 21 trace fossils 440 transdisciplinarity 342, 343 transdisciplinary 124, 177 Transformational Society 462, 466, 467, 471 transformative learning theory 340, 354, 356–9 Transiting Exoplanet Survey Satellite (TESS) 59 Treaty of Brest-Litovsk 327 Treptichnus pedum 440

Trevelyan, George Macauley 17 tunneling 443 Turchin, Peter 215–19 Tylor, Edward 115, 117, 162 U-Boats 323 Uluburun shipwreck 159 umwelt 40 unfamiliar content 381 Universal thinking 124 University of Amsterdam 10 urban population 434–5, 437–8 urban primary energy use 435, 439 Variscan Orogeny 306–8 Villanova University 11 Villers Bretonneux 330 Virgo Constellation 41 Volk, Tyler 39 Von Humboldt, Alexander 22, 173 Von Ranke, Leopold 363 weaponized gas 327 Wegener, Alfred 49 Wells, H.G. 4, 21 Western Front 325 White, Hayden 20 Wilkinson Microwave Anisotropy Probe (WMAP) 41 Wilson, Edward O. 7 Woolley, Leonard 156 World Class University (Korea) 362 World Prehistory 157, 158 writing 264 Yeats, W.B. 29

485