Rethinking Food and Agriculture: New Ways Forward 0128164107, 9780128164105

Given the central role of the food and agriculture system in driving so many of the connected ecological, social and eco

1,808 56 18MB

English Pages 476 [478] Year 2020

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Rethinking Food and Agriculture: New Ways Forward
 0128164107, 9780128164105

Table of contents :
Contents
Contributors
Gratitude and acknowledgments
Introduction • Laila Kassam and Amir Kassam
1 Setting innovation free in agriculture • Rupert Sheldrake
2 Agriculture planted the seeds of alienation from nature • Jim Mason and Laila Kassam
3 Political economy of the global food and agriculture system • Philip McMichael
4 Neocolonialism and the New Alliance for Food Security and Nutrition: A gendered analysis of the development consequences for Africa • Mark Langan and Sophia Price
5 The myth of a food crisis • Jonathan Latham
6 Animal ethics as a critique of animal agriculture, environmentalism, foodieism, locavorism, and clean meat • Robert C. Jones
7 A food system fit for the future • Tony Juniper
8 Why change the way we grow, process, and consume our food? • Hans R. Herren
9 Two paradigms of science—And two models of science-based agriculture • Colin Tudge
10 Paradigms of agriculture • Amir Kassam and Laila Kassam
11 Soil health and the revolutionary potential of Conservation Agriculture • David R. Montgomery
12 Climate change adaptability and mitigation with Conservation Agriculture • Emilio J. Gonzalez-Sanchez, Oscar Veroz-Gonzalez, Manuel Moreno-Garcia, Manuel R. Gomez-Ariza, Rafaela Ordoñez-Fernandez, Paula Trivino-Tarradas, Amir Kassam, Jesús A. Gil-Ribes, Gottlieb Basch, and Rosa Carbonell-Bojollo
13 Will gene-edited and other GM crops fail sustainable food systems? • Allison K Wilson
14 Sustaining agricultural biodiversity and heterogeneous seeds • Patrick Mulvany
15 Healthy diets as a guide to responsible food systems • Shireen Kassam, David Jenkins, Doug Bristor, and Zahra Kassam
16 Knowledge systems for inclusively responsible food and agriculture • Robert Chambers
17 Social movements in the transformation of food and agriculture systems • Nassim Nobari
18 Alternatives to the global food regime: Steps toward system transformation • Helena Norberg-Hodge
19 Cocreating responsible food and agriculture systems • Vandana Shiva
20 Toward inclusive responsibility • Laila Kassam and Amir Kassam
Index

Citation preview

Rethinking Food and Agriculture: New Ways Forward

This page intentionally left blank

Woodhead Publishing Series in Food Science, Technology and Nutrition

Rethinking Food and Agriculture New Ways Forward

Edited by

Amir Kassam

University of Reading, Reading, United Kingdom

Laila Kassam

Animal Think Tank, Lancaster, United Kingdom

An imprint of Elsevier

Woodhead Publishing is an imprint of Elsevier The Officers’ Mess Business Centre, Royston Road, Duxford, CB22 4QH, United Kingdom 50 Hampshire Street, 5th Floor, Cambridge, MA 02139, United States The Boulevard, Langford Lane, Kidlington, OX5 1GB, United Kingdom © 2021 Elsevier Inc. All rights reserved. No part of this publication may be reproduced or transmitted in any form or by any means, electronic or mechanical, including photocopying, recording, or any information storage and retrieval system, without permission in writing from the publisher. Details on how to seek permission, further information about the Publisher’s permissions policies and our arrangements with organizations such as the Copyright Clearance Center and the Copyright Licensing Agency, can be found at our website: www.elsevier.com/ permissions. This book and the individual contributions contained in it are protected under copyright by the Publisher (other than as may be noted herein). Notices Knowledge and best practice in this field are constantly changing. As new research and experience broaden our understanding, changes in research methods, professional practices, or medical treatment may become necessary. Practitioners and researchers must always rely on their own experience and knowledge in evaluating and using any information, methods, compounds, or experiments described herein. In using such information or methods they should be mindful of their own safety and the safety of others, including parties for whom they have a professional responsibility. To the fullest extent of the law, neither the Publisher nor the authors, contributors, or editors, assume any liability for any injury and/or damage to persons or property as a matter of products liability, negligence or otherwise, or from any use or operation of any methods, products, instructions, or ideas contained in the material herein. Library of Congress Cataloging-in-Publication Data A catalog record for this book is available from the Library of Congress British Library Cataloguing-in-Publication Data A catalogue record for this book is available from the British Library ISBN: 978-0-12-816410-5 ISBN: 978-0-12-816411-2 For information on all Woodhead publications visit our website at https://www.elsevier.com/books-and-journals

Publisher: Charlotte Cockle Acquisitions Editor: Nancy Maragioglio Editorial Project Manager: Emerald Li Production Project Manager: Vignesh Tamil Cover Designer: Miles Hitchen Typeset by SPi Global, India

Contents

Contributors xi Gratitude and acknowledgments xv Introduction xvii 1 Setting innovation free in agriculture Rupert Sheldrake 1.1 Introduction 1.2 The scientific priesthood 1.3 The fantasy of omniscience 1.4 The credibility crunch for materialism 1.5 The unfulfilled promises of molecular biology 1.6 Toward a more holistic approach 1.7 Setting innovation free References

1 1 2 5 7 10 13 25 26

2 Agriculture planted the seeds of alienation from nature Jim Mason and Laila Kassam 2.1 Introduction 2.2 Dominionism: Our nature-dominating worldview and its origins 2.3 Before agriculture: A world alive and ensouled 2.4 Other animals: Movers and shakers of the human mind and worldview 2.5 Agriculture: A new relationship with nature, a new social order 2.6 Misothery: The reduction of animals and nature 2.7 Herders’ legacies: The making of the Western worldview 2.8 War and colonialism: Dominating lands and others 2.9 The herder roots of property, money, and capitalism 2.10 Consequences of dominionism and misothery 2.11 Conclusion References

31

3 Political economy of the global food and agriculture system Philip McMichael 3.1 Introduction 3.2 Antecedents of the present crisis 3.3 Political economy of the food regime 3.4 Food regime complexes

53

31 31 33 35 37 39 41 44 45 46 48 49

53 53 55 57

viContents

3.5 Food system transitions 3.6 Crisis and resolution Acknowledgments References 4 Neocolonialism and the New Alliance for Food Security and Nutrition: A gendered analysis of the development consequences for Africa Mark Langan and Sophia Price 4.1 Introduction 4.2 Neocolonialism as a critique of donor aid policies in the Global South 4.3 The NAFSN: Sustainable development or a form of “neocolonial” imposition? 4.4 A gendered lens on “development”: The impact of the neocolonial NAFSN for women in Africa 4.5 Conclusion References 5 The myth of a food crisis Jonathan Latham 5.1 Introduction 5.2 The assumptions of GAPS 5.3 What are models for? Acknowledgments References Further reading 6 Animal ethics as a critique of animal agriculture, environmentalism, foodieism, locavorism, and clean meat Robert C. Jones 6.1 Introduction 6.2 Ethics basics: Moral status, moral value, and anthropocentrism 6.3 Ethical foundations of the modern animal rights movement 6.4 Animal liberation never was a triangular affair 6.5 Foodieism and locavorism: A celebration of “humane” exploitation 6.6 Will “clean meat” end animal agriculture? 6.7 Redefining speciesism 6.8 Conclusion References 7 A food system fit for the future Tony Juniper 7.1 Conversion, intensification, and degradation 7.2 Impacts of inputs 7.3 Climate change and wider food system vulnerability

64 66 70 70

77 77 78 80 85 89 90 93 93 98 103 107 107 111 113 113 113 116 119 122 128 129 130 131 135 135 137 138

Contentsvii

7.4 Drivers of the current system 7.5 What food shortage? 7.6 Is our food really that cheap? 7.7 Beware false frames 7.8 The actors 7.9 Priorities for a sustainable food system 7.10 Governments 7.11 Private sector 7.12 Consumers 7.13 The big issue References 8 Why change the way we grow, process, and consume our food? Hans R. Herren 8.1 Introduction 8.2 A bankrupt system: Dealing with symptoms rather than the causes 8.3 A new paradigm for the food system: How to change course 8.4 A framework for change References

139 140 141 142 142 143 143 144 145 146 146 149 149 154 156 158 161

9 Two paradigms of science—And two models of science-based agriculture 165 Colin Tudge 9.1 Paradigm I: The high-tech industrial route 166 9.2 Paradigm II: The path to enlightened agriculture 169 9.3 The absolute importance of agroecology 170 9.4 A cross-the-board rethink 173 9.5 Addendum I: What price livestock? Do we really need farms to be “mixed”? 175 9.6 Addendum II: The college for real farming and food culture 178 References 178 10 Paradigms of agriculture Amir Kassam and Laila Kassam 10.1 Introduction 10.2 The industrial Green Revolution agriculture paradigm 10.3 Alternative paradigms of agriculture 10.4 How ecologically sustainable are the alternative agriculture paradigms? 10.5 Consideration of structural issues 10.6 Ways forward for alternative paradigms 10.7 Future prospects for alternative agriculture paradigms in the context of inclusively responsible food and agriculture systems References

181 181 182 189 197 209 209 211 212

viiiContents

11 Soil health and the revolutionary potential of Conservation Agriculture 219 David R. Montgomery 11.1 Introduction 219 11.2 Extent of global land degradation 220 11.3 Side effects of conventional agriculture 221 11.4 Building soil health through Conservation Agriculture 223 11.5 The fifth revolution 226 References 228 12 Climate change adaptability and mitigation with Conservation Agriculture 231 Emilio J. Gonzalez-Sanchez, Oscar Veroz-Gonzalez, Manuel Moreno-Garcia, Manuel R. Gomez-Ariza, Rafaela Ordoñez-Fernandez, Paula Trivino-Tarradas, Amir Kassam, Jesús A. Gil-Ribes, Gottlieb Basch, and Rosa Carbonell-Bojollo 12.1 Introduction 231 12.2 Climate change and agriculture: Why do we need to change the agricultural paradigm? 232 12.3 Conservation Agriculture: A sustainable farming system that mitigates and adapts to climate change 233 12.4 Conclusions 243 Acknowledgments 244 References 244 13 Will gene-edited and other GM crops fail sustainable food systems? Allison K Wilson 13.1 Introduction 13.2 Impacts of HT and Bt crops 13.3 Unintended traits in GM crops 13.4 New GM traits and techniques 13.5 Sustainable agriculture and plant breeding 13.6 Conclusions: Obstacles and opportunities References 14 Sustaining agricultural biodiversity and heterogeneous seeds Patrick Mulvany 14.1 Agricultural biodiversity is intentionally heterogeneous 14.2 Distinctive features of agricultural biodiversity 14.3 Threats to agricultural biodiversity 14.4 Few crops “Feed the World”? 14.5 International governance of agricultural biodiversity and seeds 14.6 Peasant seeds: More numerous and heterogeneous 14.7 Prioritizing biodiverse and heterogeneous peasant agroecology

247 247 249 251 259 265 268 269 285 285 287 289 291 294 297 301

Contentsix

14.8 Three coalitions contesting control over agricultural biodiversity and seed systems 14.9 Heterogeneous seed and agricultural biodiversity—The basis for responsible agriculture and food systems References 15 Healthy diets as a guide to responsible food systems Shireen Kassam, David Jenkins, Doug Bristor, and Zahra Kassam 15.1 Introduction 15.2 Current disease trends 15.3 Diet patterns associated with health and longevity 15.4 Using nutrition as therapy 15.5 An approach to nutritional science that has led to misguided recommendations 15.6 Industry influence on nutrition research and guidelines 15.7 Government intervention to promote healthy diet and nutrition 15.8 Concerns about the impact of current diet trends on the environment 15.9 Conclusions and recommendations References

305 307 317 323 323 325 327 332 333 334 336 337 341 342

16 Knowledge systems for inclusively responsible food and agriculture 353 Robert Chambers 16.1 Introduction 353 16.2 Knowledge systems, words, and meanings 353 16.3 Learning from paradigmatic challenges to entrenched knowledge 356 16.4 Putting farmers and their communities and natural resources first 364 16.5 Coda 367 References 368 17 Social movements in the transformation of food and agriculture systems Nassim Nobari 17.1 Introduction 17.2 A corporate food regime for the few 17.3 Knowing the problem: Agroecology and seeing through the narrative of lack 17.4 Stepping in as an urban activist and animal liberationist: The creation of Seed the Commons 17.5 The normalization of animal agriculture in the food movement and the creation of a false dichotomy 17.6 Peeling back the layers: Regenerative grazing and the reproduction of colonial narratives 17.7 The way forward: Growing a veganic food system 17.8 Conclusion References

371 371 371 373 379 381 385 391 394 395

xContents

18 Alternatives to the global food regime: Steps toward system transformation Helena Norberg-Hodge 18.1 Globalization: The root of the problem 18.2 The costs of globalization 18.3 Shifting toward the local 18.4 Local food for our future 18.5 Making the shift 18.6 Trade treaties 18.7 Subsidies 18.8 Taxation 18.9 Health and safety regulations 18.10 Toward an economics of happiness References

399 399 401 405 406 407 408 409 409 410 410 411

19 Cocreating responsible food and agriculture systems Vandana Shiva 19.1 Two paths to the future 19.2 Rediscovering the living soil 19.3 The industrial path 19.4 Returning to the path of life 19.5 Sowing the seeds of our future References

413

20 Toward inclusive responsibility Laila Kassam and Amir Kassam 20.1 Toward holistic paradigms 20.2 Toward a narrative of abundance 20.3 Toward ecological and multifunctional paradigms of agriculture 20.4 Toward decentralizing power in the food and economic systems 20.5 Toward diets that promote human and planetary health 20.6 Toward powerful social movements and civil society 20.7 Toward an ethical framework for inclusive responsibility References

419

413 414 415 416 417 417

419 420 421 422 424 425 426 429

Index 431

Contributors

Gottlieb Basch  European Conservation Agriculture Federation (ECAF), Brussels, Belgium; Instituto Ciências Agrárias e Ambientais Mediterrânicas, Universidade de Évora, Évora, Portugal Doug Bristor Independent Researcher, Cambridge, United Kingdom Rosa Carbonell-Bojollo  Area of Ecological Production and Natural Resources, IFAPA Centro Alameda del Obispo, Cordoba, Spain Robert Chambers  Institute of Development Studies, Brighton, Sussex, United Kingdom Jesús A. Gil-Ribes  Department of Rural Engineering, School of Agricultural and Forestry Engineering, University of Cordoba; Asociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV), Cordoba, Spain Manuel R. Gomez-Ariza Asociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV), Cordoba, Spain Emilio J. Gonzalez-Sanchez Department of Rural Engineering, School of Agricultural and Forestry Engineering, University of Cordoba; European Conservation Agriculture Federation (ECAF), Brussels, Belgium; Asociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV), Cordoba, Spain Hans R. Herren Millennium Institute, Washington, DC, United States David Jenkins Department of Nutritional Sciences, Faculty of Medicine; Department of Medicine, Faculty of Medicine, University of Toronto; Clinical Nutrition and Risk Factor Modification Center; Li Ka Shing Knowledge Institute; Division of Endocrinology and Metabolism, St. Michael’s Hospital, Toronto, ON, Canada Robert C. Jones  California State University, Dominguez Hills, Carson, CA, United States Tony Juniper  Natural England, London; Institute for Sustainability Leadership, University of Cambridge, Cambridge, United Kingdom

xiiContributors

Amir Kassam University of Reading, Reading, United Kingdom Laila Kassam Animal Think Tank, Lancaster, United Kingdom Shireen Kassam King’s College Hospital, London; Winchester University, Winchester, United Kingdom Zahra Kassam Stronach Regional Cancer Centre and the Princess Margaret Cancer Centre; Department of Radiation Oncology, University of Toronto, Toronto, ON, Canada Mark Langan Newcastle University, Newcastle upon Tyne, United Kingdom Jonathan Latham The Bioscience Resource Project, Ithaca, NY, United States Jim Mason Independent Researcher, Washington, DC, United States Philip McMichael  Department of Global Development, Cornell University, Ithaca, NY, United States David R. Montgomery  Department of Earth and Space Sciences, University of Washington, Seattle, WA, United States Manuel Moreno-Garcia  Area of Ecological Production and Natural Resources, IFAPA Centro Alameda del Obispo, Cordoba, Spain Patrick Mulvany Centre for Agroecology, Water and Resilience, Coventry University, Coventry, United Kingdom Nassim Nobari Seed the Commons, San Francisco, CA, United States Helena Norberg-Hodge  Energy and Resources Group, University of California, Berkeley, CA, United States Rafaela Ordoñez-Fernandez  Asociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV); Area of Ecological Production and Natural Resources, IFAPA Centro Alameda del Obispo, Cordoba, Spain Sophia Price Leeds Beckett University, Leeds, United Kingdom Rupert Sheldrake Schumacher College, Dartington, Devon, United Kingdom Vandana Shiva  Research Foundation for Science Technology and Ecology and Navdanya, New Delhi, India

Contributorsxiii

Paula Trivino-Tarradas  Department of Rural Engineering, School of Agricultural and Forestry Engineering, University of Cordoba, Cordoba, Spain Colin Tudge Oxford Real Farming Conference and the College for Real Farming and Food Culture, Oxford, United Kingdom Oscar Veroz-Gonzalez  Asociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV), Cordoba, Spain Allison K Wilson The Bioscience Resource Project, Ithaca, NY, United States

This page intentionally left blank

Gratitude and acknowledgments

It has taken us both many combined years to start joining the dots between the multiple crises the world is facing. To ask why and how these connected crises came to be and to start finding possible ways forward that address root causes rather than mere symptoms. These years have involved much searching, critical reflection, and learning from farmers, colleagues, and civil society institutions. We continue to learn from each other, from a vast foundation of different knowledges, expertise and wisdom, as well as from nature and our animal kin. Throughout these years our sense of purpose, responsibility, and agency for contributing to the creation of a better world has grown steadily. This sense grounds our commitment to finding more sustainable and just ways forward, specifically in the areas of food and agriculture but also in the wider systems in which they are embedded and entangled. Given the complexity of the world and the crises we are facing, we consider it a blessing to have acquired our own glimpses of alternative ways forward. We feel especially fortunate to have been granted the opportunity to explore and synthesize our ideas, as a father and daughter team, in the process of developing this book. We are so grateful to have had our explorations and ideas challenged, strengthened, and developed by the authors who have contributed to this book and offered their own compelling visions of ways forward toward a better future for all. It is impossible to fully appreciate the generations of intellectual energies and practical action on which this book builds. We express our very special and sincere gratitude to all those, past and present, who have worked to make our global food and agriculture system more transparent, responsible, inclusive, just, and sustainable in their diverse ways. We also acknowledge our deep appreciation for those involved in the various ­progressive social and environmental justice movements around the world. These movements, which encompass a range of strategies from personal change to wider structural transformation, are a great source of stimulation and hope for us. These movements, and the depth of feeling they reflect, cannot and must not be ignored. We would like to recognize the enormous assistance and cooperation extended to us in the preparation of this book from a large number of colleagues and organizations. We extend our sincere thanks to the publisher Elsevier for inviting us to edit this book, especially to Nancy Maragioglio, who initially helped to conceive this book project and define the scope of the book. We also thank Michael Lutz, Emerald Li, and Vignesh Tamilselvvan from Elsevier, who actively managed different stages of the project and guided us with regards to the various agreements, timings, formats, and system management.

xvi

Gratitude and acknowledgments

Most importantly, we would like to express our deepest gratitude to every lead author and all the coauthors for their valuable contributions. Each one of you has taught us so much. We feel deeply honored and humbled to have been able to bring your inspired work together in this book. We acknowledge with many thanks the external reviewers of the book proposal and the different chapters, for their help in ensuring the book comprises the best quality material. These reviewers include: Raj Patel, Alain de Janvry, Uma Lele, Tim Wise, Per Pinstrup-Andersen, Nikos Alexandratos, Philip McMichael, Andrew MacMillan, the late Dick Harwood, Hans Gregersen, Norman Uphoff, Francis Shaxson, Thomas Welsh, Zahra Kassam, Shireen Kassam, and Parin Kassam. Finally, we would like to thank our beloved families and loved ones. Your support and encouragement for this book project and for our wider work in the world keeps us full of hope, love, and gratitude. It strengthens our conviction that a fairer, kinder, and more sustainable world is possible and that we all have our part to play.

Introduction Laila Kassama and Amir Kassamb a Animal Think Tank, Lancaster, United Kingdom b University of Reading, Reading, United Kingdom

We are living in unprecedented times. Never before have we faced such significant threats to our own and other species’ existence. These threats are of our own making. Since 1970, human activities have wiped out 60% of wildlife populations (Barrett et al., 2018). An estimated 1 million animal and plant species are now threatened with extinction (IPBES, 2019). We have impacted the natural world to such an extent that extinction rates are estimated to be 1000 times higher than they otherwise would be (Pimm et al., 2014). Nearly 90% of our ecosystems are degraded or severely degraded, while only 11% are estimated to be in reasonable condition (MEA, 2005). Human actions are driving the “sixth mass extinction” event, more aptly named by some as the “first mass extermination.” Our “annihilation” of biodiversity is an emergency that is threatening Earth’s life support systems and the ecosystems on which we all depend. This loss of biodiversity and ecosystem function is closely connected with climate breakdown. They both affect and exacerbate each other. For example, while climate breakdown has emerged as a driver of biodiversity loss, this loss is in turn reducing our resilience to the extreme weather events becoming increasingly common with climate breakdown. The risks posed by both are equally serious (IPBES, 2019). It is estimated that global heating of only half a degree beyond 1.5°C will significantly worsen the risks of drought, floods, fires, hurricanes, extreme heat, crop failure, poverty, displacement, and death for hundreds of millions of people (IPCC, 2018). These risks have been present for less industrialized, lower-income countries in the Global South for many years, and will continue to affect them disproportionately, compared to more industrialized, higher-income countries in the Global North. Less industrialized countries tend to be those least responsible for causing the problem, emitting far fewer greenhouse gases per capita over past decades than richer ones. Recent research suggests that global heating has very likely exacerbated global economic inequality, and the economic injustice of climate change has been going on for at least 60 years (Diffenbaugh & Burke, 2019). We are told that we have just over a decade to turn the existential threat of climate breakdown around (IPPC, 2018). Thus we are the last generation that can prevent this catastrophic damage to our planet, to ourselves, and to all those with whom we share the planet. According to some, however, we may already have passed the point of no return and crossed the thresholds for a series of tipping points and are in a state of “planetary emergency” (Lenton et al., 2019). Long-term climate change effects are likely to significantly impact agriculture in various ways, in particular increasing the risk to food security for the world’s most vulnerable people. Tropical and subtropical areas are expected to be the most affected by climate change. It is estimated that the frequency and magnitude of extreme climatic

xviiiIntroduction

events noted earlier, such as storms, floods, droughts, etc., will continue to increase and have serious consequences for agricultural productivity in all regions of the planet (Cline, 2007; González-Sánchez et al., 2017). There is a serious risk of future conflicts over habitable lands and natural resources, and forced human migration is expected to rise. Indeed, some analysts believe that near-term societal collapse is now inevitable (Bendell, 2018). According to Bendell (2018, p. 26): “Recent research suggests that human societies will experience disruptions to their basic functioning within less than ten years due to climate stress. Such disruptions include increased levels of malnutrition, starvation, disease, civil conflict and war – and will not avoid affluent nations.” Climate change is one of the four planetary boundaries that human activity has overshot. There are nine such boundaries that regulate the Earth System on which our societies and ecosystems depend. Two of the overshot boundaries—climate change and biosphere integrity—are “core boundaries,” which when significantly altered are likely to “drive the Earth System into a new state,” seriously affecting human wellbeing (Rockström et al., 2009; Steffen et al., 2015). Our destruction of nature takes many interrelated forms. We have lost half of the topsoil on the planet in the last 150 years (WWF, n.d.-a) and are losing 24 billion tonnes every year (UNCCD, 2017). Topsoil is eroding 13–40 times faster than nature can replenish it and if current rates of land degradation continue, all of the world’s topsoil and current agricultural croplands could be gone or severely eroded within 60 years (Montgomery, 2007). The erosion of this topsoil in runoff water and in wind, in addition to soil sediments, also carries with it agrochemicals, including fertilizers and pesticides, and microorganisms polluting water systems and the atmosphere (Juniper, 2015). Forests are also disappearing at an alarming rate. It is estimated that we have cut down 46% of trees since the start of human civilization (Crowther et al., 2015) and destroyed about 17% of the Amazonian rainforest over the past 50 years (WWF, n.d.-b). These estimates, however, do not include the destruction caused by the recent fires in the Amazon nor those in the Congo Basin, the United States, Australia, and elsewhere, the full devastation of which is still unknown. We kill an ever-increasing number of land animals for food and other products such as wool, fur, and leather. In 1961, we slaughtered around 7 billion land animals for food, and we are currently killing approximately 70 billion land animals per year (not including male chicks killed in the egg industry) (Sanders, 2018). We also kill around 80 billion farmed fish every year (Mood & Brooke, 2010). Life in the sea is also being destroyed by human activities through ocean acidification and fishing. We kill between 1 and 3 trillion wild aquatic animals every year for food (Mood & Brooke, 2010). It is estimated that if we keep fishing at the current pace the oceans will be empty of fish by 2048 (Worm et al., 2006). Given humans’ disproportionate impact on the Earth and all her inhabitants, this period in history (or new geological epoch) is increasingly being described as the “Anthropocene.” Others call it the “Capitalocene” to highlight the driving force of capital accumulation based on the creation of “cheap nature,” starting in the 15th century, as opposed to humans (Anthropos) and human nature in the crises we face (Moore, 2015). Rather, it is a certain kind of destructive human activity linked to

Introduction 

xix

Western ­capitalism, with its roots in European colonization, including of the Americas, Africa, and Asia, which is responsible. Some argue it should be called the “Necrocene” or “New Death” to reflect how endless capital accumulation “necrotizes the entire planet…devours all life…[and] leaves in its wake the disappearance of species, languages, cultures, and peoples” (McBrien, 2016, p. 97). In this way the “Necrocene” is conceived of as the Capitalocene’s “shadow double,” as capital accumulation and extinction are the same process (McBrien, 2016). At the same time, however, there has also been incredible material advancement in many fields of human endeavor, especially during the last 100 years. Socioeconomic indicators show that we have greatly improved in terms of literacy, longevity, and income. For example, the world literacy rate in 1917 was 23%, while today it stands at over 86%. Since 1900 the global average life expectancy has more than doubled and is now above 70 years. Between 1870 and 2016 the average global GDP per capita was estimated to have increased over 10-fold (from around 1263 to 14,574 international-$, respectively) (Our World in Data, 2020). As noted in the IPBES (2019) report, though, the increase in the global production of consumer goods that this growth in GDP per capita represents and the decline in all the other contributions of nature to people are directly related: The world is increasingly managed to accelerate the flow of material contributions from nature to keep up with rising demand. Since 1970, global population has doubled…, per capita consumption has increased by 45%, the value of global economic activity as measured in gross domestic product (GDP) has increased by > 300%…, global trade has increased by ~ 900%…, and the extraction of living materials from nature has increased by > 200%. (p. 2)

In addition, these material advancements have not only come at the expense of nature, but have been accompanied by and are related to ever-increasing violence, warfare, land and resource grabbing, and human and nonhuman displacement, exploitation, and killing. Overall, then, some would argue there is not much good news, except perhaps for the 26 billionaires who are estimated by Oxfam to have the same wealth as the poorest half of the planet (around 3.8 billion people). These billionaires, along with over 2000 others, saw their wealth continue to increase in 2018 by 12%, while the poorest half saw their wealth fall by almost the same proportion (11%) (Lawson et al., 2019). While these estimates may be controversial, the trend of increasing inequality is less so, and the benefits of economic growth are estimated to be captured disproportionately by the rich (between 1980 and 2016 the poorest 50% of the population had 12 cents in every dollar of global income growth, while the top 1% had 27 cents of every dollar) (Lawson et al., 2019). Despite the good news narratives of the success of the Millennium Development Goals (MDGs) in halving poverty and hunger since 1990 (MDG, 2015), independent analysis suggests that “in reality, around four billion people remain in poverty today, and around two billion remain hungry – more than ever before in history, and between two and four times what the UN would have us believe” (Hickel, 2016).

xxIntroduction

This is despite the fact that the world produces enough food to feed more than 10 billion people (Holt-Giménez, Shattuck, Altieri, Herren, & Gliessman, 2012). While hunger and malnutrition are on the rise, so is the epidemic of obesity, now recognized as a pandemic, which is not just confined to industrialized countries. In 2016, nearly 2 billion adults were overweight and 650 million were obese (WHO, 2018). According to the Lancet Commission on Obesity, the syndemic of obesity, undernutrition, and climate change represents the most important and urgent challenge for humans, the environment, and our planet (Swinburn et al., 2019).

The food and agriculture system: Part of the problem and solution Ecological and climate emergencies, environmental degradation, growing inequality, poverty, hunger, and obesity are not separate issues. They are interconnected, and the role of the food and agriculture system is central to them all. The food and agriculture system refers to all the elements (e.g., environment, humans, nonhuman animals, inputs, processes, infrastructures, and institutions) and activities related to the production, processing, distribution, preparation, and consumption of food, and the outputs of these activities, including socioeconomic and environmental outcomes. It has three core elements: (1) food supply chains; (2) food environments; and (3) consumer behavior (CFS-HLPE, 2019). According to the IPBES (2019) report, land use change (mainly land conversion for crop production, raising farmed animals, and plantations) is the strongest driver of our destruction of nature. Since WWI, the food and agriculture system has been driven increasingly by the corporate sector that developed the now dominant industrial Green Revolution paradigm of agriculture (Chapter  10). This model of agriculture production is based on intensive tillage, modern seeds, excessive use of agrochemicals and fossil fuel in monocultural, or poorly diversified cropping systems. These systems have very poor agricultural biodiversity and are not underpinned for ecological sustainability. They have suboptimal productivity and input use efficiency, deliver hardly any ecosystem societal services, and are poor in adaptability and resilience toward biotic and abiotic stresses and shocks, including those arising from climate change. They are large emitters of greenhouse gases and contribute to climate change instead of mitigating it (Brisson et al., 2010; Kassam, 2020; Kassam et  al., 2013, 2017; Kassam, Friedrich, Shaxson, & Pretty, 2009). Much of the destruction discussed earlier is directly related to industrial agriculture and the corporate food regime that is driving it (Chapter 3). It is estimated that around 60% of wildlife loss is related to agriculture and the industrial methods we use (Barrett et al., 2018). Conventional industrial agriculture has pushed 80% of insects and 93% of crop diversity to extinction (Shiva et al., 2017). Land use (including the effects of soil degradation and deforestation) is estimated to contribute to between a quarter and a third of global greenhouse gas emissions (IPCC, 2019). The food system, including feed, agrochemical manufacture, processing, and transportation, is estimated to account for over 30% of total annual global greenhouse gas emissions (Bajželj, Allwood, & Cullen, 2013; Herrero et  al., 2016; Vermeulen,

Introduction 

xxi

Campbell, & Ingram, 2012) over 60% of which come from animal agriculture. Animal agriculture is estimated to produce more greenhouse gas emissions than all forms of transportation combined (Bailey, Froggatt, & Wellesley, 2014). It is considerably more than the emissions produced by the world’s largest national economy, the United States. According to FAO (2013), farmed animals are responsible for 14.5% of total greenhouse gas emissions. Animal agriculture also accounts for at least half of all food-related greenhouse gas emissions (Herrero et al., 2016; Vermeulen et al., 2012). Worldwide, the top 20 meat and dairy corporations produce more greenhouse gas emissions than the whole of Germany (Heinrich Böll Stiftung, GRAIN, & Institute for Agriculture, & Trade Policy, 2017). Conventional tillage-based agriculture is responsible for 75% of the destruction of our soils (Montgomery, 2007). In the last 70 years it has led to the abandonment of approximately half a billion hectares of land (at an annual rate of 7–12 million hectares) due to soil erosion and degradation and loss of soil health and ecosystem functions (Gibbs & Salmon, 2015; Juniper, 2015; MEA, 2005; Montgomery, 2007; Nkonya, Mirzabaev, & von Braun, 2016). Some 25 billion tonnes of topsoil are lost every year through erosion because of intensive tillage farming (Nkonya et al., 2016) (Chapter 11). The runoff of chemical fertilizers and sewage from animal agriculture has created over 400 aquatic dead zones all over the world (Diaz & Rosenberg, 2008) and contributes to the pollution of waterways and groundwater. Agriculture is also responsible for around 80% of deforestation worldwide, of which animal agriculture (including production of animal feed) is a significant driver (Campbell et al., 2017; Hosonuma et al., 2012; Kissinger, Herold, & De Sy, 2012). In the Amazon, 70%–80% of deforested land has been converted into pasture for grazing, with much of the remaining land used to grow animal feed such as soy (Machovina & Feeley, 2014; Steinfeld et al., 2006). Animal agriculture is the most significant driver of habitat loss on the planet (Machovina, Feeley, & Ripple, 2015) and one of the biggest drivers of global biodiversity loss (Steinfeld et al., 2006). The production of meat, farmed fishes, eggs, and dairy uses around 83% of farmland while only providing 18% of calories (Poore & Nemecek, 2018). In addition, intensive industrial animal agriculture, along with the associated habitat destruction for feed production and the wildlife trade, has led to the increasing spread of novel zoonotic diseases (Morse et al., 2012). Continuation of industrial animal agriculture means further pandemics are inevitable (Ceballos, Ehrlich, & Raven, 2020). It is not, however, agricultural land use change alone that has been driving the destruction of nature, particularly since WWII. These changes, along with the industrial agriculture production systems that have developed with them, are servicing consumer demand for food products and diets that are leading not only to environmental destruction but to negative health impacts such as increased obesity, noncommunicable diseases such as cancer, heart diseases, diabetes, and general ill-health. Thus our food consumption choices also play a major role in driving land use change as well as these environmental and health impacts (Poore & Nemecek, 2018). Even so, these impacts cannot be attributed purely to consumer choices. The corporate food regime, facilitated by governments and international institutions, plays a central role in influencing consumer choices in various ways, and in driving these

xxiiIntroduction

environmental and health impacts (Chapter 3). For example, the food system is central to both the high rates of chronic hunger and the obesity pandemic, due to its inability to provide healthy and nutritious food for all, as well as its promotion of Western diet patterns high in processed food, sugar, salt, oil, and animal-based foods and low in nutritious whole plant foods (Willett et al., 2019) (Chapter 15). In turn, the increased consumption of animal products promoted by industry and society as being necessary for both human health and agricultural sustainability has led to the majority of food grain and land being allocated for animal agriculture, much of which is under intensive industrial systems. These systems are destroying forests and biodiversity, contributing to greenhouse gas emissions and climate and ecological breakdown, in some areas taking land and livelihoods away from marginalized and indigenous populations and making them homeless, all the while causing an immense amount of suffering to nonhuman animals. Ultimately, then, the pressures on the natural resource base are closely related to the size and nature of societal demand for food and feed, and how these are produced, processed, distributed, utilized, and wasted, nationally and internationally. At the same time the main political and economic drivers of the prevailing food and feed systems continue to push the agriculture system further down the industrial, intrusive, technology-oriented path. A path that has become environmentally degrading and economically unsustainable rather than one that is agroecologically oriented and based on regenerative and sustainable agroecosystem processes and outcomes (Chapters 10–12). Aside from its environmental and health impacts, agriculture is also directly related to poverty and hunger given that the vast majority of those living in poverty live in rural areas and the majority work in agriculture (Castaneda et al., 2016). Most of the world’s over 570 million farms are small and family-run farms, estimated to provide up to 80% of food in less industrialized countries on 25% of agricultural land (FAO, 2014; FAO-IFAD, 2019). Despite the fact that poor peasant farmers produce most of the world’s food, most of them are going hungry. Given that the current food and agriculture system is such a significant driver of the crises we are facing, it stands to reason that transforming it into a sustainable and democratic system is also a potentially significant part of the solution (Chapters 7 and 8). In our view, the huge role of the corporate food regime in supporting and facilitating the global neoliberal corporate capitalist economic system, which is the ultimate driver of so many of these crises, suggests its transformation holds huge potential for overall systemic transformation. Sustainable food and agriculture systems have been defined as those that “ensure food security and nutrition for all in such a way that the economic, social and environmental bases to generate food security and nutrition of future generations are not compromised” (CFS-HLPE, 2019, p. 28). Without an ecologically sustainable agriculture production paradigm at the base, such a system would not be possible. While the industrial Green Revolution production paradigm has been the dominant paradigm since WWII, several alternative, more sustainable paradigms have emerged. These alternative paradigms have developed partly in response to the destruction and degradation that industrial agriculture has been causing, as well as its unsuitability for the

Introduction 

xxiii

needs of the vast majority of resource poor small-scale farmers (Chapter 10). These alternative paradigms are operating and contributing to the food and agriculture system globally. Whereas the industrial Green Revolution paradigm is running out of steam, these alternative paradigms are gaining strength. But the production paradigm cannot be transformed without wider structural changes. Key drivers of the industrial agriculture paradigm and the destruction it is contributing to are economic globalization driven by “free trade” agreements and policies biased toward the large-scale and global (Chapter  18) and the associated corporate takeover of the global food and agriculture system (Chapters  3 and 14). Globalization, capitalism, and the corporate food regime are incompatible with sustainable development.

Mainstream responses and narratives The search for sustainable development has been an ongoing objective of the United Nations (UN), its specialist agencies, and many international and national development organizations, governments, and donors for decades. In 1972 the UN held a conference on the Human Environment that put environmental issues on the international agenda for the first time. In 1983, the UN established the World Commission on Environment and Development, chaired by Gro Bruntland, the former Norwegian Prime Minister. The Bruntland Commission aimed to reexamine the challenges posed by environmental and natural resource degradation resulting from global economic development, including agricultural development. It sought to develop action proposals to deal with these challenges and encourage countries to work toward sustainable development together. In 1987 the Commission presented its report (also known as the Bruntland Report) entitled “Our Common Future,” which included a definition of sustainable developmenta that is still often used today. Despite these initiatives, population increase, urbanization and changing lifestyles, increasing incomes, resource consumption and wastage, deforestation, and the industrial Green Revolution agriculture paradigm had caused such worrying levels of environmental and natural resource degradation globally that in 1992 the UN held a conference on Environment and Development, also known as the Earth Summit, in Rio de Janeiro, Brazil. Several conventions and treaties were established, including conventions on climate change, biodiversity, and degradation, along with Agenda 21, the “Rio Declaration on Environment and Development” and the UN Commission on Sustainable Development, to address these challenges. More recent global initiatives include the MDGs established in 2000 and the Sustainable Development Goals (SDGs) set in 2015 to be achieved by 2030. Yet despite these, and so many other, initiatives to address the challenges of environmental destruction and inequity generated by “development” (including agricultural development) nationally and internationally, things continue to get worse. As a

“development that meets the needs of the present without compromising the ability of future generations to meet their own needs.”

xxivIntroduction

we have seen, climate and ecological breakdown, overuse of our resources, poverty, and hunger continue unabated (Bruntdland, 1987; Earth Summit, 1992; Foresight, 2011; IAASTD, 2009; MEA, 2005; Montgomery, 2007; SDG, 2018). One reason for this is that many of these initiatives fail to explicitly identify the root causes of these ­crises, and so fail to identify real and effective solutions. Another reason is the view that mainstream “development” is actually just a continuation of the colonial project. Some argue that the discourse of sustainable development and food security is being used by corporations to access valuable resources and that “development” is a form of neocolonialism (Chapter 4). The UN’s SDGs are the current blueprint to address the global challenges we face, including those related to poverty, inequality, climate change, environmental degradation, peace, and justice. The achievement of all of the UN’s SDGs, like the MDGs before them, is dependent on nature and its contributions to people, either directly or indirectly (IPBES, 2019). Given the central role of the food and agriculture system, especially the conventional industrial agriculture paradigm, in the destruction of nature, and as a result of the lack of progress being made, over the last decade there have been urgent and increasing calls to move away from “business as usual” and the need for a radical transformation of the food and agriculture system (CFS-HLPE, 2019; IAASTD, 2009; IPBES, 2019; WDR, 2008). However, the response of governments, mainstream development organizations, philanthropic actors, and corporations appears to offer more of the same, based on the same paradigms of industrial agriculture and neoliberal capitalist economics that have brought us here in the first place. These reductionist paradigms have given rise to several myths and narratives that are sustaining and strengthening the current system. These myths include: hunger is due to food scarcity and overpopulation (Chapter 5), only industrial agriculture and genetically modified organisms (GMOs) can feed the world (Chapters 13 and 14); organic and ecological farming cannot feed a hungry world; the free market can end hunger; free trade is the answer; foreign aid is the best way we can help the hungry; and power is too concentrated for real change (Chapter 18) (Lappé & Collins, 2015). The myth that all the others rely on is the myth of scarcity or lack (not just scarcity as it relates to hunger and food production but also as it relates to the myth of scarce resources that our capitalist economic system relies on and artificially creates) (Chapter 17). If the scarcity myth were true, however, how can we be producing enough food to feed more than 10 billion people while over 2 billion go hungry (Holt-Giménez et al., 2012)? These myths and narratives completely overlook the fact that the multiple crises we are facing are deeply interconnected. These interconnections, along with their historical, ethical, economic, social, cultural, political, and structural drivers, need to be better understood for meaningful system-wide transformation to be possible. However, while our capitalist economic system is one of the drivers of the food and agriculture system and so many of these interconnected crises, this system itself reflects a wider paradigm or worldview based on a set of beliefs, values, and ethics. This paradigm or worldview shapes so many of our systems, including our scientific, agricultural, and knowledge systems (Chapters 1, 2, 9, and 16). Thus, in our view, the challenges we face have deeper roots. They are rooted in our separation and disconnection from and domination of nature (understood as the web of life and not just the environment devoid of humans)

Introduction 

xxv

and thus of ourselves and each other. It is this disconnection that allows us to ignore, oppress, exploit, and violate the rights of others and destroy nature (Chapters 2 and 6). In our view, then, addressing such vast levels of injustice toward and suffering of humans, other animals, and nature from the structures of oppression and violence is ultimately an issue of values and ethics. Responsible food and agriculture systems cannot be driven and sustained by profit or based on selective science. They must be shaped by ethics, equity, quality of life, and informed engagement of civil society that is connected both locally and internationally. In our view, then, we need transformation at every level and in every area of the food and agriculture system, the economic system in which it is embedded, and the ethics, worldview, and beliefs that underpin the whole paradigm we are operating in.

Scope of the book Given the central role of the food and agriculture system in driving so many of the connected ecological, social, and economic threats and challenges we currently face, this book is an attempt to “rethink” (including to review, reassess, and reimagine) the current food and agriculture system and the narrow paradigm in which it operates. This rethinking requires exploring and uncovering the historical, ethical, economic, social, cultural, political, and structural drivers and root causes of unsustainability, degradation of the agricultural environment, destruction of nature, shortcomings in science and knowledge systems, inequality, hunger and food insecurity, and disharmony. Thus this rethinking also entails reviewing past and present efforts toward “sustainable development,” including food security and production, and reassessing whether these efforts have been and/or are being implemented with adequate cultural responsibility, acceptable societal and environmental costs, and optimal engagement to secure sustainability, equity, and justice for all throughout the whole system. This rethinking ultimately demands reimagining our food and agriculture system and the efforts required to co-create a new one. This involves highlighting the many ways that farmers and their communities, civil society groups, social movements, development experts, scientists, and other relevant actors have been raising awareness of these issues, implementing solutions and forging “new ways forward,” for example, toward alternative paradigms of agriculture, natural resource management, and human nutrition, which are more sustainable and just. This book brings together some of the most experienced and forward-thinking academics, activists, development professionals, and practitioners in the field to rigorously question and “rethink” the current food and agriculture system, along with the science and political economy on which it is based. This rethinking provides the foundation for looking forward to how the food and agriculture system can be guided by the concept of “inclusive responsibility” and what such a system might look like (Chapter 20). An inclusively responsible food and agriculture system would encourage society to focus on agroecological sustainability as an integral part of overall ecosystem sustainability based on planetary boundaries. Such a system would place importance on quality of life, pluralism, equity, and justice for all. It would emphasize

xxviIntroduction

the health, wellbeing, sovereignty, dignity, and rights of farmers, consumers, and all other stakeholders, as well as of nonhuman animals and the natural world. The concept of “inclusive responsibility” is ultimately based on an understanding of the interconnectedness of nature and the place and responsibility of human society within it.

Structure of the book Each chapter looks critically at one or more aspects of the food and agriculture system and highlights new ways forward. Some chapters are more heavily weighted toward understanding the past and present actions of human societies and institutions and their consequences in terms of the direction of development of the food and agriculture system and the associated challenges described above. Other chapters are more focused on highlighting new or alternative ways forward, what they might offer, and how they could be achieved. The focus of Chapters 1–3 is on uncovering some of the key historical and philosophical roots, worldview, and ideologies that underpin our current food and agriculture system. As noted earlier, without understanding these roots we cannot really understand why and how we are in this situation, nor can we develop real or effective solutions. Chapter 1 examines the beliefs and worldview underlying the current conventional scientific paradigm within which agriculture systems operate. It explores the implications of the current paradigm, based on a materialist philosophy of nature, and a more holistic paradigm that views nature as organic and alive, for agricultural research and development. Chapter 2 explores the historical roots of our nature-dominating and anthropocentric worldview. It looks at how and when we became so alienated from nature in the first place and traces a broad range of consequences of this disconnection on humans, other animals, and the planet. Chapter 3 analyzes the political economy of the global food and agriculture system, framed through the lens of a succession of international food regimes from the mid-19th century. It addresses the historical conditions, including colonialism and capitalism, underlying the present crisis of the global food and agriculture system and its contributions to the combined climate and ecological emergency. Having explored the roots of our present food and agriculture system and the climate, ecological, and food security crises it is contributing to, Chapters  4 and 5 critically examine some examples of the narratives and solutions being offered by mainstream development institutions. These narratives are often accepted uncritically and used by the corporate sector to justify continued accumulation and expansion. Chapter 4 analyzes the example of the New Alliance for Food Security and Nutrition and the wider SDGs through the lens of neocolonialism. It also takes a gendered perspective to address the impact on women of the neocolonial relations facilitated and maintained through aid giving. Chapter 5 critically examines the narrative of a potential food crisis and the challenge of “feeding the 10 billion” that has been predicted by the models of the UN’s Food and Agriculture Organization and other international organizations. This prediction and crisis narrative are often used by corporate ­agribusiness and mainstream development organizations to justify the expansion and intensification of industrial agriculture, including the development and use of GMOs.

Introduction 

xxvii

Given the emphasis of “inclusive responsibility” on the interconnectedness of nature, and food and agriculture systems that are sustainable and just for all, Chapter 6 focuses on the issue of ethics. It examines some of the key concepts and debates in the areas of environmental, food, and animal ethics and challenges our anthropocentric ethics and worldview. This provides an ethical framing for many of the chapters that follow. Chapters 7 and 8 present an overview of the food and agriculture system. They help us better understand the magnitude of and ways in which this system, industrial agriculture in particular, has impacted, and continues to impact, the planet. They explore the environmental, health, and quality of life impacts of this system, and the connections between them. They both suggest ways forward for transforming the food and agriculture system. Chapter 7 suggests ways forward for public policy, private sector companies, and consumers. Chapter 8 outlines how the food and agriculture system can provide solutions to reduce climate change and improve the general health of ­humans, animals, and the planet. Following these overviews of the food and agriculture system, Chapters  9–12 focus specifically on the agriculture production system. Chapter  9 presents two models of agriculture—industrial versus agroecological—based on two paradigms of science (as explored in Chapter  1). Chapter  10 examines various paradigms of agriculture, including the history and impacts of the current dominant industrial Green Revolution agriculture paradigm. It also reviews and analyzes the ecological sustainability of some of the key alternative paradigms that have emerged and evolved alongside and/or in reaction to the dominant paradigm— Organic Agriculture, Agroecology, Regenerative Agriculture, and Conservation Agriculture. It suggests ways for each of these alternative paradigms to increase its ecological sustainability in the context of inclusively responsible food and agriculture systems. Chapter 11 focuses on the huge problem of soil degradation from conventional agriculture. It explores the potential of the alternative paradigm of Conservation Agriculture to help reverse this problem that has plagued societies throughout history and thus contribute to enhancing soil health for sustainable food production. Chapter 12 addresses the issue of conventional agriculture being a greenhouse gas emitter and vulnerable to climate change. It examines the various ways the alternative paradigm of Conservation Agriculture makes it possible to adapt to and mitigate climate change. Chapters 13 and 14 review and suggest ways forward for seed systems and agricultural biodiversity providing farmer-centered alternatives to the GMO and gene-­ edited solutions of big agribusiness. Chapter 13 assesses the sustainability impacts of GM crops. It summarizes the sustainability impacts of herbicide tolerant and Bacillus thuringiensis pesticidal GM crops and provides an overview of GM crop development, with a focus on the problem of unintended traits in commercial GM crops, illustrated by a case study of Golden Rice. The chapter also discusses how plant breeders can best support and promote sustainable agriculture, and thus help create sustainable food systems. Chapter 14 examines the history, scope, functioning, and governance of “agricultural biodiversity,” a vital subset of biodiversity important for agricultural production.

xxviiiIntroduction

It describes the competing paradigms of agricultural biodiversity management, especially of seeds, and the threats to their diversity. The chapter highlights the social and environmental imperative for sustaining the heterogeneous seeds, which underpin ­resilient agroecological production systems that are essential for equitable, sustainable, and responsible food systems. Chapters  15–18 move beyond the agriculture production system to explore new ways forward in other important areas—drivers and supporting aspects—of inclusive and responsible food and agriculture systems. These include healthy and sustainable diets, knowledge systems, social movements, and government policies. Chapter  15 highlights the important role of healthy diets in guiding responsible food systems. It reviews data on the impact of diet on health and illness, explores the relationship between diet and human and planetary health, and proposes a dietary pattern that optimizes human and planetary health. Chapter 16 explores the types of knowledge systems, attitudes, and behaviors we need to employ to support inclusively responsible food and agriculture systems. It shows how knowledge systems often have to confront and transform powerful interests and the embedded Newtonian professional paradigm. It suggests “canons of rigor” for learning about the complex and diverse conditions of farmers and argues for a new balance and optimality between local and dominant knowledge systems. Chapter 17 examines the role of grassroots social movements in the radical transformation of food and agriculture systems. It argues that this transformation has to come from the grassroots as the solutions promoted by corporate, governmental, and philanthropic actors tend to maintain the status quo. Chapter 18 looks at how our current food system has been produced by globalization, driven by “free trade” agreements, subsidies, and tax policies biased toward the large-scale and global. It suggests a systemic solution to globalization and steps toward transformation, highlighting new ways in which people are building prosperous local economies that reflect the desire for love and connection. Chapter 19 presents the essence of the two paths forward—the path of life or the path of death. The path of life is based on co-creation with nature and her principles of diversity, giving back and sharing the Earth’s gifts. The path of death is the industrial path based on fossil fuels and poisons, of war against the earth and her biodiversity. The concluding chapter distills some of the key themes and ways forward explored in the preceding chapters. It uses these themes to inform the further elaboration of the concept of “inclusive responsibility” in relation to food and agriculture and offers a possible vision for an inclusively responsible food and agriculture system. We would have liked the book to be comprehensive and cover every possible aspect of the food and agriculture system in depth, but it is such a vast area that this has not been possible. We have limited the scope to land-based food and agriculture production. Within that scope some obvious gaps we would have liked to have filled relate to: the causes of and ways forward for dealing with food waste; ecological restoration such as rewilding; structural solutions such as land justice, which are fundamental to transforming our food and agriculture system and alleviating poverty and hunger; the need for a shift to degrowth strategies; and the social and ecological impact of animal agriculture. Some other areas we would have liked to have explored in greater depth include: an examination of the capitalist nature of the food and agriculture system,

Introduction 

xxix

its colonial roots, and the wider neoliberal corporate capitalist economic system in which it sits; the role of international development and related institutions (such as the UN and its agencies, the World Bank and regional banks, the International Monetary Fund, the CGIAR system, and bilateral and multilateral aid communities, including foundations—both corporate and noncorporate—and “philanthro-capitalists”), in both contributing to and addressing these global crises; the role of civil society, including social and producer organizations, social movement organizations such as La Via Campesina and Extinction Rebellion, and engaged citizenship in building the social and political power needed to resist the corporate takeover of our food and agriculture system and demand transformative structural change; the role of local businesses in supporting and sustaining food sovereignty; and the role of public and private sector institutions and policy in creating an enabling environment for sustainability and progress. A chapter assessing the prospects and consequences of lab-grown food would also have been appropriate to include. Finally, the book would have greatly benefited from more contributions from the Global South.

References Bailey, R., Froggatt, A., & Wellesley, L. (2014). Livestock–Climate Change’s forgotten sector global public opinion on meat and dairy consumption. Energy, Environment, and Resources-Chatham House, https://doi.org/10.1093/acprof:oso/9780199655786.003.0034. Bajželj, B., Allwood, J. M., & Cullen, J. M. (2013). Designing climate change mitigation plans that add up. Environmental Science & Technology, 47, 8062–8069. Barrett, M., Belward, A., Bladen, S., Breeze, T., Burgess, N., Butchart, S., … de Carlo, G. (2018). Living planet report 2018: Aiming higher. Switzerland: WWF. Bendell, J. (2018). Deep adaptation: A map for navigating climate tragedy IFLA Occasional Paper 2, https://www.lifeworth.com/deepadaptation.pdf. Brisson, N., Gate, P., Gouache, D., Charmet, G., Oury, F. X., & Huard, F. (2010). Why are wheat yields stagnating in Europe? A comprehensive data analysis for France. Field Crops Research, 119, 201–212. Bruntdland, G. (1987). Our common future. World Commission on Environment and DevelopmentOxford University Press. Campbell, B. M., Beare, D. J., Bennett, E. M., Hall-Spencer, J. M., Ingram, J. S. I., Jaramillo, F., … Shindell, D. (2017). Agriculture production as a major driver of the Earth system exceeding planetary boundaries. Ecology and Society, 22(4), 8. Castaneda, A., Doan, D., Newhouse, D., Nguyen, M. C., Uematsu, H., & Azevedo, J. P. (2016). Who are the poor in the developing world? Policy Research Working Paper; No. 7844. World Bank, Washington, DC. © World Bank, https://openknowledge.worldbank.org/ handle/10986/25161. Ceballos, G., Ehrlich, P. R., & Raven, P. H. (2020). Vertebrates on the brink as indicators of biological annihilation and the sixth mass extinction. Proceedings of the National Academy of Sciences. Published online June 1st, 2020. https://doi.org/https://doi.org/10.1073/ pnas.1922686117. CFS-HLPE. (2019). Agroecological and other innovative approaches for sustainable agriculture and food systems that enhance food security and nutrition, http://www.fao.org/3/ ca5602en/ca5602en.pdf.

xxxIntroduction

Cline, W. (2007). Global warming and agricultural impact estimated by country. Columbia University Press. Crowther, T. W., Glick, H. B., Covey, K. R., Bettigole, C., Maynard, D. S., Thomas, S. M., … Tuanmu, M. N. (2015). Mapping tree density at a global scale. Nature, 525(7568), 201. Diaz, R. J., & Rosenberg, R. (2008). Spreading dead zones and consequences for marine ecosystems. Science, 321(5891), 926–929. Diffenbaugh, N. S., & Burke, M. (2019). Global warming has increased global economic inequality. Proceedings of the National Academy of Sciences, 116(20), 9808–9813. Earth Summit. (1992). United Nations Conference on Environment and Development (UNCED), Earth Summit. Rio de Janeiro, Brazil 3-14 June 1992, https://sustainabledevelopment. un.org/milestones/unced. FAO. (2013). Tackling climate through livestock: A global assessment of emissions and mitigation opportunities. Rome: FAO. FAO. (2014). The state of food and agriculture: Innovation in family farming. Rome: FAO. FAO-IFAD. (2019). United Nations decade of family farming 2019-2028. Rome: Global Action Plan. Foresight. (2011). The future of food and farming. Final Project ReportLondon: The Government Office for Science. Gibbs, H. K., & Salmon, J. M. (2015). Mapping the world’s degraded lands. Applied Geography, 57, 12–21. González-Sánchez, E. J., Moreno-García, M., Kassam, A., Holgado-Cabrera, A., TriviñoTarradas, P., Carbonell-Bojollo, R., … Basch, G. (2017). Conservation agriculture: Making climate change mitigation and adaptability real in Europe. European Conservation Agriculture Federation (ECAF). 154 pp. Heinrich Böll Stiftung, GRAIN & Institute for Agriculture & Trade Policy. (2017). Big Meat and Dairy’s supersized Climate Footprint Available at https://www.grain.org/article/ entries/5825-big-meat-and-dairy-s-supersized-climate-footprint. Herrero, M., Henderson, B., Havlík, P., et al. (2016). Greenhouse gas mitigation potentials in the livestock sector. Nature Climate Change, 6, 452–461. Hickel, J. (2016). The true extent of global poverty and hunger: Questioning the good news narrative of the Millennium Development Goals. Third World Quarterly, 37(5), 749–767. Holt-Giménez, E., Shattuck, A., Altieri, M., Herren, H., & Gliessman, S. (2012). We already grow enough food for 10 billion people and still can't end hunger. Journal of Sustainable Agriculture, 36(6), 595–598. https://www.researchgate.net/publication/241746569. Hosonuma, N., Herold, M., De Sy, V., De Fries, R. S., Brockhaus, M., Verchot, L., … Romijn, E. (2012). An assessment of deforestation and forest degradation drivers in developing countries. Environmental Research Letters, 7(4), 044009. IAASTD. (2009). Synthesis report: A synthesis of the global and sub-global IAASTD reports, agriculture at a crossroads. The International Assessment of Agricultural Knowledge, Science and TechnologyWashington, DC: Island Press. https://www.weltagrarbericht.de/ reports/Global_Report/Global_content.html. IPBES. (2019). Global assessment report of the intergovernmental science-policy platform on biodiversity and ecosystem services. New York: UN. IPCC. (2018). Global warming of 1.5°C: an IPCC special report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways. IPCC. Retrieved from https://www.ipcc.ch/sr15/. IPCC. (2019). Climate change and land: An IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems. New York: UN.

Introduction 

xxxi

Juniper, T. (2015). What has nature ever done for us? How money really does grow on trees. London: Profile Books. Kassam, A. (Ed.), (2020). Advances in conservation agriculture. Cambridge: Burleigh Dodds. Vol. 1: Systems and science, Vol. 2: Practice and benefits. Kassam, A., Basch, G., Friedrich, T., Gonzalez, E., Trivino, P., & Mkomwa, S. (2017). Mobilizing greater crop and land potentials sustainably. Hungarian Geographical Bulletin, 66(1), 3–11. Kassam, A., Basch, G., Friedrich, T., Shaxson, F., Goddard, T., Amado, T., … Mkomwa, S. (2013). Sustainable soil management is more than what and how crops are grown. In R. Lal & R. A. Stewart (Eds.), Advances in soil science: Principles of soil management in agro-ecosystems (pp. 337–400). Boca Raton, FL: CRC Press. Kassam, A. H., Friedrich, T., Shaxson, F., & Pretty, J. (2009). The spread of Conservation Agriculture: Justification, sustainability and uptake. International Journal of Agricultural Sustainability, 7(4), 292–320. Kissinger, G. M., Herold, M., & De Sy, V. (2012). Drivers of deforestation and forest degradation: A synthesis report for REDD + policymakers. Lexeme Consulting. Lappé, F. M., & Collins, J. (2015). World hunger: 10 myths. Food First Books & Institute for Food and Development PolicyNew York: Grove Press. Lawson, M., Chan, M. K., Rhodes, F., Parvez Butt, A., Marriott, A., Ehmke, E., & Gowland, R. (2019). Public good or private wealth. Oxfam Briefing Paper. Lenton, T. M., Rockström, J., Gaffney, O., Rahmstorf, S., Richardson, K., Steffen, W., & Schellnhuber, H. J. (2019). Climate tipping points—Too risky to bet against. Nature, 575, 592–595. Machovina, B., & Feeley, K. J. (2014). Meat consumption as a key impact on tropical nature: A response to Laurance et al. Trends in Ecology & Evolution, 29(8), 430–431. Machovina, B., Feeley, K. J., & Ripple, W. J. (2015). Biodiversity conservation: The key is reducing meat consumption. Science of the Total Environment, 536, 419–431. McBrien, J. (2016). Accumulating extinction: Planetary catastrophism in the Necrocene. Anthropocene or capitalocene. In J. W.  Moore (Ed.), Anthropocene or capitalocene?: Nature, history, and the crisis of capitalism (pp. 116–137). PM Press. MDG. (2015). The millennium development goals report 2015. New York: UN. MEA. (2005). Ecosystems and human well-being: Synthesis. Washington, DC: Millennium Ecosystem Assessment. Island Press. Montgomery, D. (2007). Dirt: The Erosion of civilizations. Berkeely/Los Angeles: University of California Press. Mood, A., & Brooke, P. (2010). Estimating the number of farmed fish killed in global aquaculture each year Retrieved December 12, 2019, from fishcount website: fishcount.org.uk. Moore, J. W. (2015). Capitalism in the web of life: Ecology and the accumulation of capital. Verso Books. Morse, S. S., Mazet, J. A. K., Woolhouse, M., Parrish, C. R., Carroll, D., Karesh, W. B., … Daszak, P. (2012). Prediction and prevention of the next pandemic zoonosis. The Lancet. https://doi.org/10.1016/S0140-6736(12)61684-5. Nkonya, E., Mirzabaev, A., & von Braun, J. (Eds.), (2016). Economics of land degradation and improvement—A global assessment for sustainable development. Springer Open. IFPRI and ZEF. Our World in Data (2020). Published online at OurWorldInData.org. Retrieved from: https:// ourworldindata.org/economic-growth.

xxxiiIntroduction

Pimm, S. L., Jenkins, C. N., Abell, R., Brooks, T. M., Gittleman, J. L., Joppa, L. N., … Sexton, J. O. (2014). The biodiversity of species and their rates of extinction, distribution, and protection. Science, 344(6187), 1246752. Poore, J., & Nemecek, T. (2018). Reducing food’s environmental impacts through producers and consumers. Science, 360(6392), 987–992. Rockström, J., Steffen, W., Noone, K., Persson, Å., Chapin, F. S., III, Lambin, E., … Nykvist, B. (2009). Planetary boundaries: Exploring the safe operating space for humanity. Ecology and Society, 14(2), 32. Sanders, B. (2018). Global animal slaughter statistics and charts Retrieved November 30, 2019, from Faunalytics website: https://faunalytics.org/global-animal-slaughter-statistics-and-charts/. SDG. (2018). The sustainable development goals report 2018. New York: UN. Shiva, V., Bhatt, V., Panigrahi, A., Mishra, K., Tarafdar, D., & Singh, V. (2017). Seeds of hope, seeds of resilience: How biodiversity and agroecology offer solutions to climate change by growing living carbon. New Delhi: Navdanya and RFSTE. Steffen, W., Richardson, K., Rockström, J., Cornell, S. E., Fetzer, I., Bennett, E. M., … Folke, C. (2015). Planetary boundaries: Guiding human development on a changing planet. Science, 347(6223), 1259855. Steinfeld, H., Gerber, P., Wassenaar, T. D., Castel, V., Rosales, M., Rosales, M., & de Haan, C. (2006). Livestock's long shadow: Environmental issues and options. FAO. Swinburn, B. A., Kraak, V. I., Allender, S., Atkins, V. J., Baker, P. I., Bogard, J. R., … Ezzati, M. (2019). The global syndemic of obesity, undernutrition, and climate change: The Lancet Commission report. The Lancet, 393(10173), 791–846. UNCCD. (2017). Global land outlook. Bonn: Secretariat of the United Nations Convention to Combat Desertification. Vermeulen, S. J., Campbell, B. M., & Ingram, J. S. I. (2012). Climate change and food systems. Annual Review of Environment and Resources, 37, 195–222. WDR. (2008). Agriculture for development. World Development ReportWashington, DC: World Bank. WHO. (2018). Obesity and Overweight factsheet from the WHO. Willett, W., Rockström, J., Loken, B., Springmann, M., Lang, T., Vermeulen, S., … Murray, C. (2019). Food in the Anthropocene: The EAT–Lancet Commission on healthy diets from sustainable food systems. Lancet, 6736, 3–49. https://doi.org/10.1016/ S0140-6736(18)31788-4. Worm, B., Barbier, E. B., Beaumont, N., Duffy, J. E., Folke, C., Halpern, B. S., … Sala, E. (2006). Impacts of biodiversity loss on ocean ecosystem services. Science, 314(5800), 787–790. WWF (n.d.-a). Deforestation and forest degradation. Retrieved December 12, 2019, from WWF website: https://www.worldwildlife.org/threats/deforestation-and-forest-degradation. WWF (n.d.-b). Soil erosion and degradation. Retrieved December 12, 2019, from WWF website: https://www.worldwildlife.org/threats/soil-erosion-and-degradation.

Setting innovation free in agriculture

1

Rupert Sheldrake Schumacher College, Dartington, Devon, United Kingdom

1.1 Introduction Since the late 19th century, the earth has been transformed by the applications of science through technology and modern medicine. There have also been great increases in food production through intensive agriculture and factory farming. Over the same period, capitalism has created international systems of trade, investment, and global corporations, whose job is to increase profits. Everyone is affected. Almost all agricultural systems now operate within this scientific, technological, and capitalist environment. In most parts of the world traditional agricultural methods have been replaced by modern “scientific” practices, including mechanization and the use of factory-made fertilizers, herbicides, fungicides, and pesticides. Irrigation has been extended widely, often at the expense of nonrenewable aquifers, and in some areas irrigation has degraded the soil through salinization. Large-scale deforestation to create more agricultural land has contributed to a startling loss of biodiversity. The increase in population from 1 billion in 1800 to 2 billion in 1927 to 7.7 billion in 2019 has created a much greater demand for food, as has the increase in meat consumption, which creates the need to feed billions of animals in factory farms on cereals, soya beans, and other crops. Science and economics are not theory-neutral. They are expressions of worldviews, and we need to be aware of the prevailing worldview, or else we will follow it through blind faith. In his influential book The Structure of Scientific Revolutions (1962), the historian of science Thomas Kuhn argued that, at any given time, the sciences are shaped by a particular model of reality that he called a paradigm. The paradigm defines valid ways of doing research. Most of the time, “normal science” goes on within the prevailing orthodoxy, but from time to time there is a shift in the fundamental model of reality through a scientific revolution, when a new, more inclusive paradigm supersedes the old one. Factors that help precipitate paradigm changes include anomalies, which are awkward phenomena that do not fit into the prevailing paradigm. They are usually dismissed, denied, or explained away until there is a shift in paradigm, giving a broader view of reality that enables these anomalies to be included. Contemporary science and economics are based on a materialist philosophy of nature, which asserts that all reality is material or physical. There is no reality but material reality. Matter is nonconscious. Human consciousness is a functionless by-product

Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00001-3 © 2021 Elsevier Inc. All rights reserved.

2

Rethinking Food and Agriculture

of the physical activity of the brain. God exists only as an idea in human minds, and hence in human heads. There are no purposes in nature, and evolution is purposeless. One major anomaly for this materialist paradigm is consciousness itself, the very existence of which is called “the hard problem” in the philosophy of mind. Another is purpose. Obviously, humans have purposes, and so do nonhuman animals. So, where do consciousness and purposes come from if nature is fundamentally nonconscious and purposeless? Some philosophers wrestle with these questions, but most people simply ignore them. These materialist beliefs are powerful not because most people think about them critically, but because they don’t. The facts of science are real enough, and so are the techniques that scientists use, and so are the technologies based on them. But the belief system that governs conventional scientific thinking is an act of faith, grounded in a 19th century ideology. The same materialist assumptions underlie our economic systems. And underlying these systems of thought is a vision of science as a world-­ transforming activity, led by a new priesthood, a vision dating from the 17th century, when mechanistic science was first established. The enormous successes of science and technology have indeed been world transforming and seem to have proved this vision true. But they have also led to an uncritical scientific dogmatism, which is now inhibiting free enquiry, threatening the survival of countless species, and endangering human survival.

1.2 The scientific priesthood Francis Bacon (1561–1626), a politician and lawyer who became Lord Chancellor of England, foresaw the power of organized science more than anyone else. To clear the way, he needed to show that there was nothing sinister about acquiring power over nature. When he was writing, there was a widespread fear of witchcraft and black magic, which he tried to counteract by claiming that the knowledge of nature was God given, not inspired by the devil. Science was a return to the innocence of the first man, Adam, in the Garden of Eden before the Fall. Bacon argued that the first book of the Bible, Genesis, justified scientific knowledge. He equated man’s knowledge of nature with Adam’s naming of the animals. God “brought them unto Adam to see what he would call them, and what Adam called every living creature, that was the name thereof” (Genesis 2: 19–20). This was literally man’s knowledge, because Eve was not created until two verses later. Bacon argued that man’s technological mastery of nature was the recovery of a God-given power, rather than something new. He confidently assumed that people would use their new knowledge wisely and well: “Only let the human race recover that right over nature which belongs to it by divine bequest; the exercise thereof will be governed by sound reason and true religion.”a The key to this new power over nature was organized, institutional research. In New Atlantis (1624), Bacon described a technocratic utopia in which a scientific priesthood a

Bacon (1951), p. 50.

Setting innovation free in agriculture 3

made decisions for the good of the state as a whole. The Fellows of this scientific “Order or Society” wore long robes and were treated with a respect that their power and dignity required. The head of the order traveled in a rich chariot, under a radiant golden image of the sun. As he rode in procession, “he held up his bare hand, as he went, as blessing the people.” The general purpose of this foundation was “the knowledge of causes and secret motions of things; and the enlarging of human empire, to the effecting of all things possible.” The Society was equipped with machinery and facilities for testing out explosives and armaments, experimental furnaces, gardens for plant breeding, and dispensaries.b This visionary scientific institution foreshadowed many features of institutional research, and was a direct inspiration for the founding of the Royal Society in London in 1660, and for many other national academies of science and research institutes. But although the members of these academies were often held in high esteem, none achieved the grandeur and political power of Bacon’s imaginary prototypes. In England in Bacon’s time (and still today) the Church of England was linked to the state as the established church. Bacon envisaged that the scientific priesthood would also be linked to the state through state patronage, forming a kind of established church of science. And here again he was prophetic. In nations both capitalist and communist, the official academies of science remain the centers of power of the scientific establishment. There is no separation of science and state. Scientists play the role of an established priesthood, influencing government policies on the arts of warfare, industry, agriculture, medicine, education, and research. Bacon coined the ideal slogan for soliciting financial support from governments and investors: “knowledge is power.”c But the success of scientists in eliciting funding from governments varied from country to country. The systematic state funding of science began much earlier in France and Germany than in Britain and the United States, where until the latter half of the 19th century most research was privately funded or carried out by wealthy amateurs like Charles Darwin (Kealey, 1996). In France, Louis Pasteur (1822–95) was an influential proponent of science as a truth-finding religion, with laboratories like temples through which mankind would be elevated to its highest potential: Take interest, I beseech you, in those sacred institutions which we designate under the expressive name of laboratories. Demand that they be multiplied and adorned; they are the temples of wealth and of the future. There it is that humanity grows, becomes stronger and better.d

By the beginning of the 20th century, science was almost entirely institutionalized and professionalized, and after the Second World War expanded enormously under government patronage, as well as through corporate investment.e The highest level of b

Bacon (1951), pp. 290–291. Fara (2009), p. 132. d Dubos (1960), p. 146. e Kealey (1996). c

4

Rethinking Food and Agriculture

funding is in the United States, where in 2015 the total expenditure on research and development was $495 billion, of which $121 billion came from the government.f But governments and corporations do not usually pay scientists to do research because they want innocent knowledge like that of Adam before the Fall. Naming animals, as in classifying endangered species of beetles in tropical rainforests, is a low priority. Most funding is a response to Bacon’s persuasive slogan “knowledge is power.” By the 1950s, when institutional science had reached an unprecedented level of power and prestige, the historian of science George Sarton approvingly described the situation in a way that sounds like the Roman Catholic Church before the Reformation: Truth can be determined only by the judgement of experts… Everything is decided by very small groups of men, in fact, by single experts whose results are carefully checked, however, by a few others. The people have nothing to say but simply to accept the decisions handed out to them. Scientific activities are controlled by universities, academies and scientific societies, but such control is as far removed from popular control as it possibly could be.g

Thomas Kuhn helped focus attention on the social aspect of science and reminded us that science is a collective activity. Scientists are subject to all the usual constraints of human social life, including peer-group pressure and the need to conform to the norms of the group. Kuhn’s arguments were largely based on the history of science, but sociologists of science have taken his insights further by studying science as it is actually practiced, looking at the ways that scientists build up networks of support, use resources and results to increase their power and influence, and compete for funding, prestige, and recognition. Bruno Latour’s Science in Action: How to Follow Scientists and Engineers Through Society (1987) is one of the most influential studies in this tradition. Latour observed that scientists routinely make a distinction between knowledge and beliefs. Scientists within their professional group know about the phenomena covered by their field of science, while those outside the network have only distorted beliefs. When scientists think about people outside their groups, they often wonder how they can still be so irrational: [T}he picture of non-scientists drawn by scientists becomes bleak: a few minds discover what reality is, while the vast majority of people have irrational ideas or at least are prisoners of many social, cultural and psychological factors that make them stick obstinately to obsolete prejudices. The only redeeming aspect of this picture is that if it were only possible to eliminate all these factors that hold people prisoners of their prejudices, they would all, immediately and at no cost, become as sound-minded as the scientists, grasping the phenomena without further ado. In every one of us there is a scientist who is asleep, and who will not wake up until social and cultural conditions are pushed aside.h f

Congressional Research Service, June 28, 2018, https://fas.org/sgp/crs/misc/R44307.pdf (Retrieved 1 November 2019). g Sarton (1955), pp. 16–17. h Latour (1987), pp. 184–185.

Setting innovation free in agriculture 5

For believers in the “scientific worldview,” all that is needed is to increase the public understanding of science through education and the media. Since the 19th century, a belief in materialism has indeed been propagated with remarkable success; millions of people have been converted to this “scientific” view, even though they know very little about science itself. They are, as it were, devotees of the Church of Science, or of scientism, of which scientists are the priests. This is how a prominent atheist layman, Ricky Gervais, expressed these attitudes in the Wall Street Journal in 2010, the same year that he was on the Time magazine list of the 100 most influential people in the world. Gervais is an entertainer, not a scientist or an original thinker, but he borrows the authority of science to support his worldview: Science seeks the truth. And it does not discriminate. For better or worse it finds things out. Science is humble. It knows what it knows and it knows what it doesn’t know. It bases its conclusions and beliefs on hard evidence --- evidence that is constantly updated and upgraded. It doesn’t get offended when new facts come along. It embraces the body of knowledge. It doesn’t hold on to medieval practices because they are tradition.

Gervais’ idealized view of science is hopelessly naïve in the context of the history and sociology of science. It portrays scientists as open-minded seekers of truth, not ordinary people competing for funds and prestige, constrained by peer-group pressures and hemmed in by prejudices and taboos. Bacon’s vision of a scientific priesthood has now been realized on a global scale. But his confidence that man’s power over nature would be guided by “sound reason and true religion” was misplaced.

1.3 The fantasy of omniscience The fantasy of omniscience is a recurrent theme in the history of science, as scientists aspire to a total god-like knowledge. At the beginning of the 19th century, the French physicist Pierre Simon Laplace imagined a scientific mind capable of knowing and predicting everything.i These ideas were not confined to physicists. Thomas Henry Huxley, who did so much to propagate Darwin’s theory of evolution, extended mechanical determinism to cover the entire evolutionary process: If the fundamental proposition of evolution is true, that the entire world, living and not living, is the result of the mutual interaction, according to definite laws, of the forces possessed by the molecules of which the primitive nebulosity of the universe was composed, it is no less certain the existing world lay, potentially, in the cosmic vapour, and that a sufficient intellect could, from a knowledge of the properties of the molecules of that vapour, have predicted, say, the state of the fauna of Great Britain in 1869.j i j

Laplace (1819), p. 4. Quoted in Bergson (1911), p. 40.

6

Rethinking Food and Agriculture

When the belief in determinism was applied to the activity of the human brain, it resulted in a denial of free will, on the grounds that everything about the molecular and physical activities of the brain was in principle predictable. Yet this conviction rested not on scientific evidence, but simply on the assumption that everything was fully determined by mathematical laws. Even today, many scientists assume that free will is an illusion. Not only is the activity of the brain determined by machine-like processes, but there is no nonmechanical self capable of making choices (Chivers, 2010). In 1927, with the recognition of the uncertainty principle in quantum physics, it became clear that indeterminism was an essential feature of the physical world, and physical predictions could be made only in terms of probabilities. The fundamental reason is that quantum phenomena are wavelike, and a wave is by its very nature spread out in space and time; it cannot be localized at a single point at a particular instant, or, more technically, its position and momentum cannot both be known precisely.k Quantum theory deals in statistical probabilities, not certainties. The fact that one possibility is realized in a quantum event rather than another is a matter of chance. In neo-Darwinian evolutionary theory, randomness plays a central role through the chance mutations of genes, which are quantum events. With different chance events, evolution would happen differently. T. H. Huxley was wrong in believing that the course of evolution was predictable. “Replay the tape of life,” said the evolutionary biologist Stephen Jay Gould, “and a different set of survivors would grace our planet today” (Gould, 1989). In the 20th century it became clear that not just quantum processes but almost all natural phenomena are probabilistic, including the turbulent flow of liquids, the breaking of waves on the seashore, and the weather; they show a spontaneity and indeterminism that eludes exact prediction. Weather forecasters still get it wrong in spite of having powerful computers and a continuous stream of data from satellites. This is not because they are bad scientists but because weather is intrinsically unpredictable in detail. It is chaotic, not in the everyday sense that there is no order at all, but in the sense that it is not precisely predictable. To some extent, the weather can be modeled mathematically in terms of chaotic dynamics, sometimes called “chaos theory,” but these models do not make exact predictions (Gleik, 1988). Certainty is as unachievable in the everyday world as it is in quantum physics. Even the orbits of the planets around the sun, long considered the centerpiece of mechanistic science, turn out to be chaotic over long time scales.l The belief in determinism, strongly held by many 19th- and early 20th-century scientists, turned out to be a delusion. The freeing of scientists from this dogma led to a new appreciation of the indeterminism of nature in general, and of evolution in particular. The sciences have not come to an end by abandoning the belief in determinism. Likewise, they will survive the loss of the dogmas that still bind them; they will be regenerated by new possibilities. By the end of the 19th century, the fantasy of scientific omniscience went far beyond a belief in determinism. In 1888, the Canadian-American astronomer Simon k l

Munowitz (2005), chap. 7. Horgan (1997).

Setting innovation free in agriculture 7

Newcomb wrote, “We are probably nearing the limit of all we can know about astronomy.” In 1894, Albert Michelson, later to win the Nobel Prize for Physics, declared, “The more important fundamental laws and facts of physical science have all been discovered, and these are now so firmly established that the possibility of their ever being supplanted in consequence of new discoveries is exceedingly remote . . . Our future discoveries must be looked for in the sixth place of decimals.”m And in 1900, William Thomson, Lord Kelvin, the physicist and inventor of intercontinental telegraphy, expressed this supreme confidence in an often-quoted (although perhaps apocryphal) claim: “There is nothing new to be discovered in physics now. All that remains is more and more precise measurement.” These convictions were shattered in the 20th century through quantum physics, relativity theory, nuclear fission and fusion (as in atom and hydrogen bombs), the discovery of galaxies beyond our own, and the Big Bang theory—the idea that the universe began very small and very hot some 14 billion years ago and has been growing, cooling, and evolving ever since. Nevertheless, by the end of the 20th century, the fantasy of omniscience was back again, this time fueled by the triumphs of 20th-century physics and by the discoveries of neurobiology and molecular biology. In 1997, John Horgan, a senior science writer at Scientific American, published a book called The End of Science: Facing the Limits of Knowledge in the Twilight of the Scientific Age. After interviewing many leading scientists, he advanced a provocative thesis: If one believes in science, one must accept the possibility – even the probability – that the great era of scientific discovery is over. By science I mean not applied science, but science at its purest and greatest, the primordial human quest to understand the universe and our place in it. Further research may yield no more great revelations or revolutions, but only incremental, diminishing returns.n

Horgan is surely right that once something has been discovered—like the structure of DNA—it cannot go on being discovered. But he took it for granted that the tenets of conventional science are true. He assumed that the most fundamental answers are already known. They are not, and every one of them can be replaced by more interesting and fruitful questions, as I show in my book The Science Delusion (called Science Set Free in the United States).

1.4 The credibility crunch for materialism Within biology, an extreme form of materialism took hold in the 1970s and 1980s in the form of molecular biology, which soon became the dominant approach. More holistic forms of biology were marginalized. This molecular biological paradigm had enormous effects both on medicine and also on agriculture, shifting the focus of research to the molecular level. Hundreds of billions of dollars of public and private m n

Quoted in Horgan (1997). Horgan (1997), p. 6.

8

Rethinking Food and Agriculture

funds have poured into genome projects, genetic modifications of animals and plants, gene editing techniques, and other ingenious biotechnologies. There have been impressive successes, like the technical triumph of sequencing of the human genome and the genomes of many other species, as well as some specialized applications, such as the identification of the genetic basis for rare hereditary human diseases. There are several examples of the commercially successful genetic engineering of crops. Billions of dollars in profits have flowed to corporations that own the patents on these genetically modified varieties. But this one-sided molecular approach has shifted attention away from many other possibilities in medicine and agriculture. For more than 200 years, materialists have promised that science will eventually explain everything in terms of physics and chemistry. Science will prove that living organisms are complex machines, minds are nothing but brain activity, and nature is purposeless. Believers are sustained by the faith that scientific discoveries will justify their beliefs. The philosopher of science Karl Popper called this stance “promissory materialism” because it depends on issuing promissory notes for discoveries not yet made.o Despite all the achievements of science and technology, materialism is now facing a credibility crunch that was unimaginable in the 20th century. In 1963, when I was studying biochemistry at Cambridge University, I was invited to a series of private meetings with Francis Crick and Sydney Brenner in Brenner’s rooms in King’s College, along with a few of my classmates. Crick and Brenner had recently helped to “crack” the genetic code. Both were ardent materialists and Crick was also a militant atheist. They explained that there were two major unsolved problems in biology: development and consciousness. They had not been solved because the people who worked on them were not molecular biologists—nor very bright. Crick and Brenner were going to find the answers within 10 years, or maybe 20. Brenner would take developmental biology, and Crick consciousness. They invited us to join them. Both tried their best. Brenner was awarded the Nobel Prize in 2002 for his work on the development of a tiny worm, Caenorhabditis elegans. Crick corrected the manuscript of his final paper on the brain the day before he died in 2004. At his funeral, his son Michael said that what made him tick was not the desire to be famous, wealthy, or popular, but “to knock the final nail into the coffin of vitalism” (Ridley, 2011). (Vitalism is the theory that living organisms are not fully explicable in terms of physics and chemistry alone.) Crick and Brenner failed. The problems of development and consciousness remain unsolved. Many details have been discovered, dozens of genomes have been sequenced, and brain scans are ever more precise. But there is still no proof that life and minds can be explained by physics and chemistry alone. The fundamental proposition of materialism is that matter is the only reality. Therefore consciousness is nothing but brain activity. It is either like a shadow, an “epiphenomenon,” that does nothing, or it is just another way of talking about brain activity. However, among contemporary researchers in neuroscience and consciousness studies there is no consensus about the nature of minds. Leading journals such as Behavioural and Brain Sciences and the Journal of Consciousness Studies publish many articles that o

In Popper and Eccles (1977).

Setting innovation free in agriculture 9

reveal deep problems with the materialist doctrine. The philosopher David Chalmers has called the very existence of subjective experience the “hard problem.” It is hard because it defies explanation in terms of mechanisms. Even if we understand how eyes and brains respond to red light, the experience of redness is not accounted for. In biology and psychology the credibility rating of materialism is falling. Can physics ride to the rescue? Some materialists prefer to call themselves physicalists, to emphasize that their hopes depend on modern physics, not 19th-century theories of matter. But physicalism’s own credibility rating has been reduced by physics itself, for four reasons: First, some physicists insist that quantum mechanics cannot be formulated without taking into account the minds of observers. They argue that minds cannot be reduced to physics because physics presupposes the minds of physicists (D’Espagnat, 1976). Second, the most ambitious unified theories of physical reality, string and M-theories, with 10 and 11 dimensions, respectively, take science into a completely new territory. String theories and M-theories are currently untestable, so they can only be judged by reference to other models, rather than by experiment. These theories also apply to countless other universes, none of which has ever been observed. Some physicists are deeply skeptical about this entire approach, as the theoretical physicist Lee Smolin shows in his book The Trouble With Physics: The Rise of String Theory, the Fall of a Science and What Comes Next (2006). String theories, M-theories, and “model-dependent realism,” which tests models against other models, rather than by empirical evidence, are a shaky foundation for materialism or physicalism or any other belief system. Third, since the beginning of the 21st century, it has become apparent that the known kinds of matter and energy make up only about 5% of the universe. The rest consists of “dark matter” and “dark energy.” The nature of 95% of physical reality is literally obscure. Fourth, the Cosmological Anthropic Principle asserts that if the laws and constants of nature had been slightly different at the moment of the Big Bang, biological life could never have emerged, and hence we would not be here to think about it. So, did a divine mind fine-tune the laws and constants in the beginning? To avoid a creator God emerging in a new guise, most leading cosmologists prefer to believe that our universe is one of a vast, and perhaps infinite, number of parallel universes, all with different laws and constants, as M-theory also suggests. We just happen to exist in the one that has the right conditions for us.p

This multiverse theory is the ultimate violation of Ockham’s razor, the philosophical principle that “entities must not be multiplied beyond necessity,” or in other words that we should make as few assumptions as possible. It also has the major disadvantage of being untestable.q And it does not even succeed in getting rid of God. An infinite God could be the God of an infinite number of universes.r Materialism provided a seemingly simple, straightforward worldview in the late 19th century, but 21st-century science has left it far behind. Its promises have not been fulfilled, and its promissory notes have been devalued by hyperinflation. I am convinced that the sciences in general and agricultural science in particular are being held back by assumptions that have hardened into dogmas, maintained by p

Carr (ed.) (2007). Ellis (2011). r Collins, in Carr (ed.) (2007), pp. 459–480. q

10

Rethinking Food and Agriculture

powerful taboos. These beliefs protect the citadel of established science, but act as barriers against open-minded thinking.

1.5 The unfulfilled promises of molecular biology It is hard to recall the atmosphere of exhilaration in the 1980s as new techniques enabled genes to be cloned and the sequence of “letters” in their genetic code to be discovered. This seemed like biology’s crowning moment; the genetic instructions of life itself were finally laid bare, opening up the possibility for biologists to modify plants and animals genetically, and grow richer than they could ever have imagined. Almost every week newspaper headlines reported a new breakthrough: “Scientists find genes to combat cancer,” “Gene therapy offers hope to victims of arthritis,” “Scientists find secret of aging,” and so on. The new genetics seemed so promising that soon the entire spectrum of biological researchers was busy applying its techniques to their specialties. Their remarkable progress led to a vast, ambitious vision: to spell out the full complement of genes in the human genome. As Walter Gilbert of Harvard University put it, “The search for this ‘Holy Grail’ of who we are has now reached its culminating phase. The ultimate goal is the acquisition of all the details of our genome.” The Human Genome Project was formally launched in 1990 with a projected budget of $3 billion. The Human Genome Project was a deliberate attempt to bring “Big Science” to biology. Physicists were used to huge budgets, partly as a result of the Cold War; there was enormous expenditure on missiles and hydrogen bombs, Star Wars, ­multibillion-dollar particle accelerators, the space program and the Hubble Space Telescope. Ambitious biologists suffered from physics envy. They dreamed of the days when biology would have high profile, high prestige, and multibillion-dollar international projects. The Human Genome Project was the answer. At the same time, a tide of market speculation in the 1990s led to a boom in biotechnology, reaching a peak in 2000. In addition to the official Human Genome Project, Celera Genomics carried out a private genome project, headed by Craig Venter. The company planned to patent hundreds of human genes and own the commercial rights to them. Celera Genomics’ market value, like that of many other biotechnology companies, rocketed to dizzy heights in the early months of 2000. Ironically, the rivalry between the public and private genome projects led to a bursting of the bubble before the sequencing of the genome had even been completed. In March 2000, the leaders of the public genome project publicized the fact that all their information would be freely available to everyone. This led to a statement by US President Clinton on March 14, 2000: “Our genome, the book in which all human life is written, belongs to every member of the human race…We must ensure that the profits of the human genome research are measured not in dollars, but in the betterment of human life.”s The press reported that the president planned to restrict genomic patents. The stock markets reacted dramatically. In Venter’s words, there was a “sickening s

Venter (2007), p. 299.

Setting innovation free in agriculture 11

slump.” Within 2 days, Celera’s valuation lost $6 billion, and the wider market in biotechnology shares collapsed by $500 billion.t On June 26, 2000, President Clinton and the British prime minister, Tony Blair, with Craig Venter and Francis Collins, the head of the official project, announced the publication of the first draft of the human genome. At the press conference in the White House, President Clinton said, “We are here today to celebrate the completion of the first survey of the entire human genome. Without a doubt this is the most important, most wondrous map ever produced by mankind. It will revolutionise the diagnosis, prevention and treatment of most, if not all, human diseases…Humankind is on the verge of gaining immense, new power to heal.” The British science minister, Lord Sainsbury, said, “We now have the possibility of achieving all we ever hoped for from medicine.”u One of the editors of Nature proclaimed that by the end of the 21st century, “genomics will allow us to alter entire organisms out of recognition, to suit our needs and tastes . . . [and] will allow us to fashion the human form into any conceivable shape. We will have extra limbs, if we want them, and maybe even wings to fly.”v This astonishing achievement of sequencing the human genome has indeed transformed our view of ourselves, but not as anticipated. The first surprise was that there were so few genes. Rather than the predicted 100,000 or more, the final tally of about 23,000 was very puzzling, and all the more so when compared with the genomes of other animals much simpler than ourselves. There are about 17,000 genes in a fruit fly, and about 26,000 in a sea urchin. Many species of plants have far more genes than us—rice has about 38,000, for example. In the wake of the Human Genome Project, the mood changed dramatically. The optimism that life would be understood if molecular biologists knew the “programs” of an organism gave way to the realization that there is a huge gap between gene sequences and actual human beings. In practice, the predictive value of human genomes turned out to be small. For example, in the case of height, genomes predict less than that achieved with a measuring tape. Tall parents tend to have tall children, and short parents short children. By measuring the height of parents, their children’s heights can be predicted with about 80% accuracy. In other words, height is about 80% heritable. By the year 2008, “genome-wide association studies” compared the genomes of 30,000 people and identified about 50 genes associated with tallness or shortness. To everyone’s surprise, taken together, these genes accounted for only about 5% of the inheritance of height. In other words, the “height” genes did not account for 75% of the heritability of height. Most of the heritability was missing. Many other examples of missing heritability are now known, including the heritability of many diseases, making “personal genomics” of very questionable value. Since 2008, in the scientific literature this phenomenon has been called the “missing heritability problem.” In a study published in 2011, the percentage of heritability of height that could be predicted on the basis of the genome of unrelated individuals was 15%,w an t

Venter (2007), p. 300. Quoted in Nature (2011). v Quoted in Nature (2011). w Makowsky et al. (2011). u

12

Rethinking Food and Agriculture

i­mprovement on earlier methods, but still far short of the heritability predictable by measuring the heights of relatives without sequencing any genes at all. And measuring people’s heights with a tape measure is billions of dollars cheaper than the genomic approach. The optimism of stock-market investors who hoped for big returns from molecular technologies has suffered recurring blows. After the biotech bubble burst in 2000, many biotech companies either went out of business or were taken over by pharmaceutical or chemical corporations. An article in the Wall Street Journal in 2004 entitled “Biotech’s Dismal Bottom Line: More than $40 billion in Losses”x went on to say, “Biotechnology . . . may yet turn into an engine for economic growth and cure deadly diseases. But it’s hard to argue that it’s a good investment. Not only has the biotech industry yielded negative financial returns for decades, it generally digs its hole deeper every year.”y In 2006, Harvard Business School published a detailed analysis of the industry. They found that “only a very tiny fraction” of biotechnology companies had ever made a profit, and that promises of breakthroughs had failed over and over again. Defenders of the industry argued that more time was needed, but the Harvard Business School analysis pointed to the opposite conclusion: “[G]iven the extremely poor longterm performance of the biotechnology industry in general, and specific firms in particular, capital has been, if anything, too patient.”z Nevertheless, new biotechnology companies are continually being launched. Many of them burn through hundreds of millions of dollars with nothing to show for it.aa Most promise new and highly profitable medical advances, while others are directed at producing new genetically engineered varieties of agricultural crops or farm animals. Since around 2015, a new gene-editing technique called CRISPR-Cas9 has become a popular basis for a new wave of biotech start-up companies. These techniques may well lead to some specialized medical applications, particularly for rare genetic disorders, and some niche applications in plant breeding (Jaganathan, Ramasamy, Sellamuthu, Jayabalan, & Venkataraman, 2018). A newer gene editing technique called “prime editing” allows even more precise manipulation of DNA sequences (Cohen, 2019). But precision gene editing is of very limited use because so many traits are polygenic. They are affected by large numbers of genes with very small effects, rather than one or two master genes that can be edited precisely. Despite its underwhelming business record, this vast investment in molecular biology and biotechnology has had wide-ranging effects on the practice of biology, if only by creating so many jobs. The demand for graduates in molecular biology has transformed the teaching of biology. The molecular approach now predominates in most universities and it has strongly influenced science teaching in secondary schools. Precisely because there has been such a strong emphasis on molecular biology, its limitations are becoming increasingly apparent. The sequencing of the genomes of ever more species of animals and plants, together with the determination of the structures x

Wall Street Journal, May 2, 2004. Pisano (2006), p. 184. z Pisano (2006), p. 198. aa https://www.investopedia.com/articles/fundamental-analysis/11/primer-on-biotech-sector.asp (Retrieved 29 May 2019). y

Setting innovation free in agriculture 13

of thousands of proteins, is causing molecular biologists to drown in their own data. There is practically no limit to how many more genomes they can sequence or proteins they can analyze. Molecular biologists now rely on computer specialists in the rapidly growing field of bioinformatics to store and try to make sense of this unprecedented quantity of information, sometimes called the “data avalanche” (Howe & Rhee, 2008).

1.6 Toward a more holistic approach These historical developments and philosophical movements have had an enormous impact on agriculture, along with the growth of capitalism and in particular its globalized neoliberal form. By becoming aware of these assumptions it becomes easier to see how research in agriculture could move in a more holistic direction. The paradigm of mechanistic materialism is now being challenged by a view of nature as organic and alive, a more holistic paradigm. Some philosophers of mind and neuroscientists are now adopting a panpsychist worldview instead, according to which all self-organizing systems, including atoms, have an element of mind or subjective experience (Goff, 2019). The Gaia hypothesis, the idea of the earth as a living organism, is one example of this change. But even for those who have no interest in these theoretical arguments, for practical reasons alone we need to change the way we think and act. 1. The cult of the scientific priesthood privileges knowledge of scientists from universities, and in particular scientists from universities in developed countries, over that of local farmers and agronomic practices. Traditional systems of knowledge are dismissed as superstitious or ignorant in favor of the currently fashionable views of technical experts. A more holistic approach to agriculture would not ignore traditional knowledge, locally adapted crop varieties, practices, and experience, but try to find out more about them and integrate them in agricultural development policies where appropriate. 2. The fantasy of omniscience leads to an obsession with quantitative precision, which ignores the complex ecological systems and patterns of interconnection that exist in the natural world. This reflects itself in the monopoly of monoculture, and the idea that improved varieties can be rolled out over large geographical areas because they represent the new improved scientific version of a crop as opposed the thousands of locally adapted varieties that were grown by farmers beforehand. Within plant breeding it leads to a focus on small numbers of genes that can be genetically modified, and with CRISPR-Cas9 and prime editing to a focus on single genes and small parts of single genes. Where symbiosis is recognized, as in the case of rhizobial bacteria that fix nitrogen in the root nodules of legumes, inoculants of a single defined strain are used despite the fact that in the soil there are complex mixtures of rhizobia and symbiotic mycorrhizal fungi. Moving beyond this obsession with quantitative precision would lead to a recognition of the importance of mixed cropping, soil microbiomes, the importance of many different genes, and also epigenetic effects in inheritance. 3. In economics and government planning, an obsession with quantitative results, like profits or growth in national products, leads to a tunnel vision that ignores the social and cultural life of farmers, ethical questions of land ownership and labor, animal welfare, the health of the ecosystem, effects of pollution, and long-term damage caused by soil erosion and degradation. A more holistic approach would take these factors into account instead of ignoring them.

14

Rethinking Food and Agriculture

4. Economic measurements are usually made on a short-term basis, like the annual growth in gross national product, annual profits of corporations, and annual crop yield figures. Corporations and shareholders also think in terms of annual profits, while in democracies politicians think in terms of electoral cycles of a few years. But the ecology of the earth, the health of the soil, and the sustainability of farming practices require much longer-term thinking. 5. With the increasing globalization of business, enormously powerful companies are able to influence consumer demand through advertising and government policies through lobbying. Their motives are not human welfare and health, nor long-term sustainability of the ecosystem, nor healthy populations, but short-term profits. The results include soil degradation, extinction of many species, toxic wastes, runoffs from farms, and an epidemic of obesity. A more holistic approach would inevitably take a longer-term and more sustainable perspective.

In the light of these principles, I discuss several new possibilities for agricultural research and development. Business as usual is unsustainable in the context of soil degradation, environmental pollution, deforestation, loss of biodiversity, increasing human population, climate change, limited water resources, and the obesity pandemic. New initiatives on large and small scales are needed. Here are several possibilities. Some, such as enhancing the microbiome of the soil, pragmatic plant breeding, and epigenetic inheritance, depend on a change in worldview; others make sense within both materialist and nonmaterialist frameworks of thought.

1.6.1 Rediscovering and testing traditional practices All over the world, farmers have traditional practices that developed long before the advent of mechanistic science. Some of these practices are relatively uncontroversial from an ideological point of view, like mixed cropping (including intercropping). Increasingly though, traditional practices such as mixed cropping have given way to monocultures, mainly as a result of mechanization. However, even in more industrialized countries, mixed cropping still takes place in pastures, with mixtures of grass and clover, where nitrogen fixation by the root nodules in the clover reduces the need for nitrogen fertilization. The same principles apply to traditional mixed cropping and intercropping systems in which leguminous crops are combined with cereals, as in the “three sisters” system in Mexico, with maize, climbing beans, and squash grown in clusters. Intercropping systems are more systemized in that they are planted in rows, for example, with alternate rows of cereals and legumes, such as sorghum and pigeon peas in India. In this system, the pigeon peas not only enrich the soil with nitrogen and benefit subsequent crops through their root residues and fallen leaves, but also continue growing after the harvest of the earlier-maturing sorghum, using the vacated space, and after the end of the monsoon extracting residual soil moisture through their deep roots (Sheldrake, 1984). Such intercropping systems are more efficient in terms of utilization of land and resources, show higher yields under a wide range of environmental conditions, and also reduce the risk of crop failure (Rao & Willey, 1980). By testing traditional cropping systems and other practices, useful lessons can be learned, as in the investigation of sorghum/pigeon pea intercropping systems already

Setting innovation free in agriculture 15

discussed. There may be much scope for new mixed cropping systems that can use land and water more efficiently and also reduce dangers from pests and diseases. The first place to look for suitable systems would be in traditional practices by small-scale farmers in different parts of the world. Researchers could explore how these systems could be used in modern contexts. New possibilities will be opened up by precision agricultural equipment. Unfortunately, agricultural researchers, aid agencies, and governments have often assumed that modern science knows best, and that traditional practices need to be replaced by modern farming methods, including the use of new cultivars grown in mechanized monocultures with chemical inputs of fertilizers, herbicides, and insecticides. The opposite approach is to find out what practices farmers have used traditionally, and why. This is one of the areas of research in agricultural anthropology (Rhoades & Rhoades, 2008), but this is still a neglected field of enquiry that could make an important contribution to sustainable agriculture (Sarkar, 2017).

1.6.2 Enhancing the microbiome of the soil It has been known for many years that soil microbes can help enhance the fertility of soils both through nitrogen fixation by free-living microbes and in root nodules of legumes. At the same time, mycorrhizal fungi help plants to mobilize phosphorus and nitrogen from the soil, and also increase their ability to take up water. The activities of mycorrhizae and of nitrogen-fixing microbes are suppressed when chemical fertilizers are added to the soil. The richness of the microbiome and diversity of mycorrhizae are generally highest under organic farming systems (Manoharan, Rosenstock, Williams, & Hedlund, 2017). In recent decades the emphasis has been on the use of chemical fertilizers, and relatively little attention has been paid to the ecology of soil microorganisms (Hart & Trevors, 2005). As it becomes more imperative to farm sustainably and preserve the fertility and structure of the soil, a new wave of research on the ecology of soil microbes would be very helpful. The effects of different soil management practices, cropping systems, and rotations have a major effect on the microbial ecology (Bender, Wagg, & van der Heijden, 2016). The ecosystem of the soil is severely disrupted by plowing, and one of the selling points for Conservation Agriculture is that it avoids this form of disturbance. However, insofar as conventional systems, both tillage- and nontillage based, depend on chemical herbicides, like glyphosate, they can also have adverse effects on soil ecology. Glyphosate is toxic to mycorrhizae and reduces the viability of mycorrhizal fungal spores and the colonization of roots by mycorrhizal fungi (Druille, Cabello, Omacini, & Golluscio, 2013). Organic and biodynamic farming systems avoid this problem, but still disrupt the soil microbial ecology through tillage. Conservation Agriculture generally uses less glyphosate per hectare compared to conventional tillage agriculture, and many smallholders do not use glyphosate in their Conservation Agriculture systems. A global review by Goss, Carvalho, and Brito (2017) found no evidence of negative effects of glyphosate application on mycorrhizae diversity, colonization, and function in Conservation Agriculture systems for a number of reasons, including healthy soils, no or minimal soil disturbance, m ­ aintenance of a

16

Rethinking Food and Agriculture

protective soil mulch cover, and crop and root system diversification, all of which have been shown to promote mycorrhizae. On the other hand, inversion tillage was found to have an adverse effect on mycorrhizae in the soil and their ability to inoculate subsequent crops because of the degradation of mycorrhizal habitat and soil health (Goss et al., 2017). It is tempting to think that inoculating crops with mycorrhizal fungi will solve these problems, and many companies now sell “bioinoculant” products designed to do this. However, adding exogenous mycorrhizae to the soil often has little or no effect on the growth and yield of the crop, and may even be harmful, because the added organisms are in competition with mycorrhizae that are already present in the soil, and which have evolved in those specific ecological conditions (Hart, Antunes, Chaudhary, & Abbott, 2017). The same applies to inoculation of legume crops with strains of the nitrogen-fixing bacterium Rhizobium. In laboratory experiments with plants grown in sterile soil, such inoculations can produce dramatic improvements in growth, but in the field the soil is not sterile and already contains a complex microbial ecosystem. We need sustainable agricultural practices, including crop rotations that enhance the health and effectiveness of the ecology of the soil, rather than practices that harm this ecology. Much more research is needed on the effects of crop rotations, organic and inorganic nutrient additions, and tillage or no-tillage practices on the soil microbiome and on the dynamics of mycorrhizae (Goss et al., 2017).

1.6.3 Automated equipment with artificial intelligence could help small farmers In Europe, North America, Australia, and other parts of the world, the need to cut labor costs has led to continual increases in the sizes of fields and farms, and also the size of machines. One man operating a very large machine incurs lower labor costs than several men operating smaller machines. But although the cost of labor is reduced, the capital costs rise to a level that is unaffordable for most farmers. For example, in 2018 in the United Kingdom, combine harvesters ranged in price from around £140,000 to £550,000 (about $180,000 to $720,000), depending on make and size.ab Together with large tractors and other equipment, farming involves huge capital investments or equipment leasing costs, which drive the economics of farming toward ever-larger units. Farms of 1000 acres or more are increasingly common. With the price of arable land around £8000 ($10,000) per acre, the land alone for such a farm costs around £8 million ($10.5 million). No young person could possibly become a farmer under these conditions without enormous inherited wealth or heavy financial backing. But we may be on the threshold of a major change. Rapidly improving technologies for self-driving cars and self-driving agricultural machinery, using GPS navigation systems, will enable driverless tractors and harvesters to become normal pieces of machinery. In addition, artificial intelligence (AI) allows for precision sowing, drilling, and harvesting methods. Instead of a one-size-fits-all approach, modern equipab

https://fwi-wp-assets-live.s3-eu-west-1.amazonaws.com/sites/1/Combine-buyers-guide-tables. pdf (Retrieved 5 June 2019).

Setting innovation free in agriculture 17

ment can respond to differences in soil condition and apply fertilizers appropriately in parts of fields where they are needed, recognize weeds and physically remove them, apply insecticides where plants are being attacked, and harvest fruit and vegetables selectively when they are ripe.ac Just as driverless cars can be large or small, so can automated farm machinery. Soon it may no longer be necessary to have vast farms with gigantic machines to reduce labor costs. Small farms with smaller and cheaper intelligent machines could become financially viable. Field sizes and farm sizes could be reduced, and cropping systems become more diverse. Sophisticated computerized equipment could also make it easier to use mixed cropping systems, some of which are at present unfeasible because they cannot be mechanized; without mechanization, labor costs are too high. New, sophisticated technologies using AI could enable efficient, productive, and diverse mixed cropping systems to flourish in place of monocultures.

1.6.4 Urban gardens and part-time farming In some parts of the world, significant amounts of food are grown by people who are not professional farmers, as in allotments in the United Kingdom (Acton, 2011), and community gardens, as in the United States and Canada. In many other parts of the world, including Cuba and parts of Africa, urban gardens are important sites of local food production. Altogether, an estimated 800 million people worldwide currently practice some form of urban agriculture (Edmondson, Davies, Gaston, & Leake, 2014). In terms of productivity per unit area, these gardens are often better than agricultural monocultures.ad In the United Kingdom, during the Second World War, in the “Dig for Victory” campaign, allotments and gardens provided around 10% of the food consumed, despite covering less than 1% of the area of arable cultivation. Recent research has shown that gardens and allotments can produce yields of fruit and vegetables 4–11 times greater than the same area under conventional agricultural crops (Thompson, 2014). Moreover, a study comparing the soil in allotments and in conventional arable fields in Britain showed that the allotments had on average 32% more organic carbon and 25% more total nitrogen. They were also less compacted. Unlike most arable farmers, 95% of the allotment gardeners composted biomass on site and many also added organic-based fertilizers and commercial composts. They were farming more sustainably than most farmers, and maintaining a higher soil quality (Edmondson et al., 2014). In some parts of the world, through an inheritance system whereby children share the family land, there is a continual splitting up of family farms, with the result that some people living in cities have small farms nearby that they look after on a parttime basis, combining weekend farm work with urban jobs, as in the neighborhood of Freiburg-in-Breisgau, Germany. ac

There are many YouTube videos online that show these systems in action, for example, https://www.youtube.com/watch?v=Rl77FVobxVI (Retrieved 5 June 2019). ad https://www.allotment-garden.org/allotment-information/allotment-history/ (Retrieved 11 June 2019).

18

Rethinking Food and Agriculture

There is a large potential for an increase in this kind of food production through making small farms, gardens, or orchards more available to people living in towns. In some parts of the world, there is an unsatisfied demand for opportunities to grow food in and around cities as a hobby, or as a part-time occupation. In Britain, for example, there are long waiting lists for allotments, and a chronic shortage of supply.ae Thus more food could be produced in a healthy and diversified way by making more areas for gardening or small-scale farming available in or near urban areas. But where will the land come from? Landowners near towns would be unlikely to want to sell land and thus potentially lose out on huge capital gains if the land were developed for building, but they could be willing to lease land on 5- or 10-year leases. For example, I have calculated that if landowners in Britain rented out land for family orchards and gardens in small units, say a fifth of an acre, or 800 m2, they could receive at least 20–30 times more annual rent per acre than renting it to a farmer for arable monoculture (Sheldrake, 2013). There would also probably be an improvement in the physical and mental health of the families looking after these orchards and gardens.

1.6.5 Using human wastes When I was doing research at ICRISAT in India, some of my field studies on chickpeas took place in a Himalayan village in the Lahaul Valley, near the border with Tibet. I was staying with a farming family in a traditional farmhouse. The toilet arrangement consisted of a room with an elevated floor on which there was a hole, through which all the wastes dropped down into straw, forming a rich manure. This was put back onto the family’s land each year, and thus the nutrients were directly recycled. I did not realize at the time that I was seeing, and taking part in, a traditional practice that had predominated for millennia throughout large parts of eastern Asia. As F.H. King showed in his classic book Farmers of Forty Centuries, Or, Permanent Agriculture in China, Korea and Japan (1911), nothing was wasted. Human excreta in the form of “night soil” were a valuable commodity. Public toilets were a means of collecting useful resources. As King saw for himself on a journey he made from Yokohama to Tokyo, “In such places as railway stations, provision is made for saving, not for wasting, and even along the country roads screens invite the traveller to stop, primarily for profit to the owner more than for personal convenience” (King, 1911, p. 9). By contrast, most of us have grown used to a system in which our own wastes are literally wasted, flushed away in plumbing systems that require enormous quantities of water. Meanwhile urea and other fertilizers are made in factories or mined from the earth, and then applied to the soil in excessive quantities, with the runoff causing large-scale pollution of rivers, lakes, and seas. One starting point for the recycling of human wastes would be to collect urine separately. It is easier to process and rich in nutrients, particularly urea. Humans generally excrete between 12 and 20 g of urea a day or up to 7 kg a year. If we take a moderate ae

https://www.theguardian.com/news/datablog/2011/nov/10/allotments-rents-waiting-list (Retrieved 6 June 2019).

Setting innovation free in agriculture 19

figure of 5 kg per year per person, with the world’s population around 7.5 billion, about 38 billion kg or 38 million tons of urea are excreted. In 2016, world urea production in factories was about 170 million tons a year,af about five times more than the amount produced by humans. So, urine alone cannot supply all the urea that is used. Nevertheless, urea from urine could make a significant contribution. The easiest way to start would be to collect urine from male urinals, storing it in tanks, and transporting it in tankers to enhance the composting of nitrogen-poor biological wastes, like straw, or to add to anaerobic biomass digesters, facilitating the growth of microbes that do the digesting, producing methane for generating electricity. Until recently, in Europe, as in other parts of the world, toilet facilities were in outhouses with earth closets, or simply outdoors. The human wastes went into the earth. Modern composting toilets make this process more efficient, but most people today live in cities rather than in the countryside, and often in apartment blocks. In these conditions, how can human excretions be recycled rather than literally wasted? One solution would be to have a separate plumbing system for toilets instead of combining their effluents with other aqueous wastes, such as those from kitchen sinks, baths, and showers. Nontoilet fluid waste can be treated separately and can more easily be regenerated into usable water. The toilet effluents from an urban neighborhood, or a large apartment building, could be piped directly to a local biogas digester. Food wastes could also be collected and added into it. Local biogas plants would both generate electricity and produce heat, which would be used in local heating systems. Animal wastes are already used in this way. In the Indian subcontinent there are about 2 million domestic gobar biogas digesters, which provide cooking gas for a family through the digestion of cow dung and provide valuable fertilizer from the residues. In China there are even more. In Europe, there are now many large-scale urban biogas digesters primarily using food wastes. On some farms, the slurry from cow barns is fed into these digesters and mixed with grass and other sources of biomass. These technologies could relatively easily be adapted to use human wastes, and indeed some sewage plants already include anaerobic digesters. The liquid residues from anaerobic digesters are a good organic fertilizer, rich in nitrogen and other plant nutrients, and can be applied to the land as liquid fertilizer, after suitable dilution.

1.6.6 Using weeds Weeds are by definition plants that are growing in the wrong place. They are unwanted because they compete with crops and reduce their yields. But many weeds are weeds precisely because they are hardy and grow vigorously. These could be advantages when the desired harvest is biomass. Although most weed species are useless for human consumption, or as animal feed, some might be useful crops for producing biomass for digesters. The feasibility of using weeds in biogas digesters has already been established with water hyacinth (Almoustapha, Kenfack, & Millogo-Rasolodimby, af

https://www.yara.com/siteassets/investors/057-reports-and-presentations/other/2018/fertilizer-industry-handbook-2018-with-notes.pdf/ (Retrieved 11 June 2019).

20

Rethinking Food and Agriculture

2008), which is a highly invasive water weed in many tropical and subtropical regions, and is one of the fastest growing plant species on earth.

1.6.7 Phase out the use of food crops for biofuels Partly driven by concerns about climate change, in some parts of the world food crops are used not for food but for the production of liquid fuel, mainly ethanol. In the United States, for example, about 40% of the total maize production is used for ethanol, and 36% for animal feed. Most of the rest is exported. Only a small fraction of the total maize is used as food for Americans, and much of it is in the form of high-fructose corn syrup. The energy balance of ethanol production from maize is very modest— there is only about twice as much energy in the ethanol as in the fossil fuels required to grow the maize and to produce the ethanol, compared with eight times more energy with alcohol produced from sugar cane in Brazil.ag The production of all this maize is heavily subsidized by the US government—in 2012, by about $20 billion. And much of this government money benefits the giant food companies that dominate the US market: Archer Daniels Midland and Cargill, the biggest privately owned corporations in the world. These companies in turn invest heavily in lobbying the US government and in shaping the system of subsidies that brings them so much profit (Pollan, 2006). These subsidies could be used very differently—for example, to help reduce runoff and erosion, to improve soil quality, and to promote more diverse and sustainable cropping systems. This is largely a political problem, but there is great scope for research in developing agricultural alternatives to the present system. In terms of energy production, there are already alternatives to liquid fuels, as in solar and wind power, which will become increasingly important as the shift to electric vehicles accelerates.

1.6.8 Reducing demand through human dietary changes According to an analysis in 2011, about 75% of all agricultural land, including pastureland, is dedicated to animal production. The proportion of cereal grains, soya beans, and other crops for feeding to intensively farmed animals amounts to about 24% of total crop production by mass, and about 36% by calorie content. In addition, livestock production is responsible for about 18% of total greenhouse gas emissions (Cassidy, West, Gerber, & Foley, 2013). The conversion efficiency of all this food into meat is very low. In terms of calories, only about 10%–12% of the calories fed to animals ultimately contribute to human diets through meat and other animal products (Foley, 2013). If plant foods were eaten directly by humans, far more food would be available for the world’s growing population. As one recent analysis showed, “Growing food exclusively for direct human consumption could, in principle, increase available food calories by as much as 70%, which could feed an additional 4 billion people.” Even small shifts in the allocation of ag

USDA: 2015 Energy Balance for the Corn-Ethanol Industry, https://www.usda.gov/oce/reports/energy/2015EnergyBalanceCornEthanol.pdf (Retrieved 12 June 2019).

Setting innovation free in agriculture 21

crops from animal feed could significantly increase the availability of food globally (Cassidy et al., 2013). Raising animals for meat also uses large quantities of water. Excessive meat consumption also increases the risk of several kinds of cancer, particularly colon and rectal cancer (Larsson & Wolk, 2006). In some Western countries there is already a rise in the proportion of people eating vegan, vegetarian, and “flexitarian” diets, reducing the consumption of meat, and for the European Union as a whole the consumption of meat, especially beef, is expected to decline by 2030.ah There are also a number of innovations for producing meat substitutes from fungal protein, insect proteins, and cultured animal cells, which could further help to reduce the demand for feed grains, soya beans, and oilseeds for animals in factory farms. Other dietary changes that could have large effects on overall food supply would be an increased use of drought-tolerant crops like sorghum and millet. In many parts of India, for example, where water is in short supply, much of the available water is used to grow a relatively small acreage of rice, the most water-demanding crop. But if the same water were spread more thinly, especially using sorghum or millet, crop yields would increase, overall food production would go up, and the food supply would also be more nutritious (Davis et al., 2018). Through a combination of government policies, changes in subsidy systems, marketing, and working with food industries, preferences could change, and the overall agricultural system could become more efficient through reducing the production of rice in favor of more water-efficient crops. Thus some of the most important factors shaping agriculture in the decades to come will be changes in dietary habits. These are outside the sphere of farming itself but no discussion on the future of agriculture can ignore them. And the economic forces that affect agricultures are not simply a result of humans’ dietary desires and disposable incomes; they are affected by the economic interest and lobbying power of agricultural and food businesses, food advertising, and by government policies, including taxation and subsidies. Although imposing a tax on meat or rice consumption would probably be politically difficult in most countries, governments can make major changes by altering the system of subsidies. In the United States, for example, large subsidies encourage the production of feed grains, especially maize, soya beans, and other crops used in factory farming, and thus indirectly subsidize the intensive production of meat, making it much cheaper than it would otherwise be. In India, government subsidies have favored the production of irrigated wheat and rice over other cereals, thus distorting the market in favor of the wasteful use of water.

1.6.9 Pragmatic plant breeding Long before the time of Charles Darwin and Gregor Mendel, people were breeding plants and animals by breeding from variants that appeared spontaneously, or by making crosses between promising parents and selecting among the offspring. Many of ah

EU Agricultural Outlook 2018–2030, https://ec.europa.eu/info/news/eu-agricultural-outlook-2018-2030-changing-consumer-choices-shaping-agricultural-markets-2018-dec-06_en (Retrieved 10 June 2019).

22

Rethinking Food and Agriculture

breeds of dog, including Pekinese, Afghan hounds, and sheepdogs, long predate the era of genetics. So do the many kinds of cabbage, including kale, broccoli, and Brussels sprouts, all members of the same species Brassica oleracea. Indeed, it was precisely such examples of pragmatic animal and plant breeding that provided Charles Darwin with much of the raw material for his thinking about the power of selection. His book The Origin of Species (1859) contains many examples, and he goes into more detail in The Variation of Animals and Plants Under Domestication (1868). Likewise, Mendel was able to work out some of the principles of inheritance because different varieties of peas already existed, thanks to the activities of plant breeders. The successes of plant breeding came first. Genetics followed. For most of the 20th century, governments funded the breeding of agricultural crops in agricultural institutes and universities. Their aim was to produce new varieties suitable for use by farmers, and the farmers only had to buy their seed supplies once; they could use their own seed thereafter. In Britain, for example, the Plant Breeding Institute at Trumpington, near Cambridge, funded by the UK government through the Agricultural Research Council, produced many new varieties, some still grown commercially, like Maris Piper potatoes and Maris Otter barley, used in brewing. But since the 1980s, agricultural research has been largely privatized. The Plant Breeding Institute itself was privatized in 1987 and sold to Unilever, who sold it on to Monsanto in 1998. The amount of research work declined. The buildings were demolished in 2009, and Monsanto sold the land to developers for housing.ai The concept of plant breeding and agricultural research in the spirit of public service had been replaced by the goal of maximizing corporate profits. Profits are greatest if farmers have to buy seed afresh each year, rather than saving their own from previous years. Under the older system of Plant Breeders’ Rights, farmers are allowed to use their own seed in subsequent years. But, corporations favor a system whereby they can patent seeds, and then force farmers to buy them over and over again. Monsanto successfully sued hundreds of farmers in the United States for using their own seeds on the grounds that they were infringing Monsanto’s patents, and won more than $23 million from the farmers (Harris, 2013). In 2018, the German-based Bayer Corporation acquired Monsanto. Three multinational corporations, Bayer/Monsanto, Syngenta, and DuPont, now control the majority of the world’s seed market. Coinciding with the wave of privatization in the 1980s was a rising optimism about the prospect of exploiting genes through genetic engineering. In traditional plant breeding, the selection of desirable characteristics came first, and genetic analysis followed. In genetic engineering, a gene has to be identified first and then engineered into a plant through biotechnology. Thus a gene from Bacillus thuringiensis coding for an insecticidal protein was transferred to cotton and maize, rendering them poisonous to insects. Likewise, a gene from a bacterium that can metabolize the herbicide glyphosate (marketed as Roundup by Monsanto, and now Bayer) has been transferred to a wide range of crops, most notably soya beans. By ai

http://www.trumpingtonlocalhistorygroup.org/subjects_PBIhistory.html (Retrieved 11 June 2019).

Setting innovation free in agriculture 23

making these crops “Roundup Ready” a whole field can be sprayed with Roundup to kill all plants except the Roundup Ready crops, which have enzymes that can destroy this poison. The majority of soya beans planted in the United States are now genetically modified (GM) so that they can break down Roundup. This system promotes the sales of Roundup, and creates extreme monocultures, but it does not necessarily lead to higher yields. In 1999, in an analysis of more than 8000 field trials in the United States, the Roundup Ready soya beans yielded on average 6.7% less than conventional varieties (Benbrook, 1999). This “yield drag” is probably a by-product of the technical process of genetic modification, which involves collateral damage to other genes. Monsanto claims that a new version of Roundup Ready overcomes this problem. However, the company’s own data show that its GM variety yields less than the conventional variety the trait is inserted into.aj Despite the many promises made about the transformation of agriculture by genetic engineering, the practical applications are limited by the fact that most characteristics of crops are not controlled by single genes that can be engineered into the crop. In fact genome-wide association studies in both plants and animals have shown that for most complex hereditary traits, dozens or even hundreds of genes are involved, most of which have small effects. These characteristics are “polygenic” and cannot be controlled by inserting or deleting a single gene, or even two or three genes, by genetic engineering (Sheldrake, 2020). For similar reasons, the recent technique of gene editing through the CRISPRCas9 system may have some limited uses in specialist situations, but, as in the case of genetic engineering, its highly focused approach on single genes, or even single basepairs within single genes, cannot address important characteristics in crop plants that depend on large numbers of genes. One aspect of a new pragmatism could be to breed crops as mixtures. Planting mixed varieties of a particular crop, like wheat, could reduce susceptibility to disease compared with a monoculture of a pure line. Genetically mixed crops could have survival and yield advantages, because instead of all the plants reacting in the same way to pests, diseases, and adverse climatic conditions, some respond differently and more effectively than others. For example, in a large-scale test in China, a mixture of rice varieties was sown on thousands of farms. The ravages of rice blast, the most significant fungal disease of rice, were reduced to acceptable levels without the use of any fungicide. As a report in Nature put it, “This approach is a calculated reversal of the extreme monoculture that is spreading throughout agriculture, pushed by new developments in plant genetics” (Wolfe, 2000). In summary, a combination of the molecular paradigm, biotechnology patents, and corporate empire building has distorted the field of plant breeding. Pragmatic plant breeding, aided by molecular genetic technologies where appropriate, is more likely to result in the breeding of better crops than the single gene approach. Breeding crops that can be grown as mixtures may confer further benefits in insect, disease, and drought resistance, giving greater stability of yield.

aj

http://www.aphis.usda.gov/brs/aphisdocs/06_17801p.pdf (Retrieved 15 June 2019).

24

Rethinking Food and Agriculture

1.6.10 Epigenetic inheritance and its possible applications In the 20th century, the inheritance of acquired characteristics, or “Lamarckian inheritance,” was treated as heretical in Western biology. It contravened the neo-Darwinian theory of evolution, which explicitly denied that such inheritance was possible, and focused on the natural selection of gene frequencies within interbreeding populations and on random mutations of genes. This scientific question was also heavily politicized during the period of the Cold War. In the Soviet Union, the orthodox school of biology, under the leadership of Trofim Lysenko, strongly favored the inheritance of acquired characteristics. Mendelian geneticists were persecuted. Yet, since the turn of the millennium, the inheritance of acquired characteristics has been rebranded as “epigenetic inheritance” and has become mainstream. This is an area of very active research in biology. There is now no doubt that plants sometimes inherit characteristics acquired by their parents or more remote ancestors in response to the conditions under which they were grown. In the 1960s and 1970s this was clearly demonstrated by a pioneer of what we now call epigenetic research, Alan Durrant, whose research on flax showed striking and enduring changes caused by soil fertility. For example, flax plants grown with high levels of nitrogen fertilizer became taller and less branched, and this feature appeared in subsequent generations, even without elevated levels of nitrogen in the soil (Durrant, 1962). Durrant was swimming against the neo-Darwinian tide, but recent research has revealed many other examples of epigenetic inheritance. Similar epigenetic effects occur in animals (Miska & Ferguson-Smith, 2016). Yet in agriculture, it is still generally assumed that seeds of a given variety are simply carriers of the DNA of that variety and are not influenced by the conditions under which they were grown. In the light of research on epigenetics this seems unlikely. There may well be epigenetic effects in crop plants; the conditions under which the seeds or vegetative propagules are grown may have an unsuspected influence. Indeed, recent research on epigenetic inheritance in plants points to possible useful applications (Hauser, Aufsatz, Jonak, & Luschnig, 2011). In many plant species there are inducible defense systems, whereby after an initial attack the plant produces substances that enhance its ability to withstand further attacks. One defense signaling system depends on the production and movement within the plant of jasmonic acid and related metabolites. In experiments with Arabidopsis and with tomatoes, plants were exposed to caterpillars that ate their leaves, and their descendants, grown from seed, were assessed for resistance to caterpillar attack. The growth of the caterpillars was reduced by about 40% relative to controls. This resistance carried over to the second generation, even when the first generation descended from attacked plants were not exposed to caterpillars. The resistance faded out in the third generation when there were no further attacks (Rassman et al., 2012). This epigenetically inherited resistance depended on the ability of plants to mobilize the jasmonate signaling system and to produce small interfering RNAs, which are known to play a part in the epigenetic modulation of gene expression (Henderson & Jacobsen, 2019). In other studies with Arabidopsis, the exposure of plants to drought also had epigenetically heritable effects (Zhang, Fischer, Colot, & Bossdorf, 2013). So did exposure

Setting innovation free in agriculture 25

to high temperatures. In one study with Arabidopsis, plants were exposed to mild heat (30°C) in the parental and F1 generation, and then grown at normal temperatures in the F2 generation. They were again exposed to heat in the F3 generation. In this third generation, the plants grown at 30°C produced six times more seeds when their ancestors had been exposed to heat than control plants whose ancestors had been grown at a normal temperature (Whittle, Otto, Johnston, & Krochko, 2009). In this context, it would be well worth reexamining the archives of Soviet biology, where what we now call epigenetic phenomena were widely investigated in the context of agriculture from the 1930s to the 1950s. In the West it was generally assumed that all these results must be fraudulent or pseudoscientific because the inheritance of acquired characteristics was believed to be impossible. Now that epigenetic inheritance is known to be real, it seems likely that biological and agricultural journals from the Soviet era could contain much useful information. People with a good knowledge of genetics, epigenetics, and Russian could be employed to carry out an extensive review of this literature and compile a series of review articles that might well provide starting points for new investigations. Heritable epigenetic effects often fade out over time, so the biggest effect would be in the generation following the exposure of their parents to environmental conditions that induce an adaptive response. If these effects were to be implemented practically, then this system would be rather like the production of hybrid maize or other hybrid crops. To use epigenetically improved seeds, then seed material for the next generation would be produced afresh each year, just like an F1 hybrid. Such a system might even enable organic farmers whose crops suffer from severe insect damage to recoup some of their losses. If the farmers have low seed yields from insect-damaged crops, they may be able to obtain a premium price for these seeds if they confer a greater resistance to insect attack on the next generation.

1.7 Setting innovation free Most of current agricultural research is not determined by questions from farmers or gardeners, but by developments within academic and corporate science, and above all by the prospects of patenting genetic modifications or edited genes or promoting agrochemicals. A complementary approach would be to ask farmers and gardeners what problems they face and what kinds of solutions they are hoping for. Several years ago, on a BBC radio news program, there was a discussion of science funding. An interviewer asked a woman who had an allotment what kind of question she would ask to researchers, if she had a chance to do so. She replied that she was growing carrots organically, and they suffered from carrot root fly attacks. Someone had told her that if you put grass clippings on the soil it could reduce these attacks. She wanted to know whether this was true and how she could do it most effectively. The interviewer then spoke to the head of the UK Biotechnology and Biological Sciences Research Council, a government body that channels taxpayers’ money into biological research, asking if this was the kind of question they were addressing, or could address. He

26

Rethinking Food and Agriculture

replied that they were dealing with much more fundamental and important issues like the potential for the genetic modification of crops and the sequencing of genomes. There are probably hundreds of such questions that could be answered by research. Some may already have been answered. If there was a system whereby farmers and gardeners could submit questions, receive answers if they already exist, and open up new lines of research if answers do not already exist and if the question seems worthwhile, these practical problems could be a fruitful stimulus for new research. Another approach would be to reward creativity. Many millions of people throughout the world grow food plants in gardens, allotments, or farms. Many of them have had years of experience. Some are creative and experimental. But at present, if a farmer or gardener comes up with an innovation, it is unlikely to be of interest to academics in universities or to global corporations. And there is little or no possibility for an ordinary person to communicate with academics and researchers. If systems were set up whereby innovations could be submitted as entries for prizes that rewarded the development of new crops, new cropping systems, or new horticultural or agricultural practices, then there would be an incentive for independent creativity by people who are not professional researchers. As well as innovations flowing from top down, as in the case of new crop varieties, GM plants, agrochemicals, and new machinery, innovations could flow from the bottom up, from people who are literally on the ground and who have carried out their own investigations and experiments. One of the problems with Thomas Kuhn’s analysis of paradigm change in science is that it seems to validate dictatorships. When a new paradigm comes into force, the old model is rejected or discredited. A new orthodoxy replaces the old one. This is a model of revolutions akin to political revolutions in which one authoritarian regime is replaced by another. Perhaps the next kind of scientific revolution could be different, and instead of ushering in a new orthodoxy, it could open the way to a tolerant pluralism. This is what I hope for. Innovation will be most free when no particular orthodoxy achieves a monopoly of power, or of funding, and when scientific research in general, and agricultural research in particular, are carried out pragmatically, liberated from the dogmas of materialism, molecular triumphalism, and neoliberal capitalism, and placing a major emphasis on the building up of the quality of the soil and on sustainability.

References Acton, L. (2011). Allotment gardens: A reflection of history, heritage, community and self. Papers from the Institute of Archaeology, 21, 46–58. Almoustapha, O., Kenfack, S., & Millogo-Rasolodimby, J. (2008). Biogas production using water hyacinths to meet collective energy needs in a Sahelian country. Field Actions Science Reports, 1. http://journals.openedition.org/factsreports/134. (Retrieved 5 November 2019). Bacon, F. (1951). The advancement of learning and new Atlantis. London: Oxford University Press. Benbrook, C. (1999). Evidence of the magnitude and consequences of the roundup ready soybean yield drag from university-based varietal trials in 1998. http://citeseerx.ist.psu.edu/ viewdoc/download?doi=10.1.1.41.823&rep=rep1&type=pdf. (Retrieved 11 June 2019).

Setting innovation free in agriculture 27

Bender, S. F., Wagg, C., & van der Heijden, M. G. A. (2016). An underground revolution: Biodiversity and soil ecological engineering for agricultural sustainability. Trends in Ecology and Evolution, 31, 440–452. Bergson, H. (1911). Creative evolution. London: Macmillan. Carr, B. (Ed.), (2007). Universe or multiverse?. Cambridge: Cambridge University Press. Cassidy, E. S., West, P. C., Gerber, J. S., & Foley, J. A. (2013). Redefining agricultural yields: From tonnes to people nourished per hectare. Environmental Research Letters, 8(3), 034015. Chivers, T. (2010). Neuroscience, free will and determinism: ‘I’m just a machine’. Daily Telegraph. October 12. Cohen, J. (2019). New ‘prime’ genome editor coud surpass CRISPR. Science, https://doi. org/10.1126/science.aaz9297 (Retrieved 5 November 2019). D’Espagnat, B. (1976). Conceptual foundations of quantum mechanics. Reading, MA: Benjamin. Davis, K. F., Chiarelli, D. D., Rulli, M. C., Chhatre, A., Richter, B., Singh, D., & DeFries, R. (2018). Alternative cereals can improve water use and nutrient supply in India. Science Advances, 4. eaao1108. Druille, M., Cabello, M. N., Omacini, M., & Golluscio, R. A. (2013). Glyphosate reduces spore viability and root colonization of arbuscular mycorrhizal fungi. Applied Soil Ecology, 64, 99–103. Dubos, R. (1960). Pasteur and modern science. New York: Anchor Books. Durrant, A. (1962). The environmental induction of heritable changes in Linum. Heredity, 17, 27–61. Edmondson, J., Davies, Z. D., Gaston, K. J., & Leake, J. (2014). Urban cultivation in allotments maintains soil qualities adversely affected by conventional agriculture. Journal of Applied Ecology, 51, 880–889. Ellis, G. (2011). The untestable multiverse. Nature, 469, 295. Fara, P. (2009). Science: A four thousand year history. Oxford: Oxford University Press. Foley, J. (2013). It’s time to rethink America’s corn system. Scientific American. March 6. Gleik, J. (1988). Chaos: Making a new science. London: Heinemann. Goff, P. (2019). Galileo’s error: Foundations for a new science of consciousness. London: Rider. Goss, M. J., Carvalho, M., & Brito, I. (2017). Functional diversity of mycorrhiza and sustainable agriculture: Management to overcome biotic and abiotic stress. London: Academic Press. Gould, S. J. (1989). Wonderful life: The burgess shale and the nature of history. London: Hutchinson. Harris, P. (2013). Monsanto sued small farmers to protect seed patents, report says. The Guardian. February 12. Hart, M., Antunes, P. M., Chaudhary, V. B., & Abbott, L. K. (2017). Fungal inoculants in the field: Is the reward greater than the risk? Functional Ecology, 32, 126–135. Hart, M., & Trevors, J. T. (2005). Microbe management: Application of mycorrhizal fungi in sustainable agriculture. Frontiers in Ecology and the Environment, 3, 533–539. Hauser, M. T., Aufsatz, W., Jonak, C., & Luschnig, C. (2011). Transgenerational epigenetic inheritance in plants. Biochimica et Biophysica Acta, 1809, 459–468. Henderson, I. R., & Jacobsen, S. E. (2019). Epigenetic inheritance in plants. Nature, 447, 418–424. Horgan, J. (1997). The end of science: Facing the limits of knowledge in the twilight of the scientific age. London: Little, Brown and Co. Howe, D., & Rhee, S. Y. (2008). The future of biocuration. Nature, 455, 47–48.

28

Rethinking Food and Agriculture

Jaganathan, D., Ramasamy, K., Sellamuthu, G., Jayabalan, S., & Venkataraman, G. (2018). CRISPR for crop improvement: A review. Frontiers in Plant Sciences, 9, 985. Kealey, T. (1996). The economic laws of scientific research. London: Macmillan. King, F. H. (1911). Farmers of forty centuries, or, permanent agriculture in China, Korea and Japan (p. 9). Madison, WI: King. Kuhn, T. S. (1962). The structure of scientific revolutions. Chicago: Chicago University Press. Laplace, P. (1819, Reprinted 1951). A philosophical essay on probabilities. New York: Dover. Larsson, S. C., & Wolk, A. (2006). Meat consumption and risk of colorectal cancer: A ­meta-analysis of prospective studies. International Journal of Cancer, 119, 2657–2664. Latour, B. (1987). Science in action: How to follow scientists and engineers through society. Cambridge, MA: Harvard University Press. Makowsky, R., Paieweski, N. M., Klimentidis, Y. C., Vazquez, A. I., Duarte, C. W., Allison, D. B., & de los Campos, G. (2011). Beyond missing heritability: Prediction of complex traits. PLoS Genetics, 7(4), e1002051. Manoharan, L., Rosenstock, N. P., Williams, A., & Hedlund, K. (2017). Agricultural management practices influence AMF diversity and community composition with cascading effects on plant productivity. Applied Soil Ecology, 115, 53–59. Miska, E. A., & Ferguson-Smith, A. C. (2016). Transgenerational inheritance: Models and mechanisms of non-DNA sequence-based inheritance. Science, 354, 59–63. Munowitz, M. (2005). Knowing: The nature of physical law. Oxford: Oxford University Press. Nature. (2011). Editorial: Best is yet to come. Nature, 470, 140. Pisano, G. P. (2006). Science business: The promise, the reality and the future of biotech. Boston, MA: Harvard Business School Press. Pollan, M. (2006). The omnivore’s dilemma: The search for a perfect meal in a fast-food world. London: Bloomsbury. Popper, K. R., & Eccles, J. C. (1977). The self and its brain. Berlin: Springer International. Rao, M. R., & Willey, R. W. (1980). Evaluation of yield stability in intercropping: Studies on sorghum/pigeonpea. Experimental Agriculture, 16, 105–116. Rassman, S., De Vos, M., Casteel, C. L., Tian, D., Halitsche, R., Sun, J. Y., … Jander, G. (2012). Herbivory in the previous generation primes plants for enhanced insect resistance. Plant Physiology, 158, 854–863. Rhoades, R. E., & Rhoades, V. (2008). Agricultural anthropology: A call for the establishment of a new professional specialty. Practicing Anthropology, 2, 10–12. Ridley, M. (2011). Francis crick: Discoverer of the genetic code. London: Harper Perennial. Sarkar, M. N. I. (2017). An introduction to agricultural anthropology: Pathway to sustainable agriculture. Journal of Sociology and Anthropology, 1, 47–52. Sarton, G. (1955). Introductory essay. In J. Needham (Ed.), Science, religion, and reality. New York: Brzillier. Sheldrake, R. (1984). Pigeonpea physiology. In P. R. Goldsworthy, & N. M. Fisher (Eds.), The physiology of tropical field crops. London: Wiley. Sheldrake, R. (2013). Family orchards. Resurgence. Sept/Oct https://www.resurgence.org/magazine/article3989-family-orchards.html. (Retrieved 6 June 2019). Sheldrake, R. (2020). The science delusion: Freeing the spirit of enquiry (2nd ed.). London: Coronet. Smolin, L. (2006). The trouble with physics: The rise of string theory, the fall of a science and what comes next. London: Allen Lane. Thompson, K. (2014). Soil report shows we should all grow more of our own. Daily Telegraph. April 28. Venter, C. (2007). A life decoded. London: Allen Lane.

Setting innovation free in agriculture 29

Whittle, C. A., Otto, S. P., Johnston, M. O., & Krochko, J. E. (2009). Adaptive epigenetic memory of ancestral temperature regime in Arabidopsis thaliana. Botany, 87, 650–657. Wolfe, M. S. (2000). Crop strength through diversity. Nature, 406, 681–682. Zhang, Y. Y., Fischer, M., Colot, V., & Bossdorf, O. (2013). Epigenetic variation creates potential for evolution of plant phenotypic plasticity. New Phytologist, 197, 314–322.

This page intentionally left blank

Agriculture planted the seeds of alienation from nature☆

2

Jim Masona and Laila Kassamb a Independent Researcher, Washington, DC, United States, bAnimal Think Tank, Lancaster, United Kingdom

2.1 Introduction Our worldview is the product of tens of centuries of agriculture, which is the manipulation of plants and other animals for human wants and needs. With farming as the base of human subsistence, human effort and experience has been directed toward essentially two ends: harnessing the natural processes of domesticates and subduing the natural processes of their competitors, the untamed natural world. Farmers, in other words, have striven to take the laws of nature into human hands. Generations of farmer experience have contributed dozens of views and values that have conglomerated into a worldview that sees nature as part slave, part enemy, and always a thing to be on top of. Culture carries a human society’s experience, views, and values—its worldview— from generation to generation. Our worldview, then, has come to us by way of an agrarian culture, or “agriculture.” To understand the dominant Western worldview—a worldview that many believe is the main cause of our alienation from nature and our disastrous relations with the living world and each other—we need to understand its roots in agrarian culture. Only then can we begin to radically transform our attitudes and relations toward the living world and reimagine our place within it.

2.2 Dominionism: Our nature-dominating worldview and its origins “Man’s task,” wrote Sir Keith Thomas in Man and the Natural World, “in the words of Genesis (1: 28), [is] to ‘replenish the earth and subdue it’: to level the woods, till the soil, drive off the predators, kill the vermin, plough up the bracken, drain the fens. Agriculture stood to land as did cooking to raw meat” (Thomas, 1983, p. 15). That is, it makes raw, wild nature suitable for human use and consumption. Consider agriculture’s effect on the mind and culture. For the past 100 centuries, humans have controlled, shaped, and battled plants, our animal cousins,a and natural processes—all things of the living world that we put under the word nature. The culture ☆ a

This chapter is based on Mason (1993). We prefer cousins instead of “nonhuman” when referring to other animals. “Cousins” tends to promote a sense of kinship, which we need to replace prevailing alienation. “Nonhuman” implies negativity since human is generally regarded as good and positive, as in humanity, human rights, humane, etc.

Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00002-5 © 2021 Elsevier Inc. All rights reserved.

32

Rethinking Food and Agriculture

of farming regards any lives—plant or animal—in residence on the land as unworthy of human consideration. Those lives are contemptible pests, things to be eliminated to make way for crops and “livestock” that benefit human society. Controlling nature is second nature to us, so deeply ingrained in us that we are rarely conscious of it. It is a worldview we are born to, acculturated with—unthinking and unaware. To become more aware, let us identify it and name it: dominionism is most apt. The word dominion is defined as “a supremacy in determining and directing the action of others…” (Gove, 1976, p. 672). Our agriculture’s view of the living world is ­dominionist. Indeed, the West’s creation story in Genesis tells how GODb the creator expressly grants human beings dominion over His (sic) creation, that is, all of the earth. Human mastery and control of nature is GOD’s will, the Law, the Word, and sacred beyond question. This Law, this Word, is the written culmination of the stories and beliefs passed around in agricultural societies for centuries before writing—history—began around 3000 BC. All of them from the Middle East, the “cradle of civilization”—the Western kind at any rate. Beside the Western monotheistic religions, the other main source of dominionism is the “pagan” or polytheistic classical civilizations, especially ancient Greece where the ideas of Plato and his disciple Aristotle were most influential. Aristotle’s works provided fuel for medieval Christian doctrine as well as that of the Renaissance somewhat later. For him, nature was a hierarchy of beings, and man (sic), having superior mental capacities, reigned at the top. He wrote: Plants exist for the sake of animals and brute beasts, for the sake of man—domestic animals for his use and food, wild ones (or at any rate most of them) for food and other accessories of life, such as clothing and various tools…since nature makes nothing purposeless or in vain, it is undeniably true that she has made all animals for the sake of man. (Aristotle, 1959, p. 16)

Although several prominent Roman writers denounced the exploitation of our animal cousins, their views are generally ignored by later thinkers, and Romans (whose wealth began with cattle herding and trading) took dominion over nature for granted. Cicero sums up the Roman view: We are the absolute masters of what the earth produces…The rivers and the lakes are ours. We sow the seed, and plant the trees. We fertilize the earth by overflowing it. We stop, direct, and turn the rivers: in short, by our hands we endeavor, by our various operations in this world, to make, as it were, another nature. (Walsh, 2008, p. 74)

Centuries later, St. Augustine and other Christian thinkers embraced the dominionist ideas from Greek and biblical texts and restated them in the language of Catholicism. Most influential was St. Thomas Aquinas, a Dominican priest and a key player in the forging of modern dominionism from biblical and Greek stock. His admiration for Aristotle shows: Now the order of things is such that the imperfect are for the perfect…Things, like plants which merely have life, are all alike for animals, and all animals are b

Emphasized to deny the authors’ endorsement of the conventional God myth.

Agriculture planted the seeds of alienation from nature

33

for man. Wherefore it is not unlawful if men use plants for the good of animals, and animals for the good of man, as the Philosopher states (Politics, I, 3). (Bourke, 1951, p. 352)

Around 1500, as Europe’s Renaissance bloomed and as new continents were “discovered,” the idea of human mastery over nature grew new strength. It became a guiding idea for the arts and sciences of the era as they brought Europe out of the dens of the Dark Ages. Two men in particular injected new vigor into the older body of dominionism left by the Greeks and the church. Around 1600, Francis Bacon and René Descartes, regarded as the “fathers” of modern science, provided a formula for secularizing domination of nature in a period when capitalism began eroding the cultural strength of religion. Bacon’s human chauvinism swaggered even more than the Hebrews: “Man, if we look for final causes, may be regarded as the centre of the world, insomuch that if man were taken away from the world, the rest would seem to be all astray, without aim or purpose” (Bacon, quoted in Thomas, 1983, p. 18). And: “I am come in very truth leading you to Nature with all her children to bind her to your service and make her your slave” (Bacon, quoted in Hartmann, 2001, p. 121). Descartes’ contribution completely severed any connection between humans and nature. For him, humans were not only superior to, but distinctly apart from, the living world. He explained the behavior of our animal cousins by claiming that they move mechanically, like clocks. This beast-machine view of other beings was the best possible rationalization for the way humans actually treated other animals—not only in science, but in agriculture, war, entertainment, and other endeavors that made use of them. With the extra heft of secular justification, controlling nature became not only glorious but a moral imperative; nature conquerors were now noble improvers of both nature and the human condition. This became the intellectual bandwagon of the modern age—for its scientists, technocrats, social reformers, and radicals. Marxists—for all of their revolutionary ideas—toed the party line of dominionism. “In a Communist society,” wrote Maurice Cornforth in the 1950s, “people now go forward without limit to know and control the forces of nature, to use them as servants, to remake nature, cooperating with nature to make the world a human world since humanity is nature’s highest product” (Cornforth, 1971, pp. 167–168).

2.3 Before agriculture: A world alive and ensouled Before settled agriculture, tribal peoples regarded the world around them as alive and full of souls. Anthropologists call this animism; we call it the primal worldview. This outlook grew from reality, not ideology, for primal societies lived close to nature—in and with the life around them. Primal peoples lived essentially outdoors and regularly moved about their environment. This constant exposure to nature’s forces and other living beings gave them a feeling of embeddedness, of membership in the living world. Food and materials came not by controlling plants and other animals, but by incredibly detailed knowledge about them. In the popular

34

Rethinking Food and Agriculture

terminology, this lifeway is called hunting-gathering. But hunting, as we shall see, has been exaggerated (Harari, 2014; Hart & Sussman, 2008; Nitecki & Nitecki, 1987). We prefer the term foraging. The transition from foraging to farming was gradual and uneven. It took thousands of years and some primal societies, for example, in Australia, did not participate. In any event, if we want to understand what was lost in the transition, we need to look at the primal lifeways. These are not mere obsolete things of the distant past; they are the great base of human existence. All of our ancestors and peoples all over the world lived as foragers for the millions of years after hominid divergence from the other great ape species to the beginning of domestication c.10,000 years ago. There were many variant species of hominids along the way but the evolution culminates with us, modern humans, Homo sapiens. We sprang from a spontaneous mutation in Africa some 200,000 years ago, although the number changes as new fossil finds are authenticated. Those first modern humans did not come empty headed and handed; they came with all the skills, habits, tools, and worldview of their predecessors (actually, their older relatives). They were foragers with all of the experience—culture—of primal forager societies, only smarter and more adaptable (Harari, 2014). The popular belief is that forager peoples lived lives of starvation and hardship while farmers had plenty. But it appears that ancient primal people enjoyed relatively high standards of living (Harari, 2014). Archeological evidence from the Upper Paleolithic (30,000 years ago) indicates good diets and health. Using such indicators as average height, number of missing teeth, and age at time of death, scientists can assess health. It appears that only in recent times have people enjoyed health as good as that of our Paleolithic ancestors (Cohen, 1989). Anthropologist Marvin Harris estimates the average human lifespan during the Upper Paleolithic at 28.7 years for females and 33.3 years for males (Harris, 1977). Looking at contemporary forager people, studies of southern Africa’s !Kung San women show life expectancy to be 32.5 years. Bear in mind that the life expectancy for nonwhite males in the United States did not reach 32.5 years until around 1900 (Cohen, 1989). The !Kung San are usually referred to as “hunter/gatherers” but studies from the 1970s show that 75%–80% of their diet was made up of wild vegetable foods—not hunted meat (Cohen, 1989; Harari, 2014). The !Kung’s workdays were also light: although they live in a desert ecosystem, each adult needed less than 3 hours a day to provide a diet balanced in proteins and other essential nutrients. These studies show also that the great bulk of the food—as much as 80% in most regions—was gathered largely by women and children (Cohen, 1989). Other evidence suggests that plant foods collected, prepared, and shared by women provided the bulk of the primal diet (Sanday, 1981; Tanner, 1981). Beyond being the major food providers, women’s roles as child bearers, midwives, healers, shamans, and all-around nurturers gave them considerable female power and status in primal times. Before domestication and patriarchy, to be female was to be in continuum with the major powerful mysteries: menstruation, pregnancy, childbirth, the silent but potent plant world, the fecundity of the living world. Primal female powers derived from being seen as closer to nature’s forces.

Agriculture planted the seeds of alienation from nature

35

These realities undermine the theory of man-the-mighty-hunter created by writers such as Robert Ardrey (1976) and generations of male-biased anthropology (Mason, 1993, pp. 72–86). The theory amounts to a male creation myth, which, according to Tanner, “serve(s) to interpret and justify aspects of the Western economic system and some of Western society’s particularly chauvinistic, hierarchical, and warlike characteristics” (Tanner, 1981, p. 23). The man/hunter myth offers a secular creation story—the Genesis—for the patriarchy, capitalism, and dominionism of the modern worldview. It provides a story of our beginnings that makes us accept that these characteristics are “natural” and immutable. The hunter-creation myth also helps a meat-eating society with a very troublesome problem. People, generally, are more than a little uncomfortable with killing our animal cousins for food. Most would probably not be willing to kill another animal themselves except in dire circumstances. Even hunting peoples surrounded their hunting and butchering activities with ritual—much of it to ease anxiety and discomfort (Serpell, 1986). In the modern world of factory farms, mechanized slaughter plants, and shrink-wrapped cuts of meat, isolation from these activities keeps our ­awareness— and discomfort—to a minimum.

2.4 Other animals: Movers and shakers of the human mind and worldview The most important myth of a people is its creation story, for it sets out the elements of their worldview. Virtually universal in all creation stories is the prominence of our animal cousins. The West’s is no exception, for Genesis tells us that GOD created the world full of other animals before HE created man and woman; and soon an animal charms the woman to commit the original sin that drives them from paradise. Later on, when sin corrupts humanity and the angry GOD floods the world, Noah herds animals in pairs into his ark and restarts creation. Similarly, in a primal society the creation story explains how the world began and how the people came into being. The creation story is a base for a set of other stories that give the people an identity, a sense of place in the world, and a model for their lifeways. For the three major religions today, the Torah, Bible, and Qur’an are the written compendia of these stories. Primal creation stories are surprisingly similar the world over (Campbell, 1988a; Underhill, 1965). Whatever the society’s worldview, animal cousins are the First Beings. Typically, they are animals with human abilities—in animal form but they speak and behave like humans (as the serpent in the Garden of Eden). They are the creatures who shaped the world to its present state and, in many cultures, they are the ancestors of the people. Generally, this First Being is an animal cousin in the tribe’s environment respected or feared for having power of some kind—coyote in the western plains, tiger in Asia, crocodile and lion in Africa. Common in the primal creation stories, too, is the helplessness of humans when they appear. The older, wiser First Beings teach them how to make fire, tools, find food, and live in the world. They teach them also the rituals and ceremonies necessary to keep the world in order.

36

Rethinking Food and Agriculture

It is significant that our animal cousins stand out as First Beings in so many creation stories. Apparently, they were on our minds a lot even as our minds were evolving. Other animals are obviously more like humans than are trees, rivers, and other parts of the world. They move freely and have eyes, ears, hair, and other organs like us. They sleep, eat, defecate, urinate, copulate, give birth, shiver, play, fight, bleed, die, and carry on many of the same activities of life that we do. Somehow similar to us yet somehow different, other animals puzzled us and forced comparisons, categories, and conclusions. Other animals made us think. As biologist Paul Shepard explained: “There is a profound, inescapable need for animals that is in all people everywhere” as shapers of the human mind and thought (1978, p. 2). He wrote: It is the peculiar way that animals are used in the growth and development of the human person, in those most priceless qualities which we lump together as “mind”. It is the role of animal images and forms in the shaping of personality, identity, and social consciousness. Animals are among the first inhabitants of the mind’s eye. They are basic to the development of speech and thought. (Shepard, 1978)

We can see this in the furniture of the nursery—the teddy bears, animal figures on cribs and sleepsuits—and the storybooks, songs, and bedtime stories. Children love to “play animal.” At every phase of development, children are learning about the world with animal cousins as major informants. As Shepard noted: Every child is committed to the use of animal images in the shaping of his [her] own consciousness because thought arose in the past as an interaction between different animals and between people and animals. (Shepard, 1978, p. 3)

As infants, we are becoming aware of the world around us and we try to understand the confusion, we try to put things in order by identifying, sorting, and classifying them. So did our primal ancestors. Trees, rocks, mountains, and other parts of the world offered their budding minds tangible objects to categorize and name, but the behavior of other animals stimulated thoughts of concepts and intangibles. These are even more important to the formation of human culture—that is, the carrying of learning and human experience from one generation to the next. Culture is mainly information about nonobjects, things that cannot be seen and handled physically—things like hunger, patience, courage, selfishness. These intangible, invisible things would have been “seen” by early humans in the behavior of the other animals in their environment. Then their evolving minds related these behaviors to their own experience. Our animal cousins, then, provoked deeper thoughts about the world—thoughts like Who are they? Who are we? and the usual whys and wherefores of human existence. They offered a moving feast of wonders and puzzles for the mind. Animals are “good to think,” wrote Claude Levi-Strauss (1964), the pioneer in the science of child development. From childhood on, our animal cousins give the human mind a frame of reference to the rest of the living world. When speaking of worldviews, then, we may say that animals = nature. The heavy presence of animals on the human mind is seen in art and lore. The first art painted thousands of years ago on the cave walls of Europe and Asia depicts our animal cousins almost exclusively. According to prominent art historian

Agriculture planted the seeds of alienation from nature

37

Lord Kenneth Clark: “In all of the art since the cave paintings, it is probable that animals are represented more often than any other class of things in nature” (Clark, 1977, p. 14). In religious art, other animals are often the major symbols for teachings and moral themes: the dove for peace, the lamb for innocence. The bestiaries—the moral storybooks of the Middle Ages—relied almost exclusively on images of other animals to teach. For people who could not read, a picture of an animal was truly worth a thousand words. One scholar found over 5000 examples of the use of other animals in expressions. We use animals to describe human behaviors: one “grouses” or “dogs” or “badgers.” No other set of things in nature is found so often in speech (Clark, 1968).

2.5 Agriculture: A new relationship with nature, a new social order About 13,000 years ago the Earth’s great glacial icecap was melting away because of a warming trend. No one knows exactly how and why, but by about 10,000 years ago human groups at scattered locations from northern Africa eastward to India and Southeast Asia were gathering wild plant seeds and sowing and cultivating them. At about the same time, their hunting turned into tending herds of indigenous sheep and goats, and before long, cattle and other animals familiar to the barnyard of today. Hunter-gatherers—foragers—gradually became farmers. The transition occurred in fits and starts. The first farmers still moved with the seasons and kept up many of their foraging ways. But over time, foraging gave way to farming, sedentary lifestyles, village life, and dependence on domesticated plants and animals for food. There are various theories about why foragers turned to farming, all of which boil down to creeping human population density as the prime mover (Ucko & Dimbleby, 1969). In some ecosystems—near lakes, rivers, marshes, and places where plant and animal life were abundant and diverse—human numbers and density crept upward. Over time, as some groups became more and more dependent on these areas and gradually shifted from nomadic foraging to semisedentary living, they lost many of their forager skills and knowledge of their old hunting and gathering lands. Little by little, they became circumscribed—locked in by their neighbors—and they had to intensify food procurement. They expanded their shopping list and diversified their diets and foraging activities. They learned how to weed and enlarge—cultivate—a favorite stand of wild grains such as wild wheat (Harari, 2014). Scientists believe that when people gathered the grains, the looser seeds dropped to the ground to be taken by mice, birds, and the elements. By natural variation, some plants had tighter seed heads and these were more likely to be gathered. From long forager experience, they knew about plant growth and the progression from seed to sprout to plant. They began to save some of their harvest to reseed the wild stand, and eventually the tighter-headed varieties took over the field. Circumscribed as well, hunters had to adapt. Instead of stalking far and wide to hunt fast, elusive, large animals like the red deer, they preyed on the local herds of gazelles, goats, and sheep. In the hilly regions of what is today the Kurdish region of northeastern Iraq, some tribes who specialized in hunting wild sheep and goats began to assert greater control

38

Rethinking Food and Agriculture

over the herds (Harari, 2014). They may have begun by selecting certain animals— males and younger animals. Over time this systematic selection of animals became a primitive form of herd management. Because they were still totemic—i.e., respectful of the Animal Powers of older times—perhaps mythic and ritual factors also drove the selections. A tribe may have attached itself to a particular herd, which became “its” herd to follow, cull, and live from. Moving and living with their herd, they became familiar enough that the animals’ fear of humans waned allowing a closer presence. In time, the tribe was able to assert some control over the migrations of their herd by guiding it to the best grazing and watering places. Little by little, these early herders learned the rudiments of animal husbandry—the techniques of control, mobility, diet, growth, and reproductive lives of “their” herd. In the course of centuries, they learned about the dynamics of sex, reproduction, and the roles of males and females. In time, they learned selective breeding, the selection of males and females to breed for offspring with desirable characteristics. In its earliest stages, the desired characteristics may have been coat color and pattern, horn shapes and sizes, and other visible characteristics to help an “owner” tribe distinguish its herd from those of neighboring herding tribes. Life during the change from foraging to farming was not always rosy. The new strains of plants and animals were not reliable producers of food and materials. With their genetic base narrowed by human selection and grown in monocultures, domesticated species were vulnerable to diseases and pests. Early farming methods were crude and inefficient. Some years brought good harvests, but just as often there were problems, shocks, and setbacks. In farmers’ villages, disease hit like never before. Living in one place in larger numbers meant living on top of their own refuse and wastes. Parasites, insects, rodents, and other disease carriers established their life cycles in the new human-made ecosystems of croplands, human/domestic animal hosts, and refuse heaps. With less and less variety in their diets, people fell to waves of epidemics such as cholera and typhoid as well as zoonotic diseases such as smallpox, measles, and tuberculosis (Harari, 2014), probably for the first time in human history. Regional crowding and increased trade ensured their spread from village to village. Left behind was the richly varied, nutritious diet of foragers and in its place farmer-­ villagers existed on bread, gruel, and a handful of starchy staples. Comparative studies of bones show that foragers had generally better health than farmers. About 30,000 years ago at the peak of the glacial period, adult males averaged 177 cm (5 ft, 11 in.) in height and females about 165 cm (5 ft, 6 in.). About 10,000 years ago, when farming first appeared, males had shrunk to the size of Ice-Age females and women averaged about 153 cm (5 ft). At the peak of the Ice Age, the average person died with only 2.2 teeth missing. By 6500 BCE, dental loss had risen to 3.5 teeth and by Roman times it had risen to 6.6 (Cohen, 1989). Despite poorer health, the human population grew steadily in agrarian times due to cultural changes (Cohen, 1989). For example, sedentary living probably relaxed the nomads’ taboos against frequent sexual activity and pregnancy. Children’s births did not have to be spaced years apart because they were no longer the burden they were when the forager band moved frequently (Harris, 1977).

Agriculture planted the seeds of alienation from nature

39

And all the while, nomadism and foraging were diminishing as options in the increasingly crowded Middle East. Slowly, creeping sedentism and farming took over, bringing drudgery, monotony, crowding, famine, disease, bad teeth, and lots of new grounds for violence between individuals and groups. The emerging agrarian way brought pressure to produce surpluses to carry the village through bad weather and lean years. Planters cleared more land, expanded fields, and diverted streams for irrigation. Herders expanded grazing lands and the sizes of their herds. All eliminated plants, other animals, forests, natural ecosystems—anything that hindered agricultural expansion to add to the surpluses. Wealth, in a word, brought more strains on the land and living world. As the quest for wealth took on a life of its own, it propelled farming regions into a spiral of self-generated conflict and tragedies. It pushed up all factors—human density, disease, environmental depletion, and competition for land, water, and materials. To produce more wealth, the region had inadvertently produced more environmental and social problems. Eventually, many communities collapsed—wiped out by famine, disease, drought, soil depletion, or warfare. Once the agrarian life took over, the situation was one of growing populations in an ecologically restricted space. Under pressure, people took on new strategies for survival. One strategy was intensification—increased crop and herd surpluses. The striving for and valuing of surpluses led to concepts of property and wealth, which in turn led to the emergence of classes of some people ruled by others who had more. Another strategy was expansionism, which led to mass conflicts and militarism, which in turn led to slavery, subordinate classes, and the rule of the rich and powerful. But the old ways, especially the forager tribal beliefs, died hard. Against their weight the newly emerging agrarian religious beliefs were not always strong enough to maintain social control. Raw, physical power was though, and rulers applied their wealth to building armies, which both protected the people of their realm and ensured their loyalty, which in turn preserved the wealth. It was a short step, then, to using the army to expand wealth by taking the lands and herds of a weaker neighbor. Thus warfare, conquest, and ruthlessness as means of social control and wealth building were built into the new agrarian social order in the earliest stages. The agrarian view of life as perpetual punishment was taking shape centuries before Genesis was written.

2.6 Misothery: The reduction of animals and nature Domestication and animal husbandry brought economic advantages, but they also brought about an unsettling of the old and deeply held totemic ways of seeing our animal cousins and the living world in awe. After centuries of animal husbandry, humans gained conscious control over other animals and their once-mysterious life processes. Castrated, yoked, harnessed, hobbled, penned, and shackled, with their sex lives controlled for human gain, domestic animals were thoroughly subdued. They had none of that wild, mysterious power of their wild ancestors. They were disempowered and reduced, and they came to be seen more with contempt than awe.

40

Rethinking Food and Agriculture

In reducing other animals to domestication, farmers and herders reduced our animal cousins in general and with them the living world that they had symbolized. Cropconscious farmers saw wild animals as pests and natural elements as threats. But it was the reduction of our animal cousins through husbandry that was the main driver of the radically different worldview that came with the transition from foraging to farming and herding, for it broke up the old totemic ideas of kinship and continuity with the living world. This, more than any other factory, accelerated human alienation from the living world. When herders and farmers intensified their uses of other animals, they needed some ways to resolve their beliefs in the animal spirit-powers. They needed to move away from forager beliefs in other animals as the First Beings and the souls of the living world. They also needed new ways to deal with the greater load of guilt that came with greater control and exploitation. They came up with a set of beliefs (referred to here as misothery) and cultural “distancing devices,” which replaced the older guilt-reducing rituals of the hunt and of animal sacrifice. Mason (1993) coined the word misothery because no word in the English language adequately expresses the full range of hate, contempt, loathing, disgust, fear, and all the other negative views and feelings that humans have for other animals. And since views of other animals determine views of the living world, misothery can mean hatred and contempt for nature—especially its animal-like aspects (Mason, 2017, pp. 135–151). Like misogyny, misothery reduces the power, status, and dignity of an Other. Misogyny reduces female power, status, and dignity, thus supporting male supremacy and control of women in the system we call patriarchy. Misothery reduces the power, status, and dignity of our animal cousins and nature, thus supporting human supremacy and control over the living world in a system we call dominion. Misothery, with other cultural devices, assisted in the demotion of other animals from powers and kindred beings to lowly beings and slaves. Domesticated animals lived a life of dependence upon farmers. Without emotional barriers, farmers could become too attached to “their” animals and working, driving, whipping, and slaughtering them would amount to a gross betrayal of trust and cause feelings of guilt and remorse. Farmers “have learned to cope with this dilemma using a variety of essentially dishonest techniques” (Serpell, 1986, p. 151). It is indicative of domestication’s impact on our worldview that these extend to all other animals and the entire living world. Serpell identifies four distancing devices: detachment, concealment, misrepresentation, and shifting the blame. Konrad Lorenz demonstrates detachment in his book Man Meets Dog: “Today for breakfast I ate some fried bread and sausage. Both the sausage and the lard that the bread was fried in came from a pig that I used to know as a dear little piglet. Once that stage was over, to save my conscience from conflict, I meticulously avoided any further acquaintance with that pig” (Lorenz, 1954, p. vii). Rather than give up pork and lard, Lorenz chose to give up closeness to pigs. Multiply this emotional transaction thousands of times over thousands of years and we can understand why agrarian culture alienates us from the living world.

Agriculture planted the seeds of alienation from nature

41

Concealment aids detachment by hiding stockyards, slaughterhouses, dog pounds, and the other places where the use of our animal cousins turns ugly. There is concealment in numbers: with thousands crowded and hidden in factory farms, there is no familiarity with any pig or cow. There is concealment in language to ease people’s conscience: beef, steak, pork, ham, and veal have concealed the dismemberment of other animals for their muscle tissue for centuries. Misrepresentation distorts the facts about other animals so that their sufferings and deaths seem necessary or deserved. If animals cause fear and loathing, it becomes morally easier to control, use, and kill them. And the nearer an animal cousin comes to posing an actual threat to human welfare—as do rats and wolves—the more intense the misrepresentation. The idea of animal evil is a very handy piece of misothery for agrarian society—so much so that it is kept sharpened and accessible through Western folklore. The most obvious example is the werewolf, which fed generations of Europeans with a morally righteous hatred for the “beast of waste and desolation” (Lopez, 1978). Such views have motivated centuries of herders to try to exterminate wolves both in Europe and North American, and they have very nearly succeeded. Blame shifting is a leftover from the old rituals of hunting and animal sacrifice, which shifted blame for the killing to ancestors or the gods. As part of the ritual animal sacrifice in ancient Babylonia, the priests bent down to the ear of the slaughter victim to whisper, “this deed was done by all the gods; I did not do it” (Serpell, 1986, p. 168). Holy men did the dirty work, and ever since division of labor has helped shift or diffuse the blame. Agrarian society has relegated the bloody work and emotional burdens of killing our animal cousins to butchers and slaughterhouse workers. Society tends to shun and debase these workers to make the killing seem somehow inevitable or natural because they are killers for hire and the meat eater can think: the killings will go on no matter what, so I am not to blame.

2.7 Herders’ legacies: The making of the Western worldview The Fertile Crescent of the Middle East was the epicenter of domestication of the large herd animals—sheep, goats, cattle, horses, and camels. This put a unique ingredient into the development of Western culture from the very start. As Alfred Crosby says, “The most important contrast between the Sumerians and their heirs, on the one hand, and the rest of humanity, on the other, involves the matter of livestock” (Crosby, 1986, p. 23). This ingredient created huge repercussions—results that are way out of proportion to the simple beginnings of animal domestication. As German sociologist and zoologist Richard Lewinsohn put it: the specific variations produced by domestication may be small in a zoological sense but they are enormous from the sociological point of view, for they have effected deep-reaching transformations in the history of both human beings and animals. (Lewinsohn, 1954, pp. 74–75)

42

Rethinking Food and Agriculture

Pastoralists, anthropologists’ term for herding societies, seems a misnomer as it evokes bucolic images of gentle shepherds watching over tranquil flocks and fields. As Lewinsohn says: Only in pastoral poetry were the shepherd folk peace-loving. In reality they were not far behind hunters in rapacity and belligerence. (Lewinsohn, 1954, p. 77)

The herd is everything to them, it is all their wealth, and nothing gets between the tribe’s herd and the best water and grazing lands. If anything does, the obsession turns to ruthlessness, defiance, and violence. Paul Shepard (1978) writes: Around the world common elements run through their cultures: an obsession with the goat, cow, horse, sheep, or camel so extreme that every aspect of life mediates or embellishes its image; aggressive hostility to outsiders, the armed family, feuding and raiding in a male-centered hierarchical organization, the substitution of war for hunting, elaborate arts of sacrifice, monomaniacal pride and suspicion. (p. 154)

British anthropologist B.A.L. Cranstone corroborates these views of herder culture after looking at herdspeople in Melanesia, North Africa, Syria, and Russian Turkestan. Since animals are the only form of wealth that is self-mobile, he says, their keepers need to be constantly on guard: “People who depend heavily on animals are, therefore, usually warlike because they have to be prepared to defend their herds” (Cranstone, 1969, p. 261). More warlike are those who herd camels and horses, for these are powerful animals and valuable because they can run far and fast. Less warlike are people who herd sheep, goats, and the smaller more controllable animals (Cranstone, 1969). Anthropologist and popular writer Marvin Harris has studied a variety of human cultures looking for common elements. In Cannibals and Kings, he noted that “most nomadic or seminomadic pre-state pastoral societies are expansionist and extremely militaristic” (Harris, 1977, p. 62). Anthropologist Homer Aschmann noted another dimension to the destructiveness of herding societies: their tendency to increase their herds invariably damages their rangeland. Aschmann wrote: No primarily herding society has ever achieved a stable ecological adjustment except at a lower level of productivity than the one that existed when pastoralism was introduced. (Aschmann, 1965, p. 268)

The continual need for the fresh grazing lands and water sources needed to sustain their ever-growing herds was thus also a key driver of the violence of herder peoples (Nibert, 2013). Herder subcultures were very influential in shaping the agriculture all around the Middle Eastern center. The influence of these warrior-invaders on Western culture has been written about by Riane Eisler (1988) who draws heavily from the investigations of archeologist Marija Gimbutas. They tell of several waves of pastoralists from the Eurasian steppes—“Kurgan” people—that swept across prehistoric Europe between

Agriculture planted the seeds of alienation from nature

43

4200 and 2900 BC, disrupting older farming communities and displacing egalitarian, goddess-worshipping cultures with patriarchy. A long line of Eurasian nomadic peoples carried on wars of invasion over long distances. Wherever they went, they took their horses, cattle, and male gods of war and mountains. And, wrote Eisler, “as Aryans in India, Hittites and Mittani in the Fertile Crescent, Luwians in Anatolia, Kurgans in eastern Europe, Acheans and later Dorians in Greece, they gradually imposed their ideologies and ways of life on the lands and peoples they conquered” (Eisler, 1988, p. 44). These waves of pastoralist invasions are believed by some to have shaped the violent and expansion-prone cultures of the West and increased the role of animal exploitation in the developing European economies (Nibert, 2013). Joseph Campbell says of human societies, “The earliest model was of the animal world and the hunt, where the animals slain were the sacrifice…” (1988b, p. 74). Just as animals provided a model for totemic society, so they did—though reduced—in agrarian society. Sir Keith Thomas (1983) wrote that: domestication…became the archetypal pattern for other kinds of social subordination. The model was a paternal one, with the ruler a good shepherd—like the bishop with his pastoral staff. Loyal, docile animals obeying a considerate master were an example to all employees. (p. 46)

This model and these pastoral images are pervasive in Western culture. We look reverently upon notions of the Good Shepherd and the benevolent patriarch because our agrarian culture is so heavily weighted with the herders’ legacies. The good, civilized man shepherds or husbands his family and property, his household, his community, his nation, and, by extension, the entire living world. The Good Shepherd model produced the modern secular idea of stewardship, which is supposed to be the benign control and exploitation of the living world. It provided a model for the patriarchal nation-states of early civilization: the good king tended his flock of subjects. The king had absolute authority over his people just as the shepherd ruled over this flock and the patriarch ruled over his clan. The subjects owed complete loyalty and submission to the king just as the sheep obeyed the shepherd and the clan obeyed the patriarch. Over time, the model elevated many Middle Eastern kings to the level of gods. The model grew into the West’s monotheistic religions, whose great king of heaven and earth tends over the flock of the faithful. Sociologists Gerhard and Jean Lenski find the same concept of one all-powerful god in many herding societies: “In forty of the fifty herding groups for which [George Peter] Murdock provides data, there is belief in a Supreme Deity who created the world and remains actively concerned with its affairs, especially with a people’s moral conduct” (Lenski & Lenski, 1982, p. 224). The Lenskis say the occurrence of the belief in this kind of god “varies directly with the importance of herding activities to the particular group” (Lenski & Lenski, 1982). They also note some of the other features of herding societies: “…marked social inequality…[h]ereditary slavery…raiding and warfare…military advantage over their less mobile agrarian neighbors….” (Lenski, Nolan, & Lenski, 1995, p. 227).

44

Rethinking Food and Agriculture

2.8 War and colonialism: Dominating lands and others Some may object that dictatorship, warfare, and slavery naturally accompanied the growth of the agrarian state. But in the Middle East, royal power, wars, raiding, and slavery take on extra dimensions because of the cultural influence of the herding peoples there. These large animal domesticators came from a long tradition of specialized hunters, so they had their hunter-warrior skills intact. These skills went a long way toward helping them build up their herds—their wealth and security. From the perspective of a hunter-warrior, how much more glorious (as well as easier and faster) it was to increase the herd by raiding than by slow, plodding husbandry. Warrior skills also helped keep the wealth, for a prosperous tribe had to be constantly on guard against raids by others. Herders were markedly more expansionist than ordinary farmers. Planter folks were confined to a few acres in a valley or along a river bank and they expanded their fields relatively slowly—over years. Herders, on the other hand, arrogantly dominated an entire region through military force, thus ensuring access to the best range land and water sources. In the Middle Eastern center, at least, herders were the best positioned to fill the warrior class, out of which arose the ruling elites and kings. Consequently, the entire hierarchy and culture in the region was imbued with the herder’s fierce, expansionist values. When these became integrated into the agrarian state’s religion, military, and other governing institutions, they made for ruthless nations and frequent wars of conquest. The central role of animal exploitation in enabling and promoting widespread violence, war, and colonialism is analyzed in great historical detail by sociologist David Nibert (2013). He argues that contrary to popular belief, our exploitation and oppression of other animals over the last 10,000 years has undermined the development of a just and peaceful world. Rather, it has facilitated the large-scale exploitation of and violence toward marginalized humans, especially indigenous people, and the destruction of the natural world. For example, over this period other animals such as horses have been used as instruments of war and “cattle” have been used as laborers and rations. Nibert contends that without this exploitation, the widespread violence, warfare, and empire building that Eurasian customs and institutions spread throughout much of the world would not have been possible. In fact, the exploitation of other animals has promoted such violence due to the need for ever-increasing sources of fresh water and grazing land (Nibert, 2013). Such widespread violence and destruction did not take place in regions that lacked populations of large mammals to exploit, such as the pre-Columbian Americas (Nibert, 2013). It was Christopher Columbus who first brought domesticated animals from Eurasia to the Western Hemisphere. As Nibert argues: the Spanish invasion of the Americas…was made possible only by the use of horses as instruments of war, the use of cows, pigs, sheep and goats as rations and the decimation of the local population from zoonotic diseases that the Spanish introduced into the hemisphere. (Nibert, 2012, p. 143)

During the Colonial Era, the export of the skin and body fat of other animals to Europe was very lucrative and drove the growth of ranching in the Americas, along

Agriculture planted the seeds of alienation from nature

45

with the violent displacement of indigenous peoples (Nibert, 2013). The Spanish enslaved Native Americans and later Africans to work on sugar plantations and in gold and silver mines. Other animals were used as laborers, their flesh was turned into slave rations in the form of salted “meat” and their fat was used to make candles. Without the exploitation of other animals in these ways, the colonial enterprises would not have been able to function (Nibert, 2013).

2.9 The herder roots of property, money, and capitalism Besides war mongering and empire building, herders contributed a great deal to the West’s obsession with property and money. Richard Lewinsohn (1954) argues that the concept of money—or exchangeable wealth—began with herding specifically rather than agriculture in general: The concept of property arose from the power of disposal over the herds and was older and stronger than the property concept based on ownership of land, for arable land was plentiful, tamed animals scarce. Land spells production, cattle spell consumer goods. Only consumable property has tangible worth. Animals were the first form of capital (wealth). The word “capital” stems from Latin “capita”, relating to the head count of cattle by which a man’s wealth was measured. (p. 76)

Similarly, at the root of our words impecunious and pecuniary, which pertain to money, is the Latin pecu, the old word for cattle. In the old Aryan language, the word for warfare translated literally into “a desire for more cattle.” Cattle, or more probably sheep and goats, were the first form of portable, exchangeable wealth—that is, the first form of money. Those ancient herders were the first capitalists. Nibert (2013) explores this link between animal oppression and the development of capitalism. He shows how much of the wealth and resources extracted from the Americas by European colonizers, including the huge amounts of gold and silver, paved the way for the transformation of the European feudal system to the even more exploitative capitalist economic system. He writes: The enormous influx of silver soon reduced its value, leading to inflation and prompting the land-rich, cash-poor aristocracy in Europe to look for ways to increase funds. Many sought to replace unprofitable feudal cultivation practices and used their land instead to raise sheep, whose hair was fueling the growing textile industry in Britain and the Netherlands. This more lucrative enterprise required the displacement of “peasants” whose exploited labour long was the primary source of aristocratic wealth. Increased grazing led to the conversion of the “commons”— land traditionally shared by many members of rural communities for a variety of subsistence purposes—from communal to private property. This conversion, known as the enclosure movement, occurred first in Britain and was replicated in other parts of Europe, with support and enforcement by area governments. With the gradual breakup of the feudal order in Europe and the migration of people from agricultural areas to the growing cities, the labour power of the displaced cultivators increasingly became a commodity to be sold for wages. (pp. 62–63)

46

Rethinking Food and Agriculture

Parliamentary Acts and laws facilitated this transformation. Between 1760 and 1870, around 4000 Acts of Parliament changed 7 million acres (around a sixth of the area of England) from common to enclosed land. The majority of MPs who voted on these acts were landowners (Fairlie, 2009). The law was also used to deal with noncompliance. For example, in response to peasants’ resistance to the enclosure of woodlands, the Black Act was introduced in 1723 authorizing the death penalty for offenses related to poaching, resulting in hundreds of peasants being hanged (Fairlie, 2009). Another example were the Poor Laws of England and Wales, which ensured peasants would sell their labor to capitalists by criminalizing the unemployed (HoltGiménez, 2017). An example of this is the 1834 Poor Law Amendment Act, which encouraged the large-scale development of workhouses and made relief dependent on the recipient entering the workhouse (Higginbotham, 2016). The enclosure movement would not have been able to achieve this land grab and force peasants to become wage laborers without this accompanying legislation and the support of the state (Perelman, 2000). Throughout history, and continuing to this day, the laws of society have been made by powerful elites and used to protect their interests. In fact, the state has been the prime source of power necessary for elites to establish and maintain their dominance over the rest of society and overcome resistance through institutionalizing and legalizing exploitation and oppression (Nibert, 2002, 2013). As argued by Nibert (2013): The power of the state cannot be underestimated in its capacity to harness the vast majority of societal members into a system of laws that strongly favours the interests of the privileged. Throughout most of the past ten thousand years, that power has been used largely tyrannically and oppressively. In many instances, women, children, devalued males, and other animals have been viewed as personal property, and the full weight of the state has been used to protect the economic and utilitarian uses of these “others.” (p. 148)

2.10 Consequences of dominionism and misothery The roots of our domination of nature and each other can thus be traced back to the domestication and exploitation of other animals in Eurasia approximately 10,000 years ago. Along with the corresponding shift to our dominionist, agrarian-based, Western worldview, this has had terrible consequences for humans, other animals, and the natural world. For example, through: the land theft of the colonial period, which continues to this day through land grabbing (Pearce, 2012) and “meat grabbing” (Schneider, 2014); the neocolonial and neoliberal policies of Western governments and international development and financial institutions (Nibert, 2002, 2013); and the “industrial grain-oilseed-livestock complex,” which is integral to our unequal, corporate food regime (Weis, 2013). All of these issues are connected and all contribute directly to a range of urgent and potentially catastrophic global problems. These problems include: the vastly inefficient use of land and other resources such as water, topsoil, and fossil fuel; hunger and increasing rates of chronic lifestyle diseases; climate change;

Agriculture planted the seeds of alienation from nature

47

e­ nvironmental pollution and destruction; deforestation; and mass species extinction, just to mention a few. The inefficiency of animal agriculture’s resource use is shown by recent estimates, which suggest it uses 83% of global farm land while only providing 18% of calories mainly for people in wealthy countries (Poore & Nemecek, 2018). It is also estimated that 40% of grain and 80% of soy are used to feed other animals for human consumption (Stoll-Kleemann & O'Riordan, 2015). Such inefficient use of land and other resources contributes to the fact that nearly a billion people are suffering from chronic food deprivation (FAO et al., 2018). At the same time, nearly 2 billion adults are overweight, over 650 million of whom are obese (WHO, 2018) representing a growing global health crisis. Overweight, obesity, and other chronic diseases such as cancer, diabetes, and heart disease are strongly related to Western diets, high in processed and animal-based foods. While these diseases are more prevalent among wealthier countries, the globalization of animal-based Western diets has resulted in a worldwide epidemic of chronic diseases that are intimately related to lifestyle and dietary choices (see Chapter 15). Animal agriculture is also estimated to contribute 56%–58% of food’s greenhouse gas emissions and approximately 15% of total greenhouse gas emissions (Poore & Nemecek, 2018). Even the lowest-impact animal products are estimated to have a greater impact than substitute vegetable proteins on greenhouse gas emissions, eutrophication, acidification, and land use (Poore & Nemecek, 2018). This also includes emissions generated from grass-fed ruminants in “well-managed” grazing systems, often argued to be necessary for sustainable agriculture (for example, see Sustainable Land Trust, 2015). At an aggregate level, the emissions from these grazing systems have been found to significantly outweigh their potential to offset emissions through soil carbon sequestration (Garnett et al., 2017). Overall, agriculture is responsible for around 80% of deforestation worldwide, of which the expansion of animal agriculture (including production of animal feed) is a significant driver (Campbell et al., 2017; Hosonuma et al., 2012; Kissinger, Herold, & De Sy, 2012). This is especially the case in Latin America where the greatest amount of deforestation is occurring. In the Amazon, 70%–80% of deforested land has been converted into pasture for grazing, with much of the remaining land used to grow animal feed such as soy (Machovina & Feeley, 2014; Steinfeld et al., 2006). Overall, animal agriculture is the most significant driver of habitat loss on Earth (Machovina, Feeley, & Ripple, 2015) and one of the biggest drivers of global biodiversity loss (Steinfeld et al., 2006). The rate of species extinction is so great, some scientists have called this period the “sixth mass extinction on Earth” (Barnosky et  al., 2011). However, this mass extinction is seen as even more severe when the extirpations (local extinctions) and huge declines in population are included, which will have negative cascading impacts on Earth’s life support systems. Ceballos, Ehrlich, and Dirzo (2017) describe this as “biological annihilation” to highlight the seriousness and magnitude of the situation we are facing. In short, our agrarian worldview is destroying the living world right under our noses. We are too disabled by dominionism, too alienated by misothery, to fully understand what we are doing.

48

Rethinking Food and Agriculture

2.11 Conclusion In March 1992, Vaclav Havel, then president of Czechoslovakia, wrote an opinion piece in the New York Times of the social turmoil of the modern era and of impending environmental disaster: “Man’s attitude to the world must be radically changed” (Havel, 1992, op-ed page). Twenty years earlier, California law professor Christopher D. Stone used substantially the same language in a landmark article that has become one of the “bibles” of the environmental movement: in “Should Trees Have Standing,” Stone wrote of the need for “a radical new conception of man’s relationship to the rest of nature” (Stone, 1974, p. 101). Another “bible” of environmentalism is an essay by historian Lynn White, “The Historical Roots of Our Ecological Crisis,” in which he urged a “rethinking” of “fundamentals” suggesting that we “find a new religion or rethink an old one” (White, 1967, p. 1203). So many other writers echo that what we need is a total change in our attitude of mind (Clark, 1977). And that “We face an ecological crisis compounded by a spiritual crisis. We need a radical shift in our world outlook” (Deloria, 2003, pp. 287–288). One can see some of the litany of famous names and books in A Search for Environmental Ethics, from the Smithsonian Institution. Most of its entries indict in some way Western civilization’s secular and religious traditions for our messed-up treatment of nature. Whether one reads the works of Marston Bates, David Brower, Rachel Carson, Barry Commoner, Rene Dubos, Anne and Paul Ehrlich, Aldo Leopold, John Muir, Roderick Nash, or most any environmentalist writer, the message is the same: humanity needs fundamental changes in its attitudes and relations toward the living world. Their writings are all about what intellectuals call the Nature Question. After having laid down such strong rhetoric, however, these movers and shakers stop dead in their tracks when they approach the Animal Question—the whole sticky mess of human views and uses of our animal cousins. Misothery clouds—or prevents— clear thinking on the Nature Question. The Animal Question is deemed silly, peripheral, off limits. It seems that one’s stature and seriousness as a thinker on the Nature Question depends on how well one steers clear of the Animal Question. Think about how different Professor Stone’s reputation might be today if his essay dealt with the question: “should chimpanzees have standing?” Even those who do consider the Animal Question, such as historian Yuval Harari in his bestselling book Sapiens (2014), only go so far. Harari considers a range of animal issues such as the impact of our domestication and use of other animals on society and the horrors of factory farms. However, he omits to mention the seminal ideas of Shepard (1978), one of the founders of the “Deep Ecology” movement, on the central role of animals in the evolution of the human mind, cognitive abilities, language, etc. Rather puzzlingly, Harari dismisses the discussion in just a few sentences: “The appearance of new ways of thinking and communicating, between 70,000 and 30,000  years ago, constitutes the Cognitive Revolution. What caused it? We’re not sure” (p. 23).

Agriculture planted the seeds of alienation from nature

49

The Animal Question is the very heart of the Nature Question because, as we have seen, our minds and our worldview grew up, so to speak, on the matter of other animals—for us, the embodiment of the living world around us. To exclude the matter of other animals from our discussions of the Nature Question is to exclude the most important part of the discussion. Let us pose the following question to our list of important thinkers who propose “radical” or “fundamental” changes in our worldview: What does that mean if we avoid the matter of other animals, who have, throughout our evolution as hominids, been seen as kin, significant Others, and, as such, Nature herself? Clearly, we need to stop avoiding the Animal Question if we truly want to reset our worldview. We need to do so because our misotherist attitudes about other animals and animality determine so much of our views of life, human life, other life, and our place in the living world. We need a better, healthier, more rational sense of who we are as a species and of how we ought to carry on here among all the other living beings in the world. We need to dump the prehistoric herders’ self-serving, dominionist model with its GODmade-us-and-gave-us-the-world-to-exploit mandate. A better model, or paradigm, for the human spirit must begin with biological realities. We are animals and we grew up—evolutionarily speaking—among other animals. Then and now, animals shape our worldview. If we want a better paradigm, a better worldview, we are going to have to start with that biological fact. Kinship is the biological reality here on Earth. Yet our dominionist culture, especially its misothery, denies it. It denigrates our evolutionary next of kin, makes us hate them and have contempt for them. It keeps us apart from them. It puts us all alone and over the living world, not in and of this world. It gives us a lonely station over a despicable chaos of animals and nature. No wonder we destroy the living world and suffer a malaise about the state of humanity.

References Ardrey, R. (1976). The hunting hypothesis. New York: Atheneum. Aristotle. (1959). Politics. Everyman’s library. London: J.M. Dent & Sons. Aschmann, H. (1965). In A. Leeds, & A. Vayda (Eds.), Comments on the symposium man, culture, and animals. Man, culture, and animals (pp. 259–270). Barnosky, A. D., Matzke, N., Tomiya, S., Wogan, G. O., Swartz, B., Quental, T. B., … Mersey, B. (2011). Has the Earth’s sixth mass extinction already arrived? Nature, 471(7336), 51. Bourke, V. (1951). Ethics. New York: Macmillan. Campbell, J. (1988a). Historical atlas of world mythology. The way of the animal powers, Part 1: Mythologies of the primitive hunters and gatherers: Vol. I. New York: Harper & Row. Campbell, J. (1988b). Historical atlas. Part 2: Mythologies of the great hunt: Vol. I. New York: Harper & Row. Campbell, B. M., Beare, D. J., Bennett, E. M., Hall-Spencer, J. M., Ingram, J. S. I., Jaramillo, F., … Shindell, D. (2017). Agriculture production as a major driver of the Earth system exceeding planetary boundaries. Ecology and Society, 22(4), 8.

50

Rethinking Food and Agriculture

Ceballos, G., Ehrlich, P. R., & Dirzo, R. (2017). Biological annihilation via the ongoing sixth mass extinction signaled by vertebrate population losses and declines. Proceedings of the National Academy of Sciences, 114(30), E6089–E6096. Clark, J. (1968). Beastly Folklore. Metuchen, NJ: Scarecrow Press. Clark, K. (1977). Animals and men: Their relationship as reflected in western art from prehistory to the present day. Thames and Hudson. Cohen, M. N. (1989). Health and the rise of civilization. New Haven, CT: Yale University Press. Cornforth, M. (1971). Historical materialism. Dialectical materialism: an introduction: Vol. 2. New York: International Publishers. Cranstone, B. (1969). Animal husbandry: The evidence from ethnography. In P. Ucko, & B. Dimbleby (Eds.), The domestication and exploitation of plants and animals. Chicago, IL: Aldine Publishing Co. Crosby, A. (1986). Ecological imperialism: The biological expansion of Europe 900-1900. Cambridge: Cambridge University Press. Deloria, V. (2003). God is red: A native view of religion. Fulcrum Publishing. Eisler, R. T. (1988). The chalice and the blade: Our history, our future. San Francisco, CA: Harper & Row. Fairlie, S. (2009). A short history of enclosure in Britain. The Land, 7, 16–31. FAO, IFAD, UNICEF, WFP, & WHO. (2018). The state of food security and nutrition in the world 2018. Building climate resilience for food security and nutrition. Rome: FAO. Garnett, T., Godde, C., Muller, A., Röös, E., Smith, P., de Boer, I., … van Zanten, H. (2017). Grazed and confused?: Ruminating on cattle, grazing systems, methane, nitrous oxide, the soil carbon sequestration question-and what it all means for greenhouse gas emissions. Food Climate Research Network. Gove, P. B. (1976). Webster’s third new international dictionary unabridged. Merriam-Webster. Harari, Y. N. (2014). Sapiens: A brief history of humankind. Random House. Harris, M. (1977). Cannibals and kings: The Origins of cultures. New York: Random House. Hart, D., & Sussman, R. W. (2008). Man the hunted: Primates, predators, and human evolution. Westview Press. Hartmann, T. (2001). The last hours of ancient sunlight. UK: Hodder and Stoughton. Havel, V. (March 1, 1992). The end of the modern era. The New York Times. Higginbotham, P. (2016). The workhouse. Retrieved from https://web.archive.org/ web/20090504111530/ http://institutions.org.uk/poor_law_unions/the_poor_law1.htm. Holt-Giménez, E. (2017). A foodie’s guide to capitalism: Understanding the political economy of what we eat. NYU Press. Hosonuma, N., Herold, M., De Sy, V., De Fries, R. S., Brockhaus, M., Verchot, L., … Romijn, E. (2012). An assessment of deforestation and forest degradation drivers in developing countries. Environmental Research Letters, 7(4), 044009. Kissinger, G. M., Herold, M., & De Sy, V. (2012). Drivers of deforestation and forest degradation: A synthesis report for REDD + policymakers. Lexeme Consulting. Lenski, G., & Lenski, J. (1982). Human Societies: An Introduction to Macrosociology (4th ed.). New York: McGraw-Hill. Lenski, G., Nolan, P., & Lenski, J. (1995). Human societies: An introduction to macrosociology (7th ed.). New York: McGraw-Hill. Levi-Strauss, C. (1964). Totemism (Rodney Needham, Trans.). London: Merlin Press. Lewinsohn, R. (1954). Animals, men, and myths: A history of the influence of animals on civilization and culture. Gollancz. Lopez, B. (1978). Of wolves and men. New York: Charles Scribner’s Sons. Lorenz, K. (1954). Man meets dog. London: Methuen.

Agriculture planted the seeds of alienation from nature

51

Machovina, B., & Feeley, K. J. (2014). Meat consumption as a key impact on tropical nature: A response to Laurance et al. Trends in Ecology & Evolution, 29(8), 430–431. Machovina, B., Feeley, K. J., & Ripple, W. J. (2015). Biodiversity conservation: The key is reducing meat consumption. Science of the Total Environment, 536, 419–431. Mason, J. (1993). An unnatural order: Uncovering the roots of our domination of nature and each other. New York: Simon & Schuster. Mason, J. (2017). Misothery: Contempt for animals and nature, its origins, purposes and repercussions. In L. Kalof (Ed.), The Oxford handbook of animal studies. New York: Oxford University Press. Nibert, D. (2002). Animal rights/human rights: Entanglements of oppression and liberation. Rowman & Littlefield Publishers. Nibert, D. (2012). The fire next time: The coming cost of capitalism, animal oppression and environmental ruin. Journal of Human Rights and the Environment, 3, 141. Nibert, D. (2013). Animal oppression and human violence: Domesecration, capitalism, and global conflict. Columbia University Press. Nitecki, M., & Nitecki, D. (Eds.). (1987). The evolution of human hunting. New York: Plenum Press. Pearce, F. (2012). The landgrabbers: The new fight over who owns the earth. Eden Project Books. Perelman, M. (2000). The invention of capitalism: Classical political economy and the secret history of primitive accumulation. Duke University Press. Poore, J., & Nemecek, T. (2018). Reducing food’s environmental impacts through producers and consumers. Science, 360(6392), 987–992. Sanday, P. (1981). Female power and male dominance: On the origins of sexual inequality. Cambridge: Cambridge University Press. Schneider, M. (2014). Developing the meat grab. Journal of Peasant Studies, 41(4), 613–633. Serpell, J. (1986). In the company of animals: A study of human-animal relationships. London: Basil Blackwell. Shepard, P. (1978). Thinking animals: Animals and the development of human intelligence (p. 2). New York: Viking Press. Steinfeld, H., Gerber, P., Wassenaar, T. D., Castel, V., Rosales, M., Rosales, M., & de Haan, C. (2006). Livestock’s long shadow: Environmental issues and options. FAO. Stoll-Kleemann, S., & O’Riordan, T. (2015). The sustainability challenges of our meat and dairy diets. Environment: Science and Policy for Sustainable Development, 57(3), 34–48. Stone, C. (1974). Should trees have standing?. New York: Avon Books. Sustainable Land Trust. (2015). What role for grazing livestock?. Retrieved from https://sustainablefoodtrust.org/articles/role-for-grazing-livestock-red-meat/. Tanner, N. (1981). On becoming human. Cambridge: Cambridge University Press. Thomas, K. (1983). Man and the natural world: A history of the modern sensibility. New York: Pantheon Books. Ucko, P., & Dimbleby, B. (Eds.). (1969). The domestication and exploitation of plants and animals. Chicago, IL: Aldine Publishing Co. Underhill, R. (1965). Red man’s religion: Beliefs and practices of the Indians North of Mexico. Chicago, IL: University of Chicago Press. Walsh, P. G. (Ed.). (2008). The nature of the gods OUP Oxford. Weis, T. (2013). The ecological hoofprint: The global burden of industrial livestock. Zed Books Ltd. White, L. (1967). The historical roots of our ecological crisis. Science, 155, 1203. WHO. (2018). World Health Organisation overweight and obesity fact sheet. Retrieved from https://www.who.int/news-room/fact-sheets/detail/obesity-and-overweight.

This page intentionally left blank

Political economy of the global food and agriculture system

3

Philip McMichael Department of Global Development, Cornell University, Ithaca, NY, United States

3.1 Introduction Recently, the EAT-Lancet Commission on Healthy Diets from Sustainable Food Systems summary report noted: “Global food production threatens climate stability and ecosystem resilience and constitutes the single largest driver of environmental degradation and transgression of planetary boundaries. Taken together the outcome is dire. A radical transformation of the global food system is urgently needed” (EATLancet Commission, 2019). A year before, the Inter-Academy Partnership (130 national academies of science and medicine) reported on what they identified as a “broken global food system” (The Guardian, November 28, 2018). An alternative perspective is that the global food system “is not broken … It is working precisely as a capitalist food system is supposed to work: it expands constantly, concentrating wealth in a few, powerful monopolies, while transferring all the social and environmental costs onto society” (Holt-Giménez, 2019, p. 89). Separately and together these observations address the food system crisis and its socioecological consequences. This is the focus of this chapter.

3.2 Antecedents of the present crisis The greater visibility of malnutrition, public health, and environmental degradation accompanies a cumulative increase in global temperatures and economic and environmental refugees. This moment has long-term origins in political/economic and political/ecological relationships stemming from centuries of colonization of landed cultures, with devastating impacts on indigenous habitats. Displacement of indigenous peoples and peasants from their land and water cultures accompanied conversion of these landscapes for commodity production for export to colonizing powers. Historic frontier expansions have drawn down Earth’s “ecological capital” in a process represented as the “under-reproduction of nature” (Moore, 2011, p. 28), at the expense of future sustainability. Throughout the colonial world, plantations and cash crops produced specialized tropical exports ranging from bananas to peanuts, depending on local agri-ecologies. Non-European societies were fundamentally transformed through the loss of resources and craft traditions as colonial subjects were forced to labor in mines, fields, and plantations to produce exports sustaining distant European factories and their ­growing Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00003-7 © 2021 Elsevier Inc. All rights reserved.

54

Rethinking Food and Agriculture

workforces. This was a global process, whereby slaves, peasants, and laborers in colonies provisioned European industrial classes with cheap colonial products such as sugar, tea, tropical oils, and cotton for clothing. European development was realized through a racialized process of “underdevelopment” of colonial societies, reaching its height in the late 19th century age of empire. In India, for example, the British Raj (1848–1947) used tax and irrigation policies to require farmers to produce agricultural crops for export, taxing forced commercial cropping, enabled by new irrigation canals displacing indigenous irrigation systems. A British engineer reported to the 1901 Irrigation Commission: “Canals may not protect against famine, but they may give an enormous return on your money” (quoted in Davis, 2001, pp. 332–335). Between the 1890s and 1940s, such export crops as cotton, jute, tea, peanuts, and sugar cane rose by 85%, while local food crop production declined by 7% alongside population growth at 40%, spreading hunger, famine, and social unrest (Chirot, 1977, p. 124). Meanwhile, London grain merchants deployed new telegraph technologies for the pricing of grain in hinterland villages reserved for famine and drought years. In commodifying these reserves, merchants could acquire them from complicit local chiefs, transporting the grain via new railway systems to ports for export to Britain. By 1900, 20% of Britain’s bread consumption came from such famine reserves in India. During the 1899–1900 famine, 143,000 peasants in Berar died from starvation as 747,000 bushels of grain exports left the province (Davis, 2001, pp. 29, 299, 315). Starvation from food exporting to the “mother country” was intensified by converting the “commons” into private property or state monopolies, as customary commons, forests, and pastures provided nonmarket resources (e.g., grasses for ropes and fodder, medicinal resources, wood and dung for fuel, forest debris, and dung for fertilizer). While these were common resources for all, for the poor they were “the very margin of survival” (Davis, 2001, p. 327). Expulsion from communal grasslands ruptured “the ancient ecological interdependence of pastoralists and farmers,” and traditional crop rotation and fallow practices to replenish soils disappeared as cotton and other monocrop exports expanded (Davis, 2001, pp. 328–329). Expanding European urban populations required ever-increasing imports of sugar, coffee, tea, cocoa, tobacco, and vegetable oils from the colonies. The factory system demanded ever-increasing inputs of raw materials such as cotton, timber, rubber, and jute, as commodity frontiers multiplied, dispossessing and/or enslaving native populations. From the late 18th century slave revolt presaging Haiti’s establishment as the first postcolonial society, independence movements spread through to mid-20th century as elites and masses organized against colonial rule and “underdevelopment,” purportedly joining the “modern” world. As Gilbert Rist observes: “Their right to self-determination had been acquired in exchange for the right to self-definition” (Rist, 1977, p. 79), suggesting that postcolonial elites internalized the terms of rule in an asymmetrical international political economy. Such an unequal relationship has enabled continuing dispossession of non-­ European peoples and/or their indigenous socioeconomic systems. This began with the post-WWII project modeling development as a program of “economic nationalism,” instituted internationally during the Cold War via US economic, technical, and

Political economy of the global food and agriculture system55

­ ilitary aid to safeguard Western interests. This development model presupposed rural m migration to supply labor for urban manufacturing growth (Rostow, 1960). And it has intensified with the transformation of agriculture via industrial technologies, enabling export monocultures to supply global manufacturing systems with raw materials. Today, it is generalized as a new “extractive” development model in the Global South (Connell & Dados, 2014; Petras & Veltmeyer, 2014). This current “resource grab” in turn generates waves of global migration, rising ethnopolitical tensions, and a complicated politics of nativist populism across the world today. The proverbial colonial chickens are coming home to roost.

3.3 Political economy of the food regime Situating the present crisis in relation to colonial history can be viewed through the lens of food regime analysis (Friedmann & McMichael, 1989; McMichael, 2009, 2013a, 2013b, 2013c). Food regime analysis contributes a unique perspective to the political history of world capitalism, and its progressive reorganizations of food production and circulation on a world scale. Food regimes provision cheap food globally across three episodes of international hegemony: British (1870s–1914), American (1940s–70s), and neoliberal (1980s to the present). The first food regime, organized by the British imperial state, followed abolition of the Corn Laws (1846), offshoring grains and meat production to settler states in the Americas, South Africa, and Australasia. These temperate foods complemented access to tropical foods (sugar, tea, coffee, palm oil) from the colonies. In combination, via the “imperialism of free trade,” these imports supplied cheap caloric wage-foods for an emerging industrial labor force in Britain and Europe (Hobsbawm, 1969, pp. 128–129; Mintz, 1985). By the mid-20th century the United States replaced Britain as world hegemon, modeling the nation-state as the appropriate unit of development. Instead of imperial dependence on offshore agricultures, the US economy integrated manufacturing and farm sectors domestically. This model in turn framed the mid-20th century “development project” (McMichael, 1996). This included a US public food-aid program selling food surpluses (notably cereals and dairy products) as wage-foods at concessional prices to underwrite national industrialization in strategic Third World states on Cold War perimeters. This strategy was complemented with the introduction by the United States of Green Revolution technologies on expanded acreage in key states to increase delivery of domestic wage-foods to expanding urban populations (Patel, 2013). Originating in Mexico in 1943 with material support from the US government and the Rockefeller Foundation, as an early development program in “technology transfer,” new investment priorities in Green Revolution crop acreage expanded cereal and bean production by 300% over 2 decades (Gupta, 1998, p. 53; Patel, 2013). From there it migrated to South Asia, where Pakistan and India, both substantial importers of wheat in the mid-1960s, were self-sufficient by 1968 and 1974, respectively. Here, wheat production increased more than 60% over 6 years (Akram-Lodhi, 2013, p. 88), displacing farming cultures of peasants unable to afford or risk commercial ­technologies,

56

Rethinking Food and Agriculture

often renting or vacating their lands. Alongside the expansion of bulk food production for urban consumers in parts of the postcolonial world, Green Revolution technologies have been accompanied by rising controversy (see later). Meanwhile, as the combination of cheap food imports and the Green Revolution substantially increased urban food supplies in select Third World regions, each had competitive effects (low prices and differential access to technological packages, respectively) on small-scale farmers. Resulting urban migration gave the development model the appearance of a “self-fulfilling prophecy,” with cheap labor complementing cheap wage-foods—nevertheless at the expense of rural cultures and often resilient farming systems. Within the terms of the development narrative, displacement of small farming systems was deemed a natural progression out of rural backwardness and poverty (Lewis, 1954; Moore Jr., 1967; Rostow, 1960; Sachs, 2005). The question is whether small farming systems, which currently feed up to 70% of the world (ETC, 2009), are necessarily “poor.” The development narrative’s naturalizing epistemology elides this question, projecting a modern concept of poverty onto the frugality of so-called “left-behind” peoples (such as forest dwellers, fisherfolk, peasants, nomad-pastoralists). This representation informs Walt Rostow’s five “stages of growth” vision: from traditional to high-mass consumption society (now perhaps a really terminal stage). And it informs the baseline assumption underlying Jeffrey Sachs’ The End of Poverty: The move from universal poverty to varying degrees of prosperity has happened rapidly in the span of human history. Two hundred years ago the idea that we could potentially achieve the end of extreme poverty would have been unimaginable. Just about everybody was poor, with the exception of a very small minority of rulers and large landowners. (Sachs, 2005, p. 26)

Without romanticizing Napoleonic era life, in and outside of Europe, humanity was a good deal more complex, resourceful, and diverse than simply being “poor” by our (monetary/income) standards, which ignore social and ecological wealth. In addition to expanding Green Revolution crops of wheat and rice, corn and soybeans along with sugar and oils sourced transnational animal protein and durable foods complexes, respectively, as industrial processing turned food into an increasingly profitable industry (Friedmann, 1994). These transnational complexes in turn underlay the emergence of a third, corporate food regime, and its progressive integration of commodity supply chains into worldscale production relations. Such “cross-territorial” relations are organized by global agribusiness and retailing firms, sanctioned by states under the auspices of World Trade Organization (WTO) rules in 1995. Here, “economic nationalism” began its decline with structural adjustment policies mandated by the World Bank and International Monetary Fund from the 1980s, requiring Southern states to dismantle farm sector protections and expand agro-exports to defray cumulating debt (George, 1988). In the 1990s, multilateral trade agreements liberalized agricultural investment and trade on a world scale, dramatically accelerating global depeasantization

Political economy of the global food and agriculture system57

(Araghi, 2009). A conservative Food and Agricultural Organization (FAO) estimate for 16 Southern countries claimed that between 20 and 30 million farmers exposed to agricultural trade liberalization lost their land in the half decade following establishment of the WTO in 1995 (Madeley, 2000, p. 75). This included almost 2 million Mexican campesinos, following introduction of the North American Free Trade Agreement (NAFTA) in 1994. WTO empowerment of transnational agribusiness, including agro-input and commodity traders, processors, and retailers, deepened corporate hegemony and its “ability to shape the rules under which [corporations] operate in the middle space that they occupy in the world food economy” between producers and consumers (Clapp, 2016, p. 121).

3.4 Food regime complexes This section examines the key transnational complexes (wheat, meat, and processed food) constituting the food regimes and their succession as a global food order matured. Their mutual significance emerged in the second, US-centered regime, which broadened the reach of the food regime beyond the settler relationship to increasingly encompass the “free world,” via food aid, Green Revolution, and dietary enrichment for those with purchasing power. The US-sponsored Green Revolution, centered on expanding production of staple grains, prefigured the so-called “second Green Revolution” of the third, and now global, food regime. Here, private (rather than public) investments focused on “affluent foods” such as animal protein, fruits, and vegetables, intensifying chemical agriculture, aquaculture, factory farming, and the expansion of transgenic technology (DeWalt, 1985; Patel, 2013; Weis, 2007).

3.4.1 Wheat complex To stabilize the US farm belt, postdepression, and “dust bowl,” the US government established commodity stabilization schemes, encouraging specialization in particular commodities (wheat, corn, soy, rice, sugar, dairy). Subsidies and technological support resulted in overproduction. Disposal of these farm surpluses was managed by the Public Law (PL) 480 program (1954), involving concessional food exports (predominantly wheat) to strategic Third World states such as India, South Korea, Brazil, Morocco, Yugoslavia, South Vietnam, Egypt, Tunisia, Israel, Indonesia, Taiwan, and the Philippines. Such cheap food subsidized urban consumption and wage costs, encouraging local industrial development. It also created food dependency—a longterm goal being to create new agricultural markets in the Third World as countries adopted the American diet. At the same time the PL-480 program required recipient states to pay for grain imports with “counterpart funds” deposited in local banks and used by US agencies to finance infrastructural projects such as military bases and agribusiness activities—thereby exporting the US model of industrial agriculture (Friedmann, 1990).

58

Rethinking Food and Agriculture

By 1978, the Third World was importing over three-quarters of American wheat exports, with per capita consumption of wheat rising by almost two-thirds, and a significant decline in per capita consumption of traditional root crops (Friedmann, 1990). Wheat progressively replaced rice and corn in Asian and Latin American diets. In Central America, wheat and rice imports substituted for maize, and in West Africa for millet and sorghum. Traditional starches such as cassava, potato, yams, and taro were displaced by cheap grain imports, substituting “wage-foods” for “peasant foods” (McMichael & Raynolds, 1994, p. 322). By the 1970s the United States was selling food to over 130 countries, with the Third World buying two-thirds of its wheat from the United States (Clapp, 2016, pp. 34–35). Over time, other wheat exporters (e.g., France, Canada, Australia) adopted forms of food aid, with Export Enhancement Programs to dispose of wheat surpluses. In the 21st century, most states across the world are net wheat importers, with “low-income” countries expanding their share of wheat imports from 10% in the 1970s to 40% in 2010. This pattern correlates with the convenience of bread for urban populations, and the retailing revolution, encouraging “the massive consumption of wheat-based products, relative to the traditional meals of those countries” (GonzalezEsteban, 2018, p. 104). Reward Group International, a subsidiary of a Chinese firm, invested in 900 ha of land in central France; its goal: “To put French cereals on Chinese tables,” including baguettes (de Mareschel, 2018). While the transnational “wheat complex” has subsidized industrial wages and shaped urban consumption patterns, displacing “peasant foods,” it has steadily eroded domestic farm systems, deepening hunger and food insecurity. The US export of Green Revolution technology to select Third World countries improved agricultural productivity with new seed varieties (erroneously termed “HighYielding Varieties”) and agrichemicals, to address the specter of “overpopulation” and stanch the spread of communism. The FAO provided extension services for the disposal of synthetic fertilizer, intensifying energy sector dependence (Cleaver, 1977, p. 28). Major Third World wheat-producing countries, such as India, Argentina, Pakistan, Turkey, Mexico, and Brazil, adopted Green Revolution varieties, accounting for 86% of the total Green Revolution wheat area by the 1980s. Green Revolution rice was adopted by India, Indonesia, the Philippines, Bangladesh, Burma (now Myanmar), and Vietnam, accounting for more than 87% of Green Revolution rice varieties by the 1980s (Andrae & Beckman, 1985; Dalrymple, 1985, p. 1068; Raikes, 1988). While the Green Revolution’s “miracle seeds” improved grain output—enlarging dedicated land, but not necessarily yields (Kumar, 2016; Patel, 2013; Stone, 2019), with Mexico and India reducing food imports—such high-input agriculture (with agrichemicals, irrigation, and mechanization), first tested in Mexico in the 1940s, was controversial. It weaned wealthier farmers from seed sharing and mixed farming to adopting grain monocultures, encouraging unsustainable productivist methods (Patel, 2013). Furthermore, the Green Revolution exposed rural workers to toxic chemicals. “Peasant foods” such as traditional micronutrient-rich leafy greens, redefined as “weeds,” were eliminated by herbicides to promote grain monocultures of macronutrient “wage-foods” (Shiva, 1991). While yields initially increased, they have declined since, alongside ecosystem deterioration, as detailed in the UN’s Millennium

Political economy of the global food and agriculture system59

Ecosystem Assessment (2005), and rising public health concerns. The latter are now at the center of claims of carcinogenic effects—the so-called “cancer train” in the Indian state of Punjab being a dramatic long-term effect (Das, 2016). Toxicity, and debt, deriving from the “technology treadmill,” given declining soil fertility, have been blamed for farmer suicides, especially in India. There, a farmer remarked: “chemical fertilizer makes the crop shoot up...whereas organic manure makes for strength. Without strength, no matter how much fertilizer you put, the field won’t give output” (quoted in Gupta, 1998, p. 4).

3.4.2 Meat complex Cheap grain surpluses during the food-aid regime also nurtured a transnational animal protein complex, integrated with recipient countries. Postwar, the US Marshall Plan introduced the American industrial agricultural system to Europe, the 40% of Marshall aid going to European food and agriculture concentrated on imports of feedstuff and fertilizer imports. As Friedmann notes: Despite protection, the openness to direct investment by US transnational corporations helped to integrate European and US agro-food sectors via industrial inputs and processing. Both in promoting meat intensive diets and in organizing intensive livestock production, agro-food capitals shaped agricultural reconstruction along lines similar to the US. Most important was investment in an intensive livestock sector relying on industrial feedstuffs composed from soy and maize. This linked apparently national agricultures to imported inputs. Beneath the protected surface, therefore, lay the corporate organization of a transnational agro-food complex centred on the Atlantic economy. It linked North America, especially the US, to Europe. (Friedmann, 1993)

After 1954, surpluses were redirected to Third World countries in food aid, with the US Feed Grains Council designating counterpart funds to over 400 agribusinesses to develop local livestock industries. Thus the 1970 PL-480 annual report advocated financing the “construction and operation of modern livestock feed mixing and livestock and poultry production and processing facilities. As these facilities become fully operational, they will substantially expand the market for feed-grain and other feed ingredients” (quoted in George, 1977, pp. 171–172). A global meat complex formed in this context, as Third World countries developed livestock industries, supplied with specialized feed grains (corn and soy) from the First World and middle-income countries such as Brazil and Argentina. In the 4 decades following mid-century, global soybean production expanded sixfold, with corn production transformed into a specialized, capital-intensive agro-industry, overtaking the value of the world wheat trade by a factor of six (Friedmann, 1993). Animal protein consumption multiplied as governments and elites associated meat with a modern “nutrition transition” (from starch to grain to animal protein and fresh vegetables), which catered to affluent class diets. For example, in Egypt, by the mid-1970s, animal protein consumption by the richest 27% of the urban population outstripped animal protein consumption by the poorest 27% by a factor

60

Rethinking Food and Agriculture

of four. Rising incomes, subsidized by the United States Agency for International Development and Egyptian governments, encouraged a shift from legumes and maize to wheat and meat products, subsidized in part by taxing local food grain farmers. From 1970 to 1987, livestock production exceeded crop production by 10 to 1, intensifying feed grain imports into Egypt, now the third largest importer after Japan and China (Mitchell, 1991). Following regional trade liberalization in 1961, Japan and South Korea supplemented food grain imports with feed grains such as corn and soy (Huang & Coyle, 1989, p. 42), supplied by grain traders like Cargill and large soga shosha companies (Rothacher, 1989, p. 64), from Southeast Asia and Brazil. Japanese investment in the late 1970s in Thai aquaculture supply zones accompanied corporate joint ventures with Thai agribusinesses, especially in poultry production. During the 1980s, Thailand’s share of Japan’s broiler imports rose by 41% to overtake US exports (Bishop, Christensen, Mercier, & Witucki, 1990, p. 23). By 1994, Chinese poultry production overtook that of Thailand as principal supplier to Japan, and Asia’s supplier in general. With feed costs rising, Thailand’s Charoen Pokphand invested in China, producing 300 million of the 3 billion chickens consumed by the Chinese, and exports of Chinese poultry meat to Japan almost quadrupled during 1988–93 (Handley, 1990, p. 56). With the rise of transnational animal protein complexes, corn and soybeans served as agro-industrial inputs for intensive animal protein manufacture and sale in class-­ differentiated global markets for different quality meats. Grain traders facilitate this process, watching weather patterns and harvests across the world to ensure supplies of cheap feed grains to fatten cattle, poultry, pigs, or farmed seafood. While this goal contradicted long-standing US food grain supply-management policies, feed grain producers lobbied for liberalization, as global meat production and consumption steadily increased (Winders, 2017, p. 50, and see Baines, 2015). Power accumulation in the United States by feed grain producers, in particular, anticipated the liberalization embodied in the WTO’s 1995 Agreement on Agriculture, following which the transnational animal protein complex deepened, as meat consumption exploded: “in the last quarter century, as global production of maize increased by over 100 percent and that of soybeans by over 200 percent, while global rice production only increased by 35 percent and wheat by 25 percent” (Winders, 2017, p. 134). Cows slaughtered for meat increased in the last 50 years, but the rate of increase pales in contrast to pig slaughtering: 375 million in 1961, 815 million in 1985, more than 1 billion in 1995, and 1.4 billion in 2013. For chickens the rate was more dramatic: 6.1 billion in 1961 to 61 billion in 2013. “Not only did the number of animals slaughtered for meat increase during this period, but the average size of the animals also increased—a result of selective breeding and the expansion of the industrial livestock sector” (Winders, 2017, pp. 75–76). Such expansion stems in part from the need to dispose of enormous grain surpluses, which have justified a long-standing claim, for example, by the United States Department of Agriculture to channel such surpluses into producing animal protein, as a dietary positive. Surplus disposal of worldwide Green Revolution grains metamorphosed into “surplus absorption,” where “the ultimate challenge was not to simply get rid of surpluses but to find ways of profitably absorbing them on an ongoing

Political economy of the global food and agriculture system61

basis” (Weis, 2013, p. 72). Surplus disposal of grains was an effective foreign aid strategy, but surplus absorption was a profitable strategy of “recycling them through fast-­rising populations of concentrated livestock” (Weis, 2013, p. 73). Hence, the industrialization of pork and chicken meat across the world. Development is here conflated with consumption, as consumers with purchasing power dine “up” on animal protein. Meanwhile, the working poor dine either on food-aid grains or the low end of the food chain: low-protein starchy diets (given displacement of legumes, beans, and leafy vegetables), or little at all. As noted, the transnational meat complex displaces food crops, intensifying food insecurity in addition to unequal consumption patterns (steak vs hamburger). For example, India, with almost a fifth of its population undernourished, has the largest number of cows in the world, and is the largest beef exporter (20% of all beef exports), doubling its maize and soybean feed crops between 2000 and 2015, at a time of serious food insecurity and growing water shortage (Winders, 2017, pp. 101–104). Bill Winders notes that “India’s beef exports beg comparison” to its grain exports in the late 1800s, and, net of famine, the comparison invokes “the role of the market in shaping agricultural production and trade in ways that do not lead to greater access to food for the population—just as occurred at the turn of the previous century” (Winders, 2017, p. 104). This is the market championed by the development industry. Thus a recent World Bank World Development Report on “feeding the world” projects forward from present market trends, which include the unsustainable and inequitable impacts of the animal protein complex: “To meet projected demand, cereal production will have to increase by nearly 50 percent and meat production 85 percent from 2000 to 2030” (World Bank, 2007, pp. 8, 17). Using demand as the operative metric raises questions about the sustainability of “meatification” and its “ecological hoofprint” (Weis, 2013). Agriculture is responsible for about a third of global greenhouse gas emissions, of which livestock (including feed and transport) accounts for nearly 80%. The intensive livestock complex links specialized feed crop regions with proliferating factory farms, with 70% of arable land used by the livestock industry, where feed crops occupy onethird, in a veritable “meat grab” (Schneider, 2014). Per capita demand for beef, poultry, and pig meat in China is expected to double by 2020. Since Chinese intensive meat production is sourced mainly by Brazilian soybeans and US corn, this single complex produces emissions in multiple ways, from Amazonian deforestation, fossil fuel-based transport, fertilizer for intensive grain production, to animal methane. The contributions of beef production to global warming via carbon dioxide, nitrous oxide, and methane are significant: “it takes up to 16 times more farmland to sustain people on a diet of animal protein than on a diet of plant protein…The emerging meat-eaters of the emerging economies—especially China—are driving industrial agriculture into the tropical forests of South America, sending greenhouse gases skyward in a dangerous new linkage between the palate and the warming of the planet” (Nepstad, 2006, p. 1). In addition, factory farms in the United States today annually produce well over 1 billion tons of manure, laden with chemicals, antibiotics, and hormones, which leach into rivers and water tables (Kimbrell, 2002, p. 16). The explosion of “modern industrial hog” farming in China is confirming the concerns for

62

Rethinking Food and Agriculture

environmental and public health associated with factory farming, poisoned soils and contaminated waterways, increased emissions, reductions of genetic and species diversity, and rising antibiotic resistance (Schneider, 2011, pp. 18–19). The animal protein culture, conventionally viewed as improved consumption in a “nutrition transition,” has serious environmental and health consequences. It has also displaced traditional mixed farming systems, where animals fertilize soils, consume waste, and provide protein sources, complementing, rather than displacing, food crops. Such mixed (vs factory) farming is found in marginalized locales across the world, where traditional animal husbandry may provide household needs such as wool, leather, fur, soap, eggs, milk, meat, and energy for smallholdings.a The alternative, at this “tipping point” in human history, are cropping systems supplying plant-based foods, now vital to planetary and human health, which include biomass applications such as soil mulch, cover crops, and intercropping with legumes to enhance soils and provide valuable protein (Bezner Kerr, 2010). In advocating plant-based foods as the sustainable diet, the recent EAT-Lancet Commission (2019) recommends halving global red meat consumption by 2050, given that, as reported in Science (2018): livestock production accounts for 83% of the world’s arable land, while delivering only 18% of calories. Following this logic, George Monbiot underscores the extreme inefficiency of grain-fed livestocking, arguing “if you want to eat less soya, then you should eat soya: 93% of the soya we consume, which drives the destruction of forest, savannah and marshland, is embedded in meat, dairy, eggs and fish, and most of it is lost in conversion. When we eat it directly, much less of the crop is required to deliver the same amount of protein” (Monbiot, 2019, p. 48). How to transition to sustainable, diversified farming is the challenge facing the world—especially given the power of agribusiness and its official support by governments and the development industry, enamored with the association of “meatification” with modernization and its class-based foundations. Agribusiness involves integrated power: grain trader companies partner with animal feed producers and meat packers as part of the transnational meat complex. The top four grain trading companies, ADM, Cargill (USA), COFCO (China), and Louis Dreyfus (France), control more than twothirds of the world trade in grains (corn, soy, and wheat) and oilseeds (palm oil). In the United States, for example, “where twenty feedlots feed half of the cattle, just four firms (Tyson, Cargill, Swift and Co., and National Beef Packing) account for around 85 percent of the market for beef processing. Only four firms account for 50 percent of the US market for broiler chickens, and four firms account for over 60 percent of the pork market” (Clapp, 2016, p. 107). And beyond the United States is China, consuming the majority of the world’s pork, and poised to induct millions more into the so-called “nutrition transition” as factory farming of pork and chicken intensifies. Significantly, the Chinese state-owned company Chinatex Corporation and China National Cereals, Oils and Foodstuffs Corporation (COFCO) recently reduced its dependence on American grain traders—edging out Bunge as the fourth largest grain trader, after Archer Daniel Midlands, Cargill, and Louis Dreyfus. COFCO has also a

In characterizing farming systems absent industrial inputs, Tim Flanner notes: “plant–animal interactions are at the heart of Gaia’s self-regulation. Plants capture the sun’s energy, and animals, by feeding upon plants, create and swiftly recycle nutrients that plants need in order to grow” (Flanner, 2009, pp. 86–93).

Political economy of the global food and agriculture system63

purchased the Dutch grain trader Nidera, and a 51% share in the agricultural business of the Singapore-listed Noble Group. It has multiple offices across the world, and plans to invest over US$10 billion overseas (Grimsditch, 2017, pp. 23–24). And these corporate acquisitions combine with the recent phenomenon of colonizing land offshore, circumventing markets and intermediaries to guarantee supplies of food, feed, and fuel (McMichael, 2013b).

3.4.3 Durable foods complex Wayne Roberts observes: “Westerners don’t buy food any more. They buy processed meals assembled from ingredients or inputs” (Roberts, 2008, p. 122). This is becoming a universal practice with the “retailing revolution.” Corporate global sourcing strategies bundle, rather than produce, multisited ingredients as global foods (Roberts, 2008, p. 123). French cheese farmer Jose Bové refers to these bundles of global ingredients as “food from nowhere” (Bove & Dufour, 2001, p. 55), as opposed to “food from somewhere” (McMichael, 2002). The first food regime included canned foods, but the distinctiveness of the second regime was that it converted food manufacturing into a new source of profit, beyond simply cheap basic wage-foods (Magnan, 2016, p. 15). Food industrialization opened up a new frontier of accumulation with high-­fructose corn syrup in particular, and other industrial sweeteners transforming grain markets from conversion to simple food products like bread to raw materials for processed foodstuffs. And, beyond use in the margarine industry, soya oil, along with processed soya meal, combined with hybrid corn for the intensive livestock industry. Thus these three ­complexes—wheat, meat, and durable foods—were progressively intertwined. During and beyond the second food regime, corporate food processors deployed food science and technology for transnational integration, “diversifying sources of their raw materials by substituting the products of tropical agriculture in the Third World (e.g., palm oil) for temperate inputs (soy, maize) used as generic fats and sweeteners” (Magnan, 2016, p. 16). Corn, once the foundation of a maize culture in Central America, is now the key food crop grown for industrial recombination. The global production of corn over the last quarter century has increased by over 100%, compared with 25% for wheat (Winders, 2017, p. 134). It is grown across the world as a feed crop for beef, poultry, eggs, dairy, and pork production, and is a component of sweeteners for candy, cereals, soft drinks, and other supermarket staples (Philpott, 2006). In this way, grains, vegetable oils, and other foodstuffs are woven through the manufacturing of animal proteins and convenience foods, swelling agribusiness profit and rendering diets amenable to seemingly endless financial reconstruction. Such finance-­ driven engineering of food reduces it to a fungible rather than cultural construction. Food is thereby abstracted from its organic relationship with humans (and livestock), particularly via fractionation into reconstituted “products.” The industrial food model universalized following food regime liberalization in the 1990s. While Northern corporations have dominated international trade, for example, “local sales by foreign subsidiaries of U.S. processed food firms are five times the exports of processed food from the U.S. to the rest of the world” (Reardon & Timmer, 2005, p. 28). In a global “retailing revolution,” giant retailers such as Tesco (UK),

64

Rethinking Food and Agriculture

Walmart (USA), Ahold (Netherlands), and Carrefour (France) colonize domestic markets from within, increasingly manufacturing “own brands” (Dixon, 2007). This process includes substitution of domestic foods with privately contracted fresh produce exports by supermarkets: from green beans, tomatoes, and avocados through kiwi fruit and grapes to farmed shrimp (see, e.g., Barndt, 2008). Supermarket-driven agrifood chains have proliferated in response to WTO trade rules by expanding production contracts across the Global South, “thereby widening their consumer base while improving their access to cheap inputs, including labor” (Isakson, 2014, p. 754), and sourcing year-round produce. Recent expansion of agro-estates in Southern Europe, for example, depends on a migrant labor reserve, increasingly sourced from the Middle East and North Africa (Corrado, 2017, p. 3). Low-cost and flexible migrant labor displaced by agribusiness enclosures in Morocco, Tunisia, Egypt, and Turkey enables smaller farmers to compress costs as they struggle to survive withdrawal of Common Agricultural Policy (CAP) subsidies and new market pressures exerted by powerful Northern retail chains. Such pressures take distinct forms in different spaces. Currently, over 85% of India’s food retail value is local, noncorporate retail. But in the last half dozen years, the Indian government has opened the retail sector to foreign direct investment (FDI), provided all goods sold are produced, processed, and manufactured domestically (with no restriction on destination). While proponents claim FDI in food retail would encourage agricultural diversification, reduce food waste, strengthen food security, promote entrepreneurship, and expand employment, the evidence is lacking. Corporate retailing typically restructures domestic food systems into agro-industrial estates serving value chains both domestic and global, which in India means massive farmer displacement, and “the closure of local groceries, endangering the livelihoods and food access of millions of people” (Guttal, 2018, pp. 36–37). Furthermore, the subordination of domestic farm sectors is endemic in the North, as family farming has been destabilized by the costly impact of corporate supplier and trader monopolization (Murphy & Hansen-Kuhn, 2017, p. 8). Across the world, corporate retailing promotes substantial socioecological reorganization. This includes the transformation of rural communities and long-standing relations embedded in local food systems, crop varieties and knowledge, biological processes, hydrological cycles, waste recycling systems, and social diets. The future is uncertain, as such “externalities” are not intrinsically stable. Supermarket “sales can increase, and they can destroy other means of subsistence to increase their sales in regions which rely on local markets, but they cannot create the infrastructures of daily life that underpin reproduction of workers, as well as consumers” (McMichael & Friedmann, 2007, p. 295).

3.5 Food system transitions As a constant in the history of modern capitalism, land enclosure has a new face in the 21st century, as financialization drives much “land grabbing.” Agrifood futures are now incorporated into commodity index funds, enabling speculation on land and food by wealthy investors, large financial houses and pension funds, in

Political economy of the global food and agriculture system65

addition to ­conglomerates in sectors such as oil, auto, chemicals, and agribusiness. Financialization renders land and crops (food, feed, fuel, biomass) fungible for shortterm profitability rather than long-term social needs (McMichael, 2012). Sovereign wealth funds and state enterprises have joined in, as certain food- and fuel-dependent countries (especially East Asian and Middle Eastern), concerned with the rash of food export bans and food riots during the 2008 “food crisis,” pursue “agrosecurity mercantilism,” leasing and/or purchasing land to secure offshore food supplies, rather than adhering to WTO “free market” rules (McMichael, 2013a). This, in addition to the failed 2008 Doha Round (when the G20, led by India, Brazil, and China, challenged Northern agro-export hypocrisy in retaining farm subsidies), has compromised the institutional architecture of the WTO regime. A new multipolar belt of agro-export powers (e.g., Brazil, Argentina, Chile, India, Vietnam, South Africa, Russia) weakens US/EU oligopolistic power. Perhaps in response, in 2013 G8 powers formed the New Alliance for Food Security and Nutrition (NAFSN), composed of the African Union, over 100 companies (e.g., Monsanto, Cargill, Dupont, Syngenta, Nestlé, Unilever, Yara), and several African governments such as Benin, Burkina Faso, Cote d’Ivoire, Ethiopia, Ghana, Malawi, Mozambique, Nigeria, and Tanzania. NAFSN governance mechanisms include policy commitments by African states to facilitate the integration of prime agricultural land into corporate value chains producing food for global consumers (McKeon, 2014).b This large-scale multistate project complements an earlier initiative encouraged by the World Bank at the 2008 World Food Summit in Rome, following recognition of a quarter-century neglect of small-scale producers. Public and philanthropic investment in infrastructural and credit support refocused on incorporating small producers into corporate “value chains.” In the name of improving smallholder productivity, the World Food Summit secured a pivotal agreement for the Alliance for a Green Revolution in Africa to develop a commercial seed sector in Africa, by establishing a broad (10,000) agrodealer infrastructure to encompass farmers in corporate contracts (Patel, 2013). The financial nexus associated with the spread of such contracts tends to chain producers to value relations (and farming practice) beyond their control, capture food for corporate markets, and deepen local food insecurity (McMichael, 2013c). Under such conditions, small- and medium-scale producers often resist debt relations by investing in their farm’s ecological wealth—restoring farming knowledge, biodiversity, and soil health, termed “repeasantization.” Here “ecological capital” represents an alternative “valorization” as the core of the farming enterprise (Altieri & Toledo, 2011; Da Vià, 2012; Levidow, 2015; van der Ploeg, 2018). It is a practice widespread in Europe and Latin America, parts of South Asia and Africa, and emerging on North American family farms.c Resistance to industrial agriculture’s “biophysical b

Note that NAFSN is currently stalled, and under review, since the paucity of results, and the recent withdrawal of France since it preferred to avoid land grabbing and to support small farming systems via agroecological intensification (Le Monde, February 12, 2018). Meanwhile, the African Development Bank appears to be emulating this initiative with a new African Investment Forum (2018), designed to attract pension funds and other institutional finance to reduce the risk of private investment in large-scale agriculture. c E.g., Hylton (2012) and Greenaway (2017).

66

Rethinking Food and Agriculture

override” (Weis, 2007) signals a palpable tension within the food regime, generating a substantial countermovement: reembedding agriculture in ecological relations against “agriculture without farmers” trends with corporate concentration. Corporate capture peaked in 2015, centralizing market power in each segment of the industrial food chain (ETC, 2018, p. 7). But this is old news, as ETC notes: “the bête noires of the food chain used to be Monsanto at one end and Walmart at the other” (ETC, 2018, p. 6). Across these several agribusiness mergers were investments by asset manager BlackRock. With some agribusiness equity shared among half a dozen top asset management companies, their access to information via algorithms on their Big Data platforms provides strategic knowledge of market and merger conditions, as with speculative and investment companies such as Blackstone and the Carlyle Group. Combining digitalization and financialization, global foodscapes will likely be significantly transformed under new forms of control. Obstacles to scale may dissolve as digital blockchains reduce transaction costs by removing intermediaries, enabling start-up fresh food foodies to compete with corporate food processors like Nestlé, Coca-Cola, Tyson, and Unilever. Tyson, a major global meat processor, has now invested in Beyond Meat, a plant-based fake meat start-up, and processors like Unilever, Pepsi, General Mills, and ConAgra are investing in niche market start-ups claiming green nutritional standards, even as the big processors extend their durable food frontiers in the Global South, where obesity is now rampant. Given current industry instability at this critical technological threshold, and the dominance of firms with access to industry associations to weaken standards, “the process of continual buyouts may in itself undermine firms’ ability to meet their stated commitments to sustainability” (IPES-Food, 2017, pp. 59–60).d And at the level of production relations, further displacement of humans from food landscapes may intensify in a trade-off between reductions in energy and chemical use and human touch. Aerial drones, “precision farming,” and “climate smart agriculture”e (Taylor, 2018) perpetuate agro-industrialization, fetishizing information technology, and converting “agriculture without farmers” to “agriculture without humans.”

3.6 Crisis and resolution The crisis of a “broken” global food system is expressed in multiple ways. The food price spikes of 2007–08 followed the exhaustion of productivity gains from the Green Revolution in the 1960s–1980s (Clapp, 2016, p. 173). Exporting countries ceased trading, revealing the limits of food import dependency, and world hunger rose toward a billion people, primarily in the Global South (FAO, 2008, p. 12). Biofuel mandates in the United States and the European Union raised global corn prices by at least d e

See also Howard (2016) on corporate buyouts of the organic food industry in the United States. The Global Alliance for Climate Smart Agriculture includes agribusiness lobbies (60% fertilizer) and a who’s who of food corporations, such as Coca-Cola, DuPont, Kellogg’s, Dow, Monsanto, Walmart, Tyson Foods, PepsiCo, and Unilever (IPES-Food, 2017, p. 75).

Political economy of the global food and agriculture system67

­one-third (Berthelot, 2008, p. 27), underscoring the global conversion of land from food to fuel crops. This also heralded the takeover of commodity agriculture by financial interests decoupled from the food system per se. Meanwhile, farming expenditure in the Global South, as a share of public expenditures, fell by 50% between 1980 and 2004. And the 10 countries accounting for almost 70% of the world’s hungry received only 20% of all agricultural aid, according to the Organization for Economic Cooperation and Development and FAO data (McMichael & Schneider, 2011, p. 121). Finally, the Intergovernmental Science Policy Platform on Biodiversity and Ecosystem Services (IPBES, 2019) reported, in the words of the Chair, Sir Robert Watson: “The health of ecosystems on which we and all other species depend is deteriorating more rapidly than ever. We are eroding the very foundations of our economies, livelihoods, food security, health and quality of life worldwide.” Under these circumstances, any future solutions depend on farming system principles geared to restoring ecosystem health with polycultures, organic fertilizer, and integrated pest management to replace agrochemicals, efficient water use, seed sharing, and energy conversion rather than energy consumption. As the World Bank-sponsored International Assessment of Agricultural Science and Technology for Development Report observed, markets fail to adequately value social and environmental harm, concluding: “business as usual is no longer an option” (IAASTD, 2008, p. 20). In response, IPES-Food experts propose: whether the starting point is industrial agriculture or subsistence-style farming… the agro-ecological alternative is high-tech and knowledge intensive—it requires complex synergies to be built and sustained between different crop varieties and species, and between different farming systems (mixed crop-livestock systems, for instance)…[new evidence] shows the huge potential of these systems to succeed where industrial systems are failing—namely in reconciling concerns such as food security, environmental and livelihood resistance, nutritional adequacy and social equity. (De Schutter & Frison, 2017)

This vision embodies a “multifunctionality” principle. Enshrined programmatically in the European CAP as “environmental governance,” multifunctionality via nonindustrial farming is understood and practiced as a restorative and regenerative principle, where agriculture is embedded in ecological cycles (cf. Hart, McMichael, Milder, & Scherr, 2016). Rather than designate separate spaces to conserve biodiversity and waste sinks, it integrates ecological repair and reproduction into the practice of farming itself (Perfecto, Vandermeer, & Wright, 2009). And its “labor-driven intensification emerges as a strategic…development trajectory” (van der Ploeg, 2009, p. 48). As the Coordination Paysanne Européene noted: “maintaining the number of people working in agriculture is not a sign of economic ‘backwardness’ but an added value” (2003). In other words, it well may be that to retain ecological and landscape intimacy and real sustainability, the large-scale standard of corporate agriculture will yield to smaller-scale landscape farming practices in an ecologically challenged future (cf. Hart et al., 2016). Research by van der Ploeg (2018) and Da Vià (2012) shows a European family farming “peasantry” is robust, cooperative (with seed and

68

Rethinking Food and Agriculture

labor ­sharing), and productive, with an enduring provisioning relation with urban ­consumers—embracing “food from somewhere” rather than “food from nowhere.” The international peasant coalition, under the leadership of the 200 million-strong La Vía Campesina (LVC) movement, first politicized the corporate claim for global “food security” at the 1996 World Food Summit in Rome. Noting that “the massive movement of food around the world is forcing the increased movement of people,” LVC proposed an alternative form of food security, namely “food sovereignty.” They advocated national control over food security and farming systems, in turn respecting local farming knowledges and food cuisines. The strategy was to reverse structural adjustment policies dismantling domestic farming protections, now universalized via WTO liberalization of the corporate food trade. The demand for peasant and family farmer rights to land, as well as to produce food, complemented proposals to redirect huge energy, farm, and export subsidies from agribusiness to local farming systems. Proliferation of programs and policies instituting various forms of “food sovereignty” attests to the salience of protecting and deploying local food systems to address food security and nutrition for citizens rather than relatively affluent global consumersf (Chappell, 2018; Schiavoni, 2017; Trauger, 2014; Wittman & Blesh, 2017). For example: Across the Caribbean, food imports have become a budget-busting problem, prompting one of the world’s most fertile regions to reclaim its agricultural past. But instead of turning to big agribusinesses, officials are recruiting everyone they can to combat the cost of imports, which have roughly doubled in price over the past decade. In Jamaica, Haiti, the Bahamas and elsewhere, local farm-to-table production is not a restaurant sales pitch: it is a government motto. (Cave, 2013, p. 6)

Food provisioning is the Achilles’ heel of government: “failure to provide (food) security undermines the very reason for existence of the political system” (Lagi, Bertrand, & Yaneer, 2011, p. 2). Here, the “food sovereignty” concept is no empty slogan, its power stemming from the evident shortcomings of a state system embedded in a food regime in which land and its uses are up for (extraterritorial) grabs. The peasant movement is, in effect, the mouthpiece for “food sovereignty,” given that its constituents are on the frontlines of land enclosure in the deepening “resource grab.” Its soft political power was recognized in 2018 with passage of the UN Rights of Peasants and Other Peoples Working in Rural Areas—partly stemming from its effective presence in the UN Committee on World Food Security. Here, agrarian social movements and progressive nongovernmental organizations form the Civil Society Mechanism, advocating land rights and reform/redistribution, tenure guidelines (notably security for women farmers), seed sovereignty, market access, financial ­regulation, f

“The 1986 World Bank Report ‘Poverty and Hunger’ induced an important shift in how food security was to be deployed since it redefined food security as linked, not only to national food production, but also to individual purchasing power” (Jarosz, 2009, p. 171).

Political economy of the global food and agriculture system69

agroecology, and so on (Gaarde, 2017). Such food sovereignty principles reach beyond the “peasantry” to inform practices and experiments across a global variety of rural, and urban, spaces (Andreé, Ayres, Bosia, & Massicotte, 2014). Alongside commercial farmers converting to ecological practices as agro-input costs inflate with corporate monopoly (Wise, 2019), nested markets across the world integrate local producers and consumers into solidary economies, and small farmer networks of seed and information exchange embed food sovereignty principles in the interstices of the formal economy (Da Vià, 2012; Fonte, 2013; Holt-Giménez, 2006; van der Ploeg, Jingzhong, & Schneider, 2012). This includes a plethora of local food initiatives in rural and urban spaces alike, as Food Councils form (Chappell, 2018), urban gardens proliferate,g unemployed/informal workers “return” to the land, and long-ignored indigenous territorial rights emerge as a new terrain of political struggle (Borras Jr & Franco, 2012; Mayes, 2018). Accordingly, Mexico’s Obrador (AMLO) government favors self-sufficiency in basic grains and improved livelihoods for family farmers and rural communities, following displacement of almost 5 million family farmers via NAFTA. And, following Brazil, ecological farming is gaining in India where a Zero Budget Natural Farming program is transitioning millions of farms to chemical-free agriculture (Khadse, Rosset, Morales, & Ferguson, 2018). Meanwhile, a growing consensus among researchers, practitioners, and official organizations (e.g., Adidja, Mwine, Majaliwa, & Ssekandi, 2019; Badgley et  al., 2007; FAO, 2018; FAO & IFAD, 2019; IAASTD, 2008; IPES-Food, 2018; Pretty et al., 2006; Rosset & Altieri, 2017; Smithers, 2019) attests to the resilience of agroecology and its parallel productivity to corporate agriculture. Notably, a UN study found “organic agriculture outperformed conventional production systems based on chemical-intensive farming and is thus more conducive to food security in Africa” (UN Conference on Trade and Development (UNCTAD) and UN Environment Programme (UNEP), 2008, p. 236). In consequence, organic and/or agroecological methods increasingly inform policy recommendations. At the same time, agroecology’s growing value faces appropriation by agribusiness interests to contain it (HoltGiménez, 2019). Arguably, this tension will shape future developments in the global food system, as ecological damage, food shortages, and contentious trade and land grab politics deepen. The point is that the evolving tensions between industrial and agroecological versions of food production (not to mention diversion of land to feed and fuel crops) will play out on both political and biophysical terrains—within states and their national markets. These markets are those upon which the world majority (notably low-income consumers) depends even now, however tenuously. And the politics of these markets, and their local suppliers, will become even more significant as industrial food’s fossil fuel dependence, in combination with food inflation and trade disruption, render global sourcing increasingly nonviable.

g

As much as 40% of the population of some African cities and up to 50% in some Latin American cities engage in urban or periurban agriculture.

70

Rethinking Food and Agriculture

Acknowledgments The author wishes to acknowledge the helpful input of three anonymous reviewers as well as Laila and Amir Kassam in the development of this chapter.

References Adidja, M. W., Mwine, J., Majaliwa, J. G. M., & Ssekandi, J. (2019). The contribution of agro-ecology as a solution to hunger in the world: A review. Asian Journal of Agricultural Extension, Economics & Sociology, 33(2), 1–22. Akram-Lodhi, A. H. (2013). Hungry for change. farmers, food justice and the agrarian question. Halifax and Winnipeg: Fernwood Publishing. Altieri, M., & Toledo, V. (2011). The agroecological revolution in Latin America: Rescuing nature, ensuring food sovereignty and empowering peasants. Journal of Peasant Studies, 38(3), 587–612. Andrae, G., & Beckman, B. (1985). The wheat trap: Bread and underdevelopment in Nigeria. London: Zed Books. Andreé, P., Ayres, J., Bosia, M. J., & Massicotte, M.-J. (2014). Globalization and food sovereignty. Global and local change in the new politics of food. Toronto: University of Toronto Press. Araghi, F. (2009). The invisible hand and the visible foot: Peasants, dispossession and globalization. In H. Akram-Lodhi, & C. Kay (Eds.), Peasants and globalization: Political economy, rural transformation and the agrarian question. London & New York: Routledge. Badgley, C., Moghtader, J., Quintero, E., Zakem, E., Chappell, M. J., Aviles-Vazquez, K., … Perfecto, I. (2007). Organic agriculture and the global food supply. Renewable Agriculture and Food Systems, 22(2), 86–108. Baines, J. (2015). Fuel, feed and the corporate restructuring of the food regime. Journal of Peasant Studies, 42(2), 295–321. Barndt, D. (2008). Tangled routes. Women, work and globalization on the tomato trail (2nd ed.). New York: Rowman & Littlefield. Berthelot, J. (2008). The food crisis explosion: Root causes and how to regulate them. Kurswechsel, 3, 23–31. Bezner Kerr, R. (2010). The land is changing: Contested agricultural narratives in northern Malawi. In P. McMichael (Ed.), Contesting development: Critical struggles for social change. London & New York: Routledge. Bishop, R. V., Christensen, L. A., Mercier, S., & Witucki, L. (1990). The world poultry market: Government intervention and multilateral policy reform. Washington, DC: USDA Economic Research Service. Borras, S. M., Jr., & Franco, J. (2012). A “land sovereignty” alternative? Towards a peoples’ counter-enclosure. In TNI Agrarian Justice Programme. Amsterdam, the Netherlands: Transnational Institute. 2012. Bove, J., & Dufour, F. (2001). The world is not for sale. London: Verso. Cave, D. (2013). As cost of importing food soars, Jamaica turns to the Earth. The New York Times (August 4). Chappell, J. (2018). Beginning to end hunger: Food and the environment in Belo Horizonte, Brazil, and beyond. Oakland: University of California Press.

Political economy of the global food and agriculture system71

Chirot, D. (1977). Social change in the twentieth century. New York: Harcourt, Brace, Jovanovich. Clapp, J. (2016). Food. Cambridge: Polity Press. Cleaver, H. (1977). Food, famine, and the international crisis. Zerowork, 2, 7–70. Connell, R., & Dados, N. (2014). Where in the world does neoliberalism come from? The market agenda in southern perspective. Theory and Society, 43, 117–138. Corrado, A., Corrado, A., De Castro, A., & Perrotta, D. (2017). Agrarian change and migrations in the Mediterranean from a food regime perspective. In Migration and agriculture. Mobility and change in the Mediterranean area (pp. 311–331). London & New York: Routledge. Da Vià, E. (2012). Seed diversity, farmers’ rights, and the politics of re-peasantization. International Journal of Sociology of Agriculture and Food, 19(2), 229–242. Dalrymple, D. (1985). The development and adoption of high-yielding varieties of wheat and rice in developing countries. American Journal of Agricultural Economics, 67, 1067–1073. Das, P. (2016). The shocking tale of India’s “Cancer Train”. Business Insider. June 10. Available at: https://www.businessinsider.in/the-shocking-tale-of-indias-cancer-train/articleshow/52690219.cms. Davis, M. (2001). Late Victorian holocausts. El Niño famines and the making of the third world. London: Verso. de Mareschel, E. (2018). Baguettes in Beijing? Chinese investors eating up French farm land. World Crunch. June 4. Available at: https://www.worldcrunch.com/business-finance/ baguettes-in-beijing-chinese-investors-eating-up-french-farm-land. De Schutter, O., & Frison, E. (2017). Modern agriculture cultivates climate change—We must nurture biodiversity. The Guardian (9 January). DeWalt, B. (1985). Mexico’s second green revolution: Food for feed. Mexican Studies/Estudios Mexicanos, 1, 29–60. Dixon, J. (2007). Supermarkets as new food authorities. In D. Burch, & G. Lawrence (Eds.), Supermarkets and agri-food supply chains. Transformations in the production and consumption of foods (pp. 29–50). Cheltenham, UK & Northampton, MA: Edward Elgar. EAT-Lancet Commission. (2019). Food in the anthropocene: Healthy diets from sustainable food systems. Available at: https://www.thelancet.com/journals/lancet/article/ PIIS0140-6736(18)31788-4/fulltext. ETC. (2009). Who will feed us?. www.etcgroup.org. ETC. (2018). Between BlackRock and a hard place, communique #116. www.etcgroup.org. FAO. (2008). The state of food insecurity in the world 2008: High food prices and food security—Threats and opportunities. Available at: http://www.fao.org/docrep/fao/011/i0291e/ i0291e00a.htm. FAO. (2018). FAO’s work on agroecology. A pathway to achieving the SDGs. Rome: FAO. FAO and IFAD. (2019). United Nations decade of family farming 2019-2028. Global action plan. Rome: FAO. Flanner, T. (2009). Now or never: Why we must act now to end climate change and create a sustainable future. New York: Atlantic Monthly Press. Fonte, M. (2013). Food consumption as social practice: Solidarity purchasing groups in Rome, Italy. Journal of Rural Studies, 32, 230–239. Friedmann, H. (1990). The origins of third world food dependence. In H. Bernstein, B. Crow, M. Mackintosh, & C. Martin (Eds.), The food question: Profits versus people?. London: Earthscan. Friedmann, H. (1993). The political economy of food: A global crisis. New Left Review, 197, 29–57.

72

Rethinking Food and Agriculture

Friedmann, H. (1994). Distance and durability: Shaky foundations of the world food economy. In P. McMichael (Ed.), The global restructuring of agro-food systems. Ithaca: Cornell University Press. Friedmann, H., & McMichael, P. (1989). Agriculture and the state system: The rise and fall of national agricultures, 1870 to the present. Sociologia Ruralis, 29(2), 93–117. Gaarde, I. (2017). Peasants negotiating a global policy space. La Vía Campesina in the Committee on World Food Security. London & New York: Routledge/Earthscan. George, S. (1977). How the other half dies: The real reasons for world hunger. Montclair, NJ: Allen, Osmun and Co. George, S. (1988). A fate worse than debt. New York: New Grove Press. Gonzalez-Esteban, A. L. (2018). Patterns of world wheat trade, 1945-2010. The long hangover from the second food regime. Journal of Agrarian Change, 18(1), 87–111. Greenaway, T. (2017). Where corn is king, the stirrings of a renaissance in small grains. Yale Environment 360. November 28. Available at http://e360.yale.edu/features/ where-corn-is-king-the-stirrings-of-a-small-grain-renaissance. Grimsditch, M. (2017). Chinese agriculture in Southeast Asia: Investment, aid and trade in Cambodia. Laos and Myanmar: Heinrich Boll Stiftung. Available at: https://th.boell.org/ en/2017/06/22/chinese-agriculture-southeast-asia-investment-aid-and-trade-cambodialaos-and-myanmar. Gupta, A. (1998). Postcolonial developments: Agriculture in the making of modern India. Durham: Duke University. Guttal, S. (2018). The changing face of food retail in India. In 10. Right to food and nutrition watch. When food becomes immaterial: confronting the digital age (pp. 33–39). Handley, P. (1990). Food for thought: Thailand’s Charoen Pokphand plans London listing. Far Eastern Economic Review, 25, 56–58. October. Hart, A., McMichael, P., Milder, J. C., & Scherr, S. J. (2016). Multi-functional landscapes from the grassroots? The role of rural producer movements. Agriculture and Human Values, 33(2), 305–322. Hobsbawm, E. (1969). Industry and empire. Harmondsworth: Penguin Books. Holt-Giménez, E. (2006). Campesino-á-Campesino. Voices from Latin American’s farmer to farmer movement for sustainable agriculture. Oakland: Food First Books. Holt-Giménez, E. (2019). Can we feed the world without destroying it?. Cambridge, UK & and Medford, MA: Polity Press. Howard, P. H. (2016). Concentration and power in the food system: Who controls what we eat?. New York: Bloomsbury. Huang, S., & Coyle, W. T. (1989). Structural change in east Asian agriculture. In Pacific Rim agriculture and trade report. Washington, DC: USDA Economic Research Service. Hylton, W. S. (2012). Broken heartland. The looming collapse of agriculture on the Great Plains. Harpers Magazine. July. IAASTD: International Assessment of Agricultural Knowledge, Science and Technology for Development. (2008). Executive summary of the synthesis report. Available at: www.agassessment.org/docs/SR_Exec_Sum_280508_English.pdf. IPBES. (2019). Inter-governmental Science Platform on Biodiversity and Ecosystem Services (IPBES) to launch first global assessment report on biodiversity and ecosystem services since 2005, a primer. Available at https://ipbes.net/sites/default/files/downloads/ipbes_ global_assessment_primer_english.pdf. IPES-Food. (2017). Too big to feed, report. 3. IPES-Food. (2018). Breaking away from industrial food and farming systems. In Case studies (p. 2).

Political economy of the global food and agriculture system73

Isakson, R. (2014). Food and finance: The financial transformation of agro-food supply chains. Journal of Peasant Studies, 41(4), 741–775. Jarosz, L. (2009). The political economy of global governance and world food crisis: The case of the FAO. Review XXXII, 1, 37–60. Khadse, A., Rosset, P., Morales, H., & Ferguson, B. G. (2018). Taking Agroecology to scale: The zero budget natural farming peasant movement in Karnataka, India. Journal of Peasant Studies, 45(1), 192–219. Kimbrell, A. (2002). The fatal harvest reader: The tragedy of industrial agriculture. Washington, DC: Island Press. Kumar, R. (2016). Putting wheat in its place, or why the Green Revolution wasn’t quite what it is made out to be. The Wire. 31 October. Available at: https://thewire.in/agriculture/ green-revolution-borlaug-food-security. Lagi, M., Bertrand, K. Z., & Yaneer, B. Y. (2011). The food crises and political instability in North Africa and the Middle East. Cambridge, MA: New England Complex Systems Institute. Levidow, L. (2015). European transitions towards a corporate-environmental food regime: Agroecological incorporation of contestation? Journal of Rural Studies, 40, 76–89. Lewis, A. (1954). Economic development with unlimited supplies of labour. The Manchester School, 22, 139–191. Madeley, J. (2000). Hungry for trade. London: Zed. Magnan, A. (2016). When wheat was king: The rise and fall of the Canadian-UK grain trade. Vancouver: University of British Columbia Press. Mayes, C. (2018). Unsettling food politics. agriculture, dispossession and sovereignty in Australia. London & New York: Rowman and Littlefield. McKeon, N. (2014). The new Alliance for food security and nutrition. A coup for corporate capital?. Amsterdam: Transnational Institute. McMichael, P. (1996). Development and social change. A global perspective. Thousand Oaks: Pine Forge Press. McMichael, P. (2002). La restructuration globale des systems agroalimentaires. Mondes en Developpment, 30(117), 45–54. McMichael, P. (2009). A food regime genealogy. Journal of Peasant Studies, 36(1), 139–170. McMichael, P. (2012). The land grab and corporate food regime restructuring. Journal of Peasant Studies, 39(3–4), 686–689. McMichael, P. (2013a). Food regimes and agrarian questions. Halifax, NS: Fernwood Press. McMichael, P. (2013b). Land grabbing as security mercantilism in international relations. Globalizations, 10(1), 47–64. McMichael, P. (2013c). Value-chain agriculture and debt relations: Contradictory outcomes. Third World Quarterly, 34(4), 671–690. McMichael, P., & Friedmann, H. (2007). Situating the “retailing revolution”. In D. Burch, & G. Lawrence (Eds.), Supermarkets and agri-food supply chains. Transformations in the production and consumption of foods (pp. 291–319). Cheltenham, UK & Northampton, MA: Edward Elgar. McMichael, P., & Raynolds, L. (1994). Capitalism, agriculture, and world economy. In L. Sklair (Ed.), Capitalism and development. London: Routledge and Kean Paul. McMichael, P., & Schneider, M. (2011). Food security politics and the millennium development goals. Third World Quarterly, 32(1), 119–139. Mintz, S. (1985). Sweetness and power. The place of sugar in modern history. New York: Vintage. Mitchell, T. (1991). American’s Egypt: Discourse of the development industry. MERIP, 169, 18–34.

74

Rethinking Food and Agriculture

Monbiot, G. (2019). The best way to save the planet? Drop meat and dairy. Farming livestock for food threatens all life on earth. The Guardian Weekly. June 15. Moore, J. (2011). Transcending the metabolic rift: A theory of crises in the capitalist world-­ ecology. Journal of Peasant Studies, 38(1), 1–46. Moore, B., Jr. (1967). Social origins of dictatorship and democracy. Boston: Beacon Press. Murphy, S., & Hansen-Kuhn, K. (2017). Counting the costs of agricultural dumping. Institute for Agriculture and Trade Policy. June. Nepstad, D. (2006). Diet for a hot planet. Boston Globe. 22 November. Available at: http://www.boston.com/news/globe/editorial_opinion/oped/articles/2006/11/22/ diet_for_a_hot_planet/. Patel, R. (2013). The long green revolution. Journal of Peasant Studies, 40(1), 1–63. Perfecto, I. J., Vandermeer, J., & Wright, A. (2009). Nature's matrix: Linking agriculture, conservation and food sovereignty. London: Earthscan. Petras, J., & Veltmeyer, H. (2014). The new extractivism. A post-neoliberal development model or imperialism of the twenty-first century?. London & New York: Zed Books. Philpott, T. (2006). Feeding the beast. Grist. December 13. Available at: www.grist.org. Pretty, J., Noble, A. D., Bossio, D., Dixon, J., Hine, R. E., Penning de Vries, F. W. T., & Morison, J. I. L. (2006). Resource conserving agriculture increases yields in developing countries. Environmental Science & Technology, 40(4), 1114–1119. Raikes, P. (1988). Modernising hunger: Famine, food surplus and farm policies in the EU and Africa. London: Catholic Institute for International Affairs. Reardon, T., & Timmer, C. (2005). Transformation of markets for agricultural output in developing countries since 1950: How has thinking changed? In R. Evenson, P. Pingali, & T. Schultz (Eds.), Handbook of agricultural economics—Agricultural development: Farmers, farm production and farm markets. Oxford: Elsevier Press. Rist, G. (1977). The history of development: From Western origins to global faith. London: Zed Books. Roberts, W. (2008). The no-nonsense guide to world food. Oxford: New Internationalist. Rosset, P., & Altieri, M. (2017). Agroecology: Science and politics. Halifax, NS: Fernwood Press. Rostow, W. W. (1960). The stages of economic growth: A non-communist manifesto. Cambridge, UK: Cambridge University Press. Rothacher, A. (1989). Japan’s agro-food sector. The politics and economics of excess protection. New York: St. Martin’s Press. Sachs, J. (2005). The end of poverty: Economic possibilities for our time. New York: Penguin. Schiavoni, C. M. (2017). The contested terrain of food sovereignty construction: Towards a historical, relational and interactive approach. Journal of Peasant Studies, 44(1), 1–32. Schneider, M. (2011). Feeding China’s pigs. Implications for the environment, China’s smallholder farmers and food security. Institute for Agriculture and Trade Policy. May. Schneider, M. (2014). Developing the meat grab. Journal of Peasant Studies, 41(4), 613–634. Shiva, V. (1991). The violence of the green revolution. London: Zed Books. 1991. Smithers, R. (2019). European farms could grow green and still be able to feed population. The Guardian. 20 February. Available at: https://www.theguardian.com/environment/2019/ feb/20/european-farms-could-grow-green-and-still-be-able-to-feed-population. Stone, G. D. (2019). Commentary: New histories of the Indian green revolution. The Geography Journal, 185(2), 243–250. Taylor, M. (2018). Climate-smart agriculture: What is it good for? Journal of Peasant Studies, 45(1), 89–107.

Political economy of the global food and agriculture system75

Trauger, A. (2014). Toward a political geography of food sovereignty: Transforming territory, exchange and power in the liberal sovereign state. The Journal of Peasant Studies, 41(6), 1131–1152. UN Conference on Trade and Development (UNCTAD) and UN Environment Programme (UNEP). (2008). Organic agriculture and food security in Africa. New York: United Nations. van der Ploeg, J. D. (2009). The new peasantries: Struggles for autonomy and sustainability in an era of empire and globalization. London: Earthscan. van der Ploeg, J. D. (2018). The new peasantries: Rural development in times of globalization. London, New York: Routledge. van der Ploeg, J. D., Jingzhong, Y., & Schneider, S. (2012). Rural development through the construction of new, nested, markets: Comparative perspectives from China, Brazil and the European Union. Journal of Peasant Studies, 39(1), 133–173. Weis, T. (2007). The global food economy. The battle for the future of farming. London: Zed Books. Weis, T. (2013). The ecological hoofprint. The global burden of industrial livestock. London: Zed Books. Winders, B. (2017). Grains. Cambridge: Polity Press. Wise, T. A. (2019). Eating tomorrow. Agribusiness, family farmers and the battle for the future of food. New York: The New Press. Wittman, H., & Blesh, J. (2017). Food sovereignty and Fome Zero: Connecting public food procurement programmes to sustainable rural development in Brazil. Journal of Agrarian Change, 17(1), 81–105. World Bank. (2007). World development report, 2018. Washington, DC: World Bank.

This page intentionally left blank

Neocolonialism and the New Alliance for Food Security and Nutrition: A gendered analysis of the development consequences for Africa

4

Mark Langana and Sophia Priceb a Newcastle University, Newcastle upon Tyne, United Kingdom, bLeeds Beckett University, Leeds, United Kingdom

4.1 Introduction National governments and the international community as a whole have embraced their responsibility to promote responsible and secure food systems as part of the UN Sustainable Development Goals (SDGs). The UN framework notably places a significant emphasis on food security, specifically in relation to Goal 2 on Zero Hunger. Policy-makers regularly cite the World Food Crisis of 2007–08 as an important marker in donor community understandings of the fragility of food production and food distribution networks, with particular regard to the wellbeing of poorer communities in sub-Saharan Africa. An SDG focus on food security is thereby seen to mitigate against the potential revisiting of a World Food Crisis by encouraging development partners— governments, civil society, and crucially private sector actors (in chime with Goal 8: Decent work and economic growth, of the SDG agenda)—to better plan for the sustainable production and supply of food to meet the nutritional needs of vulnerable citizenries. A key initiative hailed as part of wider SDG endeavors is a corporate-led New Alliance for Food Security and Nutrition (NAFSN). The NAFSN itself emerged in the aftermath of the World Food Crisis, with founding businesses such as Syngenta arguing that corporate leadership was necessary to boost food production in the Global South, particularly in sub-Saharan Africa. NAFSN communications have since argued that existing agricultural systems are often inefficient and underresourced, while valuable land tracts often go unutilized, exacerbating potential conditions for food insecurity and hunger in African countries. Leading corporate partners in the NAFSN, including Monsanto, Diageo, SABMiller, Unilever, and Syngenta, have thus committed themselves to assisting African governments—working alongside G8 donor partners such as the Canadian and UK governments (and civil society)—to improve agricultural production systems in Africa. In the post-2015 setting of the SDGs, this endeavor is framed in terms of a discourse of sustainable development, emphasizing Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00004-9 © 2021 Elsevier Inc. All rights reserved.

78

Rethinking Food and Agriculture

that ­ecological, social, and economic goals can be balanced through corporate-backed NAFSN initiatives in Africa to bolster food supplies and to guard against another World Food Crisis. Worryingly, however, the activities of the NAFSN—and in particular corporate-led initiatives aimed at accessing land corridors in sub-Saharan Africa—have led to accusations of land grabbing. Namely, that the discourse of sustainable development and food security is being utilized by corporate partners to legitimize (and veil) their commercial motivations in gaining access to valuable land tracts in African countries. Corporate partners are thereby seen to adopt a “development” language, whereas the tangible impact of land-grabbing can oftentimes have severely negative consequences for local subsistence farmers (denied access to land resources) and for surrounding communities (in terms of food precarity as well as pollution emanating from intensive agribusiness). In this context, certain African civil society campaigners and commentators accuse corporations and donor institutions of perpetrating “neocolonial” exploitation on African societies. Namely, productive African land tracts are being taken from local communities, often via the promises of aid (inducements) to African elites. Genuine, empirical sovereignty of African countries is therefore undermined, as domestic agricultural systems are supplanted by corporate-led production in land corridors, despite the negative consequences this may have for local communities (abandoned by their elites who are co-opted by promises of donor aid-giving to facilitate land-grabs). In this context, the chapter examines the concept of neocolonialism and the rationales of the NAFSN in further detail. It thus critically explores the discourse of sustainable development—and indeed of private sector development (PSD)—which has been utilized as a means of justifying NAFSN corporate-led activities as a contribution to responsible food systems and social justice in Africa. Moreover, the chapter then turns attention to the particularly gendered implications of neocolonial land-grabbing in terms of the situation of women producers within African agricultural systems. This draws attention to the gendered family and business norms that impact on the wellbeing of women in relation to corporate-led agribusiness activities in land corridors. The chapter then concludes with a summary of concerns regarding the NAFSN, its co-optation of “development” language, and its perpetration of a form of neocolonial incursion into African agricultural systems.

4.2 Neocolonialism as a critique of donor aid policies in the Global South Neocolonialism is a contested and now often overlooked concept for making sense of North-South power asymmetries. Most usefully defined by Kwame Nkrumah, the first president of an independent Ghana, the concept invokes the situation where a former colony’s sovereignty is compromised to the extent that its domestic elites and citizenry are no longer masters of their own political and economic trajectory (Langan, 2017). In relations of neocolonialism therefore a former colony retains legal/

Neocolonialism and new alliance for food security and nutrition

79

juridical ­sovereignty, or what Julius Nyerere memorably termed “flag independence.” However, it does not enjoy empirical sovereignty in the sense of being able to devise and implement policy agendas based on the social and economic needs of its own population (Langan, 2017). With much relevance for a contemporary critique of the UN SDGs, the NAFSN, and corporate land-grabbing, Nkrumah warned that aid monies were a primary tool of donor powers that sought to enact and perpetuate neocolonial forms of relations with countries of the Global South. Nkrumah insisted that aid monies were often given as a means of co-optation, ensuring that domestic elites would remain beholden to foreign patrons to maintain access to finances necessary for their own political survival (by lubricating their own local patronage networks). Moreover, Nkrumah highlighted that such aid monies would be tied to the economic interests of the foreign power, to the extent that the original expenditure of aid finance would be more than recouped over time by the donor (Langan, 2017, pp. 4–6). Specifically, Nkrumah warned that through aid-giving, foreign countries and their corporations would gain monopoly over the key strategic economic sectors in the neocolony. And with such economic power in place, foreign countries and corporations could then dictate political policy decisions to local elites—by threatening economic collapse and aid withdrawal should they attempt to strike a genuinely autonomous and independent political agenda. As such in the neocolonial situation, the promise of independence transformed into the illusion of sovereignty whereby the neocolony remained both subordinate and dependent on its (often former colonial) donor power (Langan, 2017). Most tragically, for Nkrumah, this situation also meant that economic and social development would not meaningfully be forthcoming in the neocolony. Domestic elites under the sway of external influence would direct their attention not to the material wellbeing of the citizenry, but instead to the perpetuation of colonial-style trade and aid systems that benefited their patron and themselves via graft. Nkrumah (1965, p. 1) lamented in these terms that: The rulers of the neo-colonial states derive their authority to govern… from the support which they obtain from their neo-colonialist masters. They have therefore little interest in developing education, strengthening the bargaining power of their workers employed by expatriate firms, or indeed of taking any step which would challenge the colonial patterns of commerce… [which] is the object of neocolonialism to preserve.

Despite the fact that neocolonialism is now often regarded as an outmoded concept—a relic of a Cold War era—the writings of Nkrumah do bear relevance for a contemporary interrogation of North-South relations, not least in terms of food security and land-grabbing associated with the current UN SDGs. While international bodies and donor nations speak the language (discourse) of sustainable development, nevertheless, further scrutiny belies their claims of promoting responsible food systems in terms of the landmark NAFSN initiative. Namely, the marriage of corporate and donor interests—as exemplified by the NAFSN—in maintaining colonial-style production and trade systems does much to verify and corroborate Nkrumah’s predictions, irrespective of the egalitarian discourse of the SDGs.

80

Rethinking Food and Agriculture

4.3 The NAFSN: Sustainable development or a form of “neocolonial” imposition? Donor nations and international bodies have asserted their responsibilities to promote responsible and secure food systems in sub-Saharan Africa in the timeframe of the UN SDGs. One of their key vehicles for achieving this outcome is the NAFSN. This landmark initiative emerged in 2012 as a corporate-led endeavor to bring together donor institutions and donor governments, agribusiness leaders such as Unilever, and African governments to (ostensibly) combat the problem of inefficient food networks in sub-Saharan Africa. To date this has involved NAFSN interventions in 10 African nations—namely, Benin, Burkina Faso, Ivory Coast, Ethiopia, Ghana, Malawi, Mozambique, Nigeria, Senegal, and Tanzania (NAFSN, 2014). Specifically, the NAFSN focuses on the concept of agricultural land corridors—meaning the creation of land tracts in which corporations take the lead in improving production, largely in the form of plantation-based systems. Corporate partners within the NAFSN sign letters of intent with African governments and donor institutions with regards to their ambitions to improve food systems (McKeon, 2014, pp. 3–5). African governments then facilitate land “releases” to the companies in question. This is underpinned by aid-giving from the donor institutions to assist the African governments with their broader “development” requirements, with a view to the UN SDGs and particularly Goal 2 on Zero Hunger. An ActionAid (2015) report makes clear that G8 donors (including the United States, Russia, Canada, Japan, and the European Union [EU]) are committing large sums of aid toward the initiative—in the hope of realizing public– private partnerships (PPPs) with corporate partners within the NAFSN: [G8 donors have] committed $4.4 billion to the 10 [African] countries of the New Alliance… the G8 support… is part of a drive to secure larger agricultural markets and sources of supply in Africa for multinational corporations. New Alliance partners such as Monsanto, Diageo, SABMiller, Unilever, Syngenta have major commercial interests in Africa and close connections with Northern governments.

Importantly, the functioning of the NAFSN aligns to broader Post-Washington Consensus (PWC) norms regarding the need for African countries to abide by ­market-friendly policies to facilitate their integration into globalized free markets. As such, African governments’ participation in NAFSN initiatives involves the formal signing of a Cooperative Framework Agreement (CFA) between the African country in question, a named donor partner given responsibility for NAFSN rollout in that African nation, and NAFSN corporate partners. The CFA ensures that the African government will abide by PWC norms regarding the importance of PSD and business activities in liberalized markets. In particular, the document emphasizes that private property rights will be respected with regards to the land corridor to be established in conjunction with NAFSN corporate partners’ investments. In some cases there is also detail about the African government’s responsibilities with respect to intellectual property rights, again with an eye to the rights of NAFSN corporate players (Oakland Institute, 2016). The divvying up of responsibilities for the

Neocolonialism and new alliance for food security and nutrition

81

o­ versight of NAFSN operations in any given African nation among G8 donors (such as the United Kingdom being responsible for Nigerian NAFSN activities) has led to accusations of a “new scramble” for Africa, and of a perpetuation of neocolonial forms of North-South relations, in which G8 donor nations utilize aid in relation to CFA policy conditionalities (Frynas & Paulo, 2007). The concept of neocolonialism here has roots back to African luminaries of the immediate era of decolonization, notably as discussed in Kwame Nkrumah (1965) who emphasizes that aid monies (among other forms of co-option and coercion) would be utilized to denude genuine empirical sovereignty in legally independent African states, and to ensure donor nations continued to wield political and economic dominance—including in terms of agricultural production (as apparently verified in terms of the NAFSN). A report for the European Parliament (De Schutter, 2015) itself highlighted, and queried, the deep reforms necessary within the CFAs under NAFSN auspices in relation to African countries’ policy space. Importantly, however, the NAFSN as an initiative articulates a legitimizing discourse surrounding its activities, which seeks to immunize it from criticisms of undue influence or neocolonialism in African polities. In particular, the NAFSN emphasizes that it seeks to modernize African agricultural systems, to make them more robust and efficient, and to thereby establish jobs as well as securing food supplies for vulnerable populations. This is justified also in terms of the language of PPPs and the ability of donor aid monies to blend with private sector capital for positive poverty reduction results in developing economies. Cooperation between corporate partners and G8 donor nations (as well as donor institutions, including the World Bank) is hailed as an innovative form of aid-giving, in keeping with Goal 8 of the SDGs to realize the role of the private sector in development, and of the primacy of economic growth strategies for poverty alleviation. Interesting in the context of the post-2015 agenda and the emergence of the SDGs themselves, the NAFSN also articulates a prominent discourse of “sustainable development” and of the importance (and possibility) within its schemes to balance social outcomes, ecological security, and economic prosperity in sub-Saharan Africa. Moreover, NAFSN activities now specifically align to Goal 2 of the SDGs with regards to the Zero Hunger pledge, thereby again presenting a legitimizing discourse of humanitarian wellbeing in relation to vulnerable communities and citizens in African countries. In one example of its pro-poor development narratives, the NAFSN insists that its interventions are necessary since governments cannot achieve food security objectives in isolation (and hence PPP initiatives under NAFSN auspices are deemed as legitimate endeavors): The New Alliance for Food Security and Nutrition is a shared commitment to achieve sustained inclusive, agriculture-led growth in Africa. Given the overwhelming importance of African agriculture in rural livelihoods and its enormous potential to bring people out of poverty, public investment in food security and agriculture has significantly increased over the last decade. While this investment is having an impact, the path to sustainable food security cannot be forged by governments alone. Agricultural transformation in Africa is a shared interest of the public and private sectors and presents a unique opportunity for a new model of partnership (NAFSN, 2019).

82

Rethinking Food and Agriculture

In a benevolent discourse of partnership and agricultural transformation leading to better livelihoods and conditions for poorer communities in Africa therefore, the NAFSN effectively distances itself away from concerns regarding the potential neocolonial implication of the marriage of donor aid monies to corporate land-grabs in the continent. In keeping with PWC norms of equitable free markets leading to social prosperity and fairer North-South relations, moreover, the NAFSN fits itself within wider donor understandings of the need for PPPs to enable developing countries to better realize the “fruits of growth” within global markets. However, when the specific allegations of land-grabbing are considered in terms of the NAFSN’s emphasis on the creation of “agricultural corridors” in which corporate partners can invest and establish plantations, it does become apparent that there are highly questionable outcomes of this initiative for poorer African citizenries. This is especially so in terms of subsistence farmers who otherwise have been dependent on access to land tracts to feed their families (albeit in terms of the NAFSN’s modernization discourse, such smallholders are often dismissed as being inefficient and unable to invest sufficient resources—fertilizers and so forth—necessary for agricultural transformation). The ActionAid report (2015, p. 13) on NAFSN activities convincingly sets out many of these concerns. It explains that these agricultural corridors (also known as staple-crop processing zones) are defined by NAFSN partners as: large areas of land that are earmarked for agribusiness. In these zones, companies are incentivised by host governments and supporting donors to establish their operations by a series of tax, regulatory and land incentives, as well as by new infrastructure (roads, railways, ports, irrigation, storage, processing facilities, etc). The projects focus mainly on agriculture, but also include forestry and mining. To ensure big business acquires these large tracts of land, governments are promoting reforms to change land tenure legislation.

Interestingly here, the very concept and phraseology of “agricultural corridors” was apparently initiated by Yara, a major corporate player in the fertilizer industry and one of the founders of the NAFSN itself (Pan Africanist Briefs, 2014). Worryingly, the land transfers that African governments have agreed to with regards to the creation of these agricultural corridors involve very substantial amounts of fertile agricultural resources. For instance, the Government of Malawi have agreed under NAFSN auspices for the transfer (or release) of 200,000 ha. The country’s National Export Strategy, furthermore, emphasizes that in fact up to 1 million hectares in total may be released to agribusiness as part of agricultural modernization strategies, as advocated within NAFSN—and broader PWC—discourse regarding efficiency and integration into globalized markets. Such scale of transfer, if brought to fruition, would account for around 26% of Malawi’s total available arable land (ActionAid, 2015, p. 1). Furthermore, there are deep concerns that the creation of these land corridors involves the forcible removal of subsistence farmers against their will (McMichael, 2015, p. 442). African governments’ own security apparatuses have been implicated in these events, bringing again to the fore the concept of neocolonialism in terms of African polities being co-opted by external nations’ commercial interests

Neocolonialism and new alliance for food security and nutrition

83

(via aid-giving). In Ethiopia, for instance, the World Development Movement (2014, p. 37) notes that: 375,000 hectares of land are being cleared to make way for sugar cane, palm oil, cotton and grain plantations… 260,000 people… are being evicted from their farmland… leaving them little option but to move to designated new villages and work on the plantations for low wages. Those people that have resisted have faced beatings, rape… intimidation, arrests and imprisonment. In order to force people to move, the military have prevented people from cultivating their land and destroyed crops and grain stores to cause hunger, then lured them to the new settlements with food aid.

Perhaps most perversely, meanwhile, a report by GRAIN et al. (2014) demonstrates that despite a legitimating NAFSN discourse of “food security,” the majority of commodities being produced on land corridors are actually defined as export cash crops. The use of narratives of food security to justify these land-grabbing processes is thus brought into question: it is clear that these firms are not interested in the kind of agriculture that will bring us food sovereignty… One farmers’ leader from Synérgie Paysanne in Benin sees these land grabs as fundamentally “exporting food insecurity” because they are about producing food for export markets, creating food insecurity for the producers. They are about answering some people's needs – for maize or money – by taking food production resources away from others (GRAIN et al., 2014, p. 16).

In addition to these concerns, Oxfam (2013) also makes clear that NAFSN donors are largely engaging African governments with dubious track records on corruption. There is the distinct implication therefore that land-grabbing has been facilitated by neocolonial patterns of political graft. Oxfam (2013, p. 3) explicitly states that: Oxfam believes that investors actively target countries with weak governance in order to maximise profits and minimise red tape. Weak governance might enable this because it helps investors to sidestep costly and time-consuming rules and regulations, which, for example, might require them to consult with affected communities. Furthermore in countries where people are denied a voice, where business regulations are weak or non-existent, or where corruption is out of control it might be easier for investors to design the rules of the game to suit themselves.

This picture is largely corroborated by Owen, Vanmulken, and Duale (2015, p. 3) in a report for the London School of Economics. They state that corporate investors in Zambia have bypassed local, traditional authorities by appealing to national government to gain land tract releases. Conversely, in Ghana, corporations have bypassed national government and have sought to negotiate land releases with chiefs directly (Owen et al., 2015). Again, questions of undue influence rise to the surface, cementing a broader view of NAFSN elites perpetrating neocolonial relations on African citizenries in pursuit of commercial gain and land.

84

Rethinking Food and Agriculture

Moreover, and with particular resonance for concerns regarding G8 nations working alongside corporations to perpetrate a form of neocolonialism with regards to land-grabbing, commentators have noted the close affiliations between personnel within official donor and corporate institutions. Kiwanga (2014), for instance, states that Unilever’s external affairs director “was previously at DFID and DFID’s director of policy used to work for Unilever.” The implication therefore is that government and corporate partners are apparently colluding in the pursuit of commercially lucrative land deals in sub-Saharan Africa, including situations in which aid is used as a lubricant for African governments’ acquiescence to such schemes. This viewpoint is shared by the World Development Movement (2014, p. 30), which underscores the close connections between corporations and the UK government in stark terms: Unilever board member Paul Walsh (chief executive of Diageo) is an advisor to the Department of Energy and Climate Change and a member of David Cameron’s Business Advisory Group. Conservative MP Malcolm Rifkind is also a current board member and former overseas development minister and now Conservative life peer… Former home secretary and trade commissioner Leon Brittan was a board member between 2000 and 2010. Former minister for trade and competitiveness David Simon, now a Labour peer, was an adviser to Unilever and was vice chairman and senior independent director between 2006 and 2009. In addition, staff have moved between the company and government.

The manner in which the concept of agricultural corridors diffused from Yara to donor policy reports in conjunction with the NAFSN further underscores the affiliations—and alliances—between major corporations and their (largely Western) government sponsors. And indeed, these affiliations underscore the concerns laid out by Nkrumah as early as the 1960s, namely that government–corporation partnerships would perpetrate forms of neocolonial trade and investment onto nominally independent African countries. The stringent conditionalities associated with the NAFSN and its marrying of aid to land releases cements Nkrumah’s perspective regarding the unequal, and unjust, facilitation of neocolonial forms of North-South relations in sub-Saharan Africa. It is important to recognize, however, that the NAFSN is not a unique or isolated venture. There are parallel concerns about land-grabbing in Africa in relation to Middle Eastern countries and their corporation entities, as well as Indian and Chinese interventions in the continent’s agrisystems. In many cases, such land-grabs are equally, if not more so, motivated by a desire to secure underground water supplies in the context of climate change (GRAIN, 2012; Robertson & Pinstrup-Andersen, 2010, p. 273). This again raises questions about the perpetration of neocolonial forms of relationships on African countries, as predicted by Nkrumah (1965). However, one dimension that Nkrumah himself largely omitted within his writings was attention to the particular gendered aspects of neocolonial imposition. Accordingly, the chapter now considers the NAFSN and wider issues of foreign intervention in African agrisystems in terms of the disproportionate impact felt for African women and their family networks.

Neocolonialism and new alliance for food security and nutrition

85

4.4 A gendered lens on “development”: The impact of the neocolonial NAFSN for women in Africa Agriculture is a deeply gendered sector. In the Global South this can take a variety of forms. In general terms, high proportions of the female population are employed in agricultural production. This reflects an ongoing feminization of agriculture, as men tend to be first to exit from the sector, thereby increasing relative levels of female participation in production. Rural women in Africa (as elsewhere) experience discrimination in accessing a range of productive resources, including land, credit, inputs, and extension services, with men having much higher access to inputs than women. De Schutter (2015, p. 30) argues that this can explain the difference in yields between male and female smallholders, although evidence also points to men being more able to command (often unremunerated) labor from family and community members. Moreover, it is important to note that the marketization of agricultural production deepens and entrenches gendered inequalities. Contracts for farming tend to be in the name of the male head of household or the landowner. When crops are produced for cash rather than for household consumption, women tend to lose decision-making control, as typically women have more input into the decisions related to the production of food for the household, but less in relation to the spending decisions around the allocation of household incomes. The marketization of agricultural production through increases in contract farming therefore can threaten to weaken rather than strengthen women’s positions in rural economies. While there has been recognition of this inequality by policy-makers at a variety of levels, support for agricultural producers has tended to be gender blind and insufficiently oriented to addressing the gendered inequalities within agricultural production and change. The NAFSN has, however, sought to address this through a specific focus on gender and particularly by removing the obstacles women face as agricultural producers. Its proponents acknowledge that contract farming and initiatives to develop the marketization of agricultural production need to incorporate a gender sensitivity. They therefore promote policies to support female smallholder farmers in terms of addressing the limitations they face in increasing their productivity and accessing markets, and in terms of expanding employment opportunities for women. The NAFSN states that it has been successful in this endeavor. It has noted that since 2012, “private investments have reached 8.2 million smallholders and created more than 21,000 jobs in 2014, over half of which were women” (NAFSN, 2019). The NAFSN’s approach thus prioritizes proactive work in relation to women’s rights, particularly in relation to property and inheritance rights, as well as power relations within households. De Schutter (2015), however, argues that this commitment to women is at a high level of generality and mainly rhetorical. For example, NAFSN country frameworks and national action plans tend not to detail tangible reforms on issues such as women’s inheritance rights, gender budgeting, the use of gender indicators in monitoring procedures, nor the inclusion of women in the design and implementation of agricultural research and development policies. In a similar vein, the Cadre de Concertation et de Cooperation des Ruraux et al. (2018) have argued that

86

Rethinking Food and Agriculture

while women are formally a focus of NAFSN policy, they have largely been ignored. The NAFSN has very few projects developed by private investors that are explicitly targeted at women producers: “no women’s organisation or cooperative operated and led by women was recruited as business partners of NAFSN, while the majority of jobs created under NAFSN went to men” (Cadre de Concertation et de Cooperation des Ruraux et al., 2018). More fundamentally, the NAFSN’s adoption of the rhetoric of gender empowerment both relies on and reinforces a particular set of neoliberal norms. This reflects dominant gender empowerment and equality frameworks that have proliferated through development policy frameworks at a variety of levels, from the global through to the local. These incorporate legal and political dimensions, with economic reform and the expansion of particular markets to women in the Global South, particularly via access to credit and associated microfinance schemes. NAFSN discourse thus conjures a vision of women as rational economic actors, whose potential is unfulfilled but, with appropriate support, can improve their entrepreneurialism and productivity though access to resources (particularly inputs for farming and access to global markets). Rural women in the Global South are identified as key agents in delivering policy outcomes through their capacity to develop productive agricultural capacity while also offering themselves, their households, and communities as new sites for the expansion of global markets and enterprise. Within this, women smallholders are framed as beneficiaries of programs, to be helped through agricultural reform and policy change, and empowered through a rights discourse. Such discourse and donor thinking implies that by correcting the decisions women make about their agricultural production and lifestyles they can be easily transformed into successful agricultural businesswomen. This relies on shifting production from household needs to cash crops for income generation, with the uncertainties for food securities that this brings. Gender initiatives in this form embody the disciplinary neoliberal feminist “Smart Economics of Empowerment,” much vaunted by key global development institutions such as the World Bank (Price, 2018). This reflects the assimilation and co-optation of feminist activists, demands, and ideas into frameworks that proliferate and embed a particular form of neoliberal reforms central to the expansion of capital accumulation. It is most clearly evident in the use of concepts of empowerment and entrepreneurialism that discursively veil the underlying political, social, and economic objectives of development and poverty alleviation strategies. The “smart economics” of neoliberal feminism, however, is more than a discursive turn, as it has deeply set material foundations and disciplinary effects (Price, 2018, p. 2). The interplay of class and gender is central to both the growth of markets and in securing the production of raw resources, while participation is disciplined through the twin forces of risk and incentive. Smallscale subsidence farmers are incentivized to increase production and engage in cash cropping and contract farming to augment household incomes, while risking the ability of families and communities to secure their ability to provide for their own needs and wellbeing. It ignores the structural limits facing women and their households in terms of embedded gender and resource inequalities, while masking the deep-rooted economic rationale for the gender sensitivity agenda, which is designed to deliver productivity improvements and incentivize increased production. While this neoliberal

Neocolonialism and new alliance for food security and nutrition

87

rhetoric relies on the delineation of women’s rights in relation to men, it provides a thin veil for the exploitative relation of private enterprise in relation to women. Explicit evidence of the limitations of the NAFSN gender policy agenda is revealed in its initiative on nutrition, which includes a specific objective on the empowerment of women. The Scaling Up Nutrition (SUN) initiative emphasizes the 1000-day window between conception and a child’s second birthday, during which adequate nutrition is vital. The NAFSN has promoted an alliance of private sector organizations, states, and civil society organizations, which they identify as the best actors to deliver such initiatives. While “exclusive breast feeding” is highlighted as good nutritional practice, the codes of practices in relation to SUN are weak, with the majority not committing governments to the full implementation of the 1981 International Code of Marketing of Breast-Milk Substitutes and the resolutions adopted by the World Health Organization Assembly on Infant and Young Children’s Nutrition. Civil society actors have voiced concerns about the role of the private sector in such alliances and initiatives, particularly given the well-publicized behavior of transnational corporations such as Nestlé in promoting breast milk substitutes. The SUN initiative therefore demonstrates the limited approach taken by the NAFSN to the impact of transnational corporations on women’s and children’s health. The core issue of land-grabbing in the Global South under NAFSN auspices further evidences the ways in which the interests of private sector actors and global finance have been prioritized over those of women, families, and households in initiatives that promote agricultural change and reform. Land-grabbing by global corporations has a particular gendered dimension and impact on women. Large-scale land acquisitions in the Global South have intensified over the last 10 years. These have involved forced evictions, human rights violations, environmental food insecurity, and the destruction of livelihoods (The Oakland Institute, 2018). In this land rush, communities have been dispossessed, families disconnected, and local farming systems destroyed as government and investors prioritize profits over people, in what Obbo (2018) describes as a new form of slavery. A prominent example is the rapid expansion of plantations, particularly for oil palm, which has destroyed the diversity of foods, resources, and medicine provided by the land. These form “important parts of the economic and cultural values that women depend on and that characterize their traditional land use” (Farmlandgrab.org, 2018). Ngobo (2018), meanwhile, further emphasizes the regressive impact of land-grabs on women with a particular emphasis on family food security and malnutrition: Lands where women have always produced food for their families are taken away from them. They are often displaced without any reasonable or lasting compensation. They are forced to travel long distances, up to tens of kilometers, to find arable lands. They often rent these lands each season to plant [sic] their crops until the soil is exhausted. Consequently, either their families do not have enough food, or they are forced to eat poor quality food.

Women additionally are increasingly forced to work as laborers in plantations as part of the economic and social changes augured by land-grabs and corporate

88

Rethinking Food and Agriculture

o­ wnership of productive agricultural tracts, largely for low wages, which limit the ability of families to secure their own social reproduction. Land-grabbing by private investors under NAFSN auspices therefore is a direct challenge and in contradiction to the rhetorical commitment to women’s empowerment and equality promoted at the institutional and policy level. Civil society initiatives such as the Rural Women’s Movement in Uganda have highlighted the impact on women who have been dispossessed and thrown into precarity. They have also demonstrated the potential for solidarity as women become collectivized in opposition and call for justice. In particular, this Ugandan movement has highlighted the impact on the climate, particularly from extractive agribusiness industries, which place a disproportionate burden on rural women. Obbo (2018) pertinently remarks in this context that “rural women are waging a climate justice fight in all dimension of their lives—on food, on energy, on health and livelihoods. The women are defending their rights, their communities and their natural resources.” Building on the success of local initiatives such as the Rural Women’s Movement in Uganda, a collective of women’s associations organized a petition and demanded an end to violence against women from the expansion of large-scale oil palm plantations. This deliberately coincided with International Women’s Day in 2018, thus raising public awareness of the plight of women in the Global South in relation to land-grabs and exploitative plantation practices. This expansion of palm oil production (facilitated by NAFSN land-grabs) has been particularly marked and rapid in Central and West Africa. It is estimated that governments have handed over more that 4 million hectares of land in oil palm concessions, which has had specific and differentiated impacts on women, including sexual violence and abuse (Farmlandgrab.org, 2018). As Ngobo (2018) makes clear, women are at increased risk of rape near plantations, they are often “searched and have their privacy violated,” and they can also be forbidden from consuming oil palm products, despite the fact that these might improve their families’ diet. Such measures are largely due to the plantation owners’ fear of thefts by local populations in relation to their crops (Ngobo, 2018). The International Women’s Day petition both highlighted the gendered dimension of land-grabbing and called for solidarity to counter the power of multinational corporations and global finance. Interestingly, meanwhile, in terms of possible alternatives to such regressive outcomes, the Cadre de Concertation et de Cooperation des Ruraux et al. (2018) posited an alternative framing of smallholders, which highlights their role as key investors in agriculture, providers of employment, and guardians of land. Furthermore, they have argued that the needs and demands of smallholder farmers should be put at the center of any new initiatives to improve food security, and within this women should have recognition as being the cornerstone of sustainable agriculture and pioneers of strategies to ensure food security. For example, they requested that the Government of Senegal strengthen rural women’s organizations and networks to provide women with more influence in food and agricultural policy making, and are supported in their efforts to process, conserve, and sell their products to local markets (Cadre de Concertation et de Cooperation des Ruraux et al., 2018). Furthermore, there should be a strengthening of the commitments to political and legal equality. Rather than a paternalistic approach in which women are regarded as beneficiaries of programs designed

Neocolonialism and new alliance for food security and nutrition

89

for them, decision making should be reliant on the incorporation and centralization of women within those processes. Women from the Global South should lead and design decision and policy making, and be responsible for the design and implementation of programs aimed at rural and agricultural development, with the ability to make choices in relation to the farming they wish to practice and whether production should be market oriented or focused on the needs of their households. In terms of the broader problem of neocolonial relations and their impact on food systems in relation to agricultural production and trade, solutions lie with pan-African endeavors—as articulated by Nkrumah and his vision of a Union of African States (Langan, 2017). It is only pan-African cooperation—now embodied in the African Union (AU)—that holds the potential to challenge donor–corporate alliances as exemplified by the NAFSN and its land-grabbing. For instance, the AU could work more effectively within the UN to challenge the efficacy of initiatives such as the NAFSN since they are discursively tied to the SDG framework, with highly dubious consequences. While the AU remains institutionally weak, however, critical actors contesting neocolonial policies can currently look to the creation of the African Continental Free Trade Area (AfCTA) with some level of potential optimism. The initiative to construct a genuinely pan-African economic bloc, bringing together Regional Economic Communities, does hold some potential for the reevaluation and renegotiation of trade arrangements, aid relationships, and corporate investments that currently work to the disadvantage of African citizenries, particularly in terms of food security and cash crop land corridors. Nevertheless, it remains to be seen whether the AfCTA will repeat the same mistakes of the AU’s New Partnership for African Development by adhering too rigorously to free market norms (espoused by the donor community) rather than open up space for developmental state policies more conducive to responsible food systems—as well as to wider economic diversification.

4.5 Conclusion It is clear from the foregoing discussion that the NAFSN discourse surrounding sustainable development, modernization, and food security should be brought into critical light and indeed critique. The NAFSN was explicitly founded in the aftermath of the World Food Crisis of 2007–08 as an ostensible contribution to resolving the problem of food insecurity in sub-Saharan Africa. Promoted by corporate founders, the Alliance promised to offer the know-how and technological resources necessary for the revitalization of African agriculture, with an eye to ensuring the proper nutrition of vulnerable African citizenries. In the context of the post-2015 consensus and the UN SDGs, this language now squarely incorporates a focus on sustainable development, with the understanding that NAFSN interventions will respect a triple bottom line in terms of promoting social wellbeing, ecological integrity, and economic prosperity. Moreover, through the involvement of G8 nation governments and their partnership with leading corporations, the NAFSN apparently demonstrates the benefits of private sector engagement and PPS for donor countries’ contributions to “development” in the international realm.

90

Rethinking Food and Agriculture

As the foregoing discussion has highlighted, however, a variety of critical coverage of NAFSN activities corroborate concerns about corporate land-grabbing. Namely, that poorer subsistence farmers are being deprived of land to make way for corporate agribusiness within so-called “land corridors” established under NAFSN auspices. Perhaps most worryingly here, the plantations and agribusiness established do not concentrate on food but rather on cash crops (such as palm oil) for export and integration into industrial supply chains. Furthermore, via an engagement with Nkrumah (1965) and the concept of neocolonialism, the functioning of the NAFSN becomes wholly problematic in terms of its marriage of aid monies to disadvantageous foreign investment, as well as the conditionalities imposed within agreements signed with African governments (whose elites often depend on foreign aid for their political survival). Finally, it is also clear that there is a disproportionate impact on women in rural communities in sub-Saharan Africa in terms of the marketization of agrisystems and land-grabbing episodes. Women are hailed by NAFSN partners as key policy beneficiaries as part of a legitimizing discourse of sustainable development conscious of gender justice. However, in practice the NAFSN fails to adequately safeguard the rights of women and in fact exposes them to greater uncertainties and jeopardies, not least in terms of the creation of plantation economies within the so-called land corridors hailed by this corporate-led “development” scheme. It is clear therefore that critical actors—not least the women’s movements that have already successfully pinpointed many of the problems associated with the NAFSN—should continue to challenge the “development” language of leading agribusiness corporations in terms of the actual impact of their activities for poorer communities in Africa. The NAFSN should not be seen as a boon to African development in light of the World Food Crisis. Instead it should be seen as a bonanza for agribusiness interests, which oftentimes ignore the social and ecological criteria to which they have ostensibly committed.

References ActionAid. (2015). New alliance, new risk of land grabs: Evidence from Malawi, Nigeria, Senegal and Tanzania, May 2015. London: ActionAid. Cadre de Concertation et de Cooperation des Ruraux, et  al. (2018). Letter addressed to the Government of Senegal, G7 member states and the African Union. Available at: https:// www.farmlandgrab.org/post/view/28219-letter-addressed-to-the-government-of-senegalg7-members-state-and-the-african-union (Accessed 7 January 2019). De Schutter, O. (2015). The new alliance for food security and nutrition in Africa. Brussels: European Parliament. Available online at: http://www.europarl.europa.eu/RegData/etudes/ STUD/2015/535010/EXPO_STU(2015)535010_EN.pdf (Accessed 18 April 2019). Farmlandgrab.org. (2018). Women demand that oil palm companies stop violence and given back community land. Available at: https://www.farmlandgrab.org/post/view/27942women-demand-that-oil-palm-companies-stop-violence-and-give-back-community-land (Accessed 20 January 2018). Frynas, J. G., & Paulo, M. (2007). A new scramble for African oil? Historical, political and business perspectives. African Affairs, 106(423), 229–251.

Neocolonialism and new alliance for food security and nutrition

91

GRAIN. (2012). Squeezing Africa dry: Behind every land grab is a water grab. Barcelona: GRAIN. GRAIN, Martinez-Alier, J., Temper, L., Munguti, S., Matiku, P., Ferreira, H., et al. (2014). The many faces of land grabbing. Cases from Africa and Latin America, 1–93. EJOLT Report No. 10. Kiwanga, G. (2014). Western corporations carve up Africa. In This is Africa. 1st April 2014. Langan, M. (2017). Neo-colonialism and the poverty of ‘development’ in Africa. New York: Palgrave. McKeon. (2014). The new alliance for food security and nutrition: A coup for corporate capital?. Amsterdam: Transnational Institute. McMichael, P. (2015). The land question in the food sovereignty project. Globalizations, 12(4), 434–451. New Alliance for Food Security and Nutrition. (2014). About. Available online at: https://www. new-alliance.org/about (Accessed 18 April 2019). New Alliance for Food Security and Nutrition. (2019). Progress. Available at: http://new-alliance.org/progress (Accessed 7 January 2019). Ngobo, M. C. (2018). Cameroon: Urban and rural activists against industrial abuses towards women. Available at: https://www.farmlandgrab.org/post/view/27943-cameroon-urban-and-rural-activists-against-industrial-plantations-abuses-towards-women (Accessed 20 January 2019). Nkrumah, K. (1965). Neo-colonialism: The last stage of imperialism. Sixth printing New York International Publishers. 1976. Oakland Institute. (2016). The unholy alliance: Five western donors shape a pro-corporate agenda for African agriculture. Oakland, CA: Oakland Institute. Obbo, B. (2018). The rural women’s movement held a feminist school, mobilizes collective power to demand climate justice. Available at: https://www.farmlandgrab.org/post/ view/28298-the-rural-womens-movement-held-a-feminist-school-mobilizes-collectivepower-to-demand-climate-justice (Accessed 20 January 2019). Owen, T., Vanmulken, M., & Duale, G. (2015). Land and political corruption in sub-saharan Africa. London: London School of Economics. Oxfam. (2013). Poor governance, good business, oxfam media briefing. Ref: 03/2013, 7th February 2013 London: Oxfam. Pan Africanist Briefs. (2014). Unilever, Monsanto take over African land and agriculture. In Pan Africanist briefs. 4th April 2014. Available at: https://www.newsghana.com.gh/unilevermonsanto-take-african-land-agriculture/ (Accessed 9 February 2017). Price, S. (2018). The risks and incentives of disciplinary Neo-liberal feminism: The case of microfinance. International Feminist Journal of Politics, 21(1), 67–88. Robertson, B., & Pinstrup-Andersen, P. (2010). Global land acquisition: Neo-colonialism or development opportunity? Food Security, 2(3), 271–283. The Oakland Institute. (2018). The highest bidder takes it all: The world bank’s scheme to privatize the commons. Available at: https://www.oaklandinstitute.org/sites/oaklandinstitute. org/files/highest-bidder-eng-high-res.pdf (Accessed 20 January 2018). World Development Movement. (2014). Carving up a continent: How the UK government is facilitating the corporate takeover of African food systems. London: World Development Movement.

This page intentionally left blank

The myth of a food crisis Jonathan Latham The Bioscience Resource Project, Ithaca, NY, United States

5

5.1 Introduction World Agriculture Towards 2030/2050 is a major report predicting global agricultural trends (Alexandratos & Bruinsma, 2012). It was produced by the economics division of the UN Food and Agriculture Organization (FAO). In its abstract the FAO authors make a prominent disclaimer. Its projections, they stress (both on p. i and p. 7), are not to be used for normative purposes; that is, their report is not a prescription of how the global food system should develop. It is merely an exploratory model; their most reliable projection of business as usual (Alexandratos & Bruinsma, 2012). In all probability this disclaimer resulted from the intense global attention that its predecessor (FAO, 2006) received. This “interim report” was cited across the globe as claiming that the world must produce 70% more food by the year 2050. This 70% number (sometimes even adjusted to a “doubling”) was almost invariably recruited to bolster a number of technological modernizing agendas for agriculture, for instance, in the promotion of genetically modified crops. Thus the UK’s chief scientist in 2009 predicted an imminent “perfect storm” of climate change and food shortages (Beddington 2009). Similar analyses were repeated in scientific articles, in generalist publications such as the Economist magazine, and by agribusiness (Peekhaus, 2010; Tomlinson, 2011; Stone & Glover, 2011). Thus FAO’s number was repeatedly taken out of context and presented as a grand challenge requiring special efforts or difficult compromises. That is, it was used normatively. Those citing FAO may not have said that: The battle to feed all of humanity is over. In the 1970s the world will undergo famines—hundreds of millions of people are going to starve to death in spite of any crash programs embarked upon now. (The Population Bomb, Paul Ehrlich, 1968).

However, the implications were comparable. But this is not merely a simple story of statistics being taken out of context. In 2016, the same FAO department described in more detail their modeling system (in 2012 it was renamed the Global Agriculture Perspectives System, or GAPS) used to derive the original prediction (Kavalleri et al., 2016). Drawing attention, in a clearly normative fashion, to its newest quantitative prediction the authors wrote: “A key finding… is that world food production should increase by some 60% from 2005/07 to 2050” (Kavalleri et al., 2016, p. 1, emphasis added). By contradicting their colleagues’ previous disclaimer, Kavalleri et al. raised the issue, since FAO is a stakeholder in the food crisis narrative, of whether the original disclaimer was sincere and whether more could not have been done to avoid the normative usages of FAO numbers. Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00005-0 © 2021 Elsevier Inc. All rights reserved.

94

Rethinking Food and Agriculture

The answer given in this chapter supports the analysis of Tomlinson that a “slide” is operating in many of the texts written by FAO, and that this slide is particularly problematic in texts written by FAO leadership (cited in Tomlinson, 2011). Given that FAO oscillates between normative and nonnormative statements, this slide might be better called a “shuffle,” but it embodies perfectly the central paradox of all quantitative models of the global food system, whether produced by FAO or by others. This paradox is that, though FAO modeling supposedly exists to “identify challenges in world food and agricultural sectors and to offer strategic policy perspectives” in an unbiased fashion (Kavalleri et al., 2016, p. 1), what GAPS does in practice is quantify food. This frames agriculture as primarily a question of production. The focus on production, no matter any disclaimer, is normative, because it marginalizes issues of poverty and access to food, ecological costs, and social costs. These are either unexamined or subsidiary. So while the titles of FAOs reports and models are broad, e.g., World Agriculture: Towards 2015/30, the focus is narrowly on the quantification of production, even though, according to the International Assessment of Agricultural Knowledge, Science and Technology for Development (IAASTD), the truly “key” questions swirling around agriculture are not about productivity (IAASTD, 2009). Productivity, concluded the IAASTD report, is a distraction. As Robert Watson, chair of IAASTD, told the press at its launch, in agriculture, “Business as usual is not an option.” The real question of agriculture is: Can we feed the people without also feeding social and ecological disasters? But farmer suicides and insect declines, salinization, dead zones, and the pollution of water bodies and other consequences of dysfunctional agriculture are absent from World Agriculture Towards 2030/2050. Moreover, World Agriculture Towards 2030/2050 even fails, ultimately, to show that productivity per se merits specific modeling attention. First, because the model it describes predicts that any necessary production increases will be solved by business as usual. Second, it also predicts “Modest reductions in the numbers undernourished” by 2050 but this reduction is dependent on continued economic growth (i.e., increases in wealth); that is, not on agricultural production (Alexandratos & Bruinsma, 2012). FAO’s previous iteration, World Agriculture Towards 2015/2030, had reached a virtually identical conclusion. It projected that out of a then total of 850 million hungry people, just 120 million would be lifted out of hunger if food production reached its target increase of 70% by 2030 (Bruinsma, 2003). So, even according to FAO’s own models, increasing production does not solve hunger. This result mirrors the conclusion of Amartya Sen in his celebrated history Poverty and Famines. When hunger and famine strike, he found, production shortfalls have virtually never been the cause (Sen, 1981). To many food system commentators this is settled beyond question (e.g., Lappé & Collins, 2015). But it is a finding that has nevertheless been disregarded by many, including FAO leaders (Tomlinson, 2011), who have instead commonly cited FAO in support of a scarcity narrative with its consequent need for a productivity focus (e.g., Conway, 2012). Thus the scarcity view, whose credibility rests almost purely on the findings of models like GAPS, finds, at best, only equivocal support there.

The myth of a food crisis95

5.1.1 Global food models and projections The purpose of this chapter, however, is to provide a quantitative critique of models like GAPS at the level of their underlying assumptions. Unless otherwise noted the focus will be on FAO’s GAPS. This focus is specifically not intended to validate the quantification of food; indeed, quantification of food as calories and weight is detrimental to a full understanding of food systems. It is rather an acknowledgment that FAO’s work is the most prominently cited and that, encouraged by FAO’s shuffle, the world has overwhelmingly interpreted this 70% number as normative. In 2012 FAO's prediction was updated to 60%, mainly to reflect a shifting baseline: we are now much closer to 2050 than we were in 2003 (Alexandratos & Bruinsma, 2012). So, at the risk of appearing to validate the general approach, it is on a purely quantitative level that these models are most transparently open to challenge. Malthus, 1798 is considered to have made the first mathematical model of a food system. His simple projection concluded that exponential population growth would eventually outstrip linear supply growth. The basic form of his model, followed ever since, was to separate food supply from food demand (McCalla & Revoredo, 2001). Besides FAO, institutions such as the International Food Policy Research Institute (IFPRI) have developed their own models (Robinson et al., 2015). In addition, special projects such as the Millennium Ecosystem Assessment, the Comprehensive Assessment of Water Management, and Agrimonde (2009) (a joint project of the French Institut Nationale de la Recherche Agronomique and the Centre de Cooperation Internationale en Recherche Agronomique) have extended the general method but with an emphasis on investigating specific questions, such as water constraints, climate impacts, and the effects of specific policy decisions (de Fraiture et al., 2007; Fischer et al., 1988; Rosegrant et al., 1995, 2001; Parry et al., 2004; Chaumet et al., 2009). All are intended to inform decision making. However, those more suited to exploring diverse potential outcomes are often called scenarios. These distinctions, along with some of the strengths and weaknesses of the models, have previously been reviewed by Reilly and Willenbockel (2010) and by Wise (2013). What these reviewers note, above all, is that there is overall a strong degree of consistency among models and scenarios that there is no need for extraordinary measures to enhance production. To quote FAO: “from the standpoint of global production potential there should be no insurmountable constraints” (Alexandratos & Bruinsma, 2012). None foresees a classic Ehrlich-style crisis, unless they expressly incorporate in their scenarios some form of mismanagement. For instance, the Millennium Assessment has as one of its four scenarios “Order through Strength” (OS) that envisages low cooperation and high trade barriers. Under OS conditions there is no overall global food shortage but there is increased malnutrition and even civil war in parts of Africa. This is a broadly reassuring conclusion, but it should nevertheless be qualified by the looming shadow of climate change (Battisti and Naylor, 2009; Nelson et al., 2010). Agreement that food production is unlikely to develop into a crisis situation (climate excepted) has not banished the alarmist narrative in the media, however (see, e.g., Hincks, 2018).

96

Rethinking Food and Agriculture

Yet, there are grounds to suppose that even this convergence, which does predict a need for increased production, is excessively pessimistic. In 2011 researchers from the World Bank Institute proposed that the world already produced enough food for 14 billion people (Herren et al., 2011). This number is well above UN population predictions, which are expected to reach 10–11 billion in 2050 and perhaps even decline thereafter (UN, 2017). Moreover, models also contradict global food price trends. Before the 2007/2008 price spike caused by changes in US and EU biofuel policies (de Gorter et al., 2015), food prices had been declining at approximately 4% per year. Since that spike, prices appear to have returned approximately to that track. This long-term decline, across every sector of agriculture, suggests strongly that food supply significantly exceeds current food demand and that the gap is if anything widening. The exact extent of this excess is not clear but the 2017 FAO estimate for global cereal stocks is 762 million tons. This amount represents approximately one-third of annual global production (FAOSTAT). Thus, independent of any modeling, there are strong grounds for supposing that even the most optimistic models are still pessimistic. They are overestimating demand or underestimating supply, or both. The overarching questions are: How does one reconcile low (and declining) food prices and persistent global commodity gluts with the projections, claimed by GAPS and other models, of the need to produce more food? Are the models flawed? If so, what are those flaws?

5.1.2 How flawed are food system models? The use of highly complex models always raises many questions of how well they represent reality (Scrieciu, 2007). But food system models are especially complex, seeking as they do to integrate biophysical, social, economic, and institutional components. Thus a criticism sometimes made of such models is their use of calories as the measure of nutrition (e.g., Herforth, 2015). Both nutritionists and those seeking a more expansive definition of food security have pointed out that calories fall well short of the definition of food security adopted at the 1996 World Food Summit: “Food security exists when all people, at all times, have physical and economic access to sufficient, safe and nutritious food to meet their dietary needs and food preferences” (e.g., Burchi et al., 2011). Thus the achievement of caloric sufficiency may ultimately be irrelevant. To frame this diversity of critiques the primary issues with quantitative models are sometimes divided into technical uncertainties, methodological uncertainties, and epistemological uncertainties (Funtowicz & Ravetz, 1990). Narrower technical concerns include the “knowledge gaps and priorities” raised by Reilly and Willenbockel (2010) and also by Wise (2013). On this level these authors agree there exist very significant problems with data quality. In many countries that extends to quantifying even the most basic inputs of the models: poverty, GDP, water availability, even simple population. The problem of questionable data is highlighted by the case of Ghana. Its national statistical agency announced in 2010 that it was revising all future GDP estimates upward by over 60%. This made Ghana into a lower-middle-income country literally overnight (Jerven, 2012). Such difficulties are

The myth of a food crisis97

acknowledged (though ultimately dismissed) in FAO’s Food balance sheets: a handbook (FAO, 2001). Problems of comparable magnitude also apply to modeling the relationships between data points. According to Reilly and Willenbockel: “more work on the validation of model components used in integrated assessment studies is required.” But these authors equally caution that validation is a two-edged sword; calibrating models to past experiences, especially in the presence of climate change and other potential abrupt changes, introduces the problem commonly known as “overfitting.” Some of these “technical”-level difficulties are acknowledged by the modelers themselves (e.g., Bruinsma, 2003). A yet further significant problem is that, because they come from many different nations, datasets in separate parts of the models are based on different scales, time periods, and conceptual schemes (FAO, 2001). Methods to reconcile such disparities are not available, however. Unsurprisingly perhaps, having discussed these limitations, Reilly and Willenbockel conclude that “model outputs should not be misinterpreted as forecasts with well-defined confidence intervals. Rather they are meant to provide quantified insights about the complex interactions in a highly interdependent system and the potential general order of size effects” (Reilly & Willenbockel, 2010). This comment raises some important issues. The first is that this limitation seems to have generally eluded those who cite these numbers. Second, without confidence intervals no one—including the modelers themselves—knows what this “general size order of effects” is. We can gain insight into how such uncertainties might affect the quality of predictions by examining an extended critique made by Thomas Hertel in his presidential address to the American Agricultural and Applied Economics Association (Hertel, 2011). A major assumption in FAO and other models, notes Hertel, concerns how they relate prices, demand, and supply. Focusing on FAO’s quantitative model (pre-GAPS) he notes that it assumed that agricultural supply hardly responds to higher prices. This assumption was introduced because measurements of how food supply responds to demand in agriculture have mostly been taken over the short term. The remit of these models is the longer term, however. Measurements taken over the long term suggest that the elasticity picture is very different. Hertel contends that if future growth in demand was sufficiently great for food prices to rise, then this would in turn stimulate supply. Thus higher agricultural prices can be expected to favor high yields, raise land prices (protecting existing land and bringing more into production), stimulate agricultural research, and reduce waste, and this is indeed what the longer-term evidence shows (Hertel, 2011). Even the declining trend in the growth of global crop yields, which according to FAO is a major determinant of future food availability, may be a function of price. To this end, Hertel quotes economist Robert Herdt: “the economics of substantially higher yields is not attractive” (International Rice Research Institute, 1979). In this connection, Hertel also quotes FAO economist Jelle Bruinsma: “given the right incentives, much of the increased demand for cereals and oilseeds in 2050 could be met using existing technology.” Hertel therefore concludes that the frequently noted long-term “slowing of yield growth may simply be due to a slowing of net demand growth.” And he summarizes: “it is not clear that the resulting models are wellsuited for the kind of long run sustainability analysis envisioned here.”

98

Rethinking Food and Agriculture

To summarize Hertel: there is strong evidence that incentives acting on farmers and other decision-makers are key to explaining agricultural productivity, but since commodity prices have been in long-term decline, FAO has been modeling a low-incentive system. With this as background the next section is devoted to analyzing four additional assumptions underlying predictive models and using GAPS as the example.

5.2 The assumptions of GAPS Key assumptions made by GAPS that strongly affect its predictions of future production needs are as follows: Assumption 1:  That biofuels are driven by “demand” Since 2002 the development of the liquid biofuel sector (ethanol and biodiesel) has been very rapid. According to US Department of Agriculture figures, the US biofuel industry consumed 127 million tons of maize in 2011, which was 15.6% of world maize production. Additionally, in Brazil over 50% of sugarcane is used for biofuels and in 2009 the EU consumed as biofuel 9 million tons of vegetable (mostly rapeseed) oil as well as 9 million tons of cereals (Alexandratos & Bruinsma, 2012). Specifically for ethanol, the 2016 figures for global production were, in millions of gallons, United States (15,250), Brazil (7295), European Union (1377), China (845), and Canada (436) (Mohanty & Swain, 2019). The initial consequences of the biofuel boom were price spikes that propagated through the global supply chain to affect most commodities (de Gorter, Drabik & Just, 2015; McMichael, 2009). These price spikes induced riots, hunger, and financial hardships and, in some countries, political upheaval. The diversion of substantial quantities of presumptive agricultural produce into biofuels is a perturbation that has, however, ultimately been absorbed by the physical supply chain. Commodities have now all but returned to their prespike prices (OECD-FAO, 2016). In the models of the FAO, biofuels are deemed a demand. That is, they reduce availability but they don’t contribute to feeding the population (except when certain residues are fed to animals). Looking to the future, FAO further assumes that biofuels will continue to expand in area, plateauing in 2020 (Alexandratos & Bruinsma, 2012, p. 97). The justification for treating biofuels as demand is that biodiesel and bioethanol are environmentally beneficial products of deliberative policy choices. This position is highly questionable. First, biofuels from palm oil or soybeans and ethanol from corn cause immense biodiversity losses. Additionally, their net greenhouse gas emissions provide no benefits or are only positive when measured over hundreds of years (e.g., Germer & Sauerborn, 2008; Danielsen et al., 2009). Second, the necessity justification is dubious since the development of biofuel policies has been driven not by climate concerns but by complex agglomerations of lobbying interests seeking to boost agricultural demand so they can sell more inputs (Baines, 2015; ActionAid, 2013). The significant point about markets driven by lobbying, rather than genuine demand, is that if demand for food were, as expected, to increase (and/or food prices were to rise)

The myth of a food crisis99

the interest in biofuels would vanish because their key political attraction is to prop up demand (Baines, 2015). In other words, land devoted to biofuels is available, if needed, for feeding populations but this opportunity is rendered invisible when models such as GAPS categorize biofuels as a demand. The amounts of food converted into biofuels are very substantial. For instance, ActionAid concluded that the G8 countries (which exclude Brazil) consumed annually enough biofuel to feed 441 million people (ActionAid, 2013). If measured today that figure would undoubtedly be much greater. It is curious that FAO modelers do not address the ethics of biofuel diversion. First, because, in a hungry world, the dilemma raised by diverting biofuels at the expense of food availability is a rather obvious one (Chakrabortty, 2008), and second, because, elsewhere in discussing their models, FAO economists do digress on the environmental and ethical implications of various actions, thus violating their supposed nonnormative stance (e.g., Alexandratos & Bruinsma, 2012, p. 131). Third, given that the premise of models like GAPS is that increasing production is an ethical imperative, why is it not equally ethically problematic to subtract from food availability by burning it for transportation? FAO’s modelers have again been inconsistent in the application of the nonnormative approach. On one level, FAO’s failure to incorporate biofuels as potential food supplies for hundreds of millions of people and instead integrate them into their model as “demand” is an error. But it also represents a further instance of inconsistency: shuffling in and out of normative modes. Is it a coincidence that FAO’s economists only take an explicitly normative stance when it aligns with powerful interest groups? At the same time, the shuffle is very much facilitated by GAPS having as its underlying paradigm the normative expectation that production trumps all. Assumption 2:  That current agricultural production systems are optimized for productivity According to FAO statistics, if we truly wanted to maximize global yield as measured by calories/ha per day we would all eat sweet potatoes (70 × 10(3) kcal/ha/day). Or, if we lived further from the equator, potatoes (54 × 10(3) kcal/ha/day) (FAO, http:// www.fao.org/docrep/t0207e/T0207E04.htm). Of course, there are reasons why we do not eat only these crops. Those reasons range from the agronomic benefits of crop rotations to the physical and cultural needs for a varied diet. However, farmers also grow many crops (such as coffee and grapes for wine) that have very limited ability to feed people. As a consequence, world agriculture has the potential to greatly enhance productivity through crop substitution. This potential exists rather obviously in developed countries such as the United States and the European Union, where subsidies and market monopolies are far more powerful drivers of production than are calories or nutrition. One might suppose, however, that countries like Bangladesh (whose population of 160 million resides in an area the size of New York state) might be different. Bangladesh has one of the highest population densities in the world and one of the highest poverty and food insecurity rates. However, although wheat yields about half that of winter season rice in Bangladeshi conditions, the market price of wheat is higher and the input costs are much lower (J. Duxbury, pers. comm.). Bangladeshi farmers therefore grow wheat on 415,000 ha (FAOSTAT, 2017). Such farmers are chasing markets not nutrition.

100

Rethinking Food and Agriculture

But perhaps the most glaring instance of suboptimal nutritional performance in agriculture is meat consumption. Historically, much meat and dairy production took advantage of marginal land that was less suited or unsuited to other forms of cropping. Increasingly, however, especially in many “developed” countries, prime land is cropped for animal feed. Even the most efficient converters of this feed (fish and chickens) yield a worse caloric return per hectare than the least nutritious vegetable, while the least efficient (beef) yields approximately fourfold less again (Cassidy et al., 2013). Thus it has been estimated that beef, in a feedlot system, has a feed conversion efficiency, measured in calories, of 3% (compared to chicken with 12%) (Cassidy et al., 2013). Clearly, this implies major opportunities for substitution since about 35% of the US corn crop goes to animal feed (Baines, 2015). In spite of these diverse opportunities, substitutions of higher-yielding crops for lower-yielding ones is a possibility neglected by GAPS. In contrast, GAPS does allow substitution in the opposite direction. This is when GAPS allows higher meat consumption following income rises. GAPS is thus again inconsistent. The standard justification for allowing this substitution is “consumer demand.” As populations become wealthier they “naturally” eat more meat, goes the theory. Predominantly vegetarian countries excepted, there is certainly a correlation between wealth and meat consumption. The caveat required is that OECD countries spend $318 billion annually on agricultural subsidies. These overwhelmingly go to supporting either meat or biofuels (OECD, 2002). Virtually none of it goes to subsidizing fruits and vegetables. Consequently, “the power and freedom of choice attributed to consumers are questionable” and so also is any straightforward expectation that other nations will closely follow this path, unless they too decide to subsidize meat production (Rivera-Ferre, 2009). How much extra food could result from crop substitutions overlooked by GAPS? Any calculation is a difficult one since much depends on what crop is substituted and what replaces it. The case studied most intensively is the impact of meat consumption. A recent estimate is that 4 billion additional people could be fed if animals were absent from the global feed chain (Cassidy et al., 2013). This study did not substitute animal rearing with the most calorific crop, however. It is worth noting that this conclusion largely assumed Western-style rearing and consumption patterns. Thus it is not applicable to those parts of the world where nomadic and seminomadic systems predominate. Another type of substitution is the growing use of mixed cropping, agroecological production systems, and conservation agriculture. These can further increase yields beyond the monocultures assumed in GAPS, sometimes dramatically (Sampson, 2018; Kassam et al., 2009). In 2015/16, conservation agriculture occupied approximately 180 million hectares of cropland globally, and since 2008/09 has expanded by over 10 million hectares a year (Kassam et al., 2018). In conclusion, models in general (and not only GAPS) are disregarding clear opportunities for crop substitutions that have the potential to feed many billions of people. Assumption 3:  The existence of maximum “yield potentials” An important assumption found in GAPS and other models is the use of the concept of “yield potential.” Yield potential describes the theory that crops have a ­genetic

The myth of a food crisis101

yield ceiling beyond which cropping systems (usually envisaged as inputs of fertilizer or chemicals) cannot lift them. Because of this ceiling, GAPS presumes that productivity increases can only come in either of two ways. Either slowly through long-term research and breeding efforts that raise the yield potential of each crop, or, alternatively, by persuading those farmers who use “suboptimal” seeds or methods and therefore operate far below this ceiling to approach the current yield potential. Thus yield potentials figure prominently in GAPS. FAO states this assumption in many places, e.g.: “There is a realization that the chances of a new Green Revolution or of one-off quantum jumps in yields, are now rather limited” (Alexandratos & Bruinsma, 2012, p. 125). Yet these yield potentials are theoretical only. They are not proven to exist and may not exist. Perhaps the leading exemplar of this is rice, the world’s most important crop. The yield potential of rice is standardly estimated at 8–10 t per hectare (Peng et al., 1999). Such high yields are assumed to occur only under agronomic conditions of very high fertilizer and chemical inputs and with ideal soil and watering regimes. Yet the world record for rice production is 22.4 t (Diwakar et al., 2012). This record was achieved with few inputs by a farmer using a method called the System of Rice Intensification (SRI). What this record implies, and peer-reviewed SRI research supports this, is that rice is far below its supposed yield potential when grown under standard “optimal” conditions (or that yield potential has no real-world meaning) (Kassam et al., 2011; Taylor & Bhasme, 2019). The implication of the yields achieved by SRI is that sustainable yields and productivity exceed those assumed by all quantitative global models by several multiples. Since theoretical yield potentials are ordinarily rarely met on real farms, 22.4 t represents in practice a potential tripling of yields over standard expectations. SRI methods have also been applied to other crops, again giving significant yield improvements (Abraham et al., 2014). In 2013, SRI was estimated to have 9.5 million practitioners (Uphoff, 2017). By 2019 this number had at least doubled (N. Uphoff, pers. comm.). By accepting the concept of yield potential and ignoring SRI, quantitative models are overlooking one of the most rapidly spreading developments in agriculture (Stoop et al., 2017). Since rice is the staple of half the globe (3.5 billion people) it can be readily appreciated that a tripling of yields, especially since SRI is a more sustainable method, represents the potential to feed perhaps a further 7 billion people (Fageria, 2007). Assumption 4:  That global food production is approximately equal to global food consumption. Unlike in most sectors of the economy, agricultural production can exceed consumption at the global or local scale. If we take cereals (wheat, rice, barley, millet, sorghum, and oats) as an example, excess production occurs even in densely populated countries such as India and China. In 2017, FAO estimated global stores of cereals at 762 million tons; this is out of a total global cereal production (in 2017) of 2595 million tons (FAOSTAT). These stocks represent an insurance against calamity. However, this 762 million tons also represents an excess of supply over global demand. An important property of these stocks is their perishability. Depending on the climate, the quality of storage,

102

Rethinking Food and Agriculture

and the crop species, they may rot or be eaten by rodents or insects; thus even if stocks are not growing, crops may still be entering them at a high rate. The second relevant property of stocks is that, if there are multiple harvests per annum, quantities of lost stocks may represent multiples of the steady-state amount of an annualized store. For example, if 33% of each rice crop is lost in storage and there are three rice storage periods, corresponding to three harvests, then 100% of the annual total stock is, in effect, lost each year. The consequence of these properties is that it is very important to count quantities entering stocks since, if needed, they are available for use. Even if stored well, stocks eventually degrade. In China, wheat stocks are considered by analysts to last maximally 3–4 years and an average of 2 years. For this reason, China, which is one of the biggest stock holders of rice and wheat, began a biofuel policy to consume excess stocks of wheat. This has steadily grown and now generates 845 million gallons of ethanol per year (Mohanty & Swain, 2019). Despite this program, Chinese wheat stocks are still growing. What this ultimately means is that it is important to understand how much is entering stocks each year. Unfortunately, for modeling purposes, FAO assumes that “At the world level production equals consumption” (Alexandratos & Bruinsma, 2012; see also FAO, 2001). This assumption is restated in the 2016 GAPS description: “The model assumes a closed world economy so that at the end of every simulation period global demand equals supply” (Kavalleri et al., 2016). Amounts entering stocks are counted, but only as the difference between opening stocks and closing stocks (FAO, 2001). The net effect is to ignore losses to insects, mold, rodents, and age. The simple way to show how this method ignores lost and degraded stocks is to work through an example. If 762 million tons was the amount in stock on January 1 and all of it was lost to rats and insects in the subsequent year and then the entire amount was replaced in the following year, FAO’s counting method would register no change to stocks and no addition to stocks. That is, it would appear that nothing had entered stocks even though 762 million tons had in reality done so. How much of the global grain supply is lost in storage? Estimating postharvest losses is difficult and seems to be a low priority. Furthermore, many estimates are not necessarily applicable to stocks but rather to postharvest in general. FAO estimates that postharvest losses in low–middle-income countries are approximately 6.4% for cereals. Most cereal and pulse loss estimates are much higher, but also highly variable and they acknowledge much uncertainty (Boxall, 2001; Kumar & Kalita, 2017; Sharon et al., 2014). Estimates include 20%–30% for maize in Africa (Tefera et al., 2011); 12% and 44% for maize in the West Cameroonian highlands during the first 6  months of storage (Tapondjou et  al., 2002); 11%–17% for rice in India, without counting storage (Alavi et al., 2012); and 35% for rice in India (Scrimshaw, 1978). Some reports estimate very high levels, for example, 59% after 90 days in sub-Saharan Africa (Kumar & Kalita, 2017). Thus FAO’s figures for postharvest losses are very much at the low end, especially since storage is often considered the stage of greatest deterioration (Kumar & Kalita, 2017). One could say that GAPS makes two errors that cancel each other out to balance the model. It ignores what enters stocks (unless they change) and it ignores losses within them. This is fine from the point of view of building a closed model but the

The myth of a food crisis103

significance of the errors for the purpose of estimating food availability is potentially very great: there is an uncounted annual excess of production, which is lost in storage, but is available to meet future demand. To state this in Malthusian terms, population is below current production and population (at least demand) can afford to grow before the two become equal. This is not to say that stocks and reserves are undesirable or unnecessary, but it is to say that a very significant unacknowledged gap likely exists. Current estimates of both stocks and stock losses are highly uncertain. However, if we accept its numbers at face value then FAO is undercounting food equivalent to the cereal needs of perhaps 1–2 billion people, even without counting the losses of more perishable (noncereal) crops. In summary, the four assumptions discussed here generate the following estimate of extra food potentially available (over and above the estimates of FAO’s GAPS). Assumption 1: 500 million; Assumption 2: 4 billion; Assumption 3: 7 billion; Assumption 4: 1–2 billion. The sum total (12.5 billion) is certainly a low-end figure because Assumption 1 underreports current biofuels because its data are old; Assumption 2 only includes the substitutions of meat by crops and not higher calorie crops by lower calories ones (because no studies exist); and Assumption 3 only includes rice, whereas SRI suggests other yield ceilings may also be similarly flawed. Nevertheless, 12.5 billion people’s worth of spare food is a very large underestimate.

5.3 What are models for? The foregoing discussion has highlighted that quantitative models such as FAO’s GAPS are highly reliant on questionable assumptions. Summarized briefly, first, GAPS does not adequately take into account the potential to substitute higher-­ yielding food species for lower-yielding ones. Second, GAPS neglects that food can and would be grown instead of biofuels if needed. Third, GAPS contains unreasonably low expectations of achievable yields of existing crops, in particular rice, which is the staple of half the global population. Fourth, GAPS neglects annual surpluses lost in storage. To this critique of GAPS and related food models should be added that of Thomas Hertel, who has asserted that FAO has underestimated the potential for food availability to rise with prices (Hertel, 2011). Thus the discrepancy that prompted this chapter to be written, which was between long-term (decreasing) global prices that demonstrate increasing overproduction and quantitative models that predict looming scarcity, can readily be resolved. The models are flawed. They underestimate actual or potential supply and they exaggerate demand. Quantitative models in science, in technology, and in economics have never been more popular (Porter, 1995). To explain this ascent, two distinct schools of thought have arisen. The standard explanation is that quantitative models are objective approximations to reality that have become more prevalent because, with computers and other developments, they have become progressively more powerful and more useful tools.

104

Rethinking Food and Agriculture

The second, rival, explanation is that the rise of such models is attributable to their usefulness as pretenses to objectivity. As British scientist Lancelot Hogben wrote in 1933, they allow the possibility of “concealing assumptions which have no factual basis behind an impressive façade of flawless algebra.” (Another formulation is attributed to ecological modeler Dick Levins: “All quantitative models are qualitative models in disguise.”) In this way, models can function as smokescreens (disguises) for either conscious or unconscious institutional biases and objectives. In the words of Theodore Porter: “Quantification is a way of making decisions without seeming to decide” (Porter, 1995). The standard explanation of the rise of modeling is surely flawed, at least in part. All modeling is a retreat from reality in which complexity is reduced to uniformity and qualities to quantities. In whittling data and simplifying interactions, choices must be made. Perhaps nowhere in all of quantitative modeling is simplification more problematic than in a representation of the food system whose component parts are biological, climatic, hydrological, economic, and social in character; that is, they are incommensurable. Thus modelers are inevitably faced with difficult choices over which of the myriad potential contributory factors to include and which to exclude. In the same way, they have to devise how to mathematically simplify the likely nature of the relations between them. These choices represent challenges to objectivity. A good example of this objectivity problem as it applies to food models is the time frame of FAO’s annual accounting. Its agricultural year runs from the beginning of January to the end of December (FAO, 2001). It is thus calibrated to the needs of Europe and North America and not those of the Southern Hemisphere or the tropics. This choice exemplifies how the values of the West, but also of international trade, and of finance are subtly present in all such models at the expense of the much less prominent or quantifiable priorities of the poorer countries (predominantly Southern and tropical) whose interests these models claim to serve. The assertion of an objective modeler is therefore untenable. This does not mean that all models are necessarily valueless. But it does suggest we ask a more subtle question about the good faith and disinterestedness of its authors and funders: What steps do the builders of such models take to minimize the potential for unconscious bias? From this perspective it is not reassuring that FAO claims to rely “as much as possible” on in-house experts (Alexandratos & Bruinsma, 2012). Nor that FAO presents its data with an implied high degree of confidence, especially in prominent fora (e.g., Diouf, 2008, cited in Tomlinson, 2011). Meanwhile, the references in FAO documents to assumptions, limitations, and uncertainties are sparse and relegated to middle pages or back pages or annexes. Sometimes these are therein even discounted without explanation (Reilly & Willenbockel, 2010; Wise, 2013). Third, the frequent shifts into the normative language (of “should”) and back again by FAO in its successive reports is again suggestive that institutional preferences exist but are not well managed. Perhaps the most instructive formulation of the disinterestedness question is to ask (if only rhetorically): If its models gave answers that were diametrically inverse to the ones presented (i.e., that food provision was ample), would FAO stand by them?

The myth of a food crisis105

5.3.1 Conflicts of interest amid the food crisis This question leads to the consideration of what is most noticeable about the four questionable assumptions identified in this chapter, that they are not neutral in their effects. Each has the consequence either of exaggerating demand or of underestimating supply, either in the present or in the future. Thus they all serve to exaggerate a Malthusian scenario. The same is true for the assumption noted by Hertel, that FAO fails to appropriately take into account the effects of price on production (Hertel, 2011). It too underestimates future food supply. Together the effect amounts to the food requirements of, minimally, many billions of people. A potential explanation for the Malthusian exaggeration is that FAO, like almost every large institution in the food/academic/philanthropic nexus, has a conflict of interest. If there were no threat of a food crisis of some kind, FAO’s institutional raison d’etre (motto: Fiat Panis) would vanish. Additionally, given these conflicts, some food crisis variants are clearly preferable, from FAO’s perspective, over others. If institutions like FAO, IFPRI, or the World Bank were to frame the problems of agriculture as resulting primarily from maldistribution of land, democratic deficits, poverty, or the excessive power of agribusiness, they would come into conflict with agribusiness, governments, and wealthy landowners, the very people who, directly or indirectly, fund or control these institutions. Far safer, politically, to blame hunger on lack of production; that is, the farmers (Food First, 2016; Sampson, 2018). In this context it is useful to recall the seemingly neglected predictions of FAO’s own models that increasing food production would have little effect on the number of people going hungry (Alexandratos & Bruinsma, 2012; Bruinsma, 2003). This prediction came true. In 2018, even though food production did subsequently increase, and food prices fell, the number of malnourished rose to 821 million. If FAO wants to solve hunger, their own model is telling them to look elsewhere than increasing production. This is a highly damaging finding to models, like GAPS, with a productivist premise. Yet, 17  years on from 2003 FAO still targets productivity, as does almost every major food security institution and philanthropic effort (Food First, 2016).

5.3.2 Final thoughts One interest group in particular benefits from the premising of models such as GAPS on the question of productivity. “Feeding the world” is the principal public relations gambit of international agribusiness. Only agribusiness has the yields to save the poor and starving, is their claim (Peekhaus, 2010; Stone & Glover, 2011). If the scarcity narrative is true, that claim is powerful. It transforms agriculture into a moral issue (Dibden et  al., 2013; Latham, 2015). Pesticides, genetically modified organisms (GMOs), and monocultures may have negative consequences, goes the narrative, but they are the necessary alternative to starvation. The sole alternative, accordingly, is merely a luxury for the privileged and of no interest to policymakers. Alternatively, if scarcity is a myth, then all pesticides are sprayed, and all GMOs exist, exclusively for profit. The destruction of the ecosphere, which is largely for the sake of agriculture, is effectively a waste (IAASTD, 2009; IPBES, 2019). The stakes are high.

106

Rethinking Food and Agriculture

Agribusiness stands or falls on this point. The reason, as George Lakoff has argued, is that humans think and act in moral terms. They wish to think of themselves as “good” (Lakoff, 2004). Thus the focus of quantitative models on productivity, which asserts a moral essence, is a gift to agribusiness since, to be effective, a credible third party has to legitimize the scarcity narrative. So, while very significant effort and attention have been directed toward creating GAPS and other models devoted to estimating and predicting global agricultural demand and production, there remains effectively no clear evidence that lack of productivity plays a pivotal role in the hunger epidemic that supposedly stimulates them (Lappé & Collins, 2015; Sen, 1981). In contrast, there is abundant, even overwhelming, evidence that agriculture, and in particular industrial monocultures, needs to be realigned to become kinder to ecosystems and more beneficial for the individuals and the communities that feed us (HLPE, 2013; IAASTD, 2009). This will not occur, however, until the scarcity narrative is set aside. The necessity of following an alternative path is acknowledged in places by FAO modelers. The foreword to FAO’s latest major modeling report, The future of food and agriculture: Alternative pathways to 2050, says: “Swift and purposeful actions are needed to ensure the sustainability of food and agriculture systems in the long run” (FAO, 2018). Yet even that report still perpetuates the scarcity narrative. It privileges a quantitative productivist focus as the ultimate arbiter of the alternative pathways it explores, thereby undermining all such possibilities. What is needed instead, from the FAO as well as others, is a focus on the broad consequences of agriculture and food systems, that is, their multifunctionality. Multifunctionality is the idea that agriculture is deeply embedded in other systems, and therefore how it is conducted has consequences well beyond the single metric of crop production by volume (IAASTD, 2009; Kremen & Merenlender, 2018). For instance, industrial agriculture is the single largest contributor to climate change and other forms of atmospheric pollution (Bauer et al., 2016; Goodland & Anhang, 2009; Steinfeld et al., 2006). Agriculture also provides livelihoods; it cleans or pollutes water (aquifers, surface waters, and oceans); it conserves or degrades biodiversity; it provides landscape value and culturally appropriate food; and so forth (IUCN Task Force on Systemic Pesticides, 2017). These contributions of agriculture, all effectively ignored by quantitative models of calorific production, should receive at least as much consideration as does productivity from decision-makers, if only because, unlike productivity, these functions are often more at risk. Having discussed some of the limitations of modeling agricultural productivity, which is a comparatively simple output, it can reasonably be questioned whether modeling of agriculture on a global level can ever perform a constructive role. Perhaps ultimately a more useful role for modeling is at more local scales, and modeling that asks questions of a more defined nature. Work on nitrogen modeling of the Paris basin constitutes an example of how modeling can have more modest aims yet be useful to policymakers and others (Billen et al., 2012). Perhaps FAO should now move its focus to assessing the relative effects of pesticides and other inputs on biodiversity or human health from different agricultural systems? This is a domain that the current regulation of pesticides is failing to address (Vandenberg et al., 2012). Such an analysis would beneficially serve to put agriculture on notice that its long-ignored externalities will in future move to center stage.

The myth of a food crisis107

Acknowledgments The author is extremely grateful to Timothy Wise, Philip McMichael, Allison Wilson, and three anonymous reviewers for reading and constructively commenting on drafts of this chapter.

References Abraham, B., Araya, H., Berthe, T., Edwards, S., Gujja, B., Khadka, R. B., … Verma, A. (2014). The system of crop intensification: Reports from the field on improving agricultural production, food security, and resilience to climate change for multiple crops. Agriculture and Food Security, 3, 4. ActionAid. (2013). Rising to the challenge: Changing course to feed the world in 2050. https:// www.actionaidusa.org/publications/feed-world-2050/. Agrimonde. (2009). Scenarios and challenges to feeding the world in 2050. In Summary report. France: CIRAD-NAR. Alavi, H. R., Htenas, A., Kopicki, R., Shepherd, A. W., & Clarete, R. (2012). Trusting trade and the private sector for food security in Southeast Asia. Washington, DC, USA: World Bank Publications. Alexandratos, N., & Bruinsma, J. (2012). World agriculture: Towards 2030/2050. ESA working paper no. 12-03 Rome: FAO. Baines, J. (2015). Fuel, feed and the corporate restructuring of the food regime. The Journal of Peasant Studies, 42(2), 295–321. Battisti, D. S., & Naylor, R. L. (2009). Historical warnings of future food insecurity with unprecedented seasonal heat. Science, 323, 240–244. Bauer, S. E., Tsigaridis, K., & Miller, R. (2016). Significant atmospheric aerosol pollution caused by world food cultivation. Geophysical Research Letters, 43, 5394–5400. Beddington, J. (2009). As reported in the guardian. In World faces ‘perfect storm’ of problems by 2030, chief scientist to warn. Wednesday 18 March 2009. Available at http://www. guardian.co.uk/science/2009/mar/18/perfect-storm-johnbeddington-energy-food-climate. (Accessed 6 2 2019). Billen, G., Garnier, J., Silvestre, M., Thieu, V., Barles, S., & Chatzimpiros, P. (2012). Localising the nitrogen imprint of Paris food supply: The potential of organic farming and changes in human diet. Biogeosciences, 9, 607–616. Boxall, R. A. (2001). Post-harvest losses to insects—A world overview. International Biodeterioration & Biodegradation, 48, 137–152. https://doi.org/10.1016/s0964-8305(01)00076-2. Bruinsma, J. (2003). World agriculture: Towards 2015/30 an FAO perspective. London: Earthscan. Burchi, F., Fanzo, J., & Frison, E. (2011). The role of food and nutrition system approaches in tackling hidden hunger. International Journal of Environmental Research and Public Health, 8(2), 358–373. https://doi.org/10.3390/ijerph8020358. Cassidy, E. S., West, P. C., Gerber, J. S., & Foley, J. A. (2013). Redefining agricultural yields: From tonnes to people nourished per hectare. Environmental Research Letters, 8, 034015. Chakrabortty, A. (2008). Secret report: Biofuel caused food crisis. https://www.theguardian. com/environment/2008/jul/03/biofuels.renewableenergy. (Accessed 6 February 2019). Chaumet, J., et al. (2009). Agrimonde: Scenarios and challenges for feeding the world in 2050. Paris: CIRAD-INRA. Conway, G. (2012). One billion hungry: Can we feed the world?. New York: Cornell University Press.

108

Rethinking Food and Agriculture

Danielsen, F., et al. (2009). Biofuel plantations on forested lands: Double jeopardy for biodiversity and climate. Conservation, 23, 348–358. de Fraiture, C., et al. (2007). Looking ahead to 2050: Scenarios of alternative investment approaches. In D. Molden (Ed.), Water for food, water for life: A comprehensive assessment of water management in agriculture (pp. 91–145). London, UK: Earthscan. de Gorter, H., Drabik, D., & Just, D. R. (2015). The economics of biofuel policies: Impacts on price volatility in grain and oilseed markets. (Palgrave studies in agricultural economics and food policy). Basingstoke, UK: Palgrave Macmillan. Dibden, J., Gibbs, D., & Cocklin, C. (2013). Framing GM crops as a food security solution. Journal of Rural Studies, 29, 59–70. Diwakar, M. C., Kumar, A., Verma, A., & Uphoff, N. (2012). Report on the world record SRI yields in Kharif season 2011 in Nalanda District, Bihar State, India. In Agriculture today. June 54–56. Ehrlich, P. (1968). The population bomb. Sierra Club/Ballantine Books. Fageria, N. K. (2007). Yield physiology of rice. Journal of Plant Nutrition, 30, 843–879. FAO. (2001). Food balance sheets: A handbook. Rome: FAO. http://www.fao.org/worldfoodsituation/foodpricesindex/en/. (Accessed 7 March 2019). FAO. (2006). World agriculture: Towards 2030/2050. In Interim report. Rome: FAO. FAO. (2018). The future of food and agriculture: Alternative pathways to 2050. Rome: FAO. FAOSTAT. (2017). http://www.fao.org/faostat/en/#home. Fischer, G., Frohberg, K., Keyzer, M. A., & Parikh, K. S. (1988). Linked national models: A tool for international food policy analysis. Dordrecht: Kluwer. Food First. (2016). The true extent of hunger: What FAO isn’t telling you. https://foodfirst.org/ wp-content/uploads/2016/06/Summer2016Backgrounder.pdf. (Accessed 10 May 2019). Funtowicz, S. O., & Ravetz, J. R. (1990). Uncertainty and quality in science for policy. Dordrecht: Kluwer. Germer, J., & Sauerborn, J. (2008). Estimation of the impact of oil palm plantation establishment on greenhouse gas balance. Environment, Development and Sustainability, 10(6), 697–716. Goodland, R., & Anhang, J. (2009). Livestock and climate change. Washington, DC: World Watch. Herforth, A. (2015). Access to adequate nutritious food: New indicators to track progress and inform action. In D. E. Sahn (Ed.), The fight against hunger and malnutrition: The role of food, agriculture and targeted policies (pp. 139–162). Oxford University Press. https://doi. org/10.1093/acprof:oso/9780198733201.003.0007. Hertel, T. W. (2011). The global supply and demand for agricultural land in 2050: A perfect storm in the making? American Journal of Agricultural Economics, 93, 259–275. Herren, H., Chihyoka, A., Scaramella, C., Mpandeli, S., Sirengo, A., & Mwendandu, R. (2011). WBI global dialogue on adaptation and food security. World Bank Institute. Hincks, J. (2018). The world is headed for a food security crisis. Here’s how we can avert it. http://time.com/5216532/global-food-security-richard-deverell/. (Accessed 9 February 2019). HLPE. (2013). Investing in smallholder agriculture for food security. In A report by the High Level Panel of Experts on Food Security and Nutrition of the Committee on World Food Security, Rome. IAASTD. (2009). Agriculture at a crossroads. Nairobi: United Nations Environment Programme (UNEP). IUCN Task Force on Systemic Pesticides. (2017). Worldwide integrated assessment of the effects of systemic pesticides on biodiversity and ecosystems. IUCN.

The myth of a food crisis109

IPBES. (2019). The IPBES global assessment report on biodiversity and ecosystem services, summary. http://bit.ly/IPBESReport. Accessed on 10.5.2019. International Rice Research Institute. (1979). Farm-level constraints to high rice yields in Asia: 1974–1977. Los Baños, Laguna, Philippines: International Rice Research Institute. Jerven, M. (2012). Poor numbers. Ithaca, USA: Cornell University Press. Kavalleri, A., Conforti, P., & van der Mensbrugghe. (2016). The global perspectives system (GAPS) version 1.0. Rome: FAO. Kassam, A., Freidrich, T., Shaxson, F., & Pretty, J. (2009). The spread of conservation agriculture: Justification, sustainability and uptake. International Journal of Agricultural Sustainability, 7(4), 292–320. Kassam, A., Stoop, W., & Uphoff, N. (2011). Review of SRI modifications in rice crop and water management and research issues for making further improvements in agricultural and water productivity. Paddy and Water Environment, 9, 163. https://doi.org/10.1007/ s10333-011-0259-1. Kassam, A., Friedrich, T., & Derpsch, R. (2018). Global spread of conservation agriculture. International Journal of Environmental Studies. https://doi.org/10.1080/00207233.2018.1 494927. Published online at. Kremen, C., & Merenlender, A. M. (2018). Landscapes that work for biodiversity and people. Science, 362. https://doi.org/10.1126/science.aau6020. eaau6020. Kumar, D., & Kalita, P. (2017). Reducing postharvest losses during storage of grain crops to strengthen food security in developing countries. Food, 6, 8. https://doi.org/10.3390/ foods6010008. Latham, J. R. (2015). How the great food war will be won. Independent science news. https:// www.independentsciencenews.org/environment/how-the-great-food-war-will-be-won/. (Accessed 8 February 2019). Lakoff, G. (2004). Don’t think of an elephant: Know your values and frame the debate. Chelsea Green Publishing. ISBN 978-1-931498-71-5. Lappé, F. M., & Collins, J. (2015). World hunger: Ten myths. Grove Press. Malthus, T. R. (1798). An essay on the principle of population. London: J. Johnson. McCalla, A. F., & Revoredo, C. L. (2001). Prospects for global food security: A critical appraisal of past projections and predictions. In Food, agriculture, and the environment discussion paper 35. Washington, DC: International Food Policy Research Institute. McMichael, P. (2009). A food regime analysis of the ‘world food crisis’. Agriculture and Human Values, 26, 281–295. https://doi.org/10.1007/s10460-009-9218-5. Mohanty, S. K., & Swain, M. R. (2019). Bioethanol production from corn and wheat: Food, fuel, and future. In R. C. Ray, & S. Ramachandran (Eds.), Bioethanol production from food crops sustainable sources, interventions, and challenges (pp. 45–59). Academic Press. Nelson, G. C., Rosegrant, M. W., Palazzo, A., Gray, I., Ingersoll, C., Robertson, R., … You, L. (2010). Food security, farming, and climate change to 2050: Scenarios, results, policy options. Washington, DC: International Food Policy Research Institute. OECD. (2002). Agricultural policies in OECD countries: A positive reform agenda. Paris: OECD. OECD-FAO. (2016). Agricultural outlook 2016-2025. Rome: FAO. Parry, M. L., Rosenzweig, C., Iglesias, A., Livermore, M., & Fischer, G. (2004). Effects of climate change on global food production under SRES emissions and socioeconomic scenarios. Global Environmental Change, 14, 53–67. https://doi.org/10.1016/j. gloenvcha.2003.10.008. Peekhaus, W. (2010). Monsanto discovers new social media. International Journal of Communication, 4, 955–976.

110

Rethinking Food and Agriculture

Peng, S., Cassman, K. G., Virmani, S. S., Sheehy, J., & Khush, G. S. (1999). Yield potential trends of tropical rice since the release of IR8 and the challenge of increasing rice yield potential. Crop Science, 39, 1552–1559. https://doi.org/10.2135/cropsci1999.3961552x. Porter, T. M. (1995). Trust in numbers the pursuit of objectivity in science and public life. Princeton: Princeton University Press. Reilly, M., & Willenbockel, D. (2010). Managing uncertainty: A review of food system scenario analysis and modelling. Philosophical Transactions of the Royal Society, 365, 3049–3063. Rivera-Ferre, M. G. (2009). Supply vs. demand of agri-industrial meat and fish products: A chicken and egg paradigm? Journal of Sociology of Agriculture, 16, 90–105. Robinson, S., et al. (2015). IFPRI discussion paper 01483. In The international model for policy analysis of agricultural commodities and trade (IMPACT). Washington, DC, USA: IFPRI. Rosegrant, M. W., Agcaoili-Sombilla, M., & Perez, N. D. (1995). Global food projections to 2020: Implications for investment. Washington, DC, USA: The International Food Policy Research Institute. Rosegrant, M. W., Paisner, M. S., Meijer, S., & Witcover, J. (2001). Global food projections to 2020: Emerging trends and alternative futures. Washington, DC, USA: The International Food Policy Research Institute. Sampson, D. (2018). Productivism, agroecology, and the challenge of feeding the world. Gastronomica: The Journal of Critical Food Studies, 18, 41–53. Scrimshaw, N. S. (1978). Global use of the instruments of scholarship for the conquest of hunger—The World Hunger Program of the United Nations University. In J. C. Somogyi (Ed.), Vol. 28. Solution of nutritional problems: The contribution of producers, distributors and nutritionists symposium organized by the Institute for Nutrition Research of the Green Meadow Foundation, Zürich (pp. 155–166). Rüschlikon-Zürich, Basel: Karger. 1979. Scrieciu, S. (2007). Commentary: The inherent dangers of using computable general equilibrium models as a single integrated modelling framework for sustainability impact assessment. A critical note on Böhringer and Löschel 2006. Ecological Economics, 60, 678–684. Sen, A. (1981). Poverty and famines. Oxford: Clarendon. Sharon, M. E. M., Abirami, C. V. K., & Alagusundaram, K. (2014). Grain storage management in India. Journal of Postharvest Technology, 02, 12–24. Steinfeld, H., et al. (2006). Livestock’s long shadow: Environmental issues and options. Rome: United Nations Food and Agriculture Organization. Stone, G. D., & Glover, D. (2011). Genetically modified crops and the ‘food crisis’: Discourse and material impacts. Development in Practice, 21, 509–516. Stoop, W. A., Sabarmatee, Sivasubramanian, P., Ravindra, A., Sen, D., Prasad, S. C., & Thakur, A. K. (2017). Opportunities for ecological intensification: Lessons and insights from the system of rice/crop intensification—Their implications for agricultural research and development approaches. CAB Reviews, 12(036), 1–19. Tapondjou, L. A., Adler, C., Bouda, H., & Fontem, D. A. (2002). Efficacy of powder and essential oil from Chenopodium ambrosioides leaves as post-harvest grain protectants against six-stored product beetles. Journal of Stored Products Research, 38, 395–402. Taylor, M., & Bhasme, S. (2019). The political ecology of rice intensification in south India: Putting SRI in its places. Journal of Agrarian Change, 19, 3–20. Tefera, T., Kanampiu, F., De Groote, H., Hellin, J., Mugo, S., Kimenju, S., … Banziger, M. (2011). The metal silo: An effective grain storage technology for reducing post-harvest insect and pathogen losses in maize while improving smallholder farmers’ food security in developing countries. Crop Protection, 30, 240–245.

The myth of a food crisis111

Tomlinson, I. (2011). Doubling food production to feed the 9 billion: A critical perspective on a key discourse of food security in the UK. Journal of Rural Studies, 29, 81–90. https://doi. org/10.1016/j.jrurstud.2011.09.001. UN. (2017). World population prospects: The 2017 revision. https://www.un.org/development/desa/publications/world-population-prospects-the-2017-revision.html. (Accessed 2 February 2019). Uphoff, N. (2017). Developments in the system of rice intensification (SRI). In T. Sasaki (Ed.), Vol. 2. Achieving sustainable rice cultivation (pp. 183–211). Cambridge, UK: Burleigh-Dodds. Vandenberg, L. N., Colborn, T., Hayes, T. B., Heindel, J. J., Jacobs, D. R., Jr., Lee, D.-H., … Myers, J. P. (2012). Hormones and endocrine-disrupting chemicals: Low-dose effects and nonmonotonic dose responses. Endocrine Reviews, 33(3), 378–455. https://doi. org/10.1210/er.2011-1050. Wise, T. (2013). Can we feed the world in 2050? A scoping paper to assess the evidence. In Global development and environment institute working paper no. 13-04. https://sites.tufts. edu/gdae/files/2019/10/13-04Wise_CanWeFeedWorld2050.pdf.

Further reading Bruinsma, J. (2009). How to feed the world in 2050. In Paper prepared for the high level expert forum. Rome 12e19 October 2009. Available at http://www.fao.org/fileadmin/templates/ wsfs/docs/expert_paper/How_to_Feed_the_World_in_ 2050.pdf. (Accessed 7 March 2019). Perfecto, I., & Vandermeer, J. (2010). The agroecological matrix as alternative to the land-­ sparing/agriculture intensification model. Proceedings of the National Academy of Sciences, 107, 5786–5791.

This page intentionally left blank

Animal ethics as a critique of animal agriculture, environmentalism, foodieism, locavorism, and clean meat

6

Robert C. Jones California State University, Dominguez Hills, Carson, CA, United States

6.1 Introduction For in whatever form it has taken through the centuries…speciesism—or the total organization of material and symbolic human life around the domination and mass killing of other sensitive beings—has throughout history served as the “primordial” substructure or organizing principle of the human project, the determining episteme and habitus of every human culture, economy, and society. (Sanbonmatsu, 2011, p. 31)

In all of human history, approximately 110 billion Homo sapiens have ever walked the planet (Curtin, 2007). However, in just one Earth year, Homo sapiens slaughter more than 70 billion land animals, while the global commercial fishing industry kills 1–3 trillion animals each year (Animal Advocacy by Numbers, 2016). Most individual land animals are raised in conditions that cause great suffering within a global system so vast that one-third of the Earth’s surface is devoted to imprisoning, transporting, and killing other beings for their milk, eggs, flesh, skin, and bones (Bland, 2012). If that were not enough, world fishing trends point to a global eradication of all taxa currently fished by the year 2048 (Worm et al., 2006). Surely, by any coherent calculus, these figures amount to a moral atrocity. Given these dire facts, the focus of this chapter is to examine some of the key concepts and debates in the areas of environmental, food, and animal ethics; to look at some of the theoretical ethical arguments for ending animal exploitation; to challenge some of the key justifications from those in the environmental, foodie, and locavore movements who are so attached to “humane” exploitation; to investigate whether so-called “clean meat” will end animal agriculture as we know it; and finally gesture toward what I see as the root cause of the suffering of countless sentient beings.

6.2 Ethics basics: Moral status, moral value, and anthropocentrism Unlike rocks and plastic straws, human beings and nonhuman animals are experiential subjects, that is, there is a subjective “what it’s like” from the inside to be a human Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00006-2 © 2021 Elsevier Inc. All rights reserved.

114

Rethinking Food and Agriculture

being or nonhuman animal. Humans and other animals are the kinds of things whose experiences matter to us, beings whose lives can go better or worse, for us, a capacity that generates interests, for example, interests in our own well-being. Having interests in this sense implies that humans and other animals can be wronged in a way that rocks or plastic straws cannot, so interests for one entity create duties and obligations for the rest of us. The fact that our interests can be violated creates duties and obligations upon certain others of us to respect those interests. In other words, experiential subjects have moral status. To say that an entity has moral status is to say that its interests matter morally for its own sake—not merely for the sake of the interests of others—and that it has entitlement to protections afforded by moral norms. To have moral status is to be the sort of thing whose interests must be considered from the moral point of view. But what about things like trees, rivers, and ecosystems? As far as we know, they are not the kinds of things that have subjective experiences and so are not the kinds of things that have interests, at least not in the same sense that human and nonhuman animals do. When we think about things like interests and well-being, it’s important to consider their relations to ethical or moral value. Put simply, values ground judgments regarding those things that we, valuers, care about; things like what is good or bad, right or wrong. Generally, ethicists identify two kinds of value. Intrinsic value refers to the value that a thing has in itself, or for its own sake. Philosophers argue that things like humans, (human) well-being, and justice possess intrinsic value. On the other hand, things that provide only extrinsic or instrumental value are not seen as valuable in-and-of-themselves but rather as valuable because of the goods they bring about. One might say, for example, that while human well-being is intrinsically valuable, the means to human well-being—proper nutrition, adequate healthcare, etc.—are merely instrumentally valuable. Related to the notion of value are the concepts of moral considerability and moral significance. An entity is said to be morally considerable just in case it is a bona fide member of the moral community in that it has robust interests and can be wronged in a morally relevant way. The fact that a being is morally considerable means that moral agents have obligations to that being. Saying that an entity is morally considerable is like saying that it’s “in the club” of things whose interests moral agents must consider. Once a being is morally considerable, however, we may then need to adjudicate questions of relative moral value between beings. That becomes a question of moral significance. Moral significance speaks to the moral value of the members once admitted to the “club.” Just because two entities are “in the club,” it does not follow that they are of equal moral value. Surely, beings like all living, sentient humans, chimpanzees, dogs, cats, deer, wolves, and birds are in the club, i.e., are morally considerable. But does that imply—all things being equal—that a human being and a sparrow have equal moral value? This a complex question worth exploring in more detail. But at this point, we can formulate a notion of moral status as follows: an entity X has moral status just in case (1) moral agents have moral obligations to X, (2) X has basic welfare interests, and (3) the moral obligations owed to X are based on X’s interests (DeGrazia, 2008).

Animal ethics

115

There are a number of theories for the grounds of moral status. The most widely held view is the anthropocentric account of moral status. The anthropocentric account holds that only being a member of the species Homo sapiens confers moral considerability and maximal moral significance. On the anthropocentric account, all and only humans have full moral status; being human (i.e., being conceived by human parents or possessing a human genetic code) is a necessary condition for entrance into the “club.” One virtue of the anthropocentric account is that it secures moral status and maximal moral significance for all humans, including infants, the severely cognitively differently abled, and those humans living with a permanent loss of consciousness (e.g., those in a persistent vegetative state). Yet, the anthropocentric account is fatally problematic for a number of reasons. According to the anthropocentric account, there exists some set of distinctively human properties that confer full moral status on all and only human beings. But for the anthropocentric account to be successful in securing full moral status for all human beings, the account must (1) identify which set of characteristics is distinctively human, (2) demonstrate that the set is possessed universally by all humans, (3) provide an account of why those particular characteristics (but not others) are the morally relevant ones, and (4) explain why those particular characteristics are sufficient to secure moral status for all and only human beings. With regard to (1), the anthropocentric account clearly, though trivially, identifies a characteristic of being human that is distinctively human, namely, being human; with regard to (2) the account trivially asserts that the characteristic of being human is possessed universally by all humans. Asserting that being human is distinctively human, and that the characteristic of being human is possessed universally by all humans tells us very little about why all and only human beings are in the moral club. Furthermore, the anthropocentric account appears to beg the question here because the very question at the heart of the anthropocentric account is this: What is it about membership in the species Homo sapiens alone that uniquely confers full moral status? To say that being human—a characteristic clearly possessed by all human beings—is the uniquely morally relevant characteristic securing moral status for all and only human beings tells us nothing whatsoever about why being human is sufficient. Were we to encounter an extraterrestrial who possessed human-like capacities for reasoning and planning, should we deny this being the same moral status that we afford full human persons? It seems clear that, contrary to the anthropocentric account, human biological properties cannot be necessary for moral status. In response, the advocate of the anthropocentric account may concede that it is not merely membership in the species Homo sapiens that is doing the moral heavy lifting here, but rather those distinctively human species-typical abilities that are necessary for entrance into the club, capacities like rationality, intelligence, language, etc. Yet, this response will not do. For the anthropocentric account to secure full moral status for all humans it must also demonstrate that the alleged distinctively human abilities are possessed universally by all humans. However, if the moral divide between h­ uman and nonhuman animals rests on the possession of some required set of uniquely human cognitive abilities possessed by all and only humans, then there will always exist

116

Rethinking Food and Agriculture

some humans who lack these characteristics. That is, for any human capacity we select, there will always exist some human who lacks this capacity, while there may exist some nonhuman animals who will possess this capacity, to varying degrees. A common anthropocentric response is to claim that though some human beings lack those cognitive abilities required for entrance into the moral club, they remain members of a species whose typical members possess the requisite capacities, something that cannot be said for any member of a nonhuman animal species. This speciestypicalism approach argues that though a permanently and severely congenitally cognitively disabled infant may forever lack the abilities required by the anthropocentric account, because she is a member of a species whose typical member possesses these abilities, she is afforded full membership in the moral community. This view is implausible since it holds that some individuals can entirely lack the properties supposedly required for moral considerability, yet still be morally considerable merely because most members of the group possess those properties. Imagine someone arguing in a like fashion that since some particular artist’s works are, by and large, excellent works of art, it therefore follows that every work of art produced by that artist must also be an excellent work of art solely in virtue of their being produced by an artist whose typical artworks are excellent (Lowe, 2014); this a terrible argument. But how then does the permanently severely congenitally cognitively disabled infant gain full membership into the moral community? Her honorary membership can’t be due to her biological species membership since, as we have seen, that view is a woefully problematic account of moral status. Furthermore, were humans able someday through gene therapy to produce a chimpanzee who developed cognitive capacities comparable to those of a human person, surely this “Superchimp” would be entitled to the same moral status due human persons despite her being an atypical member of a species whose typical members do not possess the kinds of cognitive capacities she now does (McMahan, 2002). Conferring moral status on nonparadigm members of a species based on the capacities of paradigm members is unwarranted. These and other challenges provide grounds for rejecting anthropocentrism as a basis for moral status.

6.3 Ethical foundations of the modern animal rights movement In contrast to the anthropocentric account, a number of philosophers have formulated more inclusionary accounts of moral status intended to expand the sphere of the moral world to include nonhuman animals as robust members of the moral community. Easily the most influential theory advocating increased moral status for nonhuman animals is that of Singer (1975). Singer provides what can be described as a sentientist account of moral status. To be sentient is to be the subject of experience, to have subjective experiences. In particular, sentient beings possess the capacity for joy, pleasure, pain, and suffering,

Animal ethics

117

capacities that make a difference morally, i.e., capacities that are morally relevant. According to Singer, sentience—the capacity to experience pain or pleasure—is both necessary and sufficient for having morally considerable interests. Understanding Singer’s argument for what he calls animal liberation requires understanding five notions central to Singer’s view: (1) the basic principle of equality, (2) sentience, (3) the principle of equal consideration of interests, (4) speciesism, and (5) the relevance principle. Reflecting on recent social justice movements such as women’s liberation, black liberation, LGBT+ rights, disability rights, and others, we notice that one thing that underlies and connects these movements is a belief that, in an important and profound sense, all humans are equal. This belief, the basic principle of equality, lies at the heart of Singer’s view. But what does it mean to say that all humans are equal? Given that humans differ from each other so significantly in their physical, moral, emotional, and cognitive abilities and capacities, surely, as a descriptive empirical assertion, claims of human equality in this sense are clearly factually untrue. The basic principle of equality is not intended as a factual but a normative concept: “equality is a moral idea...a prescription of how we should [emphasis added] treat human beings” (Singer, 2011, pp. 4–5). The primary central descriptive claim grounding the basic principle of equality is the fact that humans are experiential subjects. As we’ve seen, since we are sentient beings, and the kinds of beings whose experiences matter to us, we possess morally relevant interests (e.g., interests in our own well-being). The basic principle of equality, coupled with sentience, and combined with the interests that the possession of sentience provide, lead to the central principle driving Singer’s view: the principle of equal consideration of interests, the essence of which is “that we give equal weight in our moral deliberations to the like interests of all those affected by our actions” (Singer, 2011, p. 20). When engaged in any decision-making procedure concerning how we ought to treat one another morally, the principle of equal consideration of interests requires that we consider the interests of all humans equally. But since sentience provides the basis for the equality of human beings, and since human beings are not the only sentient beings, to be consistent, we must extend the principle of equal consideration of interests to all sentient creatures. The principle of equal consideration of interests requires that we weight interests not on the basis of the species (or race or gender) of an individual, but on her own merits, independently of such morally irrelevant considerations. Importantly then, the principle of equal consideration of interests emphasizes the moral salience of our interests as individuals, not as members of a particular species. Yet, equal consideration of interests is not synonymous with moral equality and equal treatment. The principle of equal consideration of interests “commits us to treating like interests in a comparable fashion, a key principle of justice, but it does not tell us what interests particular individuals have” (Garner, 2013, p. 98). To privilege the interests of humans over nonhumans solely in virtue of species membership is a form of speciesism. There are a number of important and distinct ways that scholars have characterized speciesism. I will discuss these in some detail anon. But for Singer, speciesism

118

Rethinking Food and Agriculture

is akin to other “-isms” such as racism, sexism, and ableism, and is best conceived of as a bias in favor of the interests of the members of one’s own species and against the interests of members of other species based solely or primarily on species membership. In Singer’s view, speciesism involves the belief that members of one’s own species are more valuable than and morally superior to members of another species, a prejudice that often leads to discriminatory practices and institutional oppression. Just as the wrongness of racism consists in discrimination based on a morally irrelevant trait (namely, race), the wrongness of speciesism consists in discrimination based on a different morally irrelevant trait, namely, species membership. To reiterate, equality for nonhuman animals does not entail equality of treatment, but merely the equal consideration of interests. Adjudicating differences in treatment between competing interests requires the last notion key to Singer’s view, the relevance principle. The relevance principle states that whether a difference between individuals justifies a difference in treatment depends on the kind of treatment in question. The extension of the basic principle of equality from one group to another does not imply that we must treat both groups in exactly the same way, or grant exactly the same rights to both groups. Whether we should do so will depend on the nature of the members of the two groups. The basic principle of equality does not require equal or identical treatment; it requires equal consideration. Equal consideration for different beings may lead to different treatment and different rights (Singer, 1975, p. 2). Thus equality for animals does not require, for example, that we grant pigs the right to vote, not because the interests of pigs are of less moral concern, but rather because pigs, unlike humans, have no interest in voting. On the other hand, since pigs, like humans, have an interest in not suffering, livestock production techniques that inflict suffering on pigs solely to satisfy the palates of consumers are impermissible (Cochrane, 2012, p. 5). Singer’s view, combined with his utilitarian stance, champions maximizing the overall welfare of all sentient beings and condemns practices such as industrialized livestock production as discriminatory, immoral, and clear cases of institutionalized violence and oppression. In contrast to Singer’s approach to animal liberation, philosopher Tom Regan in his influential The Case for Animal Rights (1983) rejects Singer’s utilitarian arguments for animal liberation and instead provides an account of liberation for animals that requires recognition that nonhuman animals possess moral rights. For Regan, what matters morally is the capacity to be the subject of experiences that matter to oneself. Possessing certain physiological, emotional, psychological, and cognitive capacities, over and above mere sentience, makes one what Regan calls a subject-of-a-life: To be the subject of a life…involves more than merely being alive and more than merely being conscious…Individuals are subjects-of-a-life if they have beliefs and desires; perception memory, and a sense of the future, including their own future; an emotional life together with feelings of pleasure and pain; preference and welfare-interests; the ability to initiate action in pursuit of their desires and goals; a psychophysical identity over time; and an individual welfare in the sense that their experiential life fares well or ill for them, logically independent of their being the object of anyone else’s interests. (Regan, 1983, p. 243)

Animal ethics

119

Since a great number of nonhuman animals are subjects-of-a-life whose value cannot be reduced to their utility to humans (i.e., their instrumental value), it follows for Regan that the other animals possess what he calls inherent value, a value that is intrinsic, independent of how animals are valued by—or valuable to—humans. Possession of inherent value merits the respect due a subject-of-a-life and confers upon nonhuman animals strong moral rights. For Regan, all who possess inherent value possess it equally. The possession of rights implies not merely an entitlement to equal consideration of interests, but categorical protection against being treated merely as a means to some human end. Unlike Singer’s utilitarian view, which allows for the possibility that individual interests may be trumped in cases where appeals to the general welfare outweigh the interests of the individual, in Regan’s view, rights are, for the most part, inviolable, even if trumping them, or harming the rights-bearer, would increase general welfare. Though it follows from Peter Singer’s view that practices such as the confinement conditions found in “livestock production” facilities (a.k.a., factory farms) are institutionally oppressive, immoral, and should be ameliorated (if not abolished), being that Singer’s view is largely utilitarian, it allows for the possibility of the instrumental use of other-than-human animals in cases where their interests are considered equally and such use increases aggregate welfare. However, for Regan, the possession of rights implies not merely an entitlement to equal consideration of interests, but categorical protection against being treated merely as a means to some human end. Understood this way, nonhuman animals make valid moral claims upon us to protect their interests in the face even of appeals to their general welfare. Thus Regan’s theory condemns and calls for the total abolition of every form of nonhuman animal exploitation and any instrumental use of animals, even in cases where such use improves general welfare, advocating nothing less than the “total dissolution of the animal industry as we know it” (Regan, 1983, p. 395). Specifically, the animal rights movement proper is “committed to a number of goals, including: the total abolition of the use of animals in science; the total dissolution of commercial animal agriculture; the total elimination of commercial and sport hunting and trapping” (Regan, 1985, p. 13). For Regan and his adherents, e.g., Francione (1996), true animal liberation can be achieved only through a rights-based approach, since only a rights-based approach can properly ground calls for the abolition of all forms of animal use and exploitation at the hands of humans. Though Singer and Regan provided the first extensive systematic ethical theories built upon a set of core philosophical principles aimed at expanding the moral status of nonhuman animals and overthrowing the anthropocentric view of moral status, a number of other thinkers have offered alternative accounts of animal liberation (Gruen, 2015; Ko, 2019; Nibert, 2002; Pluhar, 1995; Rachels, 1990; Rollin, 1992; Sapontzis, 1989; Taylor, 2017).

6.4 Animal liberation never was a triangular affair As we have seen, the traditional view of moral status, the anthropocentric view, holds that all and only humans have moral status; that is, only humans have intrinsic value, whereas nonhuman animals and the natural world and environment have, at best,

120

Rethinking Food and Agriculture

instrumental value. By contrast, animal liberationists like Singer argue that sentient beings (or subjects-of-a-life) have intrinsic value, while nonsentient entities have instrumental value only. In the 1970s, a different view arose that challenged this thinking, claiming distinction from and superiority to both the anthropocentric view and the animal liberation view. This view, known as environmental holism, became and continues to be the basis for modern environmentalism. In contrast to both the anthropocentric and animal liberation/sentientist/subject-of-a-life views, environmental holists see not individual beings—be they humans or nonhuman animals—as the primary objects of moral considerability, but instead see the integrity of ecosystems, the “land,” and biotic communities as the primary units of ethical concern. For environmental holists, groups and communities—not individual sentient beings—possess intrinsic value and thus deserve ethical primacy. Sentient beings by contrast have instrumental value and are thus subordinated to the well-being of the ecosystem and biotic communities. Following the publication of Animal Liberation, environmentalists went on the offensive against such theories. In his debate-defining essay “Animal liberation: A triangular affair” (Callicott, 1980), philosopher and environmental holist J. Baird Callicott lays out what remains to this day the central issue pitting animal rights folks against environmentalists. For Callicott, biotic communities have intrinsic value, with the ecological whole being the ultimate measure of moral value. On this view, the value of individual organisms lies primarily (or even solely) in their ecological function, and their well-being should be considered only inasmuch as they contribute to the ecological whole. Aldo Leopold’s Maxim captures the moral heart of environmental holism: “A thing is right when it tends to preserve the integrity, stability, and beauty of the biotic community; it is wrong when it tends otherwise” (Leopold, 1949, pp. 224–225). With such a maxim driving environmental holism, the radical egalitarianism implicit in the animal liberation view is overridden by an inegalitarianism manifest in the “land ethic” where individual nonhuman animals are assigned varying degrees of instrumental value depending on their contribution to the whole. The tension between the animal rights advocate’s primacy of individual sentient beings and the environmentalist’s holism remains the “most fundamental theoretical difference between environmental ethics and the ethics of animal liberation” (Callicott, 1980, p. 337). Callicott sees environmental holism as life-affirming in its celebration of the value of ecosystems and biotic communities, whereas he sees animal liberation as “life loathing” and “world denying” and too narrowly individualistic (Callicott, 1980, p. 333). When the needs of the “whole” clash with the interests of the individuals that comprise it, the former should trump the latter. This best explains the plight of endangered species. Is the moral badness we find in our driving species to extinction found in the suffering of its individual members, or do we decry the loss of species because of its impact on ecosystems and biotic communities? For the environmental holist, the answer is obvious. To clarify, in Callicott’s view, animal liberation is actually a “triangular affair” between (1) the individual-human-valuing anthropocentrists, (2) the individual-nonhuman-animalvaluing animal liberation individualists, and (3) the nonanthropocentric, nonindividualist environmentalists whose commitment to environmental holism makes it clearly a superior view. However, animal rights advocates have had plenty of time to reply.

Animal ethics

121

The first critique of the environmentalist position argues that the very conceptual foundation of the environmentalist’s view (namely, Leopold’s Maxim) fails not only as a land ethic, but as an ethic of any sort, for it makes absolutely no reference to ethical values. Instead, the view employs concepts of esthetic value (beauty) or of biological or ecological condition (integrity, stability). For an ethical theory to employ no moral concepts is, at best, impoverished. Second, it is difficult to accept that the moral considerability of an individual is completely determined by the role they play in an ecosystem. This seems to imply that it may be legitimate to use some individuals as mere means to promote, in this case, the alleged integrity, stability, and beauty of nature. Third, the most damning flaw of environmental holism resides in the fact that animal liberation never was a “triangular affair,” but really a difference between (1) the radical egalitarianism of animal liberation and (2) the anthropocentrism of both the anthropocentrists and the environmentalists. Allow me to explain. Many environmental holists condone sacrificing individuals for the sake of the whole—for example, by shooting rabbits to preserve plant species—but they are reluctant to sacrifice human interests in similar situations. Yet, the most abundant species destroying biotic communities is Homo sapiens. If human individuals are just another element within the larger and more important biotic community, environmental holism should call for us to “control” or “eliminate” some of these individuals for the sake of the larger whole, an implication Tom Regan calls “environmental fascism” (Regan, 1983, pp. 361–362). Accordingly, when the interests of wholes clash with the interests of individuals, the interests of individuals must be sacrificed. However, if humans cannot be sacrificed for the good of the whole, why can rabbits, deer, and wolves? Environmental holists reply by claiming that while the biotic community matters morally, it is not the only community that matters. We humans are part of various “nested” human communities, all of which have claims upon us; we are part of a tight-knit human community, but only a very loose ­human– rabbit community. Thus our obligations to the biotic community may require the culling of rabbits, but may not require the culling of humans (Callicott, 1999). But the environmentalist reply will not suffice. It would seem now that some relations within the biotic community carry more moral weight than others, an implication derived not from Leopold’s Maxim, but from the point of view of individual human members of a given biotic community. Yet, when decisions regarding the content and strength of our various community attachments and commitments are left up to individual human members of a given biotic community themselves, the door to sanctioning diverse and repugnant moral obligations opens. For example, if an individual believes that he has a much stronger community attachment and commitment to White men than Black men, does this mean that he can legitimately favor the interests of the former over the latter? If our moral commitments to the biotic community are trumped by our obligations to the human community, and if other members of the biotic community are merely instrumentally valuable, then environmental holism collapses to the anthropocentric view. And if that's the case, then animal liberation never really was a "triangular affair." Unacceptable implications of environmental holism such as these should give one pause before rejecting animal liberation as too individualistic. At a societal level, recent environmental and food movements trace their antecedents

122

Rethinking Food and Agriculture

to the philosophical foundations of anthropocentric environmental holism. Foodie and locavore culture presents a most striking and emblematic example.

6.5 Foodieism and locavorism: A celebration of “humane” exploitation The 2011 Sydney Writers’ Festival featured late British food critic A.A. Gill and late American celebrity chef Anthony Bourdain in conversation with restaurateur Tony Bilson. A significant part of that discussion involves the trio criticizing the animal rights movement (Food Fighters: AA Gill and Anthony Bourdain in conversation—YouTube, 2011). Decrying animals rights as a “false morality,” Bilson criticizes the movement for its concern for animals at a time when countless human beings are suffering (apparently he cannot imagine a person being both an animal and human rights advocate). Bilson’s comment prompts the following exchange: Gill: Well, I don’t know if it’s a false morality, I just don’t agree with it. I also don’t really care if animals suffer. If I’m perfectly honest, I don’t give a shit! Bourdain: (laughs) I’d rather not see it. Gill: Once you’ve heard one pig scream, the second one’s easier. Bourdain: And he’s right…you learn something about yourself when you kill a pig!

Putting aside such arrogance and condescension, I’d like instead to focus on their attitude, a certain indifference to the suffering and death of animals slaughtered for food. Not only their indifference, but their giddy, mocking behavior. It is that a­ ttitude—an attitude not necessarily emblematic of foodie culture, but not entirely foreign to it either—that I’d like to discuss here. In researching for this chapter, I found myself struggling to formulate a clear and precise definition of the term “foodie,” while simultaneously grappling to identify those most morally debased characteristics of foodie culture. However, as Wittgenstein (1953) taught us, trying to identify the necessary and sufficient conditions for the proper application of a given term is a fool’s errand. Nevertheless, I did exert some effort in trying to be clear on those aspects of foodie culture—specifically those related to animals as food—that I find most morally objectionable. To that end, I devised the following rough taxonomy. I formulated four, nonexhaustive, loose distinctions I refer to as moral-belief states in relation to the consumption of animal products and the plight of animals used as food. I employ (as philosophers are wont to do) a term of art in describing these four states, specifically, the term akrasia, from the Greek meaning a weakness of the will; acting contrary to one's moral values. The four moral-belief states I wish to discuss in relation to foodie culture and so-called “humane meat” are as follows: 1. Nonakratic ignorance. Individuals in this state are oblivious to the moral issues surrounding the suffering and death of animals for food. This person has never seriously questioned the morality surrounding the production and consumption of animal products, including their flesh, bodily secretions, and zygote-containing roe. Though a person of interest, this person is not my focus here.

Animal ethics

123

2. Akratic nonignorance. This is a rather common moral-belief state. Though individuals in this moral-belief state believe the production and consumption of (at least some) animal products is morally problematic, they nevertheless suffer akrasia—weakness of will—and therefore continue to consume animal products. As with the first moral-belief state, this person is not my focus. 3. Nonakratic nonignorance. Individuals in this moral-belief state believe that the production and consumption of animal products is not morally problematic, thus they suffer no akrasia or weakness of the will. This person accepts the suffering of animals as an unfortunate consequence of production and consumption. They believe the production and consumption of animal products is not morally problematic usually for at least one of four reasons (known as the “4Ns”): namely, that the consumption of animal products is (1) normal, (2) natural, (3) necessary, and/or (4) nice (Piazza et al., 2015). This, too, is a common moral-belief state. Persons in this state recognize that animals suffer and are killed in food production. While they do not think that suffering and death are necessarily good things, they do accept the suffering as a necessary part of food production and perhaps even part of the “cycle of life.” People in this moral-belief state are often opposed to “factory farms” and industrialized food production methods. They may even express this opposition by purchasing only locally produced, artisanal, “humane” animal products. This category includes people like Michael Pollan and many people who self-identify as foodies.

It’s important to acknowledge that the focus of my analysis here is on affluent Western foodie culture, and also to address briefly a critique of ethical veganism as a form of Eurocentric colonialism (Schlanger, 2018). In a nutshell, the critique goes as follows: for many poor rural communities particularly in the Global South where environmental resources are limited, animals function as an exploitable resource—indispensable as food, forms of economic security, capital, labor, etc.—that can raise standards of living and improve qualities of life. To demand of these people a commitment to ethical veganism is yet another act of colonial and cultural imperialism. While this criticism has some force, if animal ethicists are correct and the use of nonhuman animals by humans is a severe violation of interests and rights, then these differences form a legitimate case of conflicting values surrounding moral, cultural, and economic justice. Resolving such conflicts requires more than simplistic dichotomal thinking. Gruen (2001), for example, provides a sophisticated, nuanced analysis that teases out the moral complexities of such divergent conflicting values, while offering a systematic proposal for adjudicating such conflicts. Though a robust exploration of this issue is beyond the scope of this chapter, it’s safe to say that attempting to shut down discussion of veganism as morally obligatory by labeling it “colonialist” ignores the subtleties involved in such conflicts of values. Before moving onto my fourth and final category of moral-belief state in relation to the production and consumption of animal products and the plight of animals used as food, I’d like to throw out a rough-and-ready definition of “foodie.” A foodie is a kind of gourmet, a gourmand, a person who purports to have an ardent or refined interest in food, who seeks new food experiences as a kind of hobby rather than simply eating out of convenience or hunger. I’ll have more to say about this, but for now, I would like to turn to the fourth moral-belief state. 4. Sadistic nonakratic nonignorance. Like the third moral-belief state, the fourth moral-belief state involves nonakratic nonignorance, but with a twist I call sadistic nonakratic nonignorance. Like persons in the nonakratic nonignorance moral-belief state, folks in the sadistic nonakratic ­nonignorance state believe that the production and consumption of animal products is not

124

Rethinking Food and Agriculture

­ orally problematic. However, the crucial moral difference between those in the nonakratic m nonignorance moral-belief state and those in the sadistic nonakratic nonignorance moral-belief state is that persons in the latter moral-belief state reject that the suffering of animals is unfortunate. This rejection can manifest itself in myriad ways, including (1) indifference to the suffering and killing of “food animals,” (2) the mocking of the animals and their suffering and killing, and (3) in some cases, even a celebration of the suffering and killing of animals. In these cases, the knowledge of the animals’ suffering can actually add to the exotic, hedonistic, debauched dining pleasure. For these sadistic foodies, taste preference and palate satisfaction trump all competing considerations, including issues of animal suffering or even animal welfare.

For sadistic foodies, the pursuit of new food experiences is best interpreted as an artifact of affluence, and in many ways is ultimately about power, prestige, and privilege: power over the animal, the food producers, the servers, etc.; prestige with other foodies and aspiring foodies; and the privilege to afford to seek out novel and exotic food and drink. Sadistic foodies are absolved from all moral culpability since moral culpability is hidden behind the fact that these practices are culturally, socially, and legally sanctioned, encouraged, and even aspired to. In light of this discussion, it should be clear that folks like A.A. Gill and Anthony Bourdain are paradigm cases of sadistic foodies. For these reasons, sadistic foodie culture is particularly morally debased especially regarding the plight of animals used as food. Let me provide some other examples of sadistic foodies. Describing her experience cooking lobster, Julie Powell, author of the best-selling book Julie & Julia: 365 Days, 524 Recipes, 1 Tiny Apartment Kitchen, writes: Over a period of two weeks…I went on a murderous rampage. I committed gruesome, atrocious acts…If news of the carnage was not widely remarked upon in the local press, it was only because my victims were not Catholic schoolgirls or Filipino nurses, but crustaceans. This distinction means that I am not a murderer in the legal sense. But I have blood on my hands, even if it is the clear blood of lobsters. People say lobsters make a terrible racket in the pot, trying-reasonably enough-to claw their way out of the water. I wouldn’t know. I spent the next twenty minutes watching a golf game on the TV with the volume turned up…When I ventured back into the kitchen, the lobsters were very red, and not making any racket at all…Poor little beasties. (Powell, 2005, p. 194)

Commenting on these passages, B.R. Myers notes in The Atlantic: This is a prime example of foodies’ hostility to the very language of moral values. In mocking and debasing it, they exert, with Madison Avenue’s help, a baleful influence on American English as a whole. If words like “sinful” and “decadent” are now just a cutesy way of saying “delicious but fattening,” so that any serious use of them marks the speaker as a crank, and if it is more acceptable to talk of the “evils of gluten” than of the “evils of gluttony,” much of the blame must be laid at their doorstep. (Myers, 2007)

The indifference to animal suffering here is exacerbated by the mockery and sarcasm of the sadistic foodie.

Animal ethics

125

Other examples of sadistic foodie culture abound. In a 2015 article from The Guardian we learn that: Noma’s Japanese restaurant serves up a rare treat…The world’s best restaurant has opened a pop-up in Tokyo and its still-twitching, slightly gruesome menu, has critics salivating…[T]he celebrated chef behind Noma, Rene Redzepi, has upped the sushi and seasoning stakes with a creation featuring live jumbo prawns, topped with tiny black ants. At Noma Tokyo, perched on the 37th floor of the Mandarin Oriental hotel with views of Mount Fuji in the distance, the presence of half a dozen ants clinging to the wobbling flesh of each prawn is more than just a visual gimmick. With their natural reserves of formic acid, the ants give the botan ebi—or botan prawn—a sour kick. (McCurry, 2015)

In her 2014 best-selling book Anything That Moves: Renegade Chefs, Fearless Eaters, and the Making of a New American Food Culture, journalist and foodie Dana Goodyear chronicles (sadistic) foodie culture, writing: “It’s not Bacchanalian, it’s Caligulan!” the woman to my left exclaimed one night at Totoraku, an invitation-only, all-beef restaurant in Los Angeles, as course after course of raw beef came to the table. She was a member of a dining group that calls itself the Hedonists. On my right, another Hedonist, a Totoraku regular who had invited me along, was photographing each dish with a macrolens and macroflash. I felt obliged to gulp down as much raw beef throat as I could, and made sure that I was seen doing it. (Goodyear, 2014, p. 15)

As research for this chapter, I conducted an interview with Elsa Newman, a server from the exclusive Plumed Horse restaurant, a high-end, French foodie favorite in Silicon Valley. In the course of our discussion, Newman provided keen insight into the precise phenomenon that I am getting at here in sharing her views on and experiences with foodie culture: Foodieism is really a way for foodies to talk about money. It’s a disguise, a lead-in for braggadocio. They don’t talk so much about the food as much as they talk about their travels and material possessions…We offer two different kinds of caviar here. One is produced by rubbing the fish mother’s belly rather than cutting it open. That costs $200 an ounce versus $90 an ounce for the run-of-the mill caviar. When told that the difference in price is due to the fact that the belly-rubbing caviar is more humane in that it doesn’t hurt the mother, customers are turned off by this and order the eggs from the slaughtered fish. But when you tell them that the $200-an-ounce caviar has slight and unique accents of cucumber, customers fork over the $200 an ounce without hesitation. (Newman, 2015)

To reiterate, sadistic foodie culture is about more than food. It’s about intent; it’s an expression of cultural capital, economic power, power over the supply chain that

126

Rethinking Food and Agriculture

must come together to make “exceptional, special dishes.” Unlike other kinds of foodieism, it is indifferent to and even relishes in the suffering and killing of the animals it requires. And it is morally debased. Of course, a foodie might respond to these aspects of foodie culture by pointing out that sadistic foodies are a small, elite, nonrepresentative segment of foodie culture. Most foodies are of the nonakratic nonignorance type (type (3)) who, though not indifferent to the suffering and killing of food animals, nevertheless do not see the production and consumption of animal products as morally problematic. That said, I can only wonder, how many nonsadistic foodies are in reality aspiring sadistic foodies? For those that are, it would seem that the main difference between nonsadistic foodies and sadistic foodies is nothing more than wealth and access. And that’s morally troubling. And even foodies of the Michael Pollan sort—emblematic of a popular kind of foodieism—are themselves accompanied by their own troubling moral consequences. An increased awareness of the destructive nature of industrialized animal agriculture and fishing, including environmental degradation, individual and public health threats, and the atrocious conditions under which animals are raised, has led to a shift in attitudes toward meat and meat production. This acknowledgment, coupled with a sentimental nostalgia for a time when a majority of Europeans and Americans were farmers and craftspersons, has led to a booming alternative food movement. Known as locavorism, compassionate carnivorism, the sustainable meat movement, the humane meat movement, the happy meat movement, the nose-to-tail food movement, and the conscientious omnivore movement, this movement markets itself as “free range,” “grass fed,” “organic,” “natural,” and “cage free.” For those who desire to consume animal products but are ethically troubled by industrialized animal agriculture, so-called “happy” meat, eggs, and dairy purport to offer an ethical alternative both to veganism (abstaining from the exploitation, instrumental use, and consumption of animals and their products) and to the cruelty of the factory farm, ensuring happier lives (and “humane deaths”) for animals destined to become food. Measured against the vast majority of consumers whose lack of connectedness to their food enables the near-total erasure of suffering from their plates in the form of neatly shrink-wrapped, bloodless cuts of meat, so-called “compassionate carnivore” foodies deserve praise. Yet, despite this supposed concern for the animals’ lives and deaths, relatively little public attention has been paid to the experiences of their short lives or the brutality of their slaughter. In truth, an overwhelming majority of animals raised on “local” farms are sent to industrial slaughterhouses, butchered alongside their kin raised in factory farms. Animals raised in “humane” conditions routinely suffer branding, dehorning, forced impregnation, tail docking (without anesthesia), overcrowding, beak trimming, castration, tooth filing, ear notching, and nose ring piercing (Bohanec, 2013; Stănescu, 2010). In “How happy is your meat?: Confronting (dis)connectedness in the ‘Alternative’ meat industry,” geographer Kathryn Gillespie analyzes the tension between the desire for DIY butchers to forge a connection to their food by involving themselves in every step of its production (including slaughter), and the Herculean efforts they make to

Animal ethics

127

disconnect themselves from the actual animal they will butcher to avoid a sentimental or emotional attachment to the hapless subject. For many “compassionate carnivores,” the killing and eating of animals is justified by their interest in forming a consumer– food connection, where personally taking on the death of the animal acts as a means to more ethical eating, a way of honoring the subjects of slaughter while eating them. Yet, as Gillespie points out, even Michael Pollan advises DIY butchers to quickly disconnect from what it means to slaughter an animal. Gillespie characterizes this most profound disconnect, the connected disconnection, in the following way: All of the justifications for DIY slaughter as a way to connect to food, to become an artisan, to embody rusticity, and to make slaughter more humane are enlisted to conceal what the process really does. DIY slaughter connects participants to the violence against the animal, and not to the animal him/herself. This “connection” is a wholly false connection. DIY slaughter denies the actual connection we have with animals. Animals are still, in DIY slaughter, conceptualized not as individual animals but as products ready to become meat. (Gillespie, 2011, p. 120)

A further problem with both “humane” and industrial agriculture is that both rely on putting animals in the category of the edible, an ontologizing of animals as food who are made absent referents, stripped of subjectivity, individual personality, interests, and desires (including the desire not to be harmed) (Adams, 1990; Gruen, 2011; Sanbonmatsu, 2018; Vialles & Noilie, 1994). Intimately connected to foodie culture is one popular justification for eating animal flesh and secretions, namely, the claim that eating meat is a personal choice. A common response to the suggestion that nonsadistic foodies ought to go vegan goes something like this: “Though being vegan is fine for some people, vegans shouldn’t try to force their views on others. Eating meat is my personal choice.” However, though meat eating is a personal choice, it is not merely a personal choice. In claiming that meat eating is a personal choice we affirm the belief that even trivial palate pleasures trump the life and suffering of a sentient being. Selectively breeding sentient beings into existence to maintain a steady supply of future meals because we see animals as commodities is not merely a “personal choice.” Furthermore, animal agribusiness is the leading single cause of water pollution, air pollution, and climate change (Gerber et  al., 2013; Mateo-Sagasta, Zadeh, Turral, & Burke, 2017) such that, collectively, the consumption of animal products does impose and externalize the costs and consequences of such “personal choices” on others (Simon, 2013; Weis, 2013). Clearly, to argue that the consumption of animal products is merely a personal choice is to ignore and overlook important moral consequences of such choices (Grillo, 2012). But perhaps foodies’ expressions of cultural capital or locavores’ aspirations to artisan Arcadianism might all be made immaterial by a technological advancement so groundbreaking that it will entirely obviate the animal from the food production cycle. Though such an innovation may displease both foodies and locavores, would not advocates for animal liberation and ethical vegans be well advised to embrace such a development?

128

Rethinking Food and Agriculture

6.6 Will “clean meat” end animal agriculture? Produced in a laboratory using a complex bioengineering process, “clean meat” (known also as “cultured meat,” “cellular meat,” and “synthetic meat”) is the term being used to describe an emerging category of new animal meat products in which cells from a living animal are grown into muscle intended to approximate the flesh of actual animals. Proponents of clean meat—an alliance of Big Ag, venture capitalists, and even animal welfare advocates—proclaim it as the panacea for both the horrors of and environmental devastation caused by the animal agriculture industry, predicting not only a short-term reduction in the number of animals killed for food, but an end to animal agriculture as we know it within decades. In reality, not only is the clean meat movement already damaging, but there are grave reasons for rejecting such prognostications. First, the clean meat movement obscures the most direct and simplest way to create a sustainable, ethical food system, namely, a plant-based food system based on sustainable agroecological production. Furthermore, even if the clean meat phenomenon booms, the number of animals now slaughtered for food will not be reduced. A 2019 report by AT Kearney consultants predicts that, in 20  years, most protein products will consist of clean meat or plant-based meat alternatives. However, coincident with this prediction is the fact that the global meat market is expected to double in the next 20 years. Thus the overall number of animals killed will remain unchanged (AT Kearney Consultants, 2019). Furthermore, the development and marketing of clean meat technologies ensure that the clean meat industry works not as a replacement for but in concert with the conventional meat industry. Multinational agribusiness giants such as Cargill, Tyson, and Perdue invest heavily in clean meat technologies, not with an eye to transitioning away from or even reducing their factory farming operations, but rather to expanding and diversifying their “protein portfolios.” Cargill assures stockholders that while they invest in new “innovations” like cultured meats, they remain committed to “investing in, and growing” their “traditional protein” businesses, as these are “both necessary to meet [consumer] demand” (Cargill, 2017). In an effort to dominate both the clean meat/vegan and “traditional protein” markets, industrial agriculture giants are busy buying up vegan enterprises. For example, over the past 5  years, vegan food producers such as White Wave, Field Roast, Lightlife, and Daiya have been purchased by meat or pharma multinationals. These takeovers represent a kind of “catch and kill” strategy in which meat conglomerates acquire smaller vegan start-ups in an effort to control production and marketing. Despite these disturbing trends, too many animal rights advocates proclaim that the solution to the atrocity that is animal agriculture can best be found in the capitalist free market vis-à-vis clean meat, the very same system responsible for the development and implementation of massive and systemic industrial mechanisms of slaughter, colonialism, and environmental destruction (Nibert, 2013). How imprudent, how ill-considered, how treacherous an idea that a planet-annihilating capitalist free enterprise system—itself, to paraphrase Pythagoras, a ruthless destroyer of animal life—should be entrusted with solving such a colossal instance of systemic

Animal ethics

129

injustice and structural oppression. Perhaps I am being cynical; too skeptical. Is my skepticism warranted? Clearly, the philosophical foundations for establishing robust moral entitlements for nonhuman animals are sound, entitlements that make otherthan-human animals proper and legitimate subjects of justice. From the fact that nonhuman animals suffer systemic and institutional domination and oppression, it follows that animal liberation is, indeed, a social justice issue (Jones, 2015; Tuttle, 2014). But show me one social justice movement that was solved by the free market. Clean meat is a pie-in-the-sky market solution to an atrocity that requires not merely an alteration in supply–demand curves, but rather a titanic shift in the moral vision and behavior of humanity.

6.7 Redefining speciesism Though the practice of animal-based food production is at the heart of what Crist (2019) calls our “ecological demolition,” it is our values that have driven us to this place, in particular, human supremacy and speciesism. While human supremacy is a kind of unjustified privileging of human interests over those of all other nonhuman existence, speciesism has manifold forms. As we’ve seen, Peter Singer sees speciesism as a kind of belief state that places all humans morally above all nonhuman animals. However, paralleling the theoretical expansion and application of the concept of racism by sociologists in the 1960s and 1970s (Nibert, 2002), the concept of speciesism has been greatly broadened since the publication of Singer’s work. Ecofeminists criticize this kind of individualist interpretation of speciesism as blind to an explanatorily powerful feminist analysis, one that highlights the interconnectedness of racism, sexism, classism, and speciesism (Adams, 1990; Cantor, 1983; Gaard, 2002; Gruen, 2009, 2015; Kheel, 1988). On this view, speciesism is best understood as a manifestation of a deeper, pernicious ideology of domination and violence rooted in patriarchy. Adams (1990), for example, argues that patriarchy makes invisible both animals and women qua subjects, making them instead absent referents to be objectified, fragmented, and consumed. Awareness of patriarchy as the root of human supremacy—in particular, speciesism—moves one to resistance of such forces and toward a restoration of the absent referent to her proper status as subject. Similarly, Nibert (2002) argues that casting speciesism as a kind of individual prejudice obscures a more powerful, Marxist analysis of speciesism, revealing deeper social, structural, and economic origins of animal oppression. Both feminists and Marxists see speciesism as an ideology that acts to legitimate the structural oppression and exploitation of other animals, while benefiting the power structures and economic interests of capitalists, oligarchs, and the patriarchal power elite. Speciesism reveals itself as a complex set of shared beliefs—imbedded at the systemic, structural, and institutional levels through patriarchy and forces of class oppression—that inspire individual prejudice and discrimination. For Marxists, conceiving of speciesism as a form of prejudice among individuals who then create and legitimate oppressive structures, gets the causal story backward. On an ideological analysis, individual biases and prejudices are caused by the formation and implementation of the ideology.

130

Rethinking Food and Agriculture

By contrast, John Sanbonmatsu finds both the “liberal” view of Singer and the “radical” views of Marxists and feminists to be fatally impoverished (Sanbonmatsu, 2011). Against the liberal view, Sanbonmatsu argues that if speciesism were merely a set of mistaken, ill-informed beliefs, born of ignorance, then speciesism could be overcome by exposing people to facts and to better arguments. Despite the fact that this knowledge is both pervasive in the literature on animal ethics and promulgated throughout the general populace, human supremacy and speciesism remain ubiquitous. For Sanbonmatsu, the liberal model fails to account for the “sheer obstinacy” of the prejudice, which leads to the question: Why do human cultures “choose to cling to prejudicial and irrational beliefs and attitudes in the face of alternative beliefs that are logically and morally superior” (Sanbonmatsu, 2011, p. 30)? Sanbonmatsu finds the radical analysis of speciesism inadequate as well, for “if speciesism were merely the expression of the interests of a dominant class, then why does it enjoy virtually universal appeal across class, racial, national, cultural, and gender divides?” (Sanbonmatsu, 2011, p. 30). Not only elites, but “virtually all human beings in all walks of life and social positions, rich and poor, men and women, First Worlders and indigenous tribes” participate in and profit from the speciesist system. What’s missing from the radical critique is a recognition that we human beings, ourselves, constitute a dominant class (Sanbonmatsu, 2011, p. 30). Sanbonmatsu, instead, sees speciesism as a mode of production, a way of producing human life in which the bodies and minds of other beings are treated as objects for human appropriation, exploitation, and extermination. So deeply bound to the core of what it means to be human, speciesism constitutes a fundamental existential structure of human life. Borrowing from Sartre’s analysis of anti-Semitism, Sanbonmatsu argues persuasively that speciesism operates as a universal form of bad faith, a mode of self-deception—manifest as a potent social structure—that constitutes a total way of being in the world. In our refusing to refrain from violence toward other sentient beings, we both alienate ourselves from our own embodied being as animals and estrange ourselves from our own humanity. Other conceptions of speciesism include that of indigenous Cree scholar BillyRay Belcourt who sees speciesism as a vestige of colonialism, centering domesticated animal bodies as colonial subjects (Belcourt, 2015), while others like Heitzeg (2012) and Jones (2013) see speciesism as the foundational form of oppression where the relationship between humans and nature is shifted from kinship to dominion, and subsequently to domination of other humans according to class, race, gender, etc. (see also Chapter 2).

6.8 Conclusion I remain dubious that undermining systemic speciesism is possible. Yet the suffering and killing of billions of our fellow earthlings annually morally obligates us to try. Liberally paraphrasing Sartre, we do not fight speciesism because we think we’re going to win; we fight speciesism because it’s speciesist. It is imperative that we as

Animal ethics

131

individuals, at the very least, commit ourselves to vegan practice. Though capitalist markets are ineffective moral vehicles in social justice advocacy, adopting veganism from a political stance, one that involves consumer choice and effective activism, may move the moral needle away from speciesism and toward global justice (Jenkins & Stănescu, 2014; Jones, 2016; McMullen & Halteman, 2019). Educating ourselves about the role that human supremacy and speciesism play in biodiversity loss and the destruction of planetary life while raising awareness and forcing the issue into the public consciousness with an end to shifting public opinion through strategic nonviolent direct action is also required (Crist, 2019; Engler & Engler, 2016). Yet, as I hope to have made clear, we do not suffer from a lack of knowledge, or a kind of epistemological blindness. Rather, we suffer from moral bad faith. Transforming and rethinking animal agriculture does not require better science, innovative research methodologies, or conceptual arguments. That enterprise requires a kind of moral transcendence, a clear-eyed forsaking of our moral bad faith and the hubris of our unfounded human supremacy.

References Adams, C. (1990). The sexual politics of meat. New York: Crossroad Continuum. Animal Advocacy by Numbers. (2016, July 16). Faunalytics. https://faunalytics.org/ animal-advocacy-by-numbers/. AT Kearney Consultants. (2019). How will cultured meat and meat alternatives disrupt the agricultural and food industry?. Belcourt, B.-R. (2015). Animal bodies, colonial subjects: (Re)locating animality in decolonial thought. Societies, 5(1), 1–11. doi:https://doi.org/10.3390/soc5010001. Bland, A. (2012). Is the livestock industry destroying the planet?. Smithsonian Magazine. Bohanec, H. (2013). The ultimate betrayal: Is there happy meat?. iUniverse. Callicott, J. B. (1980). Animal liberation: A triangular affair. Environmental Ethics, 2(4), 311–338. Callicott, J. B. (1999). Holistic environmental ethics and the problem of ecofascism. In Beyond the land ethic: More essays in environmental philosophy (pp. 59–76). SUNY Press. Cantor, A. (1983). The club, the yoke, and the leash: What we can learn from the way a culture treats animals. Ms. Magazine[ August], (pp. 27–29). Cargill. (2017, August 23). Protein innovation: Cargill invests in cultured meats. https://www. cargill.com/story/protein-innovation-cargill-invests-in-cultured-meats. Cochrane, A. (2012). Animal rights without liberation: Applied ethics and human obligations. Columbia University Press. Crist, E. (2019). Abundant earth: Toward an ecological civilization. University of Chicago Press. Curtin, C. (2007). Do living people outnumber the dead?. doi:https://doi.org/10.1038/ scientificamerican0907-126. DeGrazia, D. (2008). Moral status as a matter of degree? The Southern Journal of Philosophy, 46(2), 181–198. doi:https://doi.org/10.1111/j.2041-6962.2008.tb00075.x. Engler, M., & Engler, P. (2016). This is an uprising: How nonviolent revolt is shaping the ­twenty-first century. Bold Type Books. Food Fighters: AA Gill and Anthony Bourdain in conversation—YouTube. (2011, February). https://www.youtube.com/watch?v=F2tjvTmbEVw&t=21s.

132

Rethinking Food and Agriculture

Francione, G. L. (1996). Animals as property. Animal L., 2, i. Gaard, G. (2002). Vegetarian ecofeminism: A review essay. Frontiers: A Journal of Women Studies, 23(3), 117–146. Garner, R. (2013). A theory of justice for animals: Animal rights in a nonideal world. Oxford University Press. Gerber, P. J., Steinfeld, H., Henderson, B., Mottet, A., Opio, C., Dijkman, J., … Tempio, G. (2013). Tackling climate change through livestock: A global assessment of emissions and mitigation opportunities. Food and Agriculture Organization of the United Nations (FAO). Gillespie, K. (2011). How happy is your meat? Confronting (dis)connectedness in the ‘alternative’ meat industry. The Brooking Review, 12(1), 100–128. Goodyear, D. (2014). Anything that moves: Renegade chefs, fearless eaters, and the making of a new American food culture. Penguin. Grillo, R. (2012, August 15). Eating animals and the illusion of personal choice. Free From Harm. https://freefromharm.org/animal-products-and-psychology/ five-reasons-why-meat-eating-cannot-be-considered-a-personal-choice/. Gruen, L. (2001). Conflicting values in a conflicted world: Ecofeminism and multicultural environmental ethics (pp. 16–18). We International. Gruen, L. (2009). Attending to nature: Empathetic engagement with the more than human world. Ethics & The Environment, 14(2), 23–38. Gruen, L. (2011). Ethics and animals: An introduction. Cambridge University Press. Gruen, L. (2015). Entangled empathy: An alternative ethic for our relationships with animals. Lantern Books. Heitzeg, N. (2012, July 18). CI: Commodities and Cages. https://www.printfriendly.com/p/g/ gaGAnM. Jenkins, S., & Stănescu, V. (2014). One struggle. Counterpoints, 448, 74–85. Jones, P. (2013, October 11). Intersectionality and animals. http://blog.bravebirds.org/ archives/1553. Jones, R. C. (2015). Animal rights is a social justice issue. Contemporary Justice Review, 18(4), 467–482. Jones, R. C. (2016). Veganisms. In Critical perspectives on veganism (pp. 15–39): Springer. Kheel, M. (1988). Animal liberation is a feminist issue. In 10. New catalyst quarterly (pp. 8–9). Ko, A. (2019). Racism as zoological witchcraft: A guide to getting out. Lantern Books. Leopold, A. (1949). A Sand County almanac. Oxford University Press. Lowe, D. (2014, March 20). Speciesism. 1000-word philosophy: An introductory anthology. https://1000wordphilosophy.com/2014/03/20/speciesism/. Mateo-Sagasta, J., Zadeh, S. M., Turral, H., & Burke, J. (2017). Water pollution from agriculture: A global review. Rome: Food and Agriculture Organization of the United Nations and the International Water Management Institute. McCurry, J. (2015, January 29). Live shrimp covered in ants, anyone? Noma’s Japanese restaurant serves up a rare treat. The Guardan. https://www.theguardian.com/world/2015/ jan/29/twitching-live-shrimp-covered-ants-dish-noma-restaurant-japan. McMahan, J. (2002). The ethics of killing: Problems at the margins of life. USA: Oxford University Press. McMullen, S., & Halteman, M. C. (2019). Against inefficacy objections: The real economic impact of individual consumer choices on animal agriculture. Food Ethics, 2(2–3), 93–110. Myers, B. R. (2007, September 1). Hard to swallow. The Atlantic. https://www.theatlantic.com/ magazine/archive/2007/09/hard-to-swallow/306123/. Newman, E. (2015). Telephone interview. March 5.

Animal ethics

133

Nibert, D. (2002). Animal rights/human rights: Entanglements of oppression and liberation. Rowman & Littlefield Publishers. Nibert, D. (2013). Animal oppression and human violence: Domesecration, capitalism, and global conflict. Columbia University Press. Piazza, J., Ruby, M. B., Loughnan, S., Luong, M., Kulik, J., Watkins, H. M., & Seigerman, M. (2015). Rationalizing meat consumption. The 4Ns. Appetite, 91, 114–128. doi:https://doi. org/10.1016/j.appet.2015.04.011. Pluhar, E. B. (1995). Beyond prejudice: The moral significance of human and nonhuman animals. Duke University Press. Powell, J. (2005). Julie and Julia: 365 days, 524 recipes, 1 tiny apartment kitchen. Little, Brown. Rachels, J. (1990). Created from animals. In The moral implications of darwinism. Oxford University Press. Regan, T. (1983). The case for animal rights. University of California Press. Regan, T. (1985). In P. Singer (Ed.), The case for animal rights in defense of animals. New York: Harper & Row Publishers. Rollin, B. E. (1992). Animal rights and human morality. Prometheus Books. Sanbonmatsu, J. (2011). The animal of bad faith: Speciesism as an existential project. In Critical theory and animal liberation (pp. 29–45). Rowman & Littlefield Publishers. Sanbonmatsu, J. (2018, March 10). Lady Macbeth at the Rotisserie: ‘Femivores,’ Violence, and the new maternalism in animal agriculture. https://www.youtube.com/ watch?v=hFRT4W6KHPk. Sapontzis, S. F. (1989). Morals, reasons, and animals. Temple University Press. Schlanger, Z. (2018, June 23). Is promoting vegetarianism a form of colonialism?. Quartz. https://qz.com/1311884/is-promoting-vegetarianism-a-form-of-colonialism/. Simon, D. R. (2013). Meatonomics: How the rigged economics of meat and dairy make you consume too much–and how to eat better, live longer, and spend smarter. Conari Press. Singer, P. (1975). Animal liberation: A new ethics for our treatment of animals. Harper Collins. Singer, P. (2011). Practical ethics. Cambridge University Press. Stănescu, V. (2010). Green’eggs and ham? The myth of sustainable meat and the danger of the local. Journal for Critical Animal Studies, 8(1–2), 8–32. Taylor, S. (2017). Beasts of burden: Animal and disability liberation. The New Press. Tuttle, W. M. (2014). Circles of compassion: Essays connecting issues of justice. Vegan Publishers. Vialles, N., & Noilie, V. (1994). Animal to edible. Cambridge University Press. Weis, T. (2013). The ecological hoofprint: The global burden of industrial livestock. Zed Books Ltd. Wittgenstein, L. (1953). Philosophical investigations. Blackwell Publishers. Worm, B., Barbier, E. B., Beaumont, N., Duffy, J. E., Folke, C., Halpern, B. S., … Watson, R. (2006). Impacts of biodiversity loss on ocean ecosystem services. Science, 314(5800), 787–790. https://doi.org/10.1126/science.1132294.

This page intentionally left blank

A food system fit for the future Tony Junipera,b,* a Natural England, London, United Kingdom, bInstitute for Sustainability Leadership, University of Cambridge, Cambridge, United Kingdom

7

7.1 Conversion, intensification, and degradation Perhaps the most fundamental pressure that arises from how we feed ourselves is seen in the process of land conversion, when natural and seminatural habitats are transformed into fields and pastures. In many parts of the world this process continues apace, especially where the remaining tropical forests and savannahs meet agricultural frontiers, where each year some 120,000 km2 of forest is mainly being replaced with different kinds of food-producing environments (WRI, 2019). The greenhouse gas emissions arising from tropical deforestation are estimated to be greater than those of the European Union and nearly equivalent to those of the United States (WRI, 2018). The fires raging across the Amazon basin in Brazil provide a recent dramatic case in point, where clearance for cattle pasture and soya production are the main ultimate drivers of forest loss. After land is cleared, in some landscapes there has been a tendency over time for farming to become ever more industrialized, with large-scale monocultures leading to further pressures being placed on any wildlife populations that survived the initial conversion of natural ecosystems to farmed landscapes. This is the principal reason why an estimated one million species are presently being driven toward the final oblivion of extinction (IPBES, 2019). Indeed, if one looks at the pressures causing wildlife declines across the world, about 60% of the loss of wildlife can be explained in relation to what we eat and how we produce it (Barrett et al., 2018). The United Kingdom is today regarded as one of the most nature-depleted countries in the world. More than half of the thousands of species examined in a 2016 study were found to be in long-term decline and one in seven deemed to be at risk of extinction in the country (RSPB et al., 2016). The main reason for this is agriculture, and especially the intensification of farming methods that began in earnest during the second half of the 20th century. The continuing decline of wildlife in Britain and much of Europe, long after the initial clearance of land to make way for fields, is down to the heavy use of pesticides, massive fertilizer application, larger-scale production of single crops, and loss of the remaining seminatural features, such as hedges, grasslands, and wetlands caused by the progressive intensification of farming. It is not only wildlife that has been impacted by the rise of industrial farming. Agricultural practices often lead to further environmental degradation, including damage to soils. Some 24 billion tonnes of topsoil leaves fields each year via air and water, * Dr Tony Juniper CBE is an environmentalist and writer. He is the former director of Friends of the Earth and is the Chair of Natural England. Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00007-4 © 2021 Elsevier Inc. All rights reserved.

136

Rethinking Food and Agriculture

much traveling along rivers, to be released into the oceans (UNCCD, 2017). As well as loss of soil from fields, these vital and complex ecological systems are also subject to compaction by machinery and hooves, salinization, and also to a loss of organic matter. Soil organic matter is comprised of living (microbes, worms, and fungi, among other things) and once-living (decaying and accumulated dead vegetation, such as is found in peatland soils) organisms, and is vital for healthy soil function, including the cycling of nutrients that enable plant growth. In addition to the loss of the forests that once grew in the soils before they were cleared of trees, the loss of organic matter from soils is causing large-scale carbon dioxide emissions (Lal, 2014). The more this carbon-rich fraction of the soil is depleted by farming, the more carbon dioxide there is in the atmosphere. The higher the original organic matter in soil, the bigger the subsequent carbon emissions that can result from some kinds of farming practices, with the biggest emissions generally associated with the cultivation of peat soils. Globally, between a quarter and a third of greenhouse gas emissions are estimated to arise from land use, including the effects of deforestation and depletion of soil organic matter (IPCC, 2019). The loss of soil organic matter is not only contributing to climate change, but also leading to soils being less resilient to the changes taking place. For example, soils with a high level of organic matter hold more water than soils where it has been depleted. Crops growing in healthier soils are thus more likely to withstand drought than those planted in more degraded soils. Also, urban and other areas adjacent to degraded areas of soils can be more susceptible to flood risk as the land tends to shed water rather than hold it. The scale of impact of farming on soil health is vast. Indeed, we’ve reached the point today where a third of farm soil is already acutely degraded and that process continues. Many tens of millions of hectares of soil have been depleted in the wake of unsustainable farming, and as degraded land is spat out from the rear end of ecologically damaging farming, so new land is being opened at its rapacious leading edge, carving swathes of new fields and pastures from forests and natural grasslands, much of it to be used in similar ways and with similar results to previously cleared land. What is perhaps most indicative of our present direction of travel is how since 2000 the area of farmed land in the world has remained about stable, despite the massive scale of land conversion that has continued since then (Alexandratos & Bruinsma, 2012). This simple fact reveals how we have embarked on a process of mining out soil health, wrecking once healthy land with unsustainable farming, before moving on to new soil. The recent fires burning across the Amazon are indicative of the trend, with more land being cleared in a region that has tens of millions of hectares of land already recently degraded by unsustainable farming (Montgomery, 2012). And while on the fundamentals of the land, let us not forget the role of the oceans in feeding people. We take some 90 million tonnes of wild seafood from the oceans each year (Hannesson, 2014) and the capture of that seafood causes some major ecological damage, including the destruction of seabed habitats by trawling gear and the bycatch of nontarget species, including endangered albatrosses, turtles, and dolphins. On top of this, in some geographies, are the ecological pressures arising from aquaculture, such as salmon, prawns, and shellfish, including pollution and habitat loss.

A food system fit for the future137

The marine environment is also impacted by our food system via plastic food packaging waste that is accumulating in the ocean. An estimated eight million tonnes of plastic is released into the sea each year, the vast bulk of it arriving via rivers (Jambeck et  al., 2015). The rise of plastic wrapping for consumer goods, including food and drink, linked with poor or nonexistent waste management facilities in those countries now most rapidly increasing their consumption of such packaged goods, adds to a rising burden of ecologically harmful waste. Some larger pieces are consumed directly by birds, mammals, and reptiles causing death, while smaller pieces enter food webs via filter-feeding organisms with consequences for marine ecosystems that are yet to be fully understood. In addition, the marine environment is also impacted by plastic and nylon material from discarded fishing gear, which is again related to our food system. For example, 46% of the plastic and nylon material in the Great Pacific Garbage Patch is from fishing nets (Lebreton et al., 2018).

7.2 Impacts of inputs Although much of the world’s agricultural soils are already degraded, yields of crops have, during the recent decades of progressive agricultural intensification, on average (and until recently at least) continued to increase. This has been down to the economies of scale derived from larger fields and bigger machines, but also higher yielding varieties, the use of pesticides and herbicides, irrigation, and also the application of fertilizers. Industrially produced nitrate fertilizers have been deployed on a vast scale. They do, however, have a significant greenhouse gas footprint, not least derived from the energy required for their manufacture, but additionally contribute to global warming as soil microbes take up fertilizer and release nitrous oxide, a very powerful greenhouse gas, as a by-product of their metabolism (GRAIN, 2015). It is, however, the progressive enrichment of the environment caused by the use of artificial fertilizers escaping from fields that is causing the most profound immediate impact. The build-up of plant-available nitrogen and phosphorus in the environment marks a massive change in fundamental conditions. Indeed, while the carbon dioxide concentration has increased in the atmosphere by about 40% since preindustrial times, the quantity of reactive nitrogen circulating on Earth has increased by around 100% (Fowler et al., 2013). That doubling of a vital plant nutrient is causing huge changes to ecosystems, and often in ways that are leading to the loss of wildlife, as, for example, rivers become choked with algae and coastal waters suffer collapses in their oxygen concentrations caused by nitrogen enrichment, creating in some cases so-called dead zones, where most wildlife has died. Nitrogen can also cause impacts on terrestrial ecosystems far away from fields via a process of dry deposition, whereby nitrogen compounds carried in the air can cause changes in grasslands and woodlands very far from where fertilizer was applied (Sutton et al., 2011). In addition to nitrogen coming from fertilizers is that arising from animal-rearing farms that release ammonia compounds from animal wastes.

138

Rethinking Food and Agriculture

Various chemical pesticides are also causing profound impacts. Many of these are often indiscriminate in their effects, including the insecticides that kill beneficial animals as well as harmful ones. Insects are of course a vital component of healthy terrestrial ecosystems, playing essential roles as pollinators but also in nutrient recycling and as food sources for many higher animals—insectivorous bats and birds included. Some pesticides are also resistant to breakdown and accumulate in food webs, occurring at vastly magnified concentrations in top predators, causing population crashes and regional extinctions (Eklöf, 2009; Isenring, 2010). This enrichment of the environment and loss of animals to toxic chemicals, including those vital for the functioning of complex food webs, is leading to an overall simplification of seminatural ecosystems, with a general tendency toward the loss of specialist animals and plants and for more common and generalist species to remain, although even these are in many cases rapidly declining. A walk through many intensively farmed landscapes in England today thus reveals few butterflies or what were once common wildflowers. There are far fewer songbirds than there were five decades ago. Often the only visible wildlife will be in the form of generalist birds such as crows and pigeons. While all these factors have taken a progressively more significant toll on ecological richness in agricultural landscapes, there has been another pressure that has been building in the background and which is now emerging as a more and more prevalent factor in determining the character of farmed areas—and that emerging new factor is climate change.

7.3 Climate change and wider food system vulnerability Climate change is not only a challenge for agriculture because of the emissions released through deforestation and soil degradation, but also because of the impacts of climate change on agriculture. Crop damage arising from more frequent and severe extreme weather, such as bursts of intense rain or hail, the devastating effects of prolonged drought and high temperature, can all impact on yield. Extreme conditions are projected to become more pronounced and severe in the decades ahead, as higher concentrations of greenhouse gases lead to further climate change. More frequent extreme weather is already leading to significant, and sometimes devastating, economic losses and can even undermine food security (Charles, Godfray, & Garnett, 2014). Additional climate change-related pressures arise from the spread of pest species to areas where they were previously limited by unsuitable conditions. Climate change impacts are not the only ecological factors that might affect future food security though. The very same forests that are being cleared to make way for fields and which are releasing carbon into the atmosphere also replenish water. Forests pump billions of tonnes of water into the atmosphere each day. As we lose more and more forests so there is an impact on the Earth’s delicately tuned hydrological cycles, in turn impacting on rainfall and thus on the viability of crop production. Some 70% of water consumption globally is for agriculture (FAO, 2017) and in some vital food-producing

A food system fit for the future139

areas water availability is a finely poised component of output (Mancosu, Snyder, Kyriakakis, & Spano, 2015). Droughts can cause major impacts and, in some cases, such as in Brazil and West Africa, recent prolonged dry periods that have impacted on food output were linked with deforestation (e.g., Nobre, 2014). It is not only climatic disruption partly caused by agriculture that is causing impacts on farming. Some of the changes to biological systems caused by farming will also lead to impacts on food production. One case in point relates to pollination. Animal-pollinated crops rely mostly on wild creatures (and mostly insects) to produce fruit and seed. The more that farming and other practices lead to pollinator declines, the more this vital service will be undermined. Some two-thirds of cultivated crop varieties depend on animal pollination (Ollerton, Winfree, & Tarrant, 2011) so this is far from a fringe issue when it comes to protecting future food security. Recent reports of large-scale insect decline suggest major profound changes taking place, with one recent estimate suggesting that 400,000 species of insect (40% of all species) are now at risk of extinction, largely due to the effects of pesticides (Goulson, 2019). Habitat loss, monoculture agriculture, and climate change are all also undoubtedly playing a part. Soil damage too is a long-term threat to food security. To date, our response to soil degradation has been to either open new areas of virgin soil from beneath forests and natural grasslands or mask the decline of soil health through the application of industrially manufactured fertilizers. In both cases there are limitations to these strategies. In the modern context, and because of among other things the need to cut ­climate-changing emissions and halt biodiversity loss, we need to be regenerating rather than clearing forests. Reducing reliance on sources of manufactured nutrients is also necessary for reasons of climate change and wider environmental protection. These threats to food security arising from environmental change, including those that are at least in part caused by agriculture itself, are reminders that of all the vital economic sectors that sustain our societies and civilizations, it is farming that is the most directly reliant upon a healthy environment. With this in mind, it is thus important to understand the reasoning that has brought us to the present situation, whereby environmentally damaging practices are not only continued but, in some cases, encouraged, for example, via public subsidy.

7.4 Drivers of the current system The quest for ever-greater levels of food production is, broadly speaking, the biggest underlying motivation that has led to the rising scale of food system environmental impact outlined earlier. That drive for greater yields has in turn largely been down to two main factors: population growth and changing dietary preferences (more on which later). The rising demand that has resulted from these factors has caused fears of food scarcity, and that in turn has led to a widespread perception of a need to increase the volume of food output. That drive for increased production has also been driven by the attraction of low prices, with “cheap” food becoming an objective for food-related policies across the world.

140

Rethinking Food and Agriculture

With the challenge of producing more food at the front of mind, there has since the 1940s, and during the period of the so-called “Green Revolution” in particular, been a strong focus on strategies to expand food production. These have already been mentioned and include a combination of plant and animal breeding to create more productive varieties, the proliferation of pesticides, the expansion of irrigation, the conversion of vast areas of once forested land, facilitating access to fertilizer, and the scaling up of machinery. As the environmental impacts of this progressive industrialization of production have become more obvious, so a mindset of “balance” became established in the discourse. This frame of reference has presented environmental damage as something that needs to be weighed against food security, with some kind of “middle ground” accommodation being the desired goal. Despite the narrative of “balance,” however, the available data confirms that some rather unbalanced outcomes have thus far been the result of the approach adopted (including deforestation, species loss, climate change, pollution, and soil damage).

7.5 What food shortage? While the popular and policy discourse has sought “balance,” the question of whether the fear of shortage is actually real has been somewhat neglected. For example, oft-­ repeated claims to the effect that production must be doubled or increased by 70% by 2050 largely go unchallenged, as do less specific claims that environmental concerns should be set aside to increase production. Yet, if food production is placed in the context of a series of related trends, then conventional assumptions as to the need for greater output at least become questionable. There are four facts that in particular should cause us to pause before embracing into the future the scarcity narrative with the certainty that it has hitherto been accompanied. The first such fact relates to food loss and waste. Around one-third of food is presently lost between harvest and consumers (Alexander et  al., 2017), suggesting that food is in fact more abundant and cheaper than some of the assumptions that still drive farming practices would hold. There are clearly very significant social and geographical differences that lie behind this general estimate, but the context is nonetheless very significant, considering how lost food contributes more than 3.3 billion tonnes of greenhouse gas emissions each year and it takes about 14 million square kilometers of land to produce (FAO, 2013). Then there is the fact of a global obesity epidemic, and the degree to which too many calories is now a greater concern for public health than too few (Hurt, Kulisek, Buchanan, & McClave, 2010). Again, very significant social and geographical factors must be taken into account, but this is another reason to probe the underlying scarcity narrative. The third fact relates to diet and the extent to which our food supply has become stretched by the explosion in consumption of meat and dairy products. These tend to place a greater strain on ecological and agricultural systems than diets more dominated by plant foods, a point underlined by the estimate that globally about 40% of crops

A food system fit for the future141

grown are being fed to our billions of farm animals (Stoll-Kleemann & O’Riordan, 2015). While we do of course eat the eggs, meats, and cheeses that are the end result, the number of people that might be fed from plant foods directly would be much greater if meat and dairy comprised a smaller proportion of average food consumption (Springmann et al., 2018). Reducing meat and dairy consumption by on average 50% would in any event reflect health advice (Tirado, Thompson, Miller, & Johnston, 2018). According to the recent Eat-Lancet Commission report, animal products should make up less than 15% of daily calories for a healthy diet (Willett et al., 2019). There is also increasing acceptance that a healthy diet does not need to contain any animal products at all (see Chapter 15). Having said this, livestock can be a component of sustainable agricultural systems, for example, some organic approaches rely on animals used in crop rotations to help aid the recovery of soil fertility. However, the vast majority of animal products are not produced using such methods. Then there is the extent to which policies have been adopted to replace diesel and petrol refined from fossil oil with alternatives made from crops, including sugar cane, maize, soya, and palm oil. Again, if policy makers feel able to justify policies to support such a diversion of food into fuel, then presumably food scarcity must be less pressing than is sometimes assumed to be in parallel discussions about food security.

7.6 Is our food really that cheap? Another driver of the mindset currently shaping agriculture has been the quest for cheap food. This has to an extent been embedded within what might be called the “must produce more” narrative, with one expected effect of abundance being lower prices. Indeed, this has been one outcome arising from an abundance of food, with consumers in many countries seeing during recent decades the proportion of their income spent on food dropping very significantly. Prices visible to consumers must, however, be seen in the context of costs that the market presently does not reveal. A new approach toward economics based on the concept of “natural capital” is helping to reveal the extent to which different human activities are depleting the very foundations of their existence, in the process depleting the economic value of the assets in question (Badura, Ferrini, Agarwala, & Turner, 2017). In some cases the values of the ecological services that underpin economic development can be quantified in monetary terms, thereby enabling full cost or full value calculations to be made. Through such analysis it is possible to, for example, calculate the costs of dealing with the climate change, soil damage, and water pollution caused by food production, in turn revealing how the costs of what many people buy in the shops is not actually as low as the labels suggest. The illusion of “cheap” food has also in some geographies (such as the European Union and United States) been shaped by the annual allocation of tens of billions of dollars’ worth of farm subsidies, designed in different ways to sustain or increase production. Thus, while food might appear a little cheaper in shops, this is in part because those very same consumers have paid for it via another route—their taxes.

142

Rethinking Food and Agriculture

7.7 Beware false frames As the ecological impacts of agriculture have become more apparent, so some new frames of reference have become prominent in the discussion about possible responses. One relates to the idea of land sparing versus land sharing. Sparing land relies on the intensification of farming to obtain more food from the same hectares, with the logic of the strategy being the obviation of the need to clear more forest and natural grasslands. The idea of sharing involves sustaining biodiversity in farmed landscapes, so that wildlife remains in mosaics of habitats where food is being produced, with the assumption being that through this route the land will be less productive and therefore the total area that must be farmed greater. While perhaps making for an interesting point for debate, there is good reason to believe that the sparing/sharing dichotomy presents a false choice. This is not least because the choice is presented in the context of the assumed need to produce more food, thus reinforcing one of the main misconceptions that lies at the heart of how we see food. If there are questions to ask in relation to the scarcity narrative, then there are also questions in relation to whether sparing versus sharing is the right issue to focus on. The sparing and sharing dichotomy also poses the risk that more fundamental challenges relating to the overall sustainability of agriculture become side-lined. For example, in any agricultural system it will be necessary to consider a range of issues such as resilience to climate change impacts, soil health, and wider food system pressures, such as those arising from high levels of food waste. In most cases it will therefore be a matter of both sparing and sharing, not either/or, and with land use decisions couched within the context of a wider system-level view of food and farming, embracing questions of food waste, the role of nonfood crops and diets, among other things.

7.8 The actors Taking this kind of system-level view of food and farming is, however, challenging, especially since this is not only a matter of finding the best pathways to find positive synergistic action across multiple agendas, but also among multiple actors, who sometimes have divergent goals. Governments design policies, standards, financial support mechanisms, and trade rules, all of which have profound implications for farming. The extent to which these are geared toward sustainable outcomes is, however, limited by the pressures upon them. Various lobby groups from supermarkets to farmer organizations, and from biofuel manufacturers to consumer organizations, sometimes advocate policies that reflect narrow and short-term goals, rather than systemic and longer-term ones. Managing the cost of living is in most countries a potent political driver, leading policy makers to focus more on shop prices than the overall sustainability of farming. The challenges to governments are not only seen individually, but also collectively in international treaty processes. Seeking common cause on climate change and ­biodiversity

A food system fit for the future143

conservation through global agreements has been a fraught decades-long process. A process that has yet to deliver decisive action on the scale required on both subjects, even when accord was reached not only among governments worldwide but also with leading companies. For example, in the wake of the New York Declaration on Forests signed in 2014, which sought to cut deforestation in half by 2020, the forest loss from 2014 to 2019 (at the time of writing) actually went up (NYDF Assessment Partners, 2019). As countries seek to accommodate their domestic political drivers and to protect their national interests and national farming and food sectors, so it becomes very challenging to incorporate climate and ecosystem sustainability. The private sector is similarly challenged. Many of the larger brand companies that have supply chains and interests linked with food and farming are locked into what are often intensely competitive situations, whereby short-term and narrow financial cost factors drive decision-making. Supermarket chains are, for example, engaged in more or less ceaseless price wars that often drive their business model innovation away from the broader sustainability dimensions and toward cost-cutting measures to achieve competitive outcomes over immediate horizons. Those that are listed on stock markets are also subject to shortterm investor pressure, with the need to prioritize financial returns in quarterly or annual reports trumping investments that would contribute to long-term sustainability. These circumstances feed through from the major consumer and retail brands into their supply chains of commodity companies, processors, and wholesaler organizations. Consumers are a diverse group, from the educated and wealthy to the poor and marginalized, and they have very diverse needs. In all cases though, the intersection points between food and sustainability are liable to create more confusion than clarity—for example, is it best to avoid palm oil ingredients or better to seek out sustainable sources? Is it best to be vegan or are there good sources of meat and dairy that actually enhance the environment, rather than degrade it? These and many other questions leave consumers unsure of what they should do to promote solutions, although some messages (for example, linked with the reduction of food waste) have wider appeal and actionability. Then of course there are the farmers themselves, caught between the powerful forces of the market, consumer culture, and government policy and regulation. The extent to which farmers can respond to environmental challenges, even if they wanted to, is thus limited by really very powerful and fundamental drivers.

7.9 Priorities for a sustainable food system In seeking to shift the food system to a sustainable one, it will be necessary to look at the roles that might be played by all of the actors. I present next my broad view of priorities for the major actors.

7.10 Governments Governments have a variety of levers at their disposal, including laws to protect important species and habitats, reduce pollution, and regulate chemical usage. These need

144

Rethinking Food and Agriculture

to be developed and used in ways that require basic standards to be observed. Beyond that, many countries have farmer incentive schemes that in most geographies can be more closely aligned with ecological goals. For example, in the United Kingdom there has during recent years been political momentum toward new farming policies that would more effectively target public money for public goods. This policy will seek to redirect billions of pounds of taxpayers’ subsidies away from direct payments to farms based on their size and instead toward protecting and restoring the environment. Paying farmers in relation to how they conserve soils, improve wildlife, and protect water will not only help agriculture to be greener but also sustain the future of farming. This might include large-scale habitat recreation, alongside more modest on-farm changes. Trade policies and agreements represent another dimension through which governments can seek to elevate environmental and sustainability goals. The European Union, for example, has a number of standards in the animal health and sustainability realm that limit the import of goods that might be environmentally harmful. This is a necessary protection for farmers in the European region who would otherwise be competing to produce food as cheaply as farmers working to lower standards elsewhere in the world, and without that protection many of them would go out of business. Governments can also work to embed farming within a wider food system policy. This might embrace consumer awareness about food waste, health advice linked with sustainability (for example, in relation to healthy levels of consumption of livestock products), and be embedded into educational curriculums.

7.11 Private sector These and other interventions by governments will steer the system in a more positive direction as regards sustainability. There are, nonetheless, many leadership actions that can be taken by private companies and financial institutions and investors. One goal that can rally action right away is to pursue zero deforestation supply chains. There is a vast area of degraded land globally and by recovering the health of this land we could produce more food as well as expand forest cover, and in the process bring a wide range of environmental benefits. For example, it is estimated that by restoring and reusing degraded land, Brazil could meet the projected global demand for soya for the next 50 years without any further deforestation (WWF, 2019), such is the opportunity to make better use of the land already cleared. Animal feed suppliers, commodity companies, consumer goods businesses, and supermarkets could all be influential in making this shift, should they insist on buying only soya that has not been produced as a result of recent deforestation. They could similarly shift behavior in relation to other commodities, including beef, palm oil, and cocoa. And while supermarkets and brands build up the supply of deforestation-free products, they can also source and sell only certified sustainable seafood as well. Companies allocating capital can also be part of the solution. For example, instead of backing companies that are engaged in deforestation as part of their business model,

A food system fit for the future145

investors can place finance into some of the new cutting-edge enterprises that are producing food indoors using hydroponics or those developing plant-based “meats.” Some of the companies in these areas are very profitable and growing fast. Another role that might be played by private sector companies is to be working with leading-edge sustainable farming businesses and connecting those with consumers. Some are already working to do this, such as UK supermarket Waitrose.

7.12 Consumers Governments, financial institutions, and food companies are ultimately reflections of wider society, voters, savers, and consumers. What we all do as individuals therefore influences the whole system. One priority that most people can pursue, and which directly impacts on the fundamental driver of overall demand, is to reduce food waste. A little thought and planning can avoid a lot of food being spoiled, and in the process not only save nature but also money—and quite a lot of it when you know that the value of the food lost and wasted in the United Kingdom alone each year is about £20 billion. Via education and culture change, the reduction in lost food could make a significant difference, including that saved by individual consumers. By taking action themselves, consumers would send messages to other actors who are in a position to also act on food waste, such as supermarkets and food-processing companies. Progress toward a sustainable food system could also be made via dietary choices. Consuming less meat and dairy foods and ensuring that what we do consume comes from sustainable production methods would make a difference. Better meat and dairy products come, for example, from animals who are improving soil health, including as part of organic farming systems. Animals can also be reared in ways that contribute to the conservation of wildlife, especially those reared by grazing and browsing, rather than imported feed crops such as soya and corn. The “rewilding” scheme at Knepp in Sussex, England, relies on the restoration of relationships between grazing animals and naturally regenerating vegetation. About 70 tonnes of meat is harvested annually, providing income from the ecologically transformative land use that has replaced unsustainable arable and dairy farming. By contrast, take the ongoing rapid loss of Brazil’s Cerrado savannahs mentioned earlier. This diverse ecosystem is being cleared mostly to grow animal feed. When one considers that about three-quarters of global output of soya is fed to livestock, rather than for direct consumption by people, it becomes clear what a huge difference can be made. One estimate suggests that if the world’s 2 billion biggest consumers cut their meat and dairy consumption by 40%, then that alone would free up an area of land twice the size of India (Searchinger et al., 2019). Consumers can also be influential in asking the brands we buy from to ensure that environmental damage is minimized in the making of the products they supply. For example, if we all insisted that companies ensure that no deforestation was caused in the supply of palm oil, cocoa, soya, or other ingredients they use, then that will have

146

Rethinking Food and Agriculture

powerful knock-on economic effects among grower and trading companies, encouraging them to change their behavior, and to, for example, invest in the recovery of degraded soil, rather than clearing more forest.

7.13 The big issue Our relationship with food increasingly defines our relationship with our planet. We know though that what we eat needn’t cost the Earth, for as we discover more and more about the impacts of our food system, so we also discover many of the solutions. By seeking system-level change through the combined actions of governments, companies, investors, and consumers, we can reorient how we feed ourselves and better align nutrition with the protection of biological diversity, climate change goals, and the conservation of critical natural resources, such as soils. In moving toward a sustainable food system, it will be necessary to dramatically expand awareness as to the impacts of our food and also to raise the profile of the remedies. This is beginning to happen and is reflected in a number of consumer trends that are now also beginning to be reflected in how markets work. For example, the elevated profile of reduced meat consumption as an environmental solution has not only led to significant culture and behavior change but also to shifts in the product ranges supplied and promoted by major supermarket chains. When coupled with not only a “less” message, but also a “better” message, then the ecological benefits that might be gained from more sustainable livestock rearing will hopefully soon be more fully realized. In the policy-making world there is too in some countries a rising awareness as to the value of investing in sustainable farming practices, not only for food security but also for the supply of clean water, reduced flood risk, and the capture of carbon in recovering soils and vegetation. This in turn is leading to discussion about the reorientation of policies and subsidies to encourage a change in farmer behavior. Food sector companies and financial institutions are increasingly aware of both changing consumer attitudes and behavior and the shifts in policy that are in part the consequence of that. Not only are these shifts powerful sources of new risk, they also reveal new opportunities, and as a consequence innovation and investment are beginning to be reoriented toward new food solutions. Campaign groups, commentators, and various specialist investigations and consultative processes have been successful in establishing new agendas and debates in the food and agriculture space. As a result, new opportunities now exist to reorient our food system toward sustainable outcomes. This will, however, require all involved to maintain sight of the fact that for success, we will need whole system change.

References Alexander, P., Brown, C., Arneth, A., Finnigan, J., Moran, D., & Rounsevell, M. D. (2017). Losses, inefficiencies and waste in the global food system. Agricultural Systems, 153, 190–200. Alexandratos, N., & Bruinsma, J. (2012). World agriculture towards 2030/2050: The 2012 revision. Rome, Italy: Food and Agriculture Organisation of the United Nations.

A food system fit for the future147

Badura, T., Ferrini, S., Agarwala, M., & Turner, K. (2017). Valuation for natural capital and ecosystem accounting. Synthesis report for the European Commission. Centre for social and economic research on the global environment, University of East Anglia. Norwich. Barrett, M., Belward, A., Bladen, S., Breeze, T., Burgess, N., Butchart, S., … de Carlo, G. (2018). Living planet report 2018: Aiming higher. Switzerland: WWF. Charles, H., Godfray, J., & Garnett, T. (2014). Food security and sustainable intensification. Philosophical Transactions of the Royal Society of London. Series B, Biological Sciences, 369(20120273), 10–1098. Eklöf, A. (2009). Species extinctions in food webs: Local and regional processes (Doctoral dissertation). Linköping University Electronic Press. FAO. (2013). Food wastage footprint: Impacts on natural resources. FAO. FAO. (2017). Water for sustainable food and agriculture. A report produced for the G20 presidency of Germany. Rome, Italy: FAO. Fowler, D., Coyle, M., Skiba, U., Sutton, M. A., Cape, J. N., Reis, S., … Vitousek, P. (2013). The global nitrogen cycle in the twenty-first century. Philosophical Transactions of the Royal Society, B: Biological Sciences, 368(1621), 20130164. Goulson. (2019). Insect declines and why they matter. Wildlife Trusts. https://www.somersetwildlife.org/sites/default/files/2019-11/FULL%20AFI%20REPORT%20WEB1_1.pdf. GRAIN. (2015). The Exxons of agriculture. GRAIN. 30 September 2015. Hannesson, R. (2014). World fisheries in crises? Marine Resource Economics, 30(3), 251–260. https://doi.org/10.1086/680443. 2015. Hurt, R. T., Kulisek, C., Buchanan, L. A., & McClave, S. A. (2010). The obesity epidemic: Challenges, health initiatives, and implications for gastroenterologists. Gastroenterology & Hepatology, 6(12), 780. IPBES. (2019). Global assessment report of the intergovernmental science-policy platform on biodiversity and ecosystem services. New York: UN. IPCC. (2019). Climate change and land: An IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems. New York: UN. Isenring, R. (2010). Pesticides and the loss of biodiversity (p. 26). London: Pesticide Action Network Europe. Jambeck, J. R., Geyer, R., Wilcox, C., Siegler, T. R., Perryman, M., Andrady, A., … Law, K. L. (2015). Plastic waste inputs from land into the ocean. Science, 347(6223), 768–771. Lal, R. (2014). Soil carbon management and climate change. Carbon Management, 4(4), 439–463. Lebreton, L., Slat, B., Ferrari, F., Sainte-Rose, B., Aitken, J., Marthouse, R., … Noble, K. (2018). Evidence that the great Pacific garbage patch is rapidly accumulating plastic. Scientific Reports, 8(1), 4666. Mancosu, N., Snyder, R. L., Kyriakakis, G., & Spano, D. (2015). Water scarcity and future challenges for food production. Water, 7(3), 975–992. Montgomery, D. R. (2012). Dirt: The erosion of civilizations. Univ of California Press. Nobre, A. D. (2014). The future climate of Amazonia, scientific assessment report. Sponsored by CCST-INPE, INPA and ARASão José dos Campos Brazil. NYDF Assessment Partners. (2019). Protecting and restoring forests: A story of large commitments yet limited progress. New York declaration on forests five-year assessment report. Climate focus (coordinator and editor). Accessible at forestdeclaration.org. Ollerton, J., Winfree, R., & Tarrant, S. (2011). How many flowering plants are pollinated by animals? Oikos, 120(3), 321–326. RSPB, et al. (2016). State of nature 2016. https://www.rspb.org.uk/globalassets/downloads/documents/conservation-projects/state-of-nature/state-of-nature-uk-report-2016.pdf.

148

Rethinking Food and Agriculture

Searchinger, T., Waite, R., Hanson, C., Ranganathan, J., Dumas, P., & Matthews, E. (2019). Creating a sustainable food future-a menu of solutions to feed nearly 10 billion people by 2050. World Resources Institute. Springmann, M., Clark, M., Mason-D’Croz, D., Wiebe, K., Bodirsky, B. L., Lassaletta, L., … Jonell, M. (2018). Options for keeping the food system within environmental limits. Nature, 562(7728), 519. Stoll-Kleemann, S., & O’Riordan, T. (2015). The sustainability challenges of our meat and dairy diets. Environment: Science and Policy for Sustainable Development, 57, 34–48. Sutton, M. A., Howard, C. M., Erisman, J. W., Billen, G., Bleeker, A., Grennfelt, P., … Grizzetti, B. (Eds.). (2011). The European nitrogen assessment: Sources, effects and policy perspectives Cambridge University Press. Tirado, R., Thompson, K. F., Miller, K. A., & Johnston, P. (2018). Less is more: Reducing meat and dairy for a healthier life and planet. Greenpeace research laboratories technical report (Review) 03-2018. ISBN: 978-1-9999978-1-6. 86 pp. UNCCD. (2017). Global land outlook. Bonn, Germany: Secretariat of the United Nations Convention to Combat Desertification. Willett, W., Rockström, J., Loken, B., Springmann, M., Lang, T., Vermeulen, S., … Murray, C. (2019). Food in the Anthropocene: The EAT–lancet commission on healthy diets from sustainable food systems. Lancet, 6736, 3–49. https://doi.org/10.1016/S0140-6736(18)31788-4. WRI. (2018). By the numbers: the value of tropical forests in the climate change equation. https:// www.wri.org/blog/2018/10/numbers-value-tropical-forests-climate-change-equation. WRI. (2019). The world lost a Belgium-sized area of primary rainforests last year. https://www. wri.org/blog/2019/04/world-lost-belgium-sized-area-primary-rainforests-last-year. WWF. (2019). Gone before we know it. Join our fight to save the Cerrado. https://support.wwf. org.uk/save-cerrado.

Why change the way we grow, process, and consume our food?

8

Hans R. Herren Millennium Institute, Washington, DC, United States “A sustainable system is any system that in its lifetime can produce more energy than it takes to establish and maintain it.” Bill Mollison.

8.1 Introduction Calls for a transformation of agriculture and the food system, or we may simply say the food system (see Box 8.1 and Fig. 8.1), are not new but have gained in strength from the request for a change in paradigm away from Green Revolution to agroecology (see Box 8.2). The formative call for this transformation by the International Assessment of Agricultural Knowledge, Science and Technology for Development (IAASTD, 2009) continues to gain urgency in the latest reports from the Intergovernmental Panel on Climate Change (IPCC) (IPCC, 2018, 2019), in which the food system has finally found a place of “honor” as both a key contributor to climate change as well as a necessary part of the solution. The long reluctant UN Food and Agriculture Organization (FAO) has finally embraced agroecology after years of denial and refusal to acknowledge the existence of the IAASTD report as well as calls from civil society groups and academics. FAO has organized a series of conferences on agroecology, both at its headquarters in Rome, as well as in Africa, Asia, Latin America, and Europe. In October 2019, the Committee on World Food Security, of which FAO is a member, presented the landmark High Level Panel of Experts on Food Security and Nutrition Report on “Agroecological and other innovative approaches for sustainable agriculture and food systems that enhance food security and nutrition.” This event is a major step forward in the transformation process for the food system. Although late, with some 10 years lost since the IAASTD report, it is “never too late to do good.” But there is now an even greater urgency to start acting on the transformation from the conventional and industrial model to the agroecological one. The IAASTD report called for a paradigm change, that “business as usual is not an option,” and that agroecology is the way ahead, a call that is even more true today. Agroecology in its newest and broadest definition (Box 8.2) is the direction for the transformation.

8.1.1 Key problems of the present food system The urgency with which we must address the key problems of the present food system cannot be overstated. Around the world, farmers are struggling to understand what they can still grow within a changing climate and adapt their farming practices Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00008-6 © 2021 Elsevier Inc. All rights reserved.

150

Rethinking Food and Agriculture

Box 8.1  Definition of food systems and industrial/conventional agriculture. Food systems

The entire range of actors and their interlinked value-adding activities involved in the production, aggregation, processing, distribution, consumption, and disposal of food products that originate from agriculture, forestry, or fisheries, and parts of the broader economic, societal, and natural environments in which they are embedded. Food systems are made of subsystems (e.g., farming system, input supply system, processing system) and interact with other key systems (e.g., energy system, trade system, health system). They hinge on actors, relationships, and networks (Fig.  8.1) and generate environmental and socioeconomic outcomes, in addition to food security and nutrition (Dury, Bendjebbar, Hainzelin, Giordano, & Bricas, 2019). Industrial/Conventional agriculture

Industrial agriculture is currently the dominant food production system in North America, Europe, parts of Asia, Australia, and in the emerging economies of South America. It is characterized by large-scale monoculture, heavy use of chemical fertilizers and pesticides, and meat production in confined animal feeding operations. The industrial approach to farming is also defined by its heavy emphasis on a few crops that overwhelmingly end up as animal feed, biofuels, and processed junk food ingredients (UCS, 2008).

a­ ccordingly. In California, pistachio growers are seeing their livelihood threatened and have initiated research to develop heat-tolerant varieties. In China’s drought-prone Fujian Province, farmers are switching from wheat to apples. In Kenya, herders are switching from cows to camels. Such examples are just a sampling from around the world (Majot, 2019). While adaptation is possible within limits, the main effort needs to be on mitigation. If the food system is redrawn according to the best science available, it can become one of the main parts of the solution, by sequestering carbon as well as being carbon neutral. While we could endlessly discuss exactly how many gigatons of carbon a full transformation to agroecological practices may yield, one thing is sure: carbon can and will be sequestered by changing the present unsustainable practices that destroy life in our soils, keep them exposed to the elements, and accelerate carbon loss to the atmosphere. Through the best possible agroecological practices, climate change can be mitigated for a great part. We know what to do and how to do it, and the time to start is now. It is likely we have passed the opportunity to keep global heating under 1.5 °C, but we must do our best to keep it below 2.0 °C. This also means that we need to watch the CO2 PPM (parts per million) level, having reached the dangerous zone of 412 PPM, when 350 PPM was meant to be the maximum acceptable level, up from preindustrial levels of 275 PPM. The task is daunting but also feasible. The best estimates by the United Nations Environment Program

Why change the way we grow, process, and consume our food?151 Policy variables Gender policies

Equity policies

Small-scale agriculture Saving

Environmental resources Temperature

Land tenure quality

Food aid

Governance

Public expenditure

Private expenditure

Investment in environmental resources

Investment in social resources

Expenditure

Climate stability Atmosphere CO2

Food import

Solar radiation Food prices Pasture area Crop area Plant and animal health Soil organic matter

Income from agriculture

Investment in economic resources

Food availability

Food access

Food and nutrition security

Food production

Water Forest area

Farmers’ organization Health Education Agriculture knowledge R&D resources Access to social Services Access to markets Access to credit

Food use

Employment

Soil nutrients Biodiversity

Social resources

Rural poverty

Income distribution

Economic resources Infrastructure Agriculture machinery Processing capacity

Input prices

Intermediate consumption

Expenditure for external inputs

Storage capacity Irrigation

External inputs Pesticides (chemical/biological)

Fertilizers synthetic/organic

External Energy seeds and feeds

Fig. 8.1  Simplified stocks and flows in the highly interconnected food system (Millennium Institute internal report).

show that agriculture, wetlands, and forestry could remove up to 12 gigatons of CO2 per year, from the needed 37 produced annually (Rogers, 2019; Rouse, 2019). These numbers do not take into consideration the CO2 already dumped into the atmosphere, which can also be sequestered underground. Considering that 40% of food produced is wasted, the potential CO2 savings of a zero-waste program are significant and need to be implemented along with underground sequestration and carbon neutral food production and processing. Parts of this equation are diet change considerations and the relocalization of the food system. The evidence that we can produce enough quality food sustainably for all is widely available. For Europe, the Poux and Aubert (2018) study shows that agroecology is a feasible option, even though there are the usual voices that call for a continuation of the present model, with more technology, chemistry, genetic engineering, larger holdings, and vertical urban farming. This study also emphasizes that diets will have to change through reducing meat consumption while increasing vegetables and fruits. From a health and environmental point of view, this makes sense, given that much of the cheap, industrial meat production is both socially and environmentally damaging and economically dependent on subsidies. Furthermore, low prices are made possible by externalizing the negative costs, which promote excessive consumption and waste. Indeed, the present conventional and industrial food system model, well established in developed and industrial countries, as well as in a number of newly industrialized

152

Rethinking Food and Agriculture

Box 8.2  Definition of agroecology, regenerative agriculture, permaculture, and the Green Revolution. Agroecology

Agroecology is a way of redesigning food systems, from the farm to the table, with a goal of achieving ecological, economic, and social sustainability. Through transdisciplinary, participatory, and change-oriented research and action, agroecology links together science, practice, and movements focused on social change (Gliessman, 2015). Regenerative agriculture

Regenerative agriculture is a system of farming principles and practices that increases biodiversity, enriches soils, improves watersheds, and enhances ecosystem services. Regenerative agriculture aims to capture carbon in soil and aboveground biomass, reversing current global trends of atmospheric accumulation. At the same time, it offers increased yields, resilience to climate instability, and higher health and vitality for farming and ranching communities (Terra Genesis International, 2019). Permaculture

Permaculture is a set of design principles centered on whole systems thinking, simulating, or directly utilizing the patterns and resilient features observed in natural ecosystems. It uses these principles in a growing number of fields from regenerative agriculture, rewilding, and community resilience (WikiPedia, 2019). The Green Revolution

The Green Revolution refers to the use of high-yielding variety (HYV) seeds, which were invented by the crop geneticist Norman Borlaug. HYVs are normally used as a part of a technological package that also includes biochemical inputs such as water, fertilizers, and pesticides, and often mechanical inputs (Encyclopedia.com, 2019).

ones (Brazil, China, and Argentina among others) is responsible for nearly 50% of greenhouse gas (GHG) emissions when accounting for these from field to fork, inclusive of the waste it generates along the value chain (Move for Hunger, 2019; Ribeiro, 2018; Xue et al., 2017). There is a lack of general understanding, and therefore disconnect, at both the level of consumers and decision makers about food production, transformation, and consumption. This lack of understanding is exploited by the agri- and food business, which promotes the virtue of cheap and (highly) processed food with year-round availability. There is, however, also a growing segment of the consumer population that is demanding quality, seasonal, chemical-, and drug-free, sustainably produced food, which in turn is helping the organic and agroecology sector to grow. This sector, however, suffers from a huge gap in R&D investments, as even to date over 98% of the total agriculture research investments still flow to the Green Revolution-type agricul-

Why change the way we grow, process, and consume our food?153

ture (new study by Biovision Foundation for Ecological Development & IPES-Food, 2020; DeLonge, Miles, & Carlisle, 2016; Pimbert & Moeller, 2018). These biased investments are an institutional issue that need to be corrected to further help organic agriculture, agroecology, and regenerative agriculture develop their full potential in support of the farmers that opt for the transformation. Roughly 75% of Earth’s land has been degraded by increasing consumption, which fuels inappropriate farming practices, urbanization, and mining (IPBES, 2018). In 2019, the dead zone in the Gulf of Mexico, courtesy of corn and soybean grower’s use of synthetic fertilizers, will reach 7000 square miles (Kling, 2019). These global dimensions of soil degradation and water pollution—along with the spiraling loss of biodiversity, rural to urban migrations, and increasing health burdens related to diets—highlight our current status of unsustainability. The report “From Uniformity to Diversity” (IPES-Food, 2017a, 2017b) described eight main blockages that stand in the way of an agroecological transformation of the food system and suggested ways to overcome them. Of particular note is the central issue of power concentration within the present economic system, which favors a maximization of profits at the expense of long-term impact in all three sustainable development dimensions (environmental, social, and economic). This problem is best represented by the mega mergers in all areas of the food sector, which work against both farmers’ and consumers’ interests. Adding now to the power concentration issue is the latest twist with data acquisition and ownership, which is the key to agriculture’s digitization. This includes blockchain technology, which could be of great benefit to small-scale farmers in all parts of the world. But farmers will again be on the losing end of these innovations, unless owned and managed by concerned stakeholders rather than external actors. Study after study published since the IAASTD report 10 years ago has emphasized the need for a radical transformation of the food system, a paradigm change from an exploitative to a regenerative system. Knowing the challenges presented by the last decade of procrastination and inaction at the political and private sector level, civil society has taken matters into its own hands. For example, farmers have transitioned to organic, ecological, and regenerative practices, despite the lack of research, which many have taken upon themselves to conduct. Consumers are now also more aware than ever about the health impact of industrial food (Wise, 2019). In its 2019 Report (Swinburn et  al., 2019) the Lancet Commission on Obesity demonstrates that the pandemics of obesity, undernutrition, and climate change represent the paramount challenge for humans, the environment, and our planet. These interacting pandemics represent The Global Syndemic with common, underlying drivers in the food, transport, urban design, and land use systems. The costs to society are out of control, with obesity affecting over 2 billion people and costing an estimated 2.8% of global GDP. In addition, 155 million children are stunted and 52 million are wasted by undernutrition. The food system that overproduces by nearly 50% still leaves over 800 million people chronically undernourished. The main issues with the present food system are beautifully visualized, described, and quantified in the compelling handout by Bendjebbar, Nicolas, and Thierry (2019). However, the policy responses to these problems have been at best inadequate, and

154

Rethinking Food and Agriculture

at worst bordering on irresponsibility. As we have already noted, in the IPES-Food report of 2017 regarding system blockages, vested interests are standing in the way of an integrated, inclusive, responsible, and holistic approach to the problems. The food system, climate change, obesity, under- and malnourishment, ecological degradation, and growing inequity are not separate events. They are also all connected, represented by the universally agreed 2030 Agenda for sustainable development, including its 17 sustainable development goals (SDGs).

8.2 A bankrupt system: Dealing with symptoms rather than the causes In a drive to assure food security and access to their citizens, many countries around the world have devised a subsidy system. This aims to fill the gap between what farmers need to have a decent income from the sale of their products and what consumers would be able, or ready, to pay for. Although laudable, the system quickly became corrupted, creating market distortion and endless squabbles at the World Trade Organization, benefiting larger, rather than small farmers. Much of the need for the subsidy system lies in the wider economic system, which promotes inequality and the concentration of power, i.e., money, in the hands of a few. When it comes to the food system, power is concentrated mainly in input providers, processors, and retailers. Farmers and consumers have little power to reverse this reality. The state of despair for many farmers in both industrial and emerging countries is reflected in the number of suicides. These are due to economic pressure and the high risks of farming, now exacerbated with climate change events (droughts, floods, pests, and diseases) and epidemics affecting large-scale livestock operations, which can ruin a farmer in a single season (Asst Editor Brewhouse, Perspective, 2015; Behere & Bhise, 2009; Swire, 2018). Another facet of the bankrupt industrial and conventional food system is its externalities, which are “socialized,” while the profits are “internalized.” Synthetic fertilizers and pesticides, which damage soil fauna and flora, have replaced essential ecosystem services. Genetic engineering is replacing locally owned, adapted land races and cultivars. Mega mechanization is promoting monocropping and landscape impoverishment. The presently prevailing food production system model is bankrupt. It is estimated that the global bill for these externalities currently amounts to a staggering $12 trillion, and will increase to $16 trillion by 2050 (Nature, 2019). Should these costs be internalized, the price of food would be beyond the reach of most consumers. It is time to revert to and revitalize the assistance from nature. Nature has developed over evolutionary times the needed support mechanisms, the ecosystem services, upon which we all depend. We have reached the end of the road of cost externalization. The growing number of organic, ecological, regenerative, and other forms of sustainable agriculture farmers is the best example that a system change is feasible. They have shown that externalities can be internalized, and the end costs to the consumer affordable, even in the short term. The longer-term benefits should also be factored in, for

Why change the way we grow, process, and consume our food?155

example, the health costs related to pesticide and pharmaceutical residues in food and water, not just for consumers but also for farmers, farm workers, and food processing plant personnel. Other benefits will come from reduced water pollution, biodiversity restoration, GHG reduction, and reduced climate change-induced risks for farmers. The food system can and must become a part of the solution to climate change. So far, however, it is still a large part of the problem, due to the blockages set in place by the agrifood complex (IPES-Food, 2017a, 2017b). The agrifood complex is not ready to yield to the sustainability paradigm, which may well make it obsolete, unless it changes course and fully applies itself to become part of the solution. The question remains if such a change is really possible under the present economic system. In a profound and fundamental transformation of the food system, it is certain that there will be winners and losers. However, by embracing new ways rapidly, the losers have a chance to be the change makers. There are signs that the transformation is taking root, starting with many farmers that are switching to sustainable practices; if not for philosophical or ideological reasons, for economic ones. They have been pushed to the brink of bankruptcy by the heavy costs of being dependent on the input and output industry. Farmland across the globe has suffered from overexploitation. The results are a depletion of nutrients and a need to have these replaced with soil life-destroying synthetic fertilizers. This effect is enhanced by the emphasis on a few commodities and a lack of crop rotations and crop diversity. Irrigation has become a necessity in many areas of the world, because the degraded soils are unable to store rainfall due to low organic matter and humus and no soil cover that minimizes evaporation. By keeping the soils bare for a large part of the year, water and wind erosion can take their toll. These practices result in a substantial loss of biodiversity, which in turn send the system further into a comprehensive degradation tailspin. Farmers are pushed into producing more from less land, again using more and more external inputs, which are unsustainable. A perfect example of a negative feedback loop, which has been at work with the support of the systemic blockages, was detailed earlier (IPES-Food, 2017a, 2017b). The impact of unsustainable industrial agriculture reaches into the socioeconomic sphere with an increased rural to urban migration. As farms grow larger to gain in efficiencies (at least in the short term), and land becomes scarcer, its value increases. This further promotes migration, while forcing remaining farmers to find ways to rationalize their operations, which often means extreme specialization, working against the needed higher crop and animal diversity that would increase biological and economical resilience. This also brings its own set of drawbacks, such as higher emissions, water pollution, and the production of unhealthy products, often with poor nutritional value. Given the strong economics and cost structure differences between small- and large-scale farmers, a coexistence will be difficult. Larger-scale farmers will always be at an unfair advantage, not least because they will have easier access to markets, credit, and subsidies. The exception here may be farmers in peri-urban settings, producing high-value vegetables and fruit crops, with easy market access and a growing, dedicated clientele. This will, however, remain a rather small fraction of today’s smallscale farmers, who will have to reinvent their business model and create cooperative producer associations to benefit from the resulting economies of scale.

156

Rethinking Food and Agriculture

The current way of dealing with these problems has been a race forward, dealing with the symptoms. This approach has been promoted by agribusiness, as it is better for business to keep selling pesticides, herbicides, fertilizers, and new seeds than dealing with the root causes of low yields, pest and disease outbreaks, low soil fertility, lack of pollination, and water management. Alternative solutions to these problems have long been known, in particular, the need to maintain healthy soils, use long crop rotations, with deep-rooted cover crops or pasture included in the rotation. Agriculture is a major intervention in the ecosystem, which to be sustainable needs to be as soft as possible. The main aim is to work in harmony with nature, the ecosystem, minimizing the disturbance and maximizing the services it offers. The present system, however, does the opposite, and then replaces these essential ecosystem services with costly and damaging external inputs. The corrupt subsidy system has rendered this possible, as it pays for these unsustainable practices and allows for the production of vast quantities of cheap food, i.e., commodities reduced to a few crops making transport, processing, and transformation more efficient. The true costs of this cheap food are astronomical, when including the social and environmental costs. A recent article in the New York Times exposed the problems with the EU agriculture subsidies and their insidious impact on farmers in Hungary, in large part due to the high corruption in this vastly uncontrollable system. The biggest losers are the farmers, along with the consumers and taxpayers who pay the annual bill of €65 billion (Apuzzo, 2019).

8.3 A new paradigm for the food system: How to change course The new food system paradigm called for by the IAASTD report was agroecology. After many conferences, side events, reports, bilateral and multilateral talks and negotiations, we now have a new report, which basically endorses the IAASTD options for action. This new report is the CFS-HLPE report (2019) entitled “Agroecological approaches and other innovations for sustainable agriculture and food systems that enhance food security and nutrition.” In addition to the already presented definitions of agroecology, one could introduce here an additional definition from an African farmer, Aisha Ali Aii Shatou, made at the occasion of the new report presentation: Agroecology allows small-scale producers a dignified life, producing affordable, healthy food in healthy conditions. It eliminates dependence on costly inputs and adopts practices which regenerate seeds and soils while mitigating and adapting to the effects of climate change.

We knew from the IAASTD report that agroecology can nourish the world population now and into the future. This message has been reinforced with new and additional science in the 2019 CFS-HLPE report. Agroecology is the only approach that will be able to do this because it deals with the root causes of the unsustainable food system.

Why change the way we grow, process, and consume our food?157

It is inclusive, holistic, and merges the three sustainable development ­dimensions and addresses all the SDGs. It helps solve the crippling climate emergency by being carbon neutral and a major carbon sink, reducing the atmospheric carbon stock, which was and still is released by conventional and industrial agriculture (crop and livestock). Agroecology will not solve the problem alone but it will make a major contribution if it is implemented in its three dimensions (social, environmental, and economic) at once and on a global scale with no further delays. The transformation is possible with immediate effect given that knowledge is sufficient to get started. R&D funds must be redirected 100% toward agroecology, organic agriculture, regenerative agriculture, and permaculture to assist farmers with new science, knowledge, and technologies, in particular because of the need for adaptation to the many different ecological, economic, and social conditions. For the remainder of this chapter, agroecology, organic agriculture, regenerative agriculture, and permaculture will be referred to as agroecology et  al., considering that agroecology is the “Chapeau” set of holistic principles that need to be implemented across the food system, as per Gliessman’s definition (Box 8.1). Some of the practitioners involved in these movements think their approach is the best and only way, while a successful transformation will result from a diversity of solutions depending on local situations and cultures. The failure of the conventional and industrial model should be sufficient motivation for food system transformation, not least from a government point of view, given the huge costs from agriculture subsidies and exploding healthcare costs. These costs are threatening national budgets at a time when there are many demands from across government departments for extra funding to meet the SDG targets. It is also increasingly clear to decision makers that the SDGs can only be met if the food system is transformed according to agroecological principles. Given that we have already overshot many planetary boundaries (Rockstrom et al., 2009; Steffen et al., 2015), mostly due to the industrial/conventional agriculture system and the wasteful food system, this transformation is even more urgent. The question of “how” this transformation needs to happen is best described in the IPES-Food reports (IPES-Food, 2017a, 2017b, 2018), the Beacons of Hope Report (Baker, Gemmill-Herren, & Leippert, 2019), and the CFS-HLPE report (CFS-HLPE, 2019). However, to be successful there is a need to emphasize that ecosystem restoration, and in particular biodiversity, above and below ground, should take high priority. Indeed, agroecology, and the practices within organic agriculture, regenerative agriculture, and permaculture, all have this main attribute. They work with nature, not against it. They use still available or restored ecosystem services to produce surplus harvestable biomass. Although we have not yet fully exploited the potential “extra” biomass production available to us humans for our nutrition, clothing, shelter, and livestock feed, there are still questions regarding how we can produce the needed biomass sustainably on the one hand, and how much we can consume on the other hand. These two sides of the equation should be the focus of our attention in designing the production systems of the future (which is now). The new paradigm therefore must be one of sufficiency and sustainability, in all three dimensions of sustainable development. As mentioned earlier, one of

158

Rethinking Food and Agriculture

the main factors that is holding back the global expansion of agroecology et al. is the belief, real or not, that we cannot “feed the world” with these types of agricultural practices and food system. There is a belief that we need an industrial approach, large-scale commodities, hyperprocessing, and global transfers to satisfy the nutritional needs of a planet with 9–10 billion people. To overcome this belief, scientists and farmers working in agroecology et  al. among others, have shown that these practices can not only produce sufficient quality nutrition for a growing population, but can do so while taking care of the health of the environment, consumers, and the economy. Calls for this transformation were the subject of the UN Secretary General’s remarks at the 2019 Climate Summit to the 2019 IPBES report (IPBES, 2019) and now also the CFS-HLPE agroecology report (CFS-HLPE, 2019). The new paradigm for our food system should also recognize and respect the cultural, social, and economic differences that exist around the globe (IAASTD, 2009). Issues like food sovereignty, land justice, seed sovereignty, etc., need greater attention, and the inclusion of all stakeholders in decision-making processes. Local communities have their own needs and wishes, and also duties to care for the land, their traditions, and their future. The need for decentralization and inclusiveness of decision-making processes, research, and the cocreation of a new sustainable development paradigm cannot be overstated. One area that has so far only been touched on at the margins in the food system transformation debate is what we will eat in 2050 or further down the road. Along with the concern about health, climate change, and planetary boundaries, a growing world population that becomes wealthier has changing tastes and needs. From synthetic meat to microbe-produced food, there are many scientists and food experts tinkering with alternatives, some needing raw materials from the field (Gunther, 2019; Thomas, 2019), others 100% of lab origin (Browne, 2019). While huge amounts of money are now being invested in the next big idea for food, attention should also be paid to the potential downsides. Food cultures are deeply rooted and slow to change. Whatever may come from these investments remains to be seen, and they should not distract us from dealing with the present urgent need to change the way we manage our food system.

8.4 A framework for change The new paradigm implementation has been hampered by the concentration of power, the lack of broad consultations, and the belief that more, bigger, and cheaper is better. This reasoning in the food system, as well as in many other economic areas, has brought us enormous challenges and problems that need fixing, not at the symptom level, but seriously at the causal level. This is hard work. It takes time and money, but it is now no longer an option, it is a must. We could follow the suggested recommendation for action from the CFS-HLPE report on agroecology, the ones from the Lancet’s Syndemic report, or go back to the IAASTD, or many others that have been published in between. The bottom line is

Why change the way we grow, process, and consume our food?159

that none of these reports is based on a comprehensive systems model, which would highlight where the leverage points are, when to activate these, and how and by whom. We need to recognize that we are dealing with ecosystems and landscapes, complex systems that do not come out of a cookie cutter. All have their idiosyncrasies and react differently to interventions. Eyhorn et  al. (2019) proposed a policy framework that sums up the actions needed very well in their Nature Sustainability journal article. It cuts through the endless debate of “for” or “against,” and shows how policies that align with the SDGs are needed to facilitate the transformation. Having noted this, there are a number of basic interventions that need to be undertaken. These will vary in scope, based on an analysis, or better an assessment of the food system situation in any given landscape. The IAASTD had recommended that every country carry out an assessment, broken down to subnational levels as may be necessary to capture the differences. Such assessments should take advantage of the latest analytical tools, such as system dynamics models, which are able to capture all the connections in the food system and other connected systems. Only with such tools can the leverage points be correctly identified, and by extension the best possible policies to move the system forward, along a roadmap to where concerned stakeholders wish to go. Such scenario models, which allow us to develop science, knowledge, and informed policies, test these in concert with other policies, and ensure coherence, do exist (Allen, Metternicht, & Wiedmann, 2016). These models visualize the synergies to take advantage of, while pointing to the negative feedbacks, both quantified over time. Such models further enable the measurement of progress against planned outcomes, highlight areas that need further attention, and help in the reporting of policy impact. The specific interventions that relate to the transformation of the food system toward the “overarching” goal of agroecology clearly go beyond the field, the factory, and the plate. They are dependent on the economic system, the available natural and human resources, financial means, and sociocultural considerations. The system models include what is known from research and practical knowledge about agriculture, forestry, fisheries, and pastoralism, the long value chain to consumption (Fig. 8.1) and its ramifications toward health, emissions, biodiversity, etc. No model is perfect, but they are a huge help when it comes to analyzing a few 1000 equations, compared to the human brain capacity to handle four or five at a time. On the blockages, it is relevant to mention that as a society we need to decide on the direction we want to take the planet, discuss what actions are needed to correct course, and then assign the actions, again as a society of equal voices. The private sector needs to step into the background, listen, and act as society decides, not the other way around as is the current situation, where vested interests decide how best to protect these interests, no matter the costs to society. There is lots of “greenwashing” from the private sector, major foundations, and their political marionettes. We cannot transform the food system without upsetting those with vested interests in maintaining the status quo. We are at the tipping point, with 3000 days to go before it is too late. Perhaps that date is much closer. Regardless, we cannot continue talking and writing. We need action and it is now coming from those who are at the receiving end of the “laissez faire” neoliberal policies, our youth, the next generation.

160

Rethinking Food and Agriculture

All the procrastination of government and the larger part of the private sector is only helping business as usual and the superficial changes to continue. Experts are calling for radical transformation, a paradigm change, nothing less. While this has a cost, it is cheaper today than tomorrow. One promising approach to support and accelerate the food system transformation is true cost accounting. Consumers react strongly to food prices. If there was a system that included the externalities of cheap food into the product’s end cost, there would be a reversal of today’s higher costs for healthy food versus cheap junk food (Pengue et al., 2018). One could elaborate at length on the recommendations made in the comprehensive CFS-HLPE report on agroecology, and of the ones in the Lancet Commission report. One could also dwell on the transformative experiences reported in the excellent Beacons of Hope report (Baker et al., 2019), all relevant and timely to move the 2030 Agenda, with all the SDG goals in mind. While the agroecology report has a long list of recommendations that cover almost every aspect for the transformation, the Beacons of Hope report has analyzed successful system transformations and the policies that made them possible. The suggestion here for policy/decision makers, policy advisors, the well-­ intentioned and forward-looking private sector, civil society interested parties, farmers, activists, and academics is: read these latest food system transformational reports, even if only the recommendations or the summaries for decision makers, and get started with what we know about agroecology et  al. Let’s invest in R&D and new skill developments that are needed across the food system. The benefits for society are huge. There will be winners and losers; this is inevitable when talking about a transformation that touches on all bases of life. The usual scare tactics from agribusiness that we shall lose jobs, not have enough to eat, and create conflicts do not apply. All studies on the subject tell us that there will be more jobs, better rewards, and overall increased benefits that are not all monetary. When the same agribusiness realizes that there is both bottom-up and top-down support and actions to get the transformation going, they are going to do two things: coopt the transformation processa and put pressure on governments. In the interest of the future of life on this planet, let’s call their bluff and move ahead regardless. A responsible society is one that respects its environment and protects it and all its living organisms. The present food system poses an existential threat to humanity. We know how to stop this. We can easily absorb some lower production in industrialized countries by reducing our substantial waste and overconsumption. As for the emerging economies, there is overwhelming evidence that productivity improves over the conventional practices as soon as agroecological practices are implemented (Khan, Midega, Pittchar, Pickett, & Bruce, 2011). Are we ready for the transformation? a

An example of this is Integrated Pest Management, which had the objective to stop calendar spraying against pests and diseases, and use pest-level intervention thresholds. The industry’s pest control advisors then simply lowered the threshold to increase pesticide applications.

Why change the way we grow, process, and consume our food?161

References Allen, C., Metternicht, G., & Wiedmann, T. (2016). National pathways to the sustainable development goals (SDGs): A comparative review of scenario modelling tools. Environmental Science & Policy, 66(2016), 199–207. https://doi.org/10.1016/j.envsci.2016.09.008. Apuzzo, M. (2019). We just wanted to talk E.U. farm policy. Why was someone always looking over our shoulders? Times insiders. The New York Times. Nov 3 https://www.nytimes. com/2019/11/03/reader-center/eu-farm-subsidy-reporting.html. Asst Editor Brewhouse, Perspective. (2015). Farmer suicides: A global phenomenon. PRAGATI The Indian National Interest Review. http://pragati.nationalinterest.in/2015/05/ farmer-suicides-a-global-phenomena/. Baker, L., Gemmill-Herren, B., & Leippert, F. (2019). Beacons of hope. Global Alliance for the Future of Food and Biovision Foundation. https://foodsystemstransformations.org/ wp-content/uploads/2019/08/BeaconsOfHope_Report_082019.pdf. Behere, P. B., & Bhise, M. C. (2009). Farmers' suicide: Across culture. Indian Journal of Psychiatry, 51(4), 242–243. https://www.ncbi.nlm.nih.gov/pmc/articles/PMC2802368/. Bendjebbar, P., Nicolas, B., & Thierry, G. (2019). Food systems at risk. Trends and challenges. In A Scientific Handout for the High Level Event of the Global Network Against Food Crises: Food & Agriculture in Times of Crisis, Brussels, 2–3 April 2019 (p. 48). Montpellier: Cirad. https://www.cirad.fr/en/news/all-news-items/articles/2019/ca-vient-de-sortir/ food-systems-at-risk-trends-and-challenges. Biovision Foundation for Ecological Development, & IPES-Food. (2020). Money Flows: What is holding back investment in agroecological research for Africa? www.biovision.ch, www.ipes-food.org. Browne, G. (2019). Microbe scientists are preparing us to eat in a post-plant world. One Zero. https://onezero.medium.com/microbe-scientists-are-preparing-us-to-eat-in-a-post-plantworld-245677693379. CFS-HLPE. (2019). Agroecological and other innovative approaches for sustainable agriculture and food systems that enhance food security and nutrition. http://www.fao.org/3/ ca5602en/ca5602en.pdf. DeLonge, M., Miles, A., & Carlisle, L. (2016). Investing in the transition to sustainable agriculture. Environmental Science & Policy, 55, 266–273. https://www.sciencedirect.com/ science/article/pii/S1462901115300812. Dury, S., Bendjebbar, P., Hainzelin, E., Giordano, T., & Bricas, N. (Eds.). (2019). Food Systems at risk: New trends and challenges. Rome, Montpellier, Brussels: FAO, CIRAD and European Commission. https://doi.org/10.19182/agritrop/00080. https://agritrop.cirad. fr/593617/1/Food_systems_at_risk.pdf. Encyclopedia.com. (2019). https://www.encyclopedia.com/food/ news-wires-white-papers-and-­books/green-revolution. Eyhorn, F., Muller, A., Reganold, J. P., Frison, E., Herren, H. R., Luttikholt, L., … Smith, P. (2019). Sustainability in global agriculture driven by organic farming. Nature Sustainability, 2, 253–255. https://rdcu.be/bvC9z. Gliessman, S. R. (2015). Agroecology: The ecology of sustainable food systems (3rd ed.). Boca Raton, FL: CRC Press/Taylor and Francis Group. https://doi.org/10.1080/21683565.2015 .1130765. Gunther, M. (2019). Science created factory farming. Science could end it. One Zero. https://onezero. medium.com/science-created-factory-farming-science-could-end-it-77f6463c7b56.

162

Rethinking Food and Agriculture

IAASTD, I.A. of A.K., Science, and Technology for Development. (2009). Synthesis report: A synthesis of the global and sub-global IAASTD reports, Agriculture at a crossroads. Washington, DC: Island Press. https://www.weltagrarbericht.de/reports/Global_Report/ Global_content.html. IPBES. (2018). In L. Montanarella, R. Scholes, & A. Brainich (Eds.), The IPBES assessment report on land degradation and restoration (p. 744). Bonn, Germany: Secretariat of the Intergovernmental Science-Policy Platform on Biodiversity and Ecosystem Services. www.ipbes.net. IPBES, & Díaz, S. (2019). In J. Settele, E. S. Brondizio, H. T. Ngo, M. Guèze, J. Agard, A. Arneth, … C. N. Zayas (Eds.), Summary for policymakers of the global assessment report on biodiversity and ecosystem services of the intergovernmental science-policy platform on biodiversity and ecosystem services. Bonn, Germany: IPBES secretariat. https://www. ipbes.net/system/tdf/ipbes_7_10_add.1_en_1.pdf?file=1&type=node&id=35329. IPCC. (2018). Summary for policymakers. In V. Masson-Delmotte, P. Zhai, H.-O. Pörtner, D. Roberts, J. Skea, P. R. Shukla, … T. Waterfield (Eds.), Global Warming of 1.5°C. An IPCC Special Report on the impacts of global warming of 1.5°C above pre-industrial levels and related global greenhouse gas emission pathways, in the context of strengthening the global response to the threat of climate change, sustainable development, and efforts to eradicate poverty (p. 32). Geneva, Switzerland: World Meteorological Organization. https://www.ipcc.ch/sr15/. IPCC. (2019). IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystems. Summary for policymakers approved draft. https://www.ipcc.ch/site/assets/ uploads/2019/08/Edited-SPM_Approved_Microsite_FINAL.pdf. IPES-Food. (2017a). Unravelling the food–health Nexus: Addressing practices, political economy, and power relations to build healthier food systems. The Global Alliance for the Future of Food and IPES-Food. http://www.ipes-food.org/_img/upload/files/Health_FullReport(1).pdf. IPES-Food. (2017b). Unravelling the food–health Nexus: Addressing practices, political economy, and power relations to build healthier food systems. The Global Alliance for the Future of Food and IPES-Food. http://www.ipes-food.org/_img/upload/files/Health_FullReport(1).pdf. IPES-Food. (2018). Breaking away from industrial food and farming systems: Seven case studies of agroecological transition. http://www.ipes-food.org/_img/upload/files/CS2_web.pdf. Khan, Z., Midega, C., Pittchar, J., Pickett, J., & Bruce, T. (2011). Push–pull technology: A conservation agriculture approach for integrated management of insect pests, weeds and soil health in Africa. International Journal of Agricultural Sustainability, 9, 162–170. https:// doi.org/10.3763/ijas.2010.0558. Kling, C. (2019). Polluting farmers should pay. Opinion, New York Times. https://www.envecon.net/2019/08/cathy-kling-.html. Majot, J. (2019). Think forward blog. IATP. https://www.iatp.org/ climate-change-and-new-­agricultural-system. Move for Hunger. (2019). About food waste. https://www.moveforhunger.org/about-food-waste/. Nature EDITORIAL. (2019). Counting the hidden $12-trillion cost of a broken food system. Nature, 574, 296. https://www.nature.com/articles/d41586-019-03117-y. Pengue, W., Muller,A., Sukhdev, P., Gemmill-Herren, B., Gundimeda, H., Hamm, M., … Sharma, K. (2018). Measuring what matters in agriculture and food systems: A synthesis of the results and recommendations of TEEB for agriculture and food's scientific and economic foundations report. https://www.researchgate.net/publication/325628199_MEASURING_WHAT_ MATTERS_IN_AGRICULTURE_AND_FOOD_SYSTEMS_A_synthesis_of_the_results_and_recommendations_of_TEEB_for_Agriculture_and_Food's_Scientific_and_ Economic_Foundations_Report.

Why change the way we grow, process, and consume our food?163

Pimbert, M. P., & Moeller, N. I. (2018). Absent agroecology aid: On UK agricultural development assistance since 2010. Sustainability, 10(2), 505. https://doi.org/10.3390/ su10020505. Poux, X., & Aubert, P.-M. (2018). An agroecological Europe in 2050: Multifunctional agriculture for healthy eating. In Findings from the ten years for agroecology (TYFA) modelling exercise, Iddri-AScA, study no. 09/18 (p. 74). Paris, France. https://www.iddri.org/sites/ default/files/PDF/Publications/Catalogue%20Iddri/Etude/201809-ST0918EN-tyfa.pdf. Ribeiro, S. (2018). Pathways out of the climate chaos. Climate crisis. Blog. https://grain.org/en/ article/6085-pathways-out-of-the-climate-chaos. Rockstrom, J., Steffen, W., Noone, K., Persson, A., Chapin, F. S., Lambin, E. F., … Foley, J. A. (2009). A safe operating space for humanity. Nature, 461, 472–475. https://www.nature. com/articles/461472a. Rogers, A. (2019). Trying to plant a trillion trees won't solve anything. Wired, science. 10.25.2019 https://www.wired.com/story/trees-regenerative-agriculture-climate-change/. Rouse, P. (2019). Could nature ‘solve’ climate change? C2G Blog. Carnegie council for ethics in international affairs. 22 October 2019. https://www.c2g2.net/ could-nature-solve-climate-change/. Steffen, W., Richardson, K., Rockström, J., Cornell, S. E., Fetzer, I., Bennett, E. M., … Sörlin, S. (2015). Planetary boundaries: Guiding human development on a changing planet. Science, 1259855. https://science.sciencemag.org/content/347/6223/1259855. Swinburn, B. A., Kraak, V. I., Allender, S., Atkins, V. J., Baker, P. I., Bogard, J. R., et al. (2019). The global syndemic of obesity, undernutrition, and climate change: The Lancet commission report. Lancet, 393(10173), P791–P846. https://www.thelancet.com/commissions/ global-syndemic. Swire, J. (2018). More than one farmer a week in the UK dies by suicide. FarmBusiness. 7th November http://www.farmbusiness.co.uk/news/more-then-one-farmer-a-week-in-the-ukdies-by-suicide-2.html. Terra Genesis International. (2019). Regenerative agriculture definition. http://www.regenerativeagriculturedefinition.com. Thomas, K. (2019). How fake meat could save the planet. One Zero. https://onezero.medium. com/how-fake-meat-could-save-the-planet-70e23b937e7b. UCS. (2008). Reports & multimedia/explainer. The hidden costs of industrial agriculture. https://www.ucsusa.org/resources/hidden-costs-industrial-agriculture. WikiPedia. (2019). Permaculture. https://en.wikipedia.org/wiki/Permaculture. Wise, T. (2019). Eating tomorrow: Agribusiness, family farmers, and the battle for the future of food. New York: The New Press. https://lccn.loc.gov/2018036746. Xue, L., Liu, G., Parfitt, J., Liu, X., Van Herpen, E., Stenmarck, Å., … Cheng, S. (2017). Missing food, missing data? A critical review of global food losses and food waste data. Environmental Science & Technology, 51(12), 6618–6633. https://pubs.acs.org/doi/ abs/10.1021/acs.est.7b00401?cookieSet=1.

This page intentionally left blank

Two paradigms of science—And two models of science-based agriculture

9

Colin Tudge Oxford Real Farming Conference and the College for Real Farming and Food Culture, Oxford, United Kingdom Colin Tudge argues that to put agriculture to rights we need to rethink everything we now take for granted from first principles, including the meaning and roles of science.

Through most of its long history (did it begin 10,000  years ago or 40,000-plus? Tudge, 1998) agriculture has been a craft, and for the most part it still is: rooted in experience, tradition, and folk knowledge and, at its best, immensely skilled and effective. From the outset, agriculture has made use of ever-more sophisticated technologies (tools of stone, wood, horn, and then of bronze and finally iron; and the power of wind, water, and animals) but only in the past couple of 100 years, a twinkling, has it made use of bona fide science—knowledge of the world based on measurements, experiments, and testable hypotheses. Out of this science has emerged an array of science-based technologies collectively known as high-tech (as opposed to the traditional artisanal technologies that grow out of craft). Justus von Liebig in Germany (1803–73) and John Bennet Lawes in England (1814–1900) in the mid-19th century were among the first true agricultural scientists and their rival versions of artificial fertilizer can be seen as the first examples of agricultural high technology. Now, agricultural practice is increasingly guided by the ideas of science, and increasingly dependent on high-tech. But the science that is now brought to bear is of two qualitatively different kinds—representing two different paradigms, which, at least for the purposes of this chapter, I will call Paradigm I and Paradigm II. Both have deep philosophical roots and huge implications—ecological, political, economic, social, psychological, moral, and indeed metaphysical. Paradigm I is increasingly and conspicuously high-tech and is what most people (including most agricultural scientists, big farmers, and policy-makers) take to be truly modern, and the basis of future policy. Paradigm II is more obviously rooted in traditional practice, less conspicuously high-tech, and at least until recent years has commonly been considered somewhat quaint, not to say retrograde. Here, though, I want to argue that Paradigm II represents true modernity, and is what the world really needs if we seriously hope to enjoy a long-term future on this Earth. Maybe the tide is turning, and perhaps it will turn in time to prevent the ecological and political collapse that is now beginning to seem inevitable. Thus:

Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00009-8 © 2021 Elsevier Inc. All rights reserved.

166

Rethinking Food and Agriculture

9.1 Paradigm I: The high-tech industrial route Paradigm I agricultural science is rooted in physics and chemistry and the “harder” forms of biology (physiology and pharmacology), and has given rise to the whole panoply of high technologies from biotechnology to IT and robotics. It is more or less taken for granted worldwide that we cannot produce all the crops we need without artificial fertilizers. Arable farmers employ an ever-expanding arsenal of pesticides, herbicides, fungicides, and every other -cide. Plant nutrition is treated as an exercize in bench-top chemistry giving absolutely predictable results, and animal nutrition aspires to the same degree of precision. Vets have a truly wondrous array of vaccines, drugs, hormones, salves, and balms to ward off all diseases and, if called upon, to control every aspect of the animals’ physiology. Plant and animal breeding become more and more precise and have been increasingly abetted first by direct manipulation of chromosomes (which provided the semidwarf cereal varieties of the Green Revolution) and then by genetic engineering and gene editing. Artificial insemination has long been with us, and now the most productive—“elite”—animals may be cloned by separating the cells of the young embryos and spread throughout the world by embryo transfer. Entire prairies of immaculate cereals are harvested with combines as big as a small cottage that can operate 24 h a day, and do not need drivers. At the other end of the scale ever-smarter robots do the fiddly stuff. The state of the farm can be and is monitored by satellites and drones. Some of the most “advanced” farms already operate virtually without human labor. For some, zero-labor agriculture is the ideal, like a modern production line run by robots. With all such technologies there are huge advantages of scale. As Earl Butz (1909–2008) advised America’s farmers—Butz was the US Secretary of State under Presidents Eisenhower, Nixon, and Ford—“Get big or get out!” There are arable farms in Ukraine and elsewhere bigger than some English counties. In the United States in particular, animals are now raised in “concentrated animal feeding operations” (CAFOs)—tens of thousands or even (pigs and poultry) a million at a time. Such agriculture can truly be called “industrial,” comparable in scale, ingenuity, and efficiency (although, as outlined later, it depends how efficiency is measured!) with the world’s greatest industries—electronics, motor cars, ship building, or what you will. Advocates of Paradigm I farming claim both that it is wonderfully successful and that it is unequivocally necessary. Certainly, the statistics are impressive. In Britain, a reasonably typical modern country, cereal yields now average more than 8 ton per ha— three times what was considered normal less than a century ago. Dairy cows in the 1950s were typically expected to yield around 3000 L (600 gal) of milk per lactation (which in effect means per year), while today’s farmers expect 6000 L. Many “elite” cows produce 10,000 L a year and some agricultural scientists dream of 20,000 L (4000 gal), with the cows milked more or less continuously by robot. The UN convened the first ever World Food Conference in 1974 in the wake of famines in Asia and Africa. The world population then was around 4 billion and was on course to double every 40 years or so and there were real fears that we would not be able to keep up. Thomas Robert Malthus’s prophecy from the early 19th century that humans would 1 day outpace their food supply seemed to be coming true. Yet present-day agriculture now supports 7.5 billion. If farming did not supply enough to feed all those people, at least up to a point, they would not be here.

Two paradigms of science—And two models of science-based agriculture167

Unquestionably, too, the underlying science is wonderful. To a biologist, the insights of molecular biology, which lie behind genetic engineering and editing, are thrilling. Some of the potential applications in all fields from medicine to animal and plant conservation are world changing—potentially for the better. But, of course, all is not so simple. Nothing is for nothing. There are serious downsides. Modern high-tech, large-scale agriculture is clearly unsustainable. It is almost entirely oil dependent, for power and transport (increasing as farming becomes more globalized) and as raw materials for agrochemicals. Runoff from agrochemistry is a major pollutant, not least of the oceans—right now posing a huge and obvious threat to Australia’s Great Barrier Reef, one of the world’s great natural wonders. CO2 from the burning of oil, methane from cattle, and oxides of nitrogen from heavily fertilized fields are major contributors to global warming. The Food and Agricultural Organization of the UN now calculates that a third of all agricultural soils worldwide are seriously degraded or eroded (FAO, 2019). In 2018, Secretary of State Michael Gove told delegates at the launch of the Sustainable Soils Alliance that some arable fields in Britain may have only 30 or 40 worthwhile harvests left in them (Alliance Website, n.d.). Modern agriculture accounts for 80% of freshwater use. Forests are felled wholesale to make way for commodity crops—another contributor both to global warming and to the mass extinction of wild species that is now well in train, and deforestation everywhere is a prime cause of flooding. At least a third of all land, including the most naturally fertile, is now used for agriculture, which, in industrial form, is increasingly hostile to wildlife—and loss of habitat is the prime cause of mass extinction. We could well be on the brink of the ecological collapse that Jared Diamond foresaw in Guns, Germs, and Steel (Diamond, 1997) and industrial agriculture in its present forms is a prime contributor. Of course, industrial agriculture does not operate in a vacuum. Nothing does. Agriculture in all its forms affects everything else that we do and is affected by everything else. In general, high-tech of every kind requires high investment and also promises high returns, so high-tech farming attracts big business. Big business and high-tech in combination lead to globalization and control by transnational corporates, which have the wealth and power to bestride the world. The corporates in turn subscribe to the rules and conventions of the form of capitalism known as neoliberalism, which is focused on the global market. Indeed, what is now perceived to be modern agriculture, and indeed to be the farming of the future, might more accurately be called Neoliberal-Industrial, or NI, agriculture. Modern governments the world over increasingly rely on corporates for their revenue. The old adage “What’s good for General Motors is good for the USA” is now in effect deemed to be true of all corporates in all countries. The global market is above all competitive (barring cartels and trade agreements). All participants in the market, which in principle at least may include all producers and traders of every kind, are required or obliged to compete head to head with all the others for profit and market share, by which they hope to attract more investment. The whole operation is entirely driven by money. The goal for all participants is and must be to maximize their own wealth. Either that or, as Earl Butz advised, “Get out!” Many argue that wealth per se must be good (Anon, n.d.). Of course, this is not necessarily

168

Rethinking Food and Agriculture

so and even if it were the means of making it are often destructive, and the inevitable concentration of wealth and hence of power as the less successful competitors fall by the wayside has obviously sinister connotations. We may ask, too, whether industrial farming is really as successful as it may seem. The UN tells us that nearly a billion of the 7.5 billion people who are now with us are chronically undernourished. The UN also tells us that more than a billion suffer from “diseases of affluence,” which means they eat too much of the wrong things (and perhaps at the same time are short of some of the right things). They may be seen as victims of a market that has virtually been given carte blanche to produce and sell whatever is most profitable in the short term, with no commensurate duty of care. The world population of diabetics, mostly diet related, now exceeds the total population of the United States (by some margin). Most cogently we may ask whether the ambitions, the goals of the high-tech industrial agriculturalists and big-business farmers, are really what the world needs. The science is brilliant, indeed, but is it well applied? The technologies are often breathtaking, but are they truly appropriate? Emphasis now is still on production, production, production. Population is still rising—the UN tells us that there could be 10 or 11 billion of us by the end of this century, and since a billion are already going hungry it seems at first sight as if we do indeed need more. A Food and Agricultural Organization report of 2009, echoed in 2011 in a report by the British government, told us that the world will need 50% more food by 2050 just to keep pace with rising numbers and increasing “demand” for meat (FAO, 2011). It seems to many, just as it did in the 1970s, that Malthus was right. Yet, Professor Hans Herren, President of the Millennium Institute in Washington, DC, points out that the world already produces enough food for 14 billion people—almost twice the present population and about 30% more than we should ever need since the UN tells us that although absolute numbers are still rising the percentage rate of increase is slowing down and by 2100 or so numbers should stabilize (at 10–11 billion). These figures are easily ratified. For the world now produces around 2.5 billion tons of cereal; 1 ton provides enough macronutrient (protein and energy) to sustain three people for a year, so 2.5 billion tons is enough for 7.5 billion people; and cereals provide only half of our food. The other half comes from noncereal grains, pulses, nuts, fruits, vegetables, meat, eggs, dairy, and fish, which between them also provide the bulk of the micronutrients (and the vital flavor and texture). Shortages and famines result in the main not from lack of capacity but from wars, income inequality, and because the wrong kind of food is produced in the wrong places. In particular, farming to feed local people has increasingly given way to commodity crops grown for export. Emphasis now should surely shift from production, production, production to food quality and provenance (including animal welfare and ecological wellbeing). This means we could, if we chose, take the heat out of farming; take the foot off the accelerator; and get off the treadmill of ever-increasing production to provide ever-­ increasing accountable wealth. We may question, too, whether NI agriculture is really as “efficient” as its advocates claim. In truth, the definition of efficiency in this context is a tautology. Efficiency is commonly measured as output-per-worker, and since NI agriculture to a large extent is designed to increase output while reducing labor (to reduce costs, at least in the

Two paradigms of science—And two models of science-based agriculture169

short term, in the present economic climate) it is bound to become more “efficient” in its own terms. However, if efficiency is measured in terms of food energy produced versus fuel energy expended, then NI agriculture is extremely inefficient. More generally we might ask whether it is really such a good idea to eliminate human labor, and to call this “efficiency.” If India, say, with more than half its people working on the land, emulated the farming of Britain, with around 1% full-time farmers, half a billion people would be robbed of their livelihood. No other industry can usefully employ people on such a scale as farming, and unemployment and all that goes with it are among the prime causes of human misery and discontent. The UN also tells us that about a billion people worldwide live in urban slums and although I know of no formal surveys on the matter (though they surely exist) it’s a fair bet that a large proportion of those slum dwellers are ex-farmers or their dependents or immediate descendants. Even if we accept (which we surely should not) that the value of human enterprise and of the biosphere must be measured primarily in money, wholecost accountancy—to include the cost of all the untoward side effects and collateral damage—would surely show that NI agriculture tends to be very profligate indeed. In short, the world clearly needs to change direction, if we really care about the state of humanity and the natural world and have any hopes of a long-term, tolerable future. High-tech, NI agriculture is not delivering what the world really needs and certainly not what will be needed in the decades to come. Yet, those who urge a shift away from high-tech industrial market-driven agriculture are commonly accused of being antiprogress, retrogressive, and “unrealistic.” They are also accused of being antiscience, which is seen to be a sin. In truth, however, some of the principal detractors from the status quo are scientists themselves who feel that science, and the high-tech to which it gives rise, are being seriously misused and abused. They urge instead a different form of science—the kind that I am calling Paradigm II. Paradigm II science leads us to embrace forms of agriculture that look traditional and indeed lean heavily on tradition, often invoking (or reinventing) ancient methods, and always consulting local farmers. So, to the uninitiated this alternative approach may seem to be old-fashioned. Yet, I suggest, Paradigm II agricultural science is far more modern, conceptually, than the flashier, high-tech kind that is seen to represent progress. Indeed, the Paradigm I, high-tech kind, firmly rooted in the ultrarationalist zeal of the 18th century, should now be seen to be seriously, conceptually, out of date.

9.2 Paradigm II: The path to enlightened agriculture The antidote to NI high-tech, high-capital agriculture is what I for about a decade or so have been calling Enlightened Agriculture (EA) sometimes known as “Real Farming.” EA is not, like NI farming, geared to the global market and hence required to compete with everyone else to maximize profit. Instead, it is loosely but adequately defined as: Farming that is expressly designed to provide everyone, everywhere, forever, with food of the highest quality, nutritionally and gastronomically, without cruelty or injustice and without wrecking the rest of the world.

170

Rethinking Food and Agriculture

Technically—given that we now produce far more food than we really need, and the world population should soon start leveling out—this ought to be well within our grasp. Certainly, I suggest, it should be our ambition. Although the expression “Enlightened Agriculture” is new (and I lay claim to it (Tudge, 2011)) the concept is based on three ideas that were developed by others and are now widely established, at least among those who take an interest. These ideas are: Agroecology Food Sovereignty Green Economic Democracy

The concept of Food Sovereignty was developed in the 1990s by the global peasant movement La Via Campesina. The details and ramifications inevitably are complex but it means, in essence, that all societies should have control of their own food supply. “Economic Democracy” is a vaguer term that again has endless ramifications but it means in general what it says: that the economy, local, national, and global, should, like democracy itself, be “of the people and for the people.” The additional “Green” means that all economies must recognize ecological limits, which right now for the most part they do not. However, in this chapter I want to concentrate on the first ingredient of EA: agroecology.

9.3 The absolute importance of agroecology The term “agroecology” was first coined in the 1920s and although, inevitably, different people interpret in different ways, I take it to mean that: All farms should be treated as ecosystems and agriculture as a whole should as far as possible make a positive contribution to the diversity, abundance, and general vitality of the biosphere.

In truth, since agriculture is designed to make other animals and plants do things they would not otherwise do, for our benefit, it seems bound to do more harm than good to the natural world. But even if farming cannot enrich the natural world, we should at least seek to minimize the damage (a point emphasized in the Jain religion). Agroecological farming in effect seeks to emulate nature. Of course, it does not do this slavishly. Nature does many things that we would not seek to emulate even if we could, including mass wipeout at intervals by tsunamis or volcanoes and the rest. What really matters, though, is that we need our agriculture not only to be productive (albeit not maximally so, but enough) but also to be sustainable (able to keep going) and resilient (able to bounce back after disaster or change direction). For this, nature

Two paradigms of science—And two models of science-based agriculture171

is a wonderful model. Earthly ecosystems have been commendably if not always maximally productive without interruption for nearly 4 billion years, despite enormous changes of climate and the peregrinations of the continents through plate tectonics. So, we might reasonably ask, how does nature manage this? The answer seems to be that nature is: Maximally diverse. Integrated (all the species interact with each other—to some extent competitively but more importantly synergistically, and with net advantage to all) with. Minimal external input (and, in particular, no fossil fuels!)

In practice, the kind of farming—agroecological farming—that results when we apply these principles is, in important respects, diametrically opposite to the NI agriculture that is now beginning to dominate the world stage. Thus in practice NI agriculture leads us not toward diversity but toward homogeneity. Elite crops are cloned or at least are highly homozygous to make sure that all the individual plants conform as closely as possible to the specification. Agroecology in contrast leads us to favor genetically diverse crops and livestock, sufficiently uniform to give reasonably consistent results but otherwise as diverse as possible. Thus some avant garde agroecological farmers are growing mixed populations of cereals, deliberately mixing different varieties or ancient varieties with modern. The genetic and phenotypic homogeneity of modern crops and livestock leaves them more vulnerable to pests and diseases: a parasite that gets a hold on any one individual can invade them all. If the population is mixed the parasite constantly faces new challenges. The need to minimize external inputs (not least of industrial chemistry) leads us toward organic farming. Organic farming at least should be the default position; what farmers naturally do unless they have very good reason to do something else. Machines on the whole are more fuel efficient when they are big. Big machines need big spaces to work in. Furthermore, despite the influx of “smart” technologies, machines on the whole do not cope well with complexity. They work best when they have just one thing to do at a time, and can operate on the biggest possible scale. Hence, high-tech, machine-dependent farming favors monoculture—uniform crops of cereal as far as the eye can see; vast herds of more or less uniform livestock. Supermarkets too demand uniformity—produce specified with the precision of an engineer. The principles of agroecology in contrast lead us to favor mixed farms (but see Addendum I, later!): different crops and different classes of livestock grown synergistically. The mixed farms that were typical of Britain in the 1950s were a good example, typically combining some grazing with some arable with dairy and pigs, and some horticulture. The animals fertilized the fields; the whey from cheese making was a fine nutritional boost for the pigs; and so on. The still extant farms typical of South-East Asia are the supreme example. Nitrogen-fixing cyanobacteria (Anabaena) between the cells of floating ferns (Azolla) provide extra fertility. Carp and ducks swim in the flooded paddy fields eating whatever grows between the cereal stems and

172

Rethinking Food and Agriculture

help to fertilize the rice. The higher ground is for horticulture—everything from yams to cabbage. Pigs and chickens sweep up the debris in the villages. Water buffalo pull the plows and generally add to the mix. All this means that agroecological systems are complex—designedly so: the absolute antithesis of industrial monoculture. Such complexity can be maintained only by thinking and skillful human beings. Agroecological farming must be skills intensive, meaning plenty of farmers and growers. In Rwanda at least until recently up to 90% of the workforce worked full time on the land, which we can all agree is far too many. This leaves too few people to do all the other jobs that need doing, and virtually no home market for the farmers to sell to so they are doomed to subsistence. On the other hand, the roughly 1% of people who now work full time on the land in the United States and Britain is very obviously far too few, if, that is, we truly aspire to provide agriculture that is sustainable and resilient. It is important to work out the proportion of land workers that would be ideal in any one country but so far as I know this is not being formally addressed. It is simply assumed that the fewer the better, with zero labor as the ideal. That is called progress. The fact that such systems cannot be sustained, and do so much damage, does not seem seriously to be considered. To replace human skill with machinery is deemed ipso facto to be progress. Why it should be considered good to make the human species obsolete is not at all obvious. In general, if any enterprise is complex and skills intensive, there is little or no advantage in scale-up, and usually much disadvantage. Therefore the ideal, agroecological farm of the future should be small to medium sized. This again is the precise opposite of the standard advice from on high: to expand and to merge small farms into big (monocultural) estates. Small farms, however, commonly operate best, to everyone’s advantage, when they form cooperatives. Cooperatives require cooperation: again the precise antithesis of the neoliberal competitiveness that is now considered de rigueur. There is no space to discuss them here but it is at least intuitively obvious that the principles of Food Sovereignty and Green Economic Democracy, the other components of Enlightened Agriculture, also lead us very clearly to favor small, mixed, preferably organic farms. Such farms too at least at times have formed the economic and social center of some of the most agreeable human societies in which everyone has a role, and although there is rivalry (of course) at least when it matters (in times of harvest or hardship) everyone pulls together. Rural life is not always like that, of course, but it certainly can be as has at times been seen the world over. It is a model worth emulating. One last and obvious caveat, however. Can agriculture based primarily or largely on small mixed organic farms “feed the world”? Advocates of the large-scale high-input kind claim that it obviously cannot. Industrial agriculture seems to be far more productive than the more traditional kinds. Most obviously without high inputs of nitrogen fertilizer we could not achieve modern yields of wheat—commonly exceeding 8 and sometimes more than 12 ton per ha. Neither would it be possible simply to add more N to traditional varieties, for they are tall plants with long stems and simply fall over (lodge) if heavily fertilized. Yet, as noted earlier, the world does not actually need such fabulous yields. About half the cereal that is now produced is for animal feed, yet we could produce all the livestock we really need without growing cereal for them at all, or very little, by

Two paradigms of science—And two models of science-based agriculture173

f­ eeding cattle and sheep on pasture and pigs and poultry on leftovers (swill) and surpluses. This would not provide as much meat as now but it would certainly provide all the animal protein that we actually need (if it were produced in the right places and more evenly distributed). Even more to the point, since all the greatest cuisines from Italy to China use meat sparingly (as garnish, stock, and for occasional feasts) such traditional feeding would provide enough to support the finest gastronomy. Finally, an increasing body of literature is showing that small, skills-intensive units, with plenty of tender loving care, integrating livestock where possible (there is always room at least for chickens) can be the most productive of all in terms of food per unit area. It is now abundantly clear, too, that most people the world over depend on smallscale, traditional farmers for their food. Indeed, a report from the ETC Group (2017) tells us that what they call the Peasant Food Web “feeds 70% of the world, including the most hungry and marginalized people.” All that, and such farms are clearly sustainable and far more wildlife friendly than the industrial kind. Enlightened Agriculture based on agroecology need hold no fears. Equally obviously, though, Enlightened Agriculture does not fit comfortably into the prevailing, neoliberal, economic paradigm, or into the general mindset, which includes Paradigm I science. So it is that peasant farmers the world over typically receive little or no help from their governments and indeed are often actively done down to make way for the supposedly more modern (and profitable!) industrial kinds. Yet, in Europe and the United States, supposed champions of the free market, industrial farmers are heavily subsidized with taxpayers’ money. Small-scale, peasant farming needs good science just as the industrial kind does, but it needs the science of agroecology: Paradigm II science, including the ever-more complex subtleties and insights of organic farming. In general, however, while Paradigm I science and its technologies are showered with grant money from big governments and industry, organic research tends to rely on the initiative of individual farmers and voluntary donations. Clearly, if we are serious about the future therefore it will not be enough simply to redesign agriculture. We need also to devise a new economy geared not to the crude multiplication of measurable wealth and the dominance of the rich, but to the real needs of humanity and the biosphere. To achieve this, we need new kinds of governments that see why this is necessary, and such governments will come to power only when and if people at large wake up to the need for wholesale change. In short, nothing can be put to rights ad hoc, and if we want Enlightened Agriculture to come about, and indeed to become the norm, then we need a complete, cross-the-board rethink.

9.4 A cross-the-board rethink What is more important—to secure the future of humanity and the biosphere, the natural world, or to protect the ideologies of neoliberal economics and top-down governance? If we believe the former is what really matters, then the choice seems open and shut. We must take agriculture very seriously, and must organize it and run it along agroecological lines.

174

Rethinking Food and Agriculture

But we cannot hope to do what is needed within the present, neoliberal economy. The problem is not that the present economy is capitalist and depends on markets. Many left-wing thinkers, including Mikhail Gorbachev no less, acknowledge that there is a place for free enterprise and for markets (Tudge, 2013). What really matters is the underlying mindset. We need to question and to expunge the moral and metaphysical conceit that says that material acquisition per se is good, that competition is good, that we ought to compete one with another for wealth and dominance, and that humanity as a whole has a right and even a duty—“man’s destiny”!—to grow richer and dominate the Earth and beyond that the galaxy (in the dreams not only of fiction writers but also some scientists and some gung-ho politicians). The mindset too includes Paradigm I science. This appears ultramodern (genetically modified organisms [GMOs] and gene editing, smart robots, remote sensing, and the rest) but it is rooted in ideas that are seriously old-fashioned. Indeed, Paradigm I science is rooted in the conceits of the 18th century Enlightenment. Essentially, the idea was and is that by applying strictly rational thought we can achieve exhaustive knowledge and understanding of life and the universe. The epitome and apotheosis of rational thinking was and is science, based on observation, measurement, analysis, and (although this was a later idea) on testable hypotheses, with maths as the overall guide and arbiter. Maths was and to a large extent is deemed to be infallible since the working is plain to see and all its errors can be corrected. Such science was and is entirely materialistic, for only the material world can be subjected to such rigor. For good measure, the logical positivists of the early 20th century declared that any idea that could not be subjected to such rigorous examination was not worth considering at all, which immediately consigned moral philosophy and metaphysics (which meant all religion), and nearly all traditional wisdom, to the realms of fantasy or indeed to the dustbin. In short, Paradigm I science seemed and seems to some to promise absolute certainty and virtual omniscience, and the high-tech that science gives rise to seems to promise omnipotence. Such powers are god-like and gods can make their own rules, can they not? However, the modern philosophy of science tells us that life is not so simple. As the zoologist Peter Medawar observed in the mid-20th century, science is and can only ever be “the art of the soluble.” Scientists perforce must tackle only those problems they think they have a chance of solving with the tools available to them, and it’s a huge leap to suggest that what can readily be studied and mathematicized is all there is. Karl Popper showed that absolute proof of ideas of a material kind is impossible. Only disproof is absolutely possible. Kurt Godel showed that mathematical equations of a complex kind always contain assumptions that cannot be proven, so even maths is to some extent “subjective” (Tudge, 2013). The material universe turned out to be innately unpredictable in detail because of the random motions of the fundamental particles of which it is composed and because we can never, even in theory, know all the factors that are at work in any one context. Thus has arisen the concept of “nonlinearity”: there is no simple relationship between cause and effect, except in the abstract. Paradigm II science takes account of all this. It does not suppose that science delivers certainty or that science tells us all we need to know. In its applied forms it accepts the principles of nonlinearity and unpredictability. In its general approach indeed it is altogether more humble than the Paradigm I kind.

Two paradigms of science—And two models of science-based agriculture175

Agroecology is rooted in the science of ecology, which attempts to engage with the realities of the real world: accepting the inescapable limits of our own knowledge and our abilities to investigate and accepting that in the end nature is beyond our ken and cannot be reshaped to fit political and economic ambition. Ecology is archetypal Paradigm II science. Ecologists acknowledge, as physicians (and engineers!) have always acknowledged, that our knowledge of the real world must always be partial and uncertain. For practical purposes, however, we have to be prepared to act despite this uncertainty. But, very obviously, we must always proceed with extreme caution, accepting that the real world is always likely to spring surprises. The superconfident gung-ho approach of Paradigm I science and its political and commercial enthusiasts puffed up with high-fallutin’ rhetoric is, all too demonstrably, very dangerous. Given that the far more benign alternative is already well established, it is also ridiculous. Far from being nonscientific, or retrogressive, Paradigm II science and the approaches to which it gives rise, including agroecology, is far more modern than the gung-ho Paradigm I science that now prevails and has its roots in the 18th century. GMOs, smart robots, and all the rest are flashy (and may in appropriate circumstances even be turned to good use) but conceptually they are seriously old-fashioned. Everyone should learn science. But science should never be taught without the philosophy of science. As present-day industrial agriculture illustrates, gung-ho Paradigm I science leads us seriously astray.

9.5 Addendum I: What price livestock? Do we really need farms to be “mixed”? Various individuals and lobbies have taken issue for various reasons with various aspects of the thesis presented here, and in particular with the idea, rooted in agroecological principles, that farms should ideally be mixed, with both plants (crops) and animals (livestock). In particular, advocates of NI agriculture take it to be more or less self-evident that farms must operate according to the tenets and norms of the “free market,” and that therefore they must above all seek to maximize profit, and that to achieve this they must be as large as possible and high-tech, with minimum labor, and that they are most likely to succeed in this if the structure and modus operandi are simplified, which leads us toward monoculture rather than toward the complexities of mixed farming. Industrialists qua industrialists do not care whether the monocultures are of plants (wall-to-wall arable or horticulture) or of livestock (the CAFO) so long as they maximize financial returns. A quite different lobby—that of the vegans—questions whether, whatever the system of farming, we should raise livestock at all. But although the vegans present many different arguments, all of which have some cogency, I still argue as most people worldwide seem to agree that it is worthwhile to keep at least some animals.

176

Rethinking Food and Agriculture

The vegan arguments are of three main kinds: Moral/metaphysical Human need and human health Ecological

Here is a (very brief) critique of each:

9.5.1 The moral metaphysical arguments Vegans commonly argue that livestock farming is cruel, which is a moral argument, and that we have no right to impose our will so decisively on other creatures, which is a moral argument with strong metaphysical connotations. These I think are the vegans’ strongest arguments. Very obviously we should not be cruel to animals any more than to people and, many argue, any restriction of their natural lives is an imposition that ipso facto borders on cruelty. I am sure, too, that we have no God-given right to use animals for our own purposes without a by-your-leave. The best defense against these arguments (or the best that I know of) is the one that small children sometimes resort to when reprimanded: “I did not ask to be born!” Indeed, we did not. But we are as we are and—despite protestations to the contrary— we have the physiology of omnivores. Like our closest relatives, the chimpanzees, we demonstrably benefit at least from some meat. Furthermore, even if we did elect to live on an all-plant diet and devise an all-plant agriculture, we would still, perforce, impinge on our fellow creatures, by depriving them of natural habitat. Every human being that elects to live at all has at least some blood on his or her hands. In practice it is very hard to improve on the Buddhist/Jainist ethic. This acknowledges that human beings have to live (now that we are here) and that we must do whatever is needed to keep ourselves alive and in health. But it also adds, commonsensically, that we should do our best to minimize the harm we do to other species and to the natural world in general. It seems to be widely accepted that a well-run farm with some livestock produces more food per unit area than an all-plant unit, so mixed farming should enable us to produce all the food we need on less land than we would need with an all-plant agriculture and so should leave a smaller ecological footprint. In short: keeping livestock is not morally ideal, but sometimes (as the Dalai Lama points out) the ideal is not possible and we are obliged simply to do the least bad thing. It seems to me, all things considered, that in our attempts to feed ourselves the least bad option is to include some livestock, provided the animals are managed in ways that complement the cultivation of crops, or to make good use of land that would otherwise be wasted, and are treated as kindly as possible.

9.5.2 Human need and human health The fact that some vegans are among the longest-lived human beings and include some of the most intelligent and in various senses “enlightened” people there have

Two paradigms of science—And two models of science-based agriculture177

ever been is proof enough that an all-plant diet is possible (Leonardo, Tolstoy, and Gandhi are commonly cited examples). It is further pointed out that we do not in fact need the high doses of animal protein that have in recent decades been considered essential. Indeed, it is often argued that excess protein and, in particular, excess animal fat are positively harmful. However, meat and offals do provide essential nutrients that are not always easily obtained from plants, such as zinc and calcium. People who live in the dry lands of Africa, for example, would very likely be severely malnourished if they had to rely on plants that grow locally. In short, although people can live well on an all-plant diet and although excesses are always likely to be bad (“moderation in all things,” said Ralph Waldo Emerson) a world with some livestock is safer, for humans, than one with none at all.

9.5.3 Ecology The ecological arguments against livestock are of various kinds: 1. Livestock can be hugely profligate. Thus: today’s (largely industrial) livestock consume about half the cereal that the world produces and more than 90% of the soya. If it were not for this no one would be talking about overpopulation. In theory at least it would be far easier to feed everyone if we all settled for a low-meat diet and indeed were vegetarian and preferably vegan.

Furthermore, it is at least a reasonable rule of thumb that land devoted to cereals or pulses typically yields about 10 times as much protein per unit area as it would if devoted to livestock. All this may be true, but the profligacy of modern livestock is not the cause of food shortage or famine. Eight hundred million still go hungry because—for political/economic reasons—the wrong foods (including commodity crops intended for animal feed) are grown in the wrong places for the wrong purposes (i.e., to maximize the profits of people who typically live far from the farms from which they extract wealth). Also: the profligacy has to do with excess, not with livestock per se. Human beings thrive on low-meat diets, which could be provided without feeding them with vast amounts of staple foods that human beings could perfectly well eat themselves (with appropriate cookery). If ruminants are raised only as was traditional on grass and other herbs and browse, and if pigs are raised exclusively on surpluses and leftovers (as was traditional), then there would not be much meat overall but there would be enough for sound nutrition and excellent cooking (given that the world’s greatest cuisines from Italy to China are all low meat). Furthermore, crops produce far more protein per unit area than livestock only if they are grown on fertile land that can be cultivated. About five-sixths of Britain’s “agricultural” land is grassland, including rough grazing, in places where it would be very difficult to raise crops at all, or to harvest them. 2. Livestock are prime contributors to global warming. In particular, cattle are blamed for producing a lot of methane, which is a potent greenhouse gas.

This seems to be true only if cattle are raised by “modern” NI means—the effect exacerbated by their high-grade cereal/soya diets raised with lashings of oil-based fertilizer and transported across the world, which adds a burden of CO2 and (from the overfertilized fields) of oxides of nitrogen.

178

Rethinking Food and Agriculture

Some evidence—although much more research is needed—seems to show that pasture grazed by well-managed cattle and sheep can be a net carbon sink, sequestering more carbon than is released. The mechanisms do not seem to be exhaustively understood but increased photosynthesis must play a part as the animals fertilize the herbage and, by their grazing, encourage new growth to replace the old. (Certainly in the Miocene through to the Pleistocene there were billions of megaherbivores worldwide, including many ruminants, and the world grew steadily cooler, culminating in the Ice Ages. The cause of this was mainly cosmological but the grazing clearly did little or nothing to halt the trend.)

9.5.4 In conclusion There is a very strong case for not raising livestock in huge numbers by NI methods but, with a few conditional clauses, including the need for high standards of welfare, there is no good case for banning livestock altogether. The world is better off with some livestock than with none at all. Mixed farming is not always the most appropriate but it should be the norm, the default position: what farmers do unless there is a very good biological reason for doing something else.

9.6 Addendum II: The college for real farming and food culture Around 2008 a group of us centered in Oxford started the Campaign for Real Farming to develop and promote the ideas of Enlightened Agriculture. Out of this in 2010 came the Oxford Real Farming Conference (ORFC), which for the past few years has attracted 1000 delegates of whom about half are farmers. At about the same time I conceived the idea of the College for Real Farming and Food Culture to continue the spirit of the ORFC throughout the year. So far—always in collaboration with likeminded institutions—we have run seminars and courses under the college banner on various aspects of Enlightened Agriculture. We also have a website (https://orfc.org. uk) although this is now in serious need of revision. We are not looking for real estate but the ideal would be to have a permanent establishment with a model farm to explore all that needs exploring from the details of husbandry to economics and politics to science and philosophy thereof to morality and metaphysics. Only such a cross-the-board rethink will do. There’s a lot of ground to make up. And fast.

References Alliance Website, n.d., See the sustainable soils alliance website. https://landworkersalliance. org.uk/. Anon n.d., President J F Kennedy used to speak of economic growth as “the rising tide that lifts all boats”. In an interview on London Weekend Television’s Weekend World Sunday Jan 6 1980 British Prime Minister Margaret Thatcher told interviewer Brian Walden that

Two paradigms of science—And two models of science-based agriculture179

“­ No-one would remember the good Samaritan if he’d only had good intentions; he had money as well.” Evidently she felt that wealth was a necessary precondition for doing good. Diamond, J. (1997). Guns, germs, and steel. W W Norton. ETC Group. (2017). Who will feed us? The peasant food web vs the industrial food chain. ETC Group. FAO. (2011). Feeding the world in 2050. November 2009 In The warning that we need 50% more food by 2050 was repeated in the British Government’s “Foresight: The future of food and farming”: The Government Office for Science. FAO. (2019). Stop soil erosion, save our future. In Report from the global symposium on soil erosion. Rome: FAO. Tudge, C. (1998). Neanderthals bandits and farmers. Orion. Tudge, C. (2011). Good food for everyone forever. Pari Publishing. Also: Six Steps Back to the Land. Green Books. 2016. Tudge, C. (2013). Why genes are not selfish and people are nice. Floris.

This page intentionally left blank

Paradigms of agriculture Amir Kassama and Laila Kassamb a University of Reading, Reading, United Kingdom, bAnimal Think Tank, Lancaster, United Kingdom

10

10.1 Introduction Awareness and questioning of the ecological and ethical foundations and consequences of our global food and agriculture system is rising. Farmers, consumers, and civil society are increasingly shifting toward a more holistic understanding of how and what we produce and consume, and the implications of these for quality of life, our relationship with nature, and those with whom we share the planet. Farmers globally are becoming increasingly aware that conventional industrial agricultural at any scale is ecologically and biologically degrading and thus environmentally, economically, and socially unsustainable. Farmers are also becoming increasingly aware of alternatives, which are being implemented by millions of small- and larger-scale farmers in all continents and types of land-based agroecologies, and are largely driven by farmers and their associations. Consumers and civil society are becoming more aware of the links between the food and agriculture system and environmental degradation and destruction, pollution and agrochemicals entering human and animal food chains. It is becoming increasingly clear both how vulnerable conventional industrial agriculture is to climate change and how much of a driver it is. Instead of mitigating climate change, agriculture and related land use contributes around 24% of global greenhouse gases (GHGs), the majority of which come from animal agriculture (Shukla et al., 2019). Furthermore, mounting evidence that animal products in our diets are a major cause of the obesity pandemic and the steady increase in noncommunicable “lifestyle” diseases such as diabetes, cancer, and heart disease is challenging the narrative on what is promoted as healthy (Willett et al., 2019; Chapter 15). Awareness is also increasing regarding the systemic injustices against both humans and other animals, linked to the current dominant industrial food and agriculture system. Analysts and development stakeholders are increasingly realizing that the world already produces enough food to feed more than the projected global population of 9.7 billion in 2050 (Berners-Lee, Kennelly, Watson, & Hewitt, 2018), or even up to twice the number of people currently on the planet (Berners-Lee et al., 2018; Buff, 2017; Holt-Giménez, Shattuck, Altieri, Herren, & Gliessman, 2012). This has led to increased questioning of the dominant mainstream narrative, pushed by corporations, international development institutions, and philanthropic organizations with vested interests, that we need to increase production through industrial agriculture, genetically modified organisms (GMOs), and more agrochemicals to “feed the world.” The

Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00010-4 © 2021 Elsevier Inc. All rights reserved.

182

Rethinking Food and Agriculture

e­ ffectiveness of the structural and governance institutions that are charged to oversee and guide the food and agriculture system’s development and sustainability are also being increasingly questioned. It is becoming clearer that our food and agriculture system is not working for the vast majority of humans, other animals, and ecosystems, and that an alternative system is needed. In the introduction to this book we proposed the need for a food and agriculture system guided by the concept of “inclusive responsibility.” Such a system, among other things, would “encourage society to focus on agroecological sustainability as an integral part of overall ecosystem sustainability based on planetary boundaries.” To achieve agroecological sustainability we need to shift from the present paradigm of industrial agriculture to one that is ecologically sustainable and contributes to the foundation of economic, social, and environmental sustainability. In this chapter we explore what such an ecologically sustainable agricultural paradigm might look like. We start by looking at the origins and impacts of the current dominant paradigm of conventional industrial agriculture. We then review some of the alternative agricultural paradigms that have developed alongside or in response, namely Organic Agriculture, Agroecology, Regenerative Agriculture, and Conservation Agriculture. We then assess how compatible they are with ecological sustainability (without which economic and social sustainability is not possible) and suggest ways to optimize their potential for sustainability. We finish by briefly discussing future prospects for alternative agriculture paradigms in the context of an inclusively responsible food and agriculture system.

10.2 The industrial Green Revolution agriculture paradigm The dominant agriculture production paradigm today is the industrial Green Revolution paradigm. The Green Revolution refers to a set of agricultural research, development, and technology transfer initiatives between the 1940s and 1970s, worldwide. The new technologies included: … new, high-yielding varieties (HYVs) of cereals, especially dwarf wheats and rices, in association with chemical fertilizers and agro-chemicals, and with controlled water-supply (usually involving irrigation) and new methods of cultivation, including mechanization. All of these together were seen as a “package of practices” to supersede “traditional” technology and to be adopted as a whole. (Farmer, 1986)

The Green Revolution in low-income countries is often attributed to crop scientist Norman Borlaug. Borlaug began conducting research in Mexico in the 1940s, funded by the Rockefeller Foundation, to develop new so-called HYVs of wheat that responded well to fertilizer and irrigation. As noted by Holt-Giménez (2019, p. 27), Borlaug’s early success:

Paradigms of agriculture183

led the Rockefeller and Ford Foundations to finance a massive research system, anchored in thirteen international agricultural research centers (IARCs) under a coordinating body called the Consultative Group for International Research (CGIAR)…the IARCs collected seeds from the rich pool of cultivars developed by peasant farmers over thousands of years. This genetic material allowed scientists to develop high-yielding hybrid varieties of irrigated wheat, maize, and rice that relied on synthetic fertilizer and pesticides that became the trademark of the Green Revolution.

10.2.1 The global spread of industrial agriculture Before the Green Revolution, farming globally was based on manual and animal-­ powered tillage, diversified cropping, including legumes, associations, rotations, and animal integration as a source of power, manure, and food. With motor power after WWI, tractors began to replace horses, mules, and bullocks as well as human farm labor. Farming in more industrialized countries became increasingly mechanized. WWII changed farming drastically. After the war, motors from tanks and armored vehicles were repurposed and mounted onto tractors, and promoted to intensify agriculture. Decommissioned Sherman tanks were used as a source of power for plowing and many were exported to low-income countries. Tillage in particular became even more intensive due to mechanization and increased farm sizes. Following WWII, American chemical companies, such as DuPont, Dow Chemical, and Monsanto, faced an excessive amount of nitrogen production capacity, which they had developed to produce bombs and explosives. They decided to create new markets for their products. They repackaged nitrate, ammonium, and phosphorus explosives as fertilizers, and chemical weapons as pesticides, for both domestic agriculture and export (Engdahl, 2007). The use of agrochemicals quickly became the norma and drove the expansion of industrial agriculture along with cheap oil and credit and new machinery. The Green Revolution was also key to opening up markets for US agribusiness to export chemicals, machinery, and new HYVs. As argued by Engdahl (2007, pp. 123–126): The global marketing of the new agri-chemicals after the war …solved the problem of finding significant new markets for the American petrochemical industry as well as the grain cartel …including Cargill, Continental Grain, Bunge and ADM. The largest grain traders were American, and their growth was a product of the development of special hybrid seeds through the spread of the Green Revolution in the 1960s and 1970s …The Green Revolution was the beginning of global control over food production, a process made complete with the Gene Revolution several decades later. The same companies, not surprisingly, were involved in both, as were the Rockefeller and other powerful US foundations …the Green Revolution introduced US agribusiness into key developing countries under the cover of promoting crop a

For example, in Britain between 1943 and 2010, N fertilizer application in winter wheat increased from 19 to 195 kg N ha− 1 and from 4 to 100 kg N ha− 1 for grass. Mineral N fertilizer application before 1940 was very low (Muhammed et al., 2018).

184

Rethinking Food and Agriculture

science and modern techniques. The new wheat hybrids in Mexico required modern chemical fertilizers, mechanized tractors and other farm equipment, and above all, they required irrigation, which meant pumps driven by oil or gas energy.

10.2.2 Impacts of the industrial Green Revolution paradigm The foregoing trends led to the industrialization of agriculture, based on the Green Revolution production paradigm. Agriculture came to rely on genetic manipulation of germplasm, monoculture production, increasing and excessive use of agrochemicals, fossil fuels, and heavy capital investment into farm machinery and equipment. Production strategies also became oriented to respond to international market demands, e.g., standardization, minimizing the number of farm workers to maximize efficiency, maximizing yield, economies of scale, profit, and power. Farming lost much of its ecological and biological resilience. This resilience had been based upon low-­ intensity soil disturbance, crop diversification (including legumes), and maintenance of habitats such as hedgerows and perennial vegetation for natural enemies of pests. In the Green Revolution paradigm, loss of top soil due to intensive tillage was partially dealt with by increased use of chemical fertilizers. Pesticides and herbicides allowed farmers to stop applying their traditional crop mixtures and rotations and other soil and water conservation practices. They focused on short-term yields while polluting and depleting freshwater sources (Nibert, 2002). As a result, industrial agriculture is currently the dominant production system in modern farming in industrialized countries such as the United Kingdom, North America, and Australia, as well as parts of South America, Asia, and Africa. The industrialization of crop agriculture and the resulting grain surpluses also led to the industrialization of animal farming and increased meat consumption, first in Western countries such as the United States and then worldwide. Traditionally cows were consumed by the more affluent due to their high production costs (Rifkin, 1992). Grain-fed cows and other farmed animals, however, increasingly became an important market for grain producers. Grain surpluses reduced feed costs significantly, making beef more affordable and increasing meat consumption hugely (Nibert, 2002; Winders & Nibert, 2004). The US government also dealt with the problem of surplus grain by promoting meat production and consumption (Winders & Nibert, 2004). In 1949 it introduced price supports for pig production and funded research to increase the efficiency of animal production, resulting in the rapid growth of industrial animal agriculture (Winders & Nibert, 2004). Between 1945 and 1960 the majority of farmed animals in the United States moved into factory farms and confined animal feeding operations. As a result, since WWII global meat production has increased by nearly five times. This increase is partly due to the massive growth in the number of chickens killed for food, which has multiplied nearly 10-fold from 6.6 to 65.8 billion between 1961 and 2016 (Sanders, 2018). Chickens are now the biggest consumers of cropbased feed globally, responsible for around 40% of feed use, followed by pigs (WWF, 2017). Since WWII, meat consumption per person has also doubled (Ritchie & Roser, 2019). However, while global per capita meat consumption is currently 43 kg/year, it

Paradigms of agriculture185

is nearly double in the United Kingdom (82 kg/year) and almost triple in the United States (115 kg/year), while only 5 kg/year in India (Ritchie & Roser, 2019). Thus much of the increase in agricultural inputs and outputs resulting from Green Revolution agriculture has gone into producing meat. Globally, 40% of grain and 80% of soya is used to feed farmed animals (Stoll-Kleemann & O’Riordan, 2015). In the United States at least 70% of grain is sold as animal feed (Nibert, 2002). This has led to significant shifts in agricultural land use. Currently, 75%–80% of agricultural land globally is used for pasture, feed grains, and fodder (Foley et  al., 2011; StollKleemann & O’Riordan, 2015). The industrialization of crop and animal agriculture driven by the Green Revolution paradigm has had serious negative environmental and social impacts, which have been well documented (Gliessman, 2015; Holt-Giménez, Altieri, & Rosset, 2006; IAASTD, 2009; Kesavan & Swaminathan, 2018; Patel, 2013). These include environmental pollution with agrochemicals, drying and contamination of aquifers, loss of biodiversity, including agricultural biodiversity, dependence on agrochemicals and fossil fuels, high costs of production, increased rural inequality and concentration of land and resources (Freebairn, 1973), and poor resilience to climate change and other stresses, both abiotic and biotic. The degradation and destruction of soils and landscapes has also been a major impact (Kassam et al., 2013; Kassam, Friedrich, Shaxson, & Pretty, 2009; Montgomery, 2007). While tillage-based agriculture has existed for centuries, the intensification of agriculture and tillage has resulted in large-scale soil and land degradation and loss of soil health. It appears that in 2015/16 approximately 87.5% of global cropland, corresponding to the area of cropland under tillage-based agriculture, experienced soil degradation due to tillage (Kassam, Fridrich, & Derpsch, 2019). Since WWII, 7–12 billion hectares of farmland has been abandoned annually (Dregne & Chou, 1992; Gibbs & Salmon, 2015; Montgomery, 2007; Pimentel et al., 1995), totaling up to half a billion hectares (Montgomery, 2007). Soil and land degradation has also occurred in areas where intensification has not happened, but where tillage has been practiced manually or with animal power (Montgomery, 2007). Industrial tillage-based agriculture has also been responsible for a significant loss of ecosystem functions and services (discussed further later) such as water cycling and the availability of clean water resources, carbon cycling and carbon sequestration, nutrient cycling, integrated biological pest control, and pollination (Kassam et  al., 2011; MEA, 2005). The economic and social costs of this are huge, and in many cases cannot be adequately costed. Once degradation or loss of function or species has occurred, they can never be recovered fully. In some instances, the loss can be total and the natural resource gone forever (Juniper, 2015). The industrialization of animal agriculture has also had devastating effects on farmed animals by significantly increasing the scale and intensity of the suffering we inflict on them. Furthermore, the use of growth hormones and antibiotics in animal agriculture, including dairy, has led to toxins accumulating up the food chain with disastrous effects. For example, all over India, the vulture population has been almost exterminated because of the high levels of antibiotic toxicity in dead cows, whom they scavenge upon (Juniper, 2015).

186

Rethinking Food and Agriculture

10.2.3 Rethinking the Green Revolution narrative The CGIAR and their donors credit the Green Revolution with “the astounding jump in basic food production that has occurred in the developing world, particularly Asia since the mid-1960s” (Baum, 1986). They also claim credit for this success. However, the initial growth rates eventually slowed down, something not widely publicized. The spread of Green Revolution agriculture in low-income countries could also not be sustained, not even in the heartlands of the Green Revolution such as India and Pakistan, where land degradation, pollution, water scarcity, ill health, and cancer began to appear in the 1980s and 1990s. Since 2000, there has been little yield increase and in most areas there has been a yield decline or suboptimal yield ceiling (Kesavan & Swaminathan, 2018). Similar situations have appeared with staple crops in industrialized countries where the Green Revolution paradigm is still promoted widely. For example, in several European countries, including the United Kingdom, national average grain yields of wheat crops have stagnated at 7–8 t ha− 1 for the past 20–25 years (Brisson et al., 2010). This ceiling is suboptimal because where UK farmers have adopted Conservation Agriculture, a non-Green Revolution approach (discussed further later), grain yields are over 11 t ha− 1, with much lower use of agrochemicals, fossil fuel, and labor (Sims & Kassam, 2011). Thus it seems that the dominant narrative of the successes of the Green Revolution is not the whole picture. This is perhaps not surprising given the involvement of various US corporations and foundations in driving it. The motivations behind the Green Revolution are also not clear-cut. According to Harris (1988) the name Green Revolution was coined to contrast with the phrase “Red revolution,” and the underlying fear of the spread of communism that had swept countries in Asia, Africa, and Latin America during the 1950s and 1960s, as the European colonial era was collapsing. Engdahl (2007) argues that the Green Revolution was about gaining control over food production in key target developing countries, promoted under the guise of free enterprise market efficiency against alleged “communist inefficiency.” The political and corporate interests involved in the Green Revolution are explored in greater detail by Patel (2013). In any case, given that the dominant narrative has the Green Revolution averting famines—especially in overpopulated India—and saving “a billion” lives, it might seem odd that the crops of initial focus were not local staple crops. For example, the Green Revolution program in Mexico focused on wheat grown by commercial farmers, while the local staple grain was maize. When the program moved to India in 1956, it initially focused on maize, which was only 3% of Indian crops. After 8 years the Rockefeller Foundation started funding wheat research, and only in 1965, during India’s drought, did rice—an actual Indian staple crop—join the program (Lele & Goldsmith, 1989). It was only after the cycle of hunger, poor weather, the negative impact of food aid, and recovery once the rains returned that it was possible for the United States to tell a story of its success in feeding the world with a bumper wheat crop (Patel, 2013; Stone, 2019). Yet, Indian production of a range of other crops all increased by 20%–30% between 1967 and 1970 without Green Revolution investment (Paddock, 1970). Crop yields were at record levels in countries like China where there was no Green Revolution (Cullather, 2010). New research by historians over the past decade forces us to rethink

Paradigms of agriculture187

200

Linear trendline (r2=0.95)

Coarse cereals

Million tons

150

First green revolution wheat crop Wheat

100

50 Rice

98

00 20

96

19

94

19

92

19

19

88

90 19

86

19

19

82

84 19

80

19

78

19

76

19

74

19

72

19

70

19

68 19

19

64

66 19

62

19

19

58

60

19

19

54

56 19

19

19

52

0

Fig. 10.1  Trends in the production of the major categories of food crops in India (Stone, 2019).

0.25

First green revolution wheat crop 0.20

Tons/person

Linear trendline (r2 =0.53) 0.15

0.10

0.05

20

00

98 19

96 19

94

92

19

90

88

19

19

19

86

84

82

19

19

80

19

78

19

19

76 19

74 19

72

70

19

19

66

68 19

64

19

19

62

60

19

19

58

56

19

54 19

19

19

52

0.00

Fig. 10.2  Trend in India’s total food grain production divided by population. Data are from the India Department of Agriculture & Co-Operation (Stone, 2019).

what happened in India and hence the wider Green Revolution narrative (Stone, 2019). This research casts doubt that there even was a famine between 1965 and 1967 in India (Figs. 10.1 and 10.2), and suggests the number of lives saved by the Green Revolution was closer to zero. According to Stone (2019, p. 7): The legendary wheat‐field triumphs came from financial incentives, irrigation, and the return of the rains, and they came at the expense of more important food crops. Long‐term growth trends in food production and food production per capita did not change, although the Green Revolution years, when separated out, actually marked a slowdown.

188

Rethinking Food and Agriculture

10.2.4 Mentality and mindset of the Green Revolution paradigm Regardless of the discrepancies in the Green Revolution narratives, what is certain is that the forces driving it were not necessarily all benign or philanthropic, nor were they always led by science and evidence. What is also certain is that in general the Green Revolution agriculture paradigm forced intensive tillage, standard agronomy, mechanization, excessive use of irrigation water, and agrochemicals, and replaced traditional cultivars with “modern varieties” and traditional cropping systems with monocropping or much-reduced diversification. For medium- and large-scale farmers, it enabled the spread of agricultural industrialization incorporating these elements. Smallholders were also pushed into adopting the Green Revolution package, including credit. Resource-poor smallholders in some parts of the world were marginalized or pushed out of farming. These trends continue to this day. The Green Revolution mindset and technologies did not have a holistic vision of sustainable agricultural production and land development. The approach was, and still is, genocentric, focused mainly on yield increases in a few food and industrial crops. For example, little attention was paid to the ecological underpinnings of production systems (especially those related to soil health and its relationship with crop root systems and biological control of crop health); cropping systems development; or sustainable natural resource management. In addition, little attention was paid to the multifunctional nature of agriculture and land management, which recognizes the “inescapable interconnectedness of agriculture’s different roles and functions” and environmental and social impacts (IAASTD, 2009). This multifunctional view encompasses agriculture’s role in enabling agricultural land to provide ecosystem, social, and economic services beyond agricultural production. In particular, agriculture’s role in facilitating ecosystem functions and services at the field, landscape, watershed, and regional levelsb that benefit rural communities and society in general was (and still is) ignored. Ecosystem functions and benefits include clean water, water storage and regulation, minimization of runoff and soil erosion, enhancement of soil health and biodiversity, avoidance of soil and land degradation, and environmental pollution, avoiding the breakdown of soil food webs and ecosystem food chains, sustaining pollination, minimizing GHGs, and so on. According to Engdahl (2007, pp. 153–155): From 1932 to 1957, the Rockefeller Foundation had handed out an impressive $90 million in grants to support the creation of the new field of molecular biology…the foundation scientists developed the idea of molecular biology from the fundamental b

At the field level, such functions and services are related to processes that sustain land and agricultural productivity such as water infiltration and its retention and supply to plants, nutrient release and its retention and supply to plants, and pollination services. At the landscape and watershed level, such services include: (1) supporting services such as soil formation, carbon, nutrient and water cycling, primary production, and atmospheric circulation; (2) regulatory services such as groundwater and stream flow regulation, water purification, climate regulation and disease regulation, and control of soil erosion; (3) provisioning services such as food, fiber and fuelwood, fresh water, biologically fixed nitrogen, genetic resources, and biochemicals; and (4) cultural services such as spiritual, recreational, educational, and cultural heritage (MEA, 2005).

Paradigms of agriculture189

assumption that almost all human problems could be “solved” by genetic and chemical manipulations”—scientific reductionism…Their methodology went back to what was termed “reductionism” by Rene Descartes, and to the methods of Charles Darwin, namely that living creatures were machines….

Thus the mentality of the Green Revolution paradigm reflects a mechanistic, reductionist view of nature that partly explains the preference for narrow technical fixes rather than systemic and holistic solutions.

10.3 Alternative paradigms of agriculture Despite the foregoing impacts, the industrial Green Revolution production paradigm is still the basis for much of the agricultural development efforts currently promoted by UN development agencies, multilateral and bilateral donor agencies, corporations, CGIAR, and many nongovernmental organizations. These ongoing efforts are directed toward smallholder farmers in low-income countries whose governments are often forced to accept this industrial corporate approach to agriculture. This has led to the widespread adoption of standardized agronomy with poor cropping system diversity and poor ecological sustainability and resilience; the establishment of national and international crop and germplasm “improvement” programs to produce and supply HYVs and/or reliance on private seed and agrochemical companies, including multinational corporations and their subsidiaries for the supply of HYVs and GMOs as well as agrochemicals; and intensification of tillage, including through the use of motorized mechanization. As a result, there have been increasing calls to address these challenges moving the global food and agriculture system away from “business as usual” (CFSHLPE, 2019; IAASTD, 2009; WDR, 2009) and toward operating sustainably within planetary boundaries. Unfortunately, it appears that the Sustainable Development Goals (SDGs) set in 2015 do not represent a shift away from business as usual. While sustainable agriculture is central to achieving most of the SDGs either directly or indirectly, little direction is offered in terms of alternative paradigms of agriculture. While the Green Revolution has been the dominant agricultural paradigm since WWII, several alternative agricultural paradigms have been conceptualized and practiced, partly as a reaction to the growth and negative environmental impacts of agrochemicals and industrial agriculture. Four of these alternative paradigms—Organic Agriculture, Agroecology, Regenerative Agriculture, and Conservation Agriculture—are reviewed nextc,d c

For more details on the origins and description of Organic Agriculture, Agroecology, and Regenerative Agriculture, see Beste (2019). d Agroforestry is not included as a separate paradigm because trees can exist within farming systems in all the foregoing paradigms. Agroforests, according to the International Centre for Research in Agroforestry’s definition (agricultural land with greater than 10% tree cover), are widespread and found on more than 40% of all agricultural land globally, or on over 1 billion hectares of land, with more than 900 million people, or 30% of the world’s rural population, living on them (Zomer et al., 2014, 2016). It is reasonable to guess that at least half of those people practice some form of “agroforestry” in the context of their overall farming system within all the paradigms described in this chapter, particularly the alternative paradigms.

190

Rethinking Food and Agriculture

10.3.1 Organic Agriculture While agriculture has been practiced for thousands of years without the use of synthetic chemical inputs, the modern organic movement has its roots in Europe and the United States in the 1920s and 1930s. Sir Albert Howard was one of the pioneers of the organic movement. He developed a system of Organic Agriculture through incorporating scientific methods with traditional farming practices based on his work in India. This system used no synthetic inputs, included crop rotations and soil management techniques, and used composts and manures. On his return to Britain in the early 1930s, he began to promote Organic Agriculture (Conford, 2001). In 1947, Howard published The Soil and Health: A Study of Organic Agriculture. Attempts were also being made in the United Kingdom during the late 1930s and early 1940s by Lady Eve Balfour to promote the idea of the interconnectedness of soil, crop, human, and animal health. She argued that this interconnectedness worked best in the organic form (Balfour, 1943). This led to the establishment of the Haughley Experiment, the first comparison of organic farming and conventional chemical-based farming, from 1939 to 1969. It also led to the establishment of the Soil Association in 1946, a leader in the promotion of Organic Agriculture in the United Kingdom. The organic movement was also influenced by Rudolf Steiner who established biodynamic farming in 1924. In the United States, J.I. Rodale published the Organic Farming and Gardening Magazine in 1941 and founded the Rodale Institute in 1947 (known then as the Soil and Health Foundation). In 2005, the International Federation of Organic Agriculture Movements (IFOAM) defined Organic Agriculture as: a production system that sustains the health of soils, ecosystems and people. It relies on ecological processes, biodiversity and cycles adapted to local conditions, rather than the use of inputs with adverse effects. Organic Agriculture combines tradition, innovation and science to benefit the shared environment and promote fair relationships and a good quality of life for all involved. (IFOAM, 2005)

This definition builds on the four aspirational principles of Organic Agriculture established by IFOAM—Health, Ecology, Fairness, and Care. IFOAM also have a set of norms and standards that form the basis of the standards used by national organic certifying bodies (IFOAM, 2018). Certification is mainly concentrated in Western, industrialized countries given the barriers, such as the price of certification, for smallscale farmers in less industrialized countries. Certified Organic Agriculture covers around 1% (70 million hectares) of total farm land area globally and is practiced by 2.9 million producers (CFSHLPE, 2019). There are also likely many millions of smallholder producers in ­low-income countries who are organic, as they have no access to agrochemicals, but not certified.

Paradigms of agriculture191

Compared to conventional industrial agriculture, the potential benefits of Organic Agriculture include organic food being much less contaminated by pesticides and with residues of much lower toxicity, higher biodiversity, increased soil organic matter, improved soil properties, and enhanced nutritional value (Gattinger et  al., 2012; Gomiero, 2018). There is also uncertainty around other dimensions such as soil health, soil mesofauna and mycorrhizae, soil erosion, soil degradation, and environmental benefits, partly due to the degradation caused by tillage (Seufert & Ramankutty, 2017). Also, conventional Organic Agriculture does not necessarily offer comparable yields and ecosystem service performance because of tillage and its inherent disadvantages, and yield gaps can be significantly high (de Ponti, Rijk, & van Ittersum, 2012; Seufert, Ramakutty, & Foley, 2012). In some cases, yield gaps can diminish over time with increased spatial stability and nutrient use efficiency (Schrama, de Haan, Kroonen, Verstegen, & Van der Putten, 2018). Also, conventional Organic Agriculture does not necessarily lower prices as organic products generally cost more (Meemken & Qaim, 2018).

10.3.2 Agroecology Since the late 1980s there has been an increasing amount of attention given to another agriculture paradigm called Agroecology. Agroecology developed through the coproduction of knowledge between farmers, mainly in Latin America, and scientists in reaction to the intensive, industrial Green Revolution paradigm (Altieri & Nicholls, 2017). As noted by the 2019 High Level Panel of Experts on Food Security and Nutrition report (CFS-HLPE, 2019) there is no consensus on the description of Agroecology. Some descriptions suggest four principles, others up to 13. However, at its foundation, Agroecology as a science is “the application of ecological concepts and principles to the design and management of sustainable agro-ecosystems” (Altieri, 1987). Altieri et al. (2017, p. 10) suggest six agroecological principles: (1) enhance the recycling of biomass, with a view to optimizing organic matter decomposition and nutrient cycling over time; (2) strengthen the “immune system” of agricultural systems through enhancement of functional biodiversity—natural enemies, antagonists, etc., by creating appropriate habitats; (3) provide the most favorable soil conditions for plant growth, particularly by managing organic matter and by enhancing soil biological activity; (4) minimize losses of energy, water, nutrients, and genetic resources by enhancing conservation and regeneration of soil and water resources and agrobiodiversity; (5) diversify species and genetic resources in the agroecosystem over time and space at the field and landscape level; and (6) enhance beneficial biological interactions and synergies among the components of agrobiodiversity, thereby promoting key ecological processes and services. Agroecology has also been difficult to define in practice. According to Wezel (2017) there is no definitive set of agroecological practices. Rather, practices can be classified along a spectrum of more or less agroecological, depending on how much they: (1) rely on ecological processes rather than agrochemical inputs; (2) are e­ quitable,

192

Rethinking Food and Agriculture

environmentally friendly, locally adapted, and controlled; and (3) adopt a systemic approach (CFS-HLPE, 2019).e Altieri (2002) suggests Agroecology, as a set of practices, seeks to design complex and resilient agroecosystems that, by “assembling crops, animals, trees, soils and other factors in spatially and temporally diversified schemes, favor natural processes and biological interactions that optimize synergies so that diversified farms are able to sponsor their own soil fertility, crop protection and productivity.” According to Altieri et al. (2017), the six principles described earlier can be applied using various practices and strategies such as crop rotations, polycultures, agroforestry systems, cover crops and mulching, and crop/“livestock” mixtures. One of the key principles is diversification, which occurs in many forms at the field level (variety mixtures, rotations, polycultures, agroforestry, and crop/“livestock” integration) and at the landscape level (hedgerows, corridors, etc.). Holt-Giménez (2019) also provides a small sample of common practices: “farmers use primarily animal manures, legumes, and cover crops to provide nutrients. Weeds are controlled by cultivating, cover-cropping, intercropping, and mulching. Pests are managed by attracting predators with companion planting, interrupting pest cycles and vectors with rotations, alley-cropping, and the use of trap crops and repellent crops” (pp. 98–99). While Agroecology started as a science and practice for sustainable agriculture, it has developed and broadened its scope from farm to watershed to the whole food system. To quote Silici (2014), Agroecology “has evolved as a scientific discipline, a set of practices and a social movement. As a science, it studies how different components of the agro-ecosystem interact. As a set of practices, it seeks sustainable farming systems that optimise and stabilise yields. As a movement, it pursues food sovereignty and new, multifunctional roles for agriculture.” By incorporating food sovereigntyf and social movements, Agroecology is actively political and seeks to challenge the power of the corporate food regime. As a way of redesigning food systems, Gliessman (2016) summarizes five stages or levels of agroecological food system transformation. The first three levels describe the steps farmers can actually take on their farms for converting from industrial or conventional agroecosystems. Two additional levels go beyond the farm to the broader food system and the societies in which they are embedded. Level 4 is to reestablish a more direct connection between those who grow our food and those who consume it. Level 5 is to build a new global food system, based on equity, participation, democracy, and e

The System of Rice Intensification (SRI) proponents consider SRI to be agroecological. It has a different crop management approach and postsowing or transplanting water management compared with conventional traditional and modern flooded rice systems (Uphoff, 2015). Conventional SRI production management, however, involves destructuring the soil through puddling. The four SRI principles can be applied to other individual crops. SRI does not confine itself to only organic inputs, although it encourages the use of organic forms of nutrients if available. The origins of SRI are in Madagascar where it was first developed by Father Henri Laulanié (Uphoff, 2015), involving the use of inorganic fertilizer. f Defined by La Via Campesina as “the right of peoples to healthy and culturally appropriate food produced through ecologically sound and sustainable methods, and their right to define their own food and agriculture systems”: https://nyeleni.org/spip.php?article290.

Paradigms of agriculture193

justice, which is not only sustainable but helps restore and protects Earth’s life support systems upon which we all depend. All five levels taken together are supposed to serve as a roadmap that outlines a process for transforming the entire global food system.

10.3.3 Regenerative Agriculture More recently, another term—Regenerative Agriculture—is being used to describe an alternative agriculture paradigm that goes beyond degeneration and sustainability toward conservation, rehabilitation, and regeneration. As with Agroecology, there is no single definition of Regenerative Agriculture. According to the Regenerative Agriculture Initiative at California State University, Chico, and The Carbon Underground (2017): “Regenerative Agriculture” describes farming and grazing practices that, among other benefits, reverse climate change by rebuilding soil organic matter and restoring degraded soil biodiversity—resulting in both carbon drawdown and improving the water cycle. Specifically, Regenerative Agriculture is a holistic land management practice that leverages the power of photosynthesis in plants to close the carbon cycle, and build soil health, crop resilience and nutrient density.

They state practices are those that “(i) contribute to generating/building soils and soil fertility and health; (ii) increase water percolation, water retention, and clean and safe water runoff; (iii) increase biodiversity and ecosystem health and resiliency; and (iv) invert the carbon emissions of our current agriculture to one of remarkably significant carbon sequestration thereby cleansing the atmosphere of legacy levels of CO2.” Practices include: no-till/minimum tillage, biologically increasing soil fertility through application of cover crops, crop rotations, compost, and animal manures; building biological ecosystem diversity with inoculation of soils with composts or compost extracts; and well-managed grazing practices. The foregoing definition and list of practices is also used by Regeneration International,g an international network established in 2015 with the mission to: “promote, facilitate and accelerate the global transition to regenerative food, farming and land management for the purpose of restoring climate stability, ending world hunger and rebuilding deteriorated social, ecological and economic systems.” In general, the principles of Regenerative Agriculture include taking a holistic systems approach and improving whole agroecosystems of soil, water, and biodiversity, with an emphasis on increasing soil fertility. Regenerative Agriculture on smaller farms and gardens is often based on systems such as Permaculture, Agroecology, Agroforestry, and Holistic Management, while larger farms often use no-till and/or “reduced till” practices. The Rodale Institute started using the term “Regenerative Agriculture” in the early 1980s but hardly used it after the late 1980s. In 2014 Rodale released a report entitled “Regenerative Organic Agriculture and Climate Change,” which claims we could g

https://regenerationinternational.org/about-us/.

194

Rethinking Food and Agriculture

s­ equester over 100% of current annual CO2 emissions by switching to “regenerative organic agriculture,” which works to maximize carbon fixation and minimize carbon loss when carbon is returned to the soil (Rodale Institute, 2014, p. 2). The report states that “Regenerative organic agriculture” “takes advantage of the natural tendencies of ecosystems to regenerate when disturbed” unlike many other forms of agriculture. It works toward “closed nutrient loops, greater diversity in the biological community, fewer annuals and more perennials, and greater reliance on internal rather than external resources.” The report further states: In practical terms, regenerative organic agriculture is foremost an organic system…designed to build soil health…comprised of organic practices including (at a minimum): cover crops, residue mulching, composting and crop rotation. Conservation tillage, while not yet widely used in organic systems, is a regenerative organic practice integral to soil‑carbon sequestration…these practices are not intended to be judged or implemented in isolation. Regenerative agriculture is, above all else, a holistic systems approach to appropriate farming in context. (Rodale Institute, 2014, p. 2)

An important aspect of Regenerative Agriculture is Rotational Grazing. Rotational Grazing is defined in Rodale’s standards for Regenerative Organic Certification (ROC) under “soil and crop management” as: a livestock production system where livestock graze in one portion (a paddock) of a pasture that has been divided into several paddocks. Livestock are systematically moved from paddock to paddock based on the stage of growth of the forages and on the objectives of the grazing system. While one paddock is being grazed, the rest of the pasture rests. This rest and recovery time maintains forage plants and builds soil organic matter. (Regenerative Organic Alliance, 2018, p. 6). Integration of crops and animals are also included in the ROC standards under “regenerative practices.”

The reported benefits of Regenerative Agriculture include increased soil organic carbon stocks, decreased GHG emissions, climate change adaptation and mitigation, maintenance and increased resilience of yields, improved water retention and plant uptake, increased profitability, revitalized traditional farming communities, enhanced biodiversity, and resilience of ecosystem services (Rodale Institute, 2014).

10.3.4 Conservation Agriculture Another alternative agriculture paradigm is Conservation Agriculture. Conservation Agriculture developed in the 1940s and 1950s in response to the “Dust Bowl” and all the issues it reflected, including increased soil erosion, soil and land degradation, and loss of soil health; droughts; environmental degradation and pollution; and costs of production. The “Dust Bowl” refers to the Southern Plains region of the United States that suffered severe dust storms during a multiyear drought in the 1930s. The “Dust Bowl” was caused by intensive mechanical tillage of large areas of the ­prairies.

Paradigms of agriculture195

Tillage killed the natural grasses that kept soil in place and turned the topsoil into dust. The storms killed thousands of humans and other animals and led to crop failures across the region (Montgomery, 2007). These events revealed the vulnerability of conventional tillage-based production systems. This vulnerability is due to the way tillage destroys soil health and soil functions and reduces the water infiltration and moisture retention capacity of the soil. This was further exacerbated after WWII with the intensification of tillage and use of agrochemicals (as discussed earlier). The destruction of soil health and function through tillage, along with the lack of soil organic matter and soil surface protection, became known by many scientists as the root causes of soil degradation (Kassam et al., 2013; Montgomery, 2007). The need to address these root causes has influenced a shift in thinking regarding agricultural sustainability. As with the other alternative paradigms, Conservation Agriculture also developed in response to the conventional tillage-based Green Revolution agriculture and the resulting loss of ecosystem services, its poor ability to adapt to and mitigate climate change, and its unsuitability for low-income countries, particularly for smallholder farmers (FAO, 2011). The Conservation Agriculture global community of practice defines Conservation Agriculture as an ecosystem approach to regenerative sustainable agriculture and land management, based on the context-specific, locally adapted practical application of the three interlinked principles (Kassam et al., 2019; www.fao.org/conservation-agriculture). These principles are: 1. Continuous no or minimum mechanical soil disturbance (no-till seeding/planting and weeding, and minimum soil disturbance with all other farm operations, including harvesting);

2. Permanent maintenance of soil mulch cover (crop biomass, stubble, and cover crops); and 3. Diversification of cropping systems (economically, environmentally, and socially adapted rotations and/or sequences and/or associations involving annuals and/or perennials, including legumes and cover crops), along with other complementary good agricultural production and land management practices.

This description of Conservation Agriculture was first adopted in 1998 at a Food and Agriculture Organization (FAO) regional workshop in Harare, Zimbabwe, and promoted by FAO as a sustainable way of intensifying agriculture for food security, livelihoods, and economic development. Conservation Agriculture was globalized by FAO and the European Conservational Agriculture Federation through the First World Congress on Conservation Agriculture in Spain in 2001. The development of Conservation Agriculture over the years has been driven and led by farmers in both the Global North and Global South (Kassam et al., 2019). The main benefits of Conservation Agriculture systems cover many areas and contribute to several SDGs. They have been elaborated in the literature (Farooq & Siddique, 2014; Goddard et  al., 2008; Kassam, 2020a; Kassam et  al., 2009, 2013; Kassam, Saidi, & Friedrich, 2017; Jat, Sahrawat, & Kassam, 2014), and include: ●



Increased biomass, productivity, and profit Decreased soil erosion and land degradation

196 ●













Rethinking Food and Agriculture

Improved soil health and reduced use of fertilizer Reduced or minimized use of pesticides and herbicides Reduced machinery, energy, and labor use and costs Greater adaptability to climate change Greater contribution to climate change mitigation Enhanced ecosystem services to society Rehabilitation or regeneration of degraded lands and ecosystem services.

There has been some confusion about what Conservation Agriculture is and about its field performance. This is mainly due to the imprecise use of language (often related to types of tillage) and the Conservation Agriculture definition. Conservation Tillage, Reduced Tillage, and Minimum Tillage are not part of Conservation Agriculture, nor is No-Till on its own (Derpsch et  al., 2014; Kassam et  al., 2013, 2019; Reicosky, 2015). No-Till can only be referred to as a Conservation Agriculture practice if it is part of a Conservation Agriculture system based on the application of the three interlinked principles. Conservation Agriculture is being practiced in rainfed and irrigated systems, annual, perennial and mixed cropping systems, orchards and plantation systems, agroforestry systems, pasture systems, organic and nonorganic systems, and rice-based systems (Kassam, 2020a, 2020b). In 2015/16, Conservation Agriculture covered around 180 million hectares of annual cropland globally (equivalent to 12.5% of global cropland, in all continents and agroecologies in different climates, altitudes, and soils) (Kassam et al., 2019). Since 2008/09, Conservation Agriculture has been spreading at an annual rate of around 10.5 million hectares, suggesting the current area may be around 200 million hectares. It is important to note that all of these alternative agriculture paradigms are overlapping and the distinctions between them are sometimes unclear. In the real world of agriculture on the ground, we are often dealing with complex continua rather than distinct paradigms. These different paradigms seem to be increasingly borrowing and sharing ideas from each other, making some of these definitions somewhat imprecise. In Agroecology (defined as a science and practice), all systems are organic, while not all Organic Agriculture systems can be considered compatible with Agroecology, especially conventional Organic Agriculture systems producing monocultures. Most Regenerative Agriculture is also organic and could also be considered as Agroecology; however, Agroecology does not share Regenerative Agriculture’s emphasis on perennials and Holistic Grazing. Conservation Agriculture and Agroecology define themselves similarly in terms of applying ecological principles to agroecosystems. While Conservation Agriculture is not generally included in the Agroecology paradigm as it does not reject agrochemicals, organic Conservation Agriculture systems are compatible with Agroecology. Regenerative Agriculture systems in some cases (e.g., that described by the Regenerative Agriculture Initiative at California State University, Chico, and The Carbon Underground report) borrow from Conservation Agriculture. The principles of no-till, soil cover, and diversity of Conservation Agriculture also overlap with the ecological foundations for Regenerative Agriculture systems such as Permaculture. On the other hand, Organic Agriculture is not necessarily included under Regenerative Agriculture unless it also

Paradigms of agriculture197

incorporates certain regenerative ­principles, hence the creation of the ROC framework. At the same time, Agroecology in its fullest definition (as a science, practice, and movement) would overlap with the alternative paradigms only if they align with Agroecology’s production principles as well as its emphasis on transforming power relations in the food system.

10.4 How ecologically sustainable are the alternative agriculture paradigms? In this section we assess how compatible the alternative agriculture paradigms reviewed earlier are with ecological sustainability. We start by looking at how nature functions to distill some key ecological principles and processes. We then examine how far the alternative paradigms incorporate these into their production systems.

10.4.1 Natural land-based ecosystem processes: How well are these incorporated in the agricultural paradigms? In nature, ecosystem processes, functions, and services work to maintain healthy soils and landscapes. Many of these ecosystem functions and services are soil mediated, such as clean groundwater and regulated ground and surface stream flows as part of the normal water cycle; build-up and mobilization of nutrient stocks as part of the nutrient cycle; and build-up and sequestration of soil carbon as part of the carbon cycle. Healthy soils and landscapes deliver life supporting, regulatory, provisioning, and cultural services (MEA, 2005). In the natural world, biodiversity and ecosystem functions are sustained as an integral part of a life-sustaining, solar-powered system, without any interference from humans. All the ecosystem processes work in harmony, and provisioning services related to tangible products such as yield are just some of many outputs from a multifunctional ecosystem. Through these processes, nature has a built-in capacity to repair, regenerate, and sustain herself, restore habitats when they are damaged, and produce vegetation, water, and other ecosystem products such biologically fixed nitrogen and sequestered carbon in the form of glycoproteins such as glomalin. When soil water is available and when temperatures are conducive to plant growth, vegetation establishes and perpetuates itself without mechanical tillage.h The ground is kept covered and protected with vegetal biomass, and there is diversification in the mix of plant species. Thus there is little or no erosion of soil, which enables most if not all rainwater to infiltrate, and the soil is protected against abiotic stresses. Under such conditions, the soil is rich in biodiversity of microorganisms, including algae, protozoa, bacteria, fungi, nematodes, earthworms, arthropods and soil-­ inhabiting insects, spiders, etc. This web of organisms maintains healthy soil life and h

Mechanical tillage refers to soil disturbance carried out using tillage equipment such as moldboard, disc, or chisel plows operated by motorized or animal power, or hoes operated manually.

198

Rethinking Food and Agriculture

habitats, with soil biopores and aggregate structure and porosity that allow high water infiltration, moisture retention and through drainage, avoiding waterlogging, flooding, runoff, and erosion. It is these conditions and biological processes that allow soil formation to occur. Soil physical particles come from weathering of parent material, but for weathered particles to form into soil at the top requires help from biological organisms and vegetation, including root material, under undisturbed conditions. Natural land-based ecosystems operate and sustain themselves with no external inputs, such as intensive mechanical tillage or agrochemicals, and without exploiting domesticated animals. Wild animal density and grazing pressure in natural land-based ecosystems tend to be very low and extensive, with an animal food chain that operates to maintain an ecological equilibrium. Grazing animals, mesofauna (soil animals), and microorganisms all process plant biomass in various forms, and when incorporated into the soil it can serve several purposes, including improving the physical condition of the soil, serving as a source of nutrients, providing substrate for carbon sequestration, etc. Furthermore, under such conditions, there is ecosystem resilience due to minimum soil erosion and degradation, enhancement of soil and landscape health and functions, and adaptability to climate change. Healthy soils and landscapes sustain a dynamic web of microorganisms, including mesofauna in the soil that maintain balance in favor of a disease-free environment. At and above the ground surface, nature ensures that the biodiversity includes a food chain that maintains a dynamic pest and predator population that is able to keep pests and diseases in check most of the time. This natural environment includes groundlevel and above ground-level habitats for insect pests and predators and birds of prey who all contribute to maintaining a landscape that sustains high-biodiversity food chains and webs. In light of the foregoing, it would seem that for a land-based agricultural operation or a farm to be economically, socially, and environmentally sustainable, it must first be ecologically sustainable. Thus it would appear that an agricultural production system that could offer long-term ecological sustainability, upon which to build and manage economical social and environmental sustainability, would reflect as much as possible the functioning of natural ecosystems. Such a system, including its intensification in terms of output, would operate based on the following four dimensions: (i) An ecosystems approach, meaning it would be holistic in design and practice, optimize not just production and yield but all other multifunctional processes of the ecosystem and thus also have the ability to address ecosystem issues by harnessing the rehabilitation, regeneration, and other life-giving processes of nature. (ii) Ecological underpinnings for sustainability based on a foundation of interlinked ecological principles and practices of no or minimum soil disturbance, permanent soil cover, and crop diversification upon which to build a sustainable production system to deliver biological products and ecosystem services in a regenerative manner. ( iii) Minimizing use of external inputs, including agrochemicals, seeds, animal manure, water, energy, time, and machinery, based on maximizing input use efficiency and output factor productivity.

Paradigms of agriculture199

(iv) Maintaining agroecosystem resilience by sustaining crop health and productivity, soil and landscape health and functions, and offering best climate change adaptability and mitigation.

While it is clear that the Green Revolution agricultural paradigm falls short of all of these aspects, what about the paradigms that have emerged in response. How closely do they align with and/or contribute to the ecological processes found in natural landbased ecosystems? This question is analyzed in relation to the foregoing four sets of characteristics of sustainability in the following sections.

10.4.1.1 Ecosystems approach Unlike the Green Revolution paradigm, all of the alternative paradigms described previously view agriculture as multifunctional. Their principles reflect holistic ecosystem approaches to designing agroecosystems. In practice, however, there are some areas of concern, especially in the Organic Agriculture paradigm. For example, when organic systems are producing certified products for supermarkets, they tend to focus on single commodities produced in monocultures. This goes against one of nature’s key principles of diversity. This also leads to an inability to fully protect soil health and functions due to the need to till intensively. These systems also tend not to promote permanent soil cover/soil mulching with biomass due to the difficulty of this in tillage-based systems (more on tillage and soil cover later). Thus many such certified systems are organic versions of their conventional nonorganic counterpart systems. While they may replace agrochemicals with organic inputs, these systems are still basically similar to the conventional industrial paradigm in their overall approach to system design. The sometimes uncritical inclusion of Organic Agriculture within the Agroecology paradigm (e.g., CFS-HLPE, 2019) without differentiating between conventional and more holistic organic systems is thus also a concern. An important outcome of taking a holistic, ecosystem approach to agriculture production is the ability of agroecosystems to continue to facilitate ecosystem functions and services as they would occur in nature. Given how central soil is to mediating ecosystem services, this ability refers especially to maintaining and regenerating soil health and the related ability to mitigate climate change through soil carbon sequestration. However, insofar as any of these paradigms continue to till, they cannot optimally facilitate soil-mediated ecosystem services. Clean air and water and sustainable food production are not possible without healthy soil. It is estimated that the vast majority (90%) of CO2 from crop fields is released from soil through tillage with the remaining 10% coming from fossil fuel used in tractors for farm operations (González-Sánchez et al., 2017; Lal, 2010). It is also estimated that 24% of the CO2 in the world’s atmosphere from the terrestrial pool comes from agricultural soil (Lal, 2010; MEA, 2005; Montgomery, 2007). As explored more later, Organic Agriculture and Agroecology are for the most part tillage based. Regenerative Agriculture promotes reduced and no-till to varying degrees, while Conservation Agriculture requires no-till.

200

Rethinking Food and Agriculture

10.4.1.2 Ecological underpinnings for sustainability For agroecosystems to reflect natural ecosystems they need to be ecologically underpinned with clear principles and practices that can provide a foundation for sustainable production systems capable of delivering the multifunctional objectives of agriculture. While each of the alternative paradigms is built on ecological principles, the way these principles are reflected in and put into practice is variable. For example, the IFOAM standards are very outcome oriented. They do not provide clarity on “how” the aspirational organic principles can be achieved by farmers, and often provide a range of contradictory practices from which to choose. For example, under “Soil and Water Conservation” it states: Operators shall take defined and appropriate measures to prevent erosion and minimize loss of topsoil. Such measures may include, but are not limited to: minimal tillage, contour plowing, crop selection, maintenance of soil plant cover and other management practices that conserve soil. (IFOAM, 2018, pp. 33–34)

While it is important to ensure that possible practices are not too prescriptive to enable flexibility and adaptability to local conditions, we would argue that this range of practices cannot contribute to fully achieving the ecological underpinning of the production system for sustainability. Minimal tillage (different to no-till) and contour plowing—both forms of tillage—are quite at odds with each other, and with soil health and thus ecological sustainability. The other practices mentioned cannot optimally protect and enhance or regenerate soil health and functions, given the inherent mechanical soil disturbance built into these tillage practices, and that ecological sustainability would need to be underpinned by several practices working together. Based on our review of the IFOAM standards it appears that there is no specific set of practical principles that comprise the ecological core of Organic Agriculture. This also seems to be the case with Agroecology. Many of Agroecology’s principles and processes indicate preferred outcomes rather than practices giving clear guidance on “how” farmers can achieve these outcomes. And as noted earlier, while these principles can be applied through a range of various practices and strategies, there is no consensus on what these are and if any are required. Thus it is unclear how well the principles provide the necessary practical ecological underpinnings for production systems when by default the production system must be organic but is based on tillage. The ROC standards from the Regenerative Agriculture paradigm provide clearer guidance on soil management. However, all three certification grades include tillage to varying degrees (explored more later), so still do not fully protect against ecological unsustainability. The set of four bundles of practices of Regenerative Agriculture from the Regenerative Agriculture Initiative and The Carbon Underground also provide guidance on soil management, and includes the practice of no-till/minimum tillage as the first set. However, we would argue that the description of the four set of practices is not always clear. For example, the first practice deals with no-till/minimum tillage. Many would consider minimum tillage to have no place in sustainable soil

Paradigms of agriculture201

­ anagement. In fact, minimum tillage, which disturbs 100% of the soil surface, often m up to 10 cm, can cause irreparable damage to the soil through pulverization and erosion, including wind erosion, as well as surface sealing when wet, and loss of carbon in the soil surface. Also, there is no explicit mention of the development of soil mulch cover through the return of crop biomass and stubbles to protect the soil surface and to serve as substrate for soil microorganisms. Furthermore, it is not clear why there is an insistence on including farmed animals and well-managed grazing practices, as we do not believe that this component is essential for ecological sustainability of production systems, which in many cases may not include farmed animals. Conservation Agriculture is built around the application of a clear set of three required interlinked ecological principles. These principles underpin the functioning of natural land-based ecosystems and in our view provide the necessary underpinnings for an ecologically sustainable production system. They are aligned with the six Agroecology principles described earlier. They also give clear guidance on the types of practices that reflect these principles. The practical application of these principles is broad enough to be locally formulated and adapted. In this way, Conservation Agriculture removes some of the main ecological weaknesses in the conventional industrial agriculture paradigm of the Green Revolution, i.e., tillage, lack of soil biomass cover, and poor crop biodiversity. It also significantly reduces the use of inputs of seeds, agrochemicals, fuel, time, and machinery. Furthermore, Conservation Agriculture can be and is practiced without agrochemicals by many thousands of smallholders (e.g., Lalani, Dorward, Kassam, & Dambiro, 2017). Also, organic Conservation Agriculture-based annual cropping systems are now spreading on large farms in North America and Europe based on planting green using roller crimpers and no-till seeding. Perennial organic Conservation Agriculture systems such as orchards, plantations, and agroforestry, also exist in all continents (Kassam, 2020a). The three Conservation Agriculture principles of continuous no or minimal soil disturbance, permanent ground cover, and crop diversification, and their level of incorporation in the various alternative paradigms, are explored more.

Continuous no or minimum mechanical soil disturbance When humans established agriculture, in most places mechanical soil disturbance through tillage became a core practice to prepare the seed bed for planting and crop establishment, control weeds, and aerate the soil. While tillage has short-term benefits, in the longer term it guaranteed the destruction of soil health and function. As noted earlier, tillage is destructive to soil health and functions. In natural systems, earthworms produce a network of biopores in the soil that permits aeration, deeper drainage of water, and facilitates root growth to deeper depths. These biopores, along with the network of mycorrhizae, and the mesofauna are destroyed in tillage-­ based agriculture, whether in the conventional industrial or alternative paradigms. Soil fungi, including mycorrhizae, produce glycoproteins such as glomalin, which provide the inert carbon cement for the formation of micro and macro soil aggregates, and for soil aggregate stability and carbon sequestration. Thus soils in ­tillage-based systems, whether conventional or alternative, are relatively poor in ­structure and

202

Rethinking Food and Agriculture

a­ggregate ­stability, drainage, sequestered carbon, and functional soil biodiversity (Goss, Carvalho, & Brito, 2017). Maintaining and regenerating soil health is directly related to maintaining and regenerating soil organic matter. Conventional tillage such as plowing progressively destroys soil organic matter by oxidation and releases GHG emissions from the soil, and in mechanized systems from the use of fossil fuels. So far, both Organic Agriculture and Agroecology tend to rely on intensive tillage and mechanical cultivation to control weeds and help release nutrients from composts and other organic sources through oxidation and mineralization (contrary to the analysis presented by Migliorini and Wezel (2017), e.g., see Soil Association (2018)). While there is generally higher organic matter content in organically managed tilled soils when compared to soils in the conventional paradigm, soil structure and health still remain in a suboptimal state, with habitats of most soil-inhabiting organisms destroyed or damaged. Thus despite moving away from the conventional industrial paradigm by rejecting chemical inputs and introducing other ecological management practices, if systems in Organic Agriculture and Agroecology rely on tillage, they still retain one of the core weaknesses of the industrial paradigm. However, no-till organic systems are being practiced and would undoubtedly lead to improvements in soil health and biodiversity. Unlike the Organic Agriculture and Agroecology paradigms, Regenerative Agriculture has a clear focus on minimizing tillage, though this seems to vary in practice. For example, minimum soil disturbance or no-till is not included in the minimum requirements of the ROC standards. Rather, Conservation Tillage is promoted.i For gold certification, tillage is restricted to “one tillage operation (no deeper than 10 inches) every three years” (p. 13). This is still likely to degrade the soil. Even disturbing the top 2–3 in. of soil increases runoff (Reicosky, 2015) and therefore erosion. Surface mechanical disturbance by chiseling of top soil (0–5 cm), where most of the carbon is found, leads to the loss of enough carbon that would require the return of about 10 t ha− 1 biomass to compensate (Sá, Sá, Santos, & Ferreira, 2008), and reduced infiltration and surface sealing. Similarly, while the principles of Regenerative Agriculture from the Regenerative Agriculture Initiative and The Carbon Underground (2017) include no-till/minimum tillage, they also suggest that some soils benefit from interim ripping to break apart hardpans, and certain low-level chiseling. Both of these practices are destructive and unnecessary for maintaining ecological sustainability. To remove soil compaction and hardpans should be a one-time investment. There is no need to make ripping or chiseling an ongoing or interim practice. Conservation Agriculture is the only alternative paradigm, which insists on continuous no or minimum soil disturbance as a principle. In practice this is implemented by no-till seeding or broadcasting of crop seeds and direct placing of planting material into untilled soil; no-till weeding; and causing minimum soil disturbance from any cultural operation, harvest operation, or farm traffic. i

While Conservation Tillage is often confused with no-till, it is an umbrella term that includes no-till plus a broad set of practices such as minimum tillage, reduced tillage, contour plowing, bunding, and terracing. These practices are still destructive to the soil and have multiple definitions of their own (Reicosky, 2015).

Paradigms of agriculture203

Permanent ground cover In nature, permanent ground cover protects soil from erosion, provides soil microorganisms and mesofauna with a food source, and improves soil structure. Tillage-based systems in Organic Agriculture and Agroecology cannot keep the ground sufficiently covered as it is difficult to maintain soil mulch cover if tillage is practiced. Organic standards suggest but do not require that the soil surface must be kept protected with mulch. Similarly, Agroecology promotes cover cropping and mulching, but mulching cannot provide permanent soil cover in tillage-based systems. The Rodale 2014 report on Regenerative Agriculture and the ROC framework emphasize ground cover as a regenerative practice. However, the ROC framework also uses the terms mulching and composting (including manure) interchangeably when they are quite different. Mulching is a permanent layer of biomass mulch cover, while composting is used for building fertility in the soil, not for ground cover. Mixing compost into the soil and mechanically disturbing it to hasten the release of plant nutrients through mineralization leads to a great deal of organic matter loss, as noted earlier. Mulching as such is not mentioned in the description of practices by the Regenerative Agriculture Initiative and The Carbon Underground. However, cover crops are included. Conservation Agriculture insists on a permanent mulch cover on the soil surface, which in practice it means retaining crop biomass on the soil surface, using stubble mulching and cover cropping, and using other sources of biomass from ex situ sources.

Crop species diversification Species diversification is a common feature of natural ecosystems. For agriculture, increasing cropping and agroecosystem diversity is key for improving soil quality and function, plant health, crop productivity, and the resilience of production systems. Crop rotations are often suggested in Organic Agriculture as a way to practice diversification. However, rotations on small-scale farms are often difficult, which is why smallholder farmers often practice crop associations and sequences, which can be just as effective (Brooker et  al., 2015; Mousavi & Eskandari, 2011). The key principle is crop diversification rather than rotations. Often organic standards do not include cropping system diversification and their practice protocols are based on individual crops rather than the cropping or farming systems level (e.g., see UK Soil Association standards). This suggests that cropping system diversity is not necessarily built into Organic Agriculture systems. Diversification is a key principle of Agroecology. This occurs in many forms at the field level (including cropping system diversification and crop/“livestock” integration) and at the landscape level, as described earlier. Diversification in Agroecology also includes enhancing soil biodiversity. However, this is not practically feasible with tillage. Diversification is also important in Regenerative Agriculture. For example, crop rotations for both annual and perennial systems are a requirement at the silver and gold levels of ROC standards. However as noted previously, for small farmers, rotations can be difficult and there are different ways to promote diverse cropping systems. Practices such as crop rotations, full-time planting of multiple crop i­ntercrop ­plantings, and

204

Rethinking Food and Agriculture

multispecies cover crops are also included in the Regenerative Agriculture Initiative principles, which also focus heavily on soil biodiversity. Conservation Agriculture systems require cropping system diversification through rotations and/or sequences and/or associations, involving annual and perennial crops, including legumes. By requiring no or minimum soil disturbance with no-till seeding and weeding, enhancing soil biodiversity as well as the overall biodiversity is also a core element in Conservation Agriculture.

10.4.1.3 Minimizing the use of external inputs The Green Revolution agriculture paradigm promotes external inputs such as agrochemicals, modern seeds, and fossil fuel. Nature functions without external inputs and alternative paradigms should ideally minimize their use and optimize their productivity. Next, we look at a few of the external inputs currently used in the alternative paradigms: agrochemicals and domesticated animals and their products (often used to replace agrochemicals).

Agrochemicals Organic Agriculture, Agroecology, and the vast majority of Regenerative Agriculture systems reject the use of agrochemicals for plant nutrition and protection. This is an important issue in these systems due to the toxicity of these chemicals and the wide range of serious negative impacts they have, including air and water pollution, reduced soil fertility, death of nontargeted animals and plants, toxic residues in food, reduced self-reliance, autonomy and profits for farmers, etc. (Shiva, 2016). However, conventional organic systems often rely on biological or organic versions of agrochemicals, which can also impact on human, animal, and environmental health (Avery, 2006; Bahlai, Xue, McCreary, Schaafsma, & Hallett, 2010; Bengtsson, Ahnstrom, & Weibull, 2005; Bradberry, Cage, Proudfoot, & Vale, 2005; Hudson et al., 2014; Macan, Varnai, & Turk, 2006; Walters et al., 2009). While there are both organic and nonorganic Conservation Agriculture systems, Conservation Agriculture is the only alternative paradigm that does not require a rejection of agrochemicals. However, it does enable a significant reduction in inputs compared to conventional industrial Green Revolution systems. In general Conservation Agriculture systems use 30%–50% fewer agrochemicals and seeds, up to 40% less water, and up to 70% less fossil fuel (Carvalho et  al., 2010; Friedrich, Kassam, & Shaxson, 2009; González-Sánchez et al., 2017; Jat et al., 2014; Kassam et al., 2019). Application of agrochemicals such as glyphosate in Conservation Agriculture systems also has significantly less impact on soil health than in conventional tillage systems due to the emphasis on no-till, soil mulch cover, and crop and root system diversity, which promote soil fungi, including mycorrhizae, and soil biodiversity (Goss et al., 2017). However, this does not address some of the other issues noted earlier, e.g., toxic residues on food, other than reducing their application and impact. Conservation Agriculture has often been blamed for promoting herbicide-resistant GMOs and glyphosate. However, as shown previously it is the conventional industrial Green Revolution paradigm that is promoting the use of GMOs and agrochemicals.

Paradigms of agriculture205

Conservation Agriculture predates GMOs, and globally more GMOs are used in conventional tillage agriculture than in Conservation Agriculture systems. In Europe and in many parts of Asia and Africa, Conservation Agriculture systems have no GMO crops. Even in the Americas and Australia, both GMO and non-GMO crops make up Conservation Agriculture cropping systems. Similarly, glyphosate globally is used largely by conventional industrial tillage-based agriculture in which more glyphosate is applied per hectare and per unit of output than in Conservation Agriculture systems. Many conventional tillage farmers apply glyphosate as a preharvest desiccant. Many smallholder Conservation Agriculture farmers do not use any glyphosate (Khan, Midega, Pittchar, Murage, & Pickett, 2017; Lalani et  al., 2017; Owenya, Mariki, Kienzle, Friedrich, & Kassam, 2011) nor do some of the perennial Conservation Agriculture systems (Kassam, 2020a). Often when Conservation Agriculture is blamed for nonperformance or for negative impact, it turns out that the systems referred to are not in fact Conservation Agriculture systems (Derpsch et al., 2014; Reicosky, 2015). Presently, there is significant land area and millions of smallholders who practice uncertified organic Conservation Agriculture particularly in the Global South. Examples include a Conservation Agriculture scaling-up project involving 3600 smallholder households in Tanzania (Owenya et al., 2011), a Conservation Agriculture ­scaling-up project involving over 30,000 smallholder households in Mozambique (Lalani et al., 2017), Conservation Agriculture agroforestry systems in Zambia, Malawi, and Burkina Faso, oil palm and rubber plantation systems in Malaysia, and olive orchard systems in Spain (Kassam, 2020a). In the Americas, on-farm research is showing the possibility of using cover crops and planting green techniques in Conservation Agriculture systems to avoid or further minimize the use of glyphosate (Duiker, 2017; Gullickson, 2018). Planting green has been practiced successfully in South America for quite some time (Pieri, Evers, Landers, O’Connell, & Terry, 2002). Organic production in Conservation Agriculture has been more difficult in largescale farming with annual cropping systems, although increasingly fewer agrochemicals are being used to enhance soil health and protect crops and habitats for agrobiodiversity (Kassam, 2020a; Lindwall & Sonntag, 2010; Sims & Kassam, 2011). Organic Agriculture systems generally offer lower yields as well as suboptimal ecosystem services as discussed previously. Thus it is not necessarily the case that organic automatically means sustainable. We would argue that Conservation Agriculture systems that use no or low levels of agrochemicals would still be more sustainable than tillage-based certified conventional Organic Agriculture systems. In recent years, the introduction in Conservation Agriculture systems with complex cover crops to control weeds and enhance soil health, followed by planting green without glyphosate by using roller crimpers or knife rollers, is permitting the development of high-output systems with low inputs of agrochemicals in both large- and smallscale systems (Duiker, 2017; Gullickson, 2018). These Conservation Agriculture systems optimize the potential for integrated pest management (IPM) through the application of the three ecological principles, and supplement these with other practices such as push/pull practices that allow for integrated management of weeds, insects, and pathogens (Khan et al., 2017; Khan, Murage, Pittchar, & Midega, 2020). More work on this is needed but extrapolating from comparisons of conventional versus

206

Rethinking Food and Agriculture

Conservation Agriculture systems, the advantages would also be available under organic Conservation Agriculture systems of all farm sizes.

Domesticated animals and their products An often repeated and promoted notion is that of domesticated ruminant animals being essential to ecological sustainability of production system. One argument is that animals permit nutrient cycling by eating fodder biomass from pasture and crops, which could not be eaten by humans. Another argument is the need to emulate the diversification found in nature to improve the ecological and economic resilience of production systems. These may be valid arguments, but they do not mean that annual cropland and perennial systems cannot function sustainably without the integration of domesticated animals, nor do they mean that these systems require animal integration to function sustainably or optimally. Organic Agriculture, Agroecology, and Regenerative Agriculture systems are heavily reliant on animal manure as a source of plant nutrients, and thus animals are included in these systems by default. Manure is a rich source of nitrogen and other nutrients and can also support soil functions such as water retention and microbial activity. However, nutrients contained in plant biomass do not need to pass through the guts of ruminant and nonruminant farmed animals for the biomass organic matter to become a source of plant nutrients or to be incorporated into the soil. Manure is essentially plant biomass in the form of fodder and feed ingested by animals. If crop biomass is left on the soil surface, as it is in natural ecosystems, it can serve as an organic substrate for mesofauna such as earthworms and microorganisms. Earthworms ingest surface biomass, mix it with mucus, gums, and soil particles, add nitrogen to it from the nitrogen-fixing bacteria in their guts, and expel the muddy mixture as worm casts or worm manure. Thus, while it may be possible for agricultural systems, which include crop/“livestock” integration, to be sustainable, this integration is not necessary. This is shown by natural ecosystems as well as by those organic Conservation Agriculture systems where farmed animals and/or their products are not used. It is also shown by the fact that the benefits provided by Conservation Agriculture are not contingent upon integrating farmed animals. Thus the dichotomy that is promoted by some alternative agriculture paradigms of the need for either manure or chemical fertilizer is misleading. In our view there is a superior option of managing production systems based on the application of Conservation Agriculture principles and relying on the nutrient cycling processes used in nature, as have many smallholders in Africa and Asia (Khan et  al., 2017, 2020; Kassam et  al., 2019; Lalani et  al., 2017). Recent work by Johnson (2018) shows that it is possible to use fungal-dominant soil biology in Conservation Agriculture systems to mobilize natural sources of plant nutrients for plant nutrition, as well as to control weeds. However, if biological products, both crops and animals, are taken out of production systems on a regular basis, as they normally are in agricultural systems, then it would be difficult to close the nutrient cycling loop completely and maintain nutrient self-sufficiency in perpetuity. Topping up or replacement of some soil mineral nutrients would be necessary at some stage to maintain soil mineral stocks to optimum levels of availability. This will depend on a

Paradigms of agriculture207

range of factors, including geology, soil parent material, and soil depth. This is the case with all agriculture systems and is not overcome through crop/animal integration, unless the animals are not removed, and their corpses are left to reintegrate into the soil as they would in nature. In some Regenerative Agriculture systems, rotational grazing (also known as mob grazing) is considered to be an important component. Rotational Grazing is argued to be a beneficial or even necessary practice for the regeneration of grassland, soil health, and carbon sequestration (e.g., by Allan Savory and his Holistic Management approach). There have been criticisms by some that statements about Rotational Grazing (and Holistic Management) are not supported by empirical evidence and experimental work (Briske, Ash, Derner, & Huntsinger, 2014; Carter, Jones, O’Brien, Ratner, & Wuerthner, 2014), while others have defended particular types of locally adapted Rotational Grazing (Teague, 2014). However, even if there were ecological benefits to Rotational Grazing, the climate impact must also be considered. An extensive review by Food Climate Research Network found: “The contribution of grazing ruminants to soil carbon sequestration is small, time-limited, reversible and substantially outweighed by the greenhouse gas emissions they generate” (Garnett et al., 2017). Proponents of Regenerative Agriculture also argue that shifting away from grassfed meat to diets rich in grains, oils, and sugars, as well as chickens and pigs whose feed depends on arable crops, will lead to more carbon being released due to tillage of more land. However, as argued previously, there is no need for tillage in agroecosystems and the significant potential for soil carbon sequestration of cropland under Conservation Agriculture should also be considered (see Chapter 12). Going further and reforesting land currently used for animal agriculture would make a huge contribution not only to carbon sequestration but also to wildlife habitats and the reduction of nitrogen and phosphorus pollution (from both manure and chemical fertilizers applied to feed crops) (Harwatt & Hayek, 2019). In our view, grazing is not necessary for ecologically sustainable soil and nutrient management. The integration of farmed animals in any system is thus due to the desire for animal production for other reasons, e.g., for income, consumption, and/or capital.

10.4.1.4 Agroecosystem resilience Resilient production systems have no or minimum soil erosion and soil degradation; can regenerate and sustain soil health and functions, including when damaged; can adapt to climatic variability, stresses, and extreme events; and can protect themselves against weeds, insect pests, and pathogens. Resilient production systems are also efficient in terms of input use and factor productivity. In the Green Revolution paradigm, great reliance is placed on the genetic make-up of modern cultivars to cope with abiotic and biotic stresses in production systems. Under a standardized agronomy with crowded plant densities and excessive agrochemical application, plants remain weak and vulnerable to stresses. With poor soil health, biomass management, and agrobiodiversity in the production system, abiotic and biotic stresses appear frequently, and crop plants succumb more quickly to abiotic

208

Rethinking Food and Agriculture

stresses of drought, heat, waterlogging, and to biotic stresses from weeds, insect pests, and pathogens. All of the alternative paradigms are more resilient than conventional Green Revolution agriculture to varying degrees. Aside from conventional tillage-based Organic Agriculture, all of the alternate paradigms generate system resilience by integrating and promoting biodiversity. However, given that tillage destroys soil health, functions, and biodiversity, and leads to soil erosion and degradation, only no-till systems (e.g., Conservation Agriculture, some Regenerative Agriculture systems, no-till organic) can minimize this (see Chapter 11). Conservation Agriculture (and no-till based Regenerative Agriculture) systems pay particular attention to enhancing soil health and establishing dynamic ecological conditions in the whole soil/plant/landscape system. This offers resilient performance with maximum productivity (output and efficiency) and flow of ecosystem and societal services. This is possible because these systems enhance soil health by nurturing the soil and landscape as living biological systems, promote good functional relationships between root systems and soil microorganisms, and generate system resilience by integrating and promoting biodiversity in the cropping system, in the soil, and in the landscape (Kassam et al., 2013). No-till systems such as Conservation Agriculture are also considered to be better able to mitigate climate change through carbon sequestration (FAO, 2012), and to be more adaptable to increased climatic variations in rainfall and temperature (GonzálezSánchez et al., 2017, 2019; Chapter 12). This also applies in varying degrees to Organic Agriculture, Agroecology, and Regenerative Agriculture but to a lesser extent where such systems do not promote the full ecological sustainable underpinnings discussed earlier. Crop health protection in Organic Agriculture, especially conventional systems, is generally weak. There is a limited possibility of developing IPM without maintaining a permanent layer of living or dead biomass or mulch cover on the soil surface, and a diversified cropping system. Organic Agriculture’s reliance on tillage rules out this possibility. This also applies to tillage-based Agroecology and Regenerative Agriculture systems. Water use efficiency and water productivity in no-till systems such as Conservation Agriculture tend to be higher than in tillage-based systems because untilled agricultural soils have a higher rate of infiltration and more of the water is retained in the bulk soil throughout the growing season (FAO, 2008). Soil evaporation losses are also reduced by mulch cover. This is not the case with tillage-based Organic Agriculture and Agroecology systems that do not maintain permanent soil cover. In Regenerative Agriculture systems with no-tillage and ground cover, water use efficiency and water productivity would be higher. No-till-based systems also benefit from fossil fuel savings. For example, Conservation Agriculture systems can save up to 70% of fossil fuel energy used by tillage (Friedrich et al., 2009; Friedrich & Kassam, 2012; Sims & Kassam, 2011). Similarly, in manual or animal-powered production systems, there is some 50% decrease in energy and time required due to no-till and less time spent on weeding (Friedrich et al., 2009; Owenya et al., 2011).

Paradigms of agriculture209

10.5 Consideration of structural issues In addition to the ecological aspects of sustainability elaborated earlier, it is also important to look at the ability of alternative paradigms to incorporate wider structural issues in the food and agriculture system, which are driving the destructive industrial Green Revolution paradigm. Without considering these social, economic, and political issues, namely where power lies in the food and agriculture system, it will not be possible to transform or replace the Green Revolution agriculture paradigm. Examination of these structural issues is beyond the scope of this chapter (but addressed in other chapters, e.g., Chapters 3, 18, and the final chapter). However, it is worth noting that of all these alternative paradigms, the one that most explicitly and comprehensively integrates the sociopolitical dimensions of food sovereignty, local autonomy, and community control of land, water, and agricultural biodiversity into its vision, definition, principles, and practice is Agroecology. Agroecology is the only paradigm that actively seeks to challenge the structural root causes of the environmental and social crisis of industrial agriculture, i.e., capitalism. It does this by questioning capitalist relations of production and allying itself with agrarian peasant social movements, which are resisting the advancement of the corporate food system, industrial agriculture, and neoliberal policies. This political dimension of Agroecology is slowing its spread in the industrialized world (Altieri et  al., 2017; Holt-Giménez, 2017, 2019). Insofar as paradigms such as Organic, Regenerative, and Conservation Agriculture are based on practices that increase the efficiency of input use or substitute organic inputs for agrochemicals, but that do not challenge monoculture and reliance on external inputs or address the sociopolitical dimensions and context, they cannot transform the food and agricultural system at the local and global levels (Altieri et al., 2017; Gliessman, 2016). However, increasing the efficiency of input use, conserving resources, and increasing productivity for small-scale farmers can also be viewed as helping to ameliorate the impacts of the corporate food regime on the most marginalized. As such, paradigms like Conservation Agriculture can be seen as pro-poor and go some way to challenging the dependency on external inputs. All alternative paradigms are resource conserving as they rely relatively less on purchased inputs than the Green Revolution paradigm for agricultural intensification.

10.6 Ways forward for alternative paradigms While it is clear that the industrial Green Revolution agriculture paradigm is unsustainable by any measure, the alternative paradigms that have developed in response or alongside have both strengths and weaknesses. While they have all turned agroecosystem development back around toward nature to varying degrees, none of them seem to go quite far enough. In our view, no production system or paradigm, which depends on intensive mechanical soil disturbance through tillage, does not treat permanent soil mulch cover (living or dead) as essential, and prefers to rely on farmed animal

210

Rethinking Food and Agriculture

i­ntegration and animal manure for plant nutrients or regeneration rather than in situ recycling of nutrients and managing a positive nutrient balance, will be able to fully harness the desired range of supporting, regulatory, and provisioning ecosystem services that are necessary for society and nature. Similarly, no paradigm that allows the inclusion of agrochemicals for crop nutrition and protection, even in reduced amounts, or that allows GMOs (central to the Green Revolution 2.0), can be seen as truly sustainable or challenging the conventional paradigm and the corporate control of the food and agriculture system. Some international development agencies such as FAO, a couple of CGIAR centers such as CIMMYT and ICARDA, and a few national governments do support alternative paradigms. However, these efforts do not seem to be organized in a consistent manner as part of long-term organization-wide strategies. In general, alternative paradigms have been driven by farmer-based organizations, civil society interest groups, social movements and organizations, and development activists in the fields of research, education, and extension. Both IAASTD (2009) and the CFS-HLPE (2019) as well as grassroots movements such as La Via Campesina have promoted Agroecology, including Organic Agriculture, as an alternative paradigm for agriculture. They see Agroecology as representing a transformation of the whole food system, with the ability to end hunger and poverty and effectively mitigate and adapt to climate change. However, promoting paradigms that do not remove the root causes of the degradation and destruction of soil and landscapes, and hence soil- and landscape-mediated ecosystem functions and services, cannot be ecologically sustainable. Integrating a sociopolitical dimension does not remove the weakness. Given that ecological sustainability can be achieved without farmed animals, the reliance on tillage, manure, and integration of farmed animals to varying degrees in Organic Agriculture, Agroecology, and Regeneration Agriculture paradigms reveals a lack of understanding of how to optimize soil health management without farmed animals and deliver all the on-farm, landscape, and ecosystem benefits that agriculture is capable of delivering. In a neoliberal capitalist economic system, any paradigm that does not work to explicitly challenge the power relations within the food and agriculture system and actively reject the corporate influence and control of the food system is vulnerable to co-option by vested interests, be they corporate, international organization or philanthropic actors. This co-option has happened to a certain extent with Organic Agriculture (conventional tillage-based organic) and Conservation Agriculture and is also starting to happen with Agroecology. As tempting as it is to propose an entirely new paradigm that synthesizes the best of each, we understand the value in embracing the plurality and diversity of these paradigms and the many more that exist within and outside those we have explored, while also recognizing that each one has work to do to move toward being both ecologically sustainable as well as including an analysis of the sociopolitical context. In the context of an inclusively responsible food and agriculture system, which includes recognition of the growing trend toward diets that are healthful and sustainable, both for humans and the environment, as well as the need for justice for human and

Paradigms of agriculture211

nonhuman animals, in our view the most promising way forward would be for each of these paradigms to: ●







be ecologically underpinned by the practical application of the three interlinked Conservation Agriculture principles; be driven biologically to end their reliance on agrochemicals and farmed animals and their manure; meet the increasing demands for sustainable and nutritious diets and for climate change adaptability and mitigation; and integrate into the wider movements resisting the corporate food regime and fighting for local autonomy, food sovereignty, and land and seed justice.

10.7 Future prospects for alternative agriculture paradigms in the context of inclusively responsible food and agriculture systems Future prospects for the transformation of the dominant agriculture paradigm are improving continuously. Knowledge, skills, and support are becoming increasingly available to farmers through farmers’ own participatory activities as well as national and international programs. A key driver of this transformation will continue to be the fact that conventional tillage-based agriculture is unable to deliver the multifunctional sustainable agriculture required by society and nature for the future. Conventional tillage agriculture in the form of the Green Revolution paradigm, in particular the industrial version, is a highly degrading land use paradigm. The agriculture of the future must be truly ecological, multifunctional, and regenerative. All four alternative agriculture paradigms reviewed here represent a work in progress. Those paradigms that continue to use intensive tillage and/or agrochemicals will not be optimal in terms of ecological sustainability, productivity, or ecosystem services. Much still needs to be discovered and innovated to ensure that our future food and agriculture system is sustainable and productive, whether under the Organic, Agroecological, Regenerative, or Conservation Agriculture labels. It must also ensure access to and availability of sustainable, plant-based diets that meet quality of life standards in terms of human health, longevity, and ethics and cause minimum harm to the environment. To ensure our relationship to our fellow human and nonhuman animals is inclusive, responsible, and just. At the production system level, in our view organic Conservation Agriculture currently offers the most promising way forward. It combines the possibility of higher output and productivity with ecological sustainability, including the flow of ecosystem services, input use efficiency, and resilience. Small- and large-scale organic Conservation Agriculture systems already exist. More farms will adopt organic systems if technologies to enhance and manage soil health and productivity as well as to protect the crops become readily available and adoptable. Practices such as cover crops, planting green, and push/pull IPM, as well as practices that manipulate fungi/ bacteria ratios in the soil to suppress weeds and mobilize plant nutrients, point to

212

Rethinking Food and Agriculture

a future where we will have Conservation Agriculture systems that are biological, either totally organic or mostly organic, and regenerative in the sense that they are self-protecting, self-repairing, and self-sufficient, requiring minimal external inputs and intervention. Transforming the current paradigm into one based on the principles of organic Conservation Agriculture is a radical shift. The challenges that must be addressed to achieve the transformation are formidable. This is partly because within our neoliberal capitalist system, change attracts attention only if it is good for the market. However, it makes no market sense if the land use system of the future is unsustainable environmentally, economically, and socially, and is unable to cope with climate change. As understood so well by those working within the Agroecology paradigm (HoltGiménez, 2017), we also doubt that under the current economic system we can have a truly sustainable food and agriculture system. A capitalist food system that puts infinite growth and profit before people and planet, leads to obesity and ill health as well as hunger and malnutrition in equal measure, produces unnecessary food surpluses and wastage, relies on the use and oppression of nonhuman animals, exploits workers, marginalizes and dispossesses smallholders, works against democratic community control of food production and food sovereignty, and is controlled and driven by multinational corporations facilitated by governments and international institutions is bound to fail, at least for the vast majority of the world. As such, shifting the agricultural production paradigm is necessary but not sufficient to transform our food system into one that is inclusively responsible, sustainable, and just for all. Wider social, political, and economic changes are also needed to transform the corporate capitalist food system that is the ultimate driver of the increasingly destructive and irresponsible practices of industrial agriculture.

References Altieri, M. A. (1987). Agroecology: The scientific basis of alternative agriculture. Boulder: Westview Press. Altieri, M. A. (2002). Agroecology: The science of natural resource management for poor farmers in marginal environments. Agriculture Ecosystems & Environment, 93(1–3), 1–24. https://doi.org/10.1016/S0167-8809(02)00085-3. Altieri, M. A., & Nicholls, C. I. (2017). Agroecology: A brief account of its origins and currents of thought in Latin America. Agroecology and Sustainable Food Systems, 41(3–4), 231–237. Altieri, M., Nicholls, C., & Montalba, R. (2017). Technological approaches to sustainable agriculture at a crossroads: An agroecological perspective. Sustainability, 9(3), 349. Avery, A. A. (2006). Organic pesticide use: What we know and don’t know about use, toxicity, and environmental impacts. In A. S. Felsot, & K. D. Racke (Eds.), ACS symposium series: Vol. 947. Crop protection products for organic agriculture: Environmental, health and efficacy assessment (pp. 58–77): ACS Publications. Chapter 5. https://doi.org/10.1021/bk2007-0947.ch005. Bahlai, C. A., Xue, Y., McCreary, C. M., Schaafsma, A. W., & Hallett, R. H. (2010). Choosing organic pesticides over synthetic pesticides may not effectively mitigate environmental risk in soybeans. PLoS One, 5(6), 1–7. https://doi.org/10.1371/journal.pone.0011250.

Paradigms of agriculture213

Balfour, E. B. (1943). The living soil and the haughley experiment. Universe Books. Baum, W. C. (1986). Partners against hunger: Consultative group on international agricultureal research. Washington, DC: The World Bank. Bengtsson, J., Ahnstrom, J., & Weibull, A.-C. (2005). The effects of organic agriculture on biodiversity and abundance: A meta-analysis. Journal of Applied Ecology, 42, 261–269. Berners-Lee, M., Kennelly, C., Watson, R., & Hewitt, C. N. (2018). Current global food production is sufficient to meet human nutritional needs in 2050 provided there is radical societal adaptation. Elementa: Science of the Anthropocene, 6(1), 52. https://doi.org/10.1525/ elementa.310. Beste, A. (2019). Comparing organic, agroecological and regnerative farming. In ARC2020 Newsletter. 26 November 2019. Checked on 4 January 2020. Part  1—Organic http:// www.arc2020.eu/organic-agroecological-and-regenerative-whats-the-diff-organic/. Part  2—Agroecology http://www.arc2020.eu/comparing-organic-agroecological-and-­ regenerative-farming-part-2-agroecology/. Part 3—Regenerative Ag http://www.arc2020. eu/comparing-organic-agroecological-and-regenerative-farming-part-3-regenerative/. Bradberry, S. M., Cage, S. A., Proudfoot, A. T., & Vale, J. A. (2005). Poisoning due to pyrethroids. Toxicological Reviews, 24(2), 93–106. Briske, D. D., Ash, A. J., Derner, J. D., & Huntsinger, L. (2014). Commentary: A critical assessment of the policy endorsement for holistic management. Agricultural Systems, 125, 50–53. Brisson, N., Gate, P., Gouache, D., Charmet, G., Oury, F. X., & Huard, F. (2010). Why are wheat yields stagnating in Europe? A comprehensive data analysis for France? Field Crops Research, 119(1), 201–212. https://doi.org/10.1016/j.fcr.2010.07.012. Brooker, R. W., Bennett, A. E., Cong, W. F., Daniell, T. J., George, T. S., Hallett, P. D., … Li, L. (2015). Improving intercropping: A synthesis of research in agronomy, plant physiology and ecology. New Phytologist, 206(1), 107–117. Buff, E. (2017). Can we solve world hunger and feed 9 billion people just by eating less meat? One green planet. https://www.onegreenplanet.org/environment/world-hunger-­populationgrowth-ditching-meat/ visited 29 November 2019. Carter, J., Jones, A., O’Brien, M., Ratner, J., & Wuerthner, G. (2014). Holistic management: Misinformation on the science of grazed ecosystems. International Journal of Biodiversity, 2014, 1–10. Carvalho, M., Basch, G., Barros, J., Calado, J., Freixial, R., Santos, F., & Brandao, M. (2010). Strategies to improve soil organic matter under Mediterranean conditions and its consequences on the wheat response to nitrogen fertilization. In ECAF (Ed.), Proceedings of the European congress on conservation agriculture: Towards agro-environmental climate and energetic sustainability (pp. 303–308). Spain: Madrid. CFS-HLPE. (2019). Agroecological and other innovative approaches for sustainable agriculture and food systems that enhance food security and nutrition. FAO. http://www.fao. org/3/ca5602en/ca5602en.pdf. Conford, P. (2001). The origins of the organic movement. Floris Books. Cullather, N. (2010). The hungry world. Harvard University Press. de Ponti, T., Rijk, B., & van Ittersum, M. K. (2012). The crop yield gap between organic and conventional agriculture. Agricultural Systems, 108, 1–9. Derpsch, R., Franzluebbers, A. J., Duiker, S. W., Reicosky, D. C., Koeller, K., Friedrich, T., … Weiss, K. (2014). Why do we need to standardize no-tillage research? Soil and Tillage Research, 137, 16–22. Dregne, H. E., & Chou, N. T. (1992). Global desertification dimensions and costs. In H. E. Dregne (Ed.), Degradation and restoration of arid lands (pp. 73–92). Lubbock, USA: Texas Technical University.

214

Rethinking Food and Agriculture

Duiker, S. W. (2017). Planting green—A new cover crop management technique. Field Crop News. https://extension.psu.edu/planting-green-a-new-cover-crop-management-technique. (Accessed 15 May 2018). Engdahl, F. W. (2007). Seeds of destruction: The hidden agenda of genetic manipulation. Global Research. FAO. (2008). Investing in sustainable crop Intensification: The case for soil health. In Report of the international technical workshop, FAO, Rome, July. Integrated crop management, Vol. 6. Rome: FAO. http://www.fao.org/ag/ca/. FAO. (2011). Save and grow: A policymaker’s guide to the sustainable intensification of smallholder crop production. FAO. FAO. (2012). Soil organic carbon accumulation and carbon budget in conservation agriculture. Integrated crop management: Vol. 16. Rome: FAO. Farmer, B. (1986). Perspectives on the ‘Green Revolution’ in South Asia. Modern Asian Studies, 20(1), 175–199. Retrieved from www.jstor.org/stable/312485. Farooq, M., & Siddique, K. H. M. (Eds.). (2014). Conservation agriculture. Cham, Switzerland: Springer International Publishing. Foley, J. A., et al. (2011). Solutions for a cultivated planet. Nature, 478, 337–342. Freebairn, D. K. (1973). Income disparities in the agricultural sector: Regional and institutional stresses. In Food population and employment: The impact of the green revolution (pp. 97–119). New York: Praeger. Friedrich, T., & Kassam, A. (2012). No-till farming and the environment: Do no-till systems require more chemicals? Outlook on Pest Management, 23(4), 153–157. https://doi. org/10.1564/23aug02. Friedrich, T., Kassam, A., & Shaxson, F. (2009). Conservation agriculture. In Final report, STOA project: Agricultural Technology for Developing Countries. European Technology Assessment Group. Garnett, T., Godde, C., Muller, A., Röös, E., Smith, P., De Boer, I. J. M., & Van Zanten, H. H. E. (2017). Grazed and confused?: Ruminating on cattle, grazing systems, methane, nitrous oxide, the soil carbon sequestration question-and what it all means for greenhouse gas emissions. FCRN. Gattinger, A., Muller, A., Haeni, M., Skinner, C., Fliessbach, A., Buchmann, N., … Niggli, U. (2012). Enhanced top soil carbon stocks under organic farming. Proceedings of the National Academy of Sciences of the United States of America, 109(44), 18226–18231. Gibbs, H. K., & Salmon, J. M. (2015). Mapping the world’s degraded lands. Applied Geography, 57, 12–21. Gliessman, S. R. (2015). Agroecology: The ecology of sustainable food systems (3rd ed.). Boca Raton, FL: CRC Press/Taylor and Francis Group. https://doi.org/10.1080/21683565.2015.1130765. Gliessman, S. R. (2016). Transforming food systems with agroecology. Agroecology and Sustainable Food Systems, 40(3), 187–189. https://doi.org/10.1080/21683565.2015.1130765. Goddard, T., Zoebisch, M. A., Gan, Y. T., Ellis, W., Watson, A., & Sombatpanit, S. (Eds.). (2008). No-till farming systems (pp. 179–195). Bangkok: World Association of Soil and Water Conservation (WASWC). Special Publication No. 3. Gomiero, T. (2018). Food quality assessment in organic vs. conventional agricultural produce: Findings and issues. Applied Soil Ecology, 123, 714–728. https://doi.org/10.1016/j. apsoil.2017.10.014. González-Sánchez, E. J., Moreno-García, M., Kassam, A., Holgado-Cabrera, A., TriviñoTarradas, P., Carbonell-Bojollo, R., … Basch, G. (2017). Conservation agriculture: Making climate change mitigation and adaptation real in Europe. European Conservation Agriculture Federation (ECAF).

Paradigms of agriculture215

González-Sánchez, E. J., Veroz-Gonzalez, O., Conway, G., Moreno-Garcia, M., Kassam, A., Mkomwa, S., … Carbonell-Bojollo, R. (2019). Meta-analysis on carbon sequestration through Conservation Agriculture in Africa. Soil & Tillage Research, 190, 22–30. Goss, M. J., Carvalho, M., & Brito, I. (2017). Functional diversity of Mycorrhiza and sustainable agriculture: Management to overcome biotic and abiotic stress. London: Academic Press. Gullickson, G. (2018). Planting into green cover crops. Successful Farming. https://www.agriculture.com/crops/soybeans/planting-into-green-cover-crops. Accessed 15 May 2018. Harris, J. (1988). Capitalism and peasant production: The Green Revolution in India. Penguin Books. Harwatt, H., & Hayek, M. (2019). Eating away at climate change with negative emissions: Repurposing UK agricultural land to meet climate goals. Harvard Law School. Holt-Giménez, E. (2017). A foodie’s guide to capitalism. NYU Press. Holt-Giménez, E. (2019). Can we feed the world without destroying it?. Polity Press. Holt-Giménez, E., Altieri, M. A., & Rosset, P. (2006). Food first policy brief no. 12: Ten reasons why the Rockefeller and the Bill and Melinda Gates Foundations’ alliance for another green revolution will not solve the problems of poverty and hunger in Sub-Saharan Africa. Oakland: Institute of Food and Development Policy. Holt-Giménez, E., Shattuck, A., Altieri, M., Herren, H., & Gliessman, S. (2012). We already grow enough food for 10 billion people and still can't end hunger. Journal of Sustainable Agriculture, 36(6), 595–598. https://www.researchgate.net/publication/241746569. Hudson, N. L., Kasner, E. J., Beckman, J., Mehler, L., Schwartz, A., Higgins, S., … Mitchell, Y. (2014). Characteristics and magnitude of acute pesticide-related illnesses and injuries associated with pyrethrin and pyrethroid exposures—11 states, 2000-2008. American Journal of Industrial Medicine, 57(1), 15–30. IAASTD. (2009). Synthesis report: A synthesis of the global and sub-global IAASTD reports, agriculture at a crossroads. Washington, DC: Island Press. https://www.weltagrarbericht. de/reports/Global_Report/Global_content.html. IFOAM. (2005). Definition of organic agriculture. Retrieved November 30, 2019 IFOAM. website https://www.ifoam.bio/en/organic-landmarks/definition-organic-agriculture. IFOAM. (2018). The IFOAM norms for organic production and processing, version 2014. IFOAM-Organics International. https://www.ifoam.bio/sites/default/files/ifoam_norms_ july_2014_t.pdf. Jat, R. A., Sahrawat, K. L., & Kassam, A. H. (2014). Conservation agriculture: Global prospects and challenges. Wallingford, UK: CABI. Johnson, D. C. (2018). Regenerating the diversity of life in soils—Hope for farming, ranching and climate. Keynote speaker at the conference on regenerative agriculture. Chico, CA, USA: California State University. 7 September 2018 https://media.csuchico.edu/ media/0_ysqa9svw. Juniper, T. (2015). What has nature ever done for us? How money really does grow on trees. London, UK: Profile Books. Kassam, A. (Ed.). (2020a). Advances in conservation agriculture: Systems and science. Cambridge, UK: Burleigh Dodds Scientific Publishing. Kassam, A. (Ed.). (2020b). Advances in conservation agriculture: Practice and benefits. Cambridge, UK: Burleigh Dodds Scientific Publishing. Kassam, A., Basch, G., Friedrich, T., Shaxson, F., Goddard, T., Amado, T., … Mkomwa, S. (2013). Sustainable soil management is more than what and how crops are grown. In R. Lal, & R. A. Stewart (Eds.), Advances in soil science. Principles of soil management in agro-ecosystems. 2013: CRC Press.

216

Rethinking Food and Agriculture

Kassam, A., Fridrich, T., & Derpsch, R. (2019). Global spread of conservation agriculture. International Journal of Environmental Studies, 76(1), 29–51. https://doi.org/10.1080/00 207233.2018.1494927. Kassam, A. H., Friedrich, T., Shaxson, T. F., & Pretty, J. N. (2009). The spread of conservation agriculture: Justification, sustainability and uptake. International Journal of Agriculture Sustainability, 7(4), 292–320. Kassam, A., Friedrich, T., Shaxson, T. F., Reeves, T., Pretty, J., & de Moraes Sà, J. C. (2011). Production systems for sustainable intensification: Integrated productivity with ecosystem services. Technikfolgenabschatzung—Theorie und Praxis, 2, 39–45. Kassam, A., Saidi, M., & Friedrich, T. (2017). Conservation agriculture for Africa: Building resilient farming systems in a changing climate. Wallingford: CABI. Kesavan, P. C., & Swaminathan, M. S. (2018). Modern technologies for sustainable food and nutrition security. Current Science, 115(10), 1876. Khan, Z. R., Midega, C. A. O., Pittchar, J. O., Murage, A., & Pickett, J. (2017). Climate-smart pushpull: A conservation agriculture technology for food security and environmental sustainability in Africa. In A. Kassam, S. Mkomwa, & T. Friedrich (Eds.), Conservation agriculture for africa: Building resilient farming syststems in a changing climate. Wallingford, UK: CABI. Khan, Z. R., Murage, A. W., Pittchar, J. O., & Midega, C. A. O. (2020). Insect pest and disease management practice and benefits in Conservation Agriculture systems: A case of ­push-pull practice. In A. Kassam (Ed.), Practice and benefits: 2. Advances in conservation agriculture (pp. 143–168). Cambridge, UK: Burleigh Dodds. Lal, R. (2010). Managing soils and ecosystems for mitigating anthropogenic carbon emissions and advancing global food security. Bioscience, 60(9), 708–721. https://doi.org/10.1525/ bio.2010.60.9.8. Lalani, B., Dorward, P., Kassam, A., & Dambiro, J. (2017). Innovation systems and farmer perceptions regarding conservation agriculture in Cabo Delgado, Mozambique. In A. Kassam, S. Mkomwa, & T. Friedrich (Eds.), Conservation agriculture for Africa: Building resilient farming systems in a changing climate. Wallingford, UK: CABI. Lele, U., & Goldsmith, A. A. (1989). The development of national agricultural research capacity: India's experience with the Rockefeller Foundation and its significance for Africa. Economic Development and Cultural Change, 37(2), 305–343. Lindwall, C. W., & Sonntag, B. (Eds.). (2010). Landscape transformed: The history of conservation tillage and direct seeding. Knowledge impact in society. Saskatoon: University of Saskatchewan. Macan, J., Varnai, V. M., & Turk, R. (2006). Health effects of pyrethrins and pyrethroids. Arhiv za Higijenu Rada i Toksikologiju, 57(2), 237–243. MEA. (2005). Ecosystems and human well-being: Synthesis. Millennium ecosystem assessment. Washington, DC: Island Press. Meemken, E.-M., & Qaim, M. (2018). Organic agriculture, food security, and the environment. Annual Review of Resource Economics, 10, 39–63. https://doi.org/10.1146/ annurev-resource-100517-023252. Migliorini, P., & Wezel, A. (2017). Converging and diverging principles and practices of organic agriculture regulations and agroecology. A review. Agronomy for Sustainable Development, 37, 63. https://doi.org/10.1007/s13593-017-0472-4. Montgomery, D. R. (2007). Dirt: The erosion of civilizations. Berkeley and Los Angeles: University of California Press. 285 pp. Mousavi, S. R., & Eskandari, H. (2011). A general overview on intercropping and its advantages in sustainable agriculture. Journal of Applied Environmental and Biological Sciences, 1(11), 482–486.

Paradigms of agriculture217

Muhammed, S. E., Coleman, K., Wu, L., Bell, V. A., Davies, J. A., Quinton, J. N., … Naden, P. S. (2018). Impact of two centuries of intensive agriculture on soil carbon, nitrogen and phosphorus cycling in the UK. Science of the Total Environment, 634, 1486–1504. Nibert, D. (2002). Animal rights/human rights: Entanglements of oppression and liberation. Rowman & Littlefield Publishers. Owenya, M. Z., Mariki, W. L., Kienzle, J., Friedrich, T., & Kassam, A. (2011). Conservation agriculture (CA) in Tanzania: The case of the Mwangaza B CA farmer field school (FFS), Rhotia Village, Karatu District, Arusha. International Journal of Agricultural Sustainability, 9(1), 145–152. https://doi.org/10.3763/ijas.2010.0557. Paddock, W. C. (1970). How green is the green revolution? Bioscience, 20(16), 897–902. Patel, R. (2013). The long green revolution. The Journal of Peasant Studies, 40(1), 1–63. https:// doi.org/10.1080/03066150.2012.719224. Pieri, C., Evers, G., Landers, J., O’Connell, P., & Terry, E. (2002). No-till farming for sustainable rural development. Agriculture and rural development working paper. Washington, DC: World Bank. Pimentel, D., Harvey, C., Resosudarmo, P., Sinclair, K., Kurz, D., & McNair, M. (1995). Environmental and economic costs of soil erosion and conservation benefits. Science, 267(5201), 1117–1123. Regenerative Agriculture Initiative, & The Carbon Underground. (2017). What is regenerative agriculture?. Retrieved November 30, 2019, from: http://2igmzc48tf4q88z3o24qjfl8.wpengine. netdna-cdn.com/wp-content/uploads/2017/02/Regen-Ag-Definition-2.23.17-1.pdf. Regenerative Organic Alliance. (2018). Framework for regenerative organic certification. https://regenorganic.org/pdf/ROC-Framework.pdf. Reicosky, D. C. (2015). Conservation tillage is not conservation agriculture. Journal of Soil and Water Conservation, 70(5), 103A–108A. Rifkin, J. (1992). Beyond beef: The rise and fall of the cattle culture. Thorsons. Ritchie, H., & Roser, M. (2019). Meat and dairy production. In Our World in data. Retrieved November 30, 2019, from: https://ourworldindata.org/meat-production. Rodale Institute. (2014). Regenerative organic agriculture and climate change. Rodale Institute. https://rodaleinstitute.org/assets/WhitePaper.pdf. Sá, J. C. M., Sá, M. F. M., Santos, J. B., & Ferreira, A. O. (2008). Dinâmica da Matéria Orgânica nos Campos Gerais (Org.). In S. de Gabriel Araújo, S. da Leandro Souza, L. P. Canellas, & F. A. O. de Camargo (Eds.), Vol. 1. Fundamentos da Matéria Orgânica do Solo: Ecossistemas tropicais & subtropicais (2nd ed., pp. 443–461). Metropole: Porto Alegre. Sanders, B. (2018). Global animal slaughter statistics and charts. Retrieved November 30, 2019 Faunalytics. website https://faunalytics.org/global-animal-slaughter-statistics-and-charts/. Schrama, M., de Haan, J. J., Kroonen, M., Verstegen, H., & Van der Putten, W. H. (2018). Crop yield gap and stability in organic and conventional farming systems. Agriculture, Ecosystems & Environment, 256, 123–130. Seufert, V., Ramakutty, N., & Foley, J. (2012). Comparing the yields of organic and conventional agriculture. Nature, 485(7397), 229–232. doi:https://doi.org/10.1038/nature11069. Seufert, V., & Ramankutty, N. (2017). Many shades of gray—The context-dependent performance of organic agriculture. Science Advances, 3(3), e1602638. Shiva, V. (2016). The violence of the green revolution: Third world agriculture, ecology, and politics. University Press of Kentucky. Shukla, P. R., Skea, J., Slade, R., van Diemen, R., Haughey, E., Malley, J., … Portugal Pereira, J. (Eds.). (2019). Technical summary of the special report on climate change and land. In Climate Change and Land: an IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes

218

Rethinking Food and Agriculture

in terrestrial ecosystems (P.R. Shukla, J. Skea, E. Calvo Buendia, V. Masson-Delmotte, H.-O. Pörtner, D. C. Roberts, P. Zhai, R. Slade, S. Connors, R. van Diemen, M. Ferrat, E. Haughey, S. Luz, S. Neogi, M. Pathak, J.): WMO and UNEP. Silici, L. (2014). Agroecology—What it is and what it has to offer. IIED, London: IIED Issues Paper. Sims, B., & Kassam, A. (2011). East Anglia branch visit to Tony Reynolds: No-till farming in the Lincolnshire fens. Special issue on urban and peri-urban agriculture Agriculture for Development, 26, 51–52. Soil Association. (2018). To plough or not to plough Tillage and soil carbon sequestration. Policy briefing. UK: Soil Association. November 2018. Stoll-Kleemann, S., & O’Riordan, T. (2015). The sustainability challenges of our meat and dairy diets. Environment: Science and Policy for Sustainable Development, 57, 34–48. Stone, G. D. (2019). Commentary: New histories of the Indian green revolution. The Geographical Journal, 185(2), 243–250. Teague, R. (2014). Deficiencies in the Briske et al. rebuttal of the Savory method. Rangelands, 36(1), 37–39. Uphoff, N. (2015). The System of Rice Intensification (SRI): Responses to frequently asked questions. Ithaca, NY: SRI-Rice, Cornell University. Walters, J. K., Boswell, L. E., Green, M. K., Heumann, M. A., Karam, L. E., Morrissey, B. F., & Waltz, J. E. (2009). Pyrethrin and pyrethroid illnesses in the Pacific northwest: A five-year review. Public Health Reports, 124(1), 149–159. WDR. (2009). Agriculture for development. World development report. Washington, DC: World Bank. Wezel, A. (2017). Agroecological practices for sustainable agriculture: Principles, applications, and making the transition. World Scientific Publishing. 485 pp. Willett, W., Rockström, J., Loken, B., Springmann, M., Lang, T., Vermeulen, S., … Murray, C. (2019). Food in the Anthropocene: The EAT–Lancet Commission on healthy diets from sustainable food systems. Lancet, 6736, 3–49. https://doi.org/10.1016/ S0140-6736(18)31788-4. Winders, B., & Nibert, D. (2004). Consuming the surplus: Expanding “meat” consumption and animal oppression. International Journal of Sociology and Social Policy, 24(9), 76–96. WWF. (2017). Appetite for destruction. WWF. https://www.wwf.org.uk/sites/default/ files/2017-11/WWF_AppetiteForDestruction_Full_Report_Web_0.pdf. Zomer, R. J., Neufeldt, H., Xu, J., Ahrends, A., Bossio, D., Trabucco, A., … Wang, M. (2016). Global tree cover and biomass carbon on agricultural land: The contribution of agroforestry to global and national carbon budgets. Nature Scientific Reports, 6, 29987. https:// doi.org/10.1038/srep29987. Zomer, R. J., Trabucco, A., Coe, R., Place, F., van Noordwijk, M., & Xu, J. C. (2014). Trees on farms: An update and reanalysis of agroforestry’s global extent and socio-ecological characteristics. Working paper 179 Bogor, Indonesia: World Agroforestry Centre (ICRAF) Southeast Asia Regional Program.

Soil health and the revolutionary potential of Conservation Agriculture

11

David R. Montgomery Department of Earth and Space Sciences, University of Washington, Seattle, WA, United States

11.1 Introduction Soil loss and degradation have undermined agricultural civilizations time and again around the world (Montgomery, 2007a). Global recognition of the problem of soil erosion increased in the aftermath of the Dust Bowl era, leading to substantial societal investment in soil conservation efforts. Yet, the problem is not just ancient history (Montgomery, 2007b). Less than half a century ago, Brink, Densmore, and Hill (1977) warned about the potential for soil deterioration to collide with growing world demand for food. Pimentel et al. (1995) subsequently highlighted this problem in estimating that a third of the world’s cropland had been degraded since the Second World War and that another 0.5%–1% (about 12 million hectares) was being lost to food production each year. More recently, a 2015 United Nations Food and Agriculture Organization report on the status of the world’s soil resources (FAO, 2015) estimated that each year soil erosion and degradation reduced global crop yields by another 0.3%, enough to reduce global harvests 10% by 2050. And in 2018 the United Nations reported that global land degradation already negatively impacts the well-being of at least 3.2 billion people—more than a third of humanity (IPBES, 2018). The combination of a growing human population and ongoing soil degradation ensures that it will become progressively more challenging to feed the world in the coming decades. Although grazing and croplands cover more than a third of Earth’s land surface, the amount of arable land per capita has decreased from about 0.45 ha in 1960 to 0.32 ha in 1980, and is projected to be down to just 0.22 ha in 2020 (FAO, 2015). Average erosion rates from a global compilation of studies confirmed that conventionally plowed agricultural fields erode one to two orders of magnitude faster than the natural pace of soil production, whereas erosion rates under no-till farming are close to natural soil production rates (Montgomery, 2007b). So far, conventional practices relying on regular tillage and intensive nitrogen fertilizer use helped deplete about half the organic matter in America’s agricultural soils (Baumhardt, Stewart, & Sainju, 2015). In some regions, historical land degradation has already resulted in long-lasting societal effects. Consider, for example, the contrast between the impoverished agronomic state of Syria and Libya today and their enviably productive agriculture in Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00011-6 © 2021 Elsevier Inc. All rights reserved.

220

Rethinking Food and Agriculture

Roman times. Or how by the time the disaster of the 1930s Dust Bowl focused societal attention on soil erosion, colonial farming practices had already done substantial damage to upland terrain across the American Southeast. Most of the Piedmont region lost at least 10 cm of soil—virtually the entire topsoil—due to postcolonial soil erosion that peaked in the 19th century (Meade, 1982). The societal consequences of widespread soil erosion across the region helped propel the American drive west (Montgomery, 2007a). Visiting the region today one finds subsoil exposed at the surface on farms now dependent on chemical fertilizers to maintain commercial harvests. This style of land degradation does not show up in global assessments that account only for land taken out of agricultural production, making estimates of global land degradation all the more sobering.

11.2 Extent of global land degradation Estimates for how much of the world’s 15 billion hectares of land has been degraded range from less than 1 billion hectares to over 6 billion hectares (Gibbs & Salmon, 2015). The wide range of such estimates reflects different methods and assumptions used in various assessments of global land degradation. Assessments have been based on expert opinion, satellite-derived estimates of net primary productivity (the rate at which plants build biomass), numerical models of biophysical systems, and estimates of the area of historically abandoned cropland (Gibbs & Salmon, 2015). Each of these approaches offers a particular perspective on land degradation based on different limitations and ways to define varying degrees of degradation. The Global Assessment of Soil Degradation was the first attempt to map the global impact of soil degradation (Oldeman, 1994). The approach relied on integrating the opinions of several hundred national experts in coarse-scale assessments of the type, extent, degree, rate, and causes of soil degradation. It estimated that 1.2 billion hectares of arable land had been degraded. Subsequently, a similar assessment based on expert opinion came to the conclusion that roughly twice as much land was degraded, including half of Earth’s rain-fed croplands and almost a third of irrigated croplands (Dregne & Chou, 1992). While such expert opinion-based assessments have been criticized as subjective, they provide the only uniform global assessments of soil degradation. The FAO’s Global Assessment of Lands Degradation and Improvement project relied on satellite-derived measurements of differences in the normalized difference vegetation index from 1981 to 2003 to assess vegetation condition and productivity as a proxy for net primary productivity (Bai, Dent, Olsson, & Schaepman, 2008). The approach revealed a declining trend in productivity across 2.7 billion hectares of land. However, it also revealed an increasing trend on roughly the same amount of land. Importantly, this approach cannot detect soil degradation that is masked by increased fertilizer use—it only measures net changes in productivity. So, lands on which agronomic methods degrade the soil but maintain yields over the period of measurement do not show up as degrading the soil.

Soil health and the revolutionary potential of Conservation Agriculture 221

The extent of abandoned agricultural lands offers another way to estimate the extent of global land degradation. Analyses of historical data found that over the past three centuries, 269 million hectares of cropland and 479 million hectares of pasture were abandoned due to land degradation, economic factors, or water scarcity (Campbell, Lobell, Genova, & Filed, 2008). However, the magnitude of historical land abandonment is far greater if one considers soil degradation dating from earlier civilizations (Montgomery, 2007a).

11.3 Side effects of conventional agriculture Primary agricultural drivers of soil loss and degradation include accelerated soil erosion, loss of soil organic matter, and disrupted soil life due to the effects of tillage, prolonged chemical fertilizer use that results in pH changes (acidification), and failure to return crop residues and manure to the soil. While tillage serves farmers well for preparing the ground for planting and for weed control, over the long run it degrades soil structure and soil organic matter. It leaves the soil bare and vulnerable to erosion, exposed to the action of wind and rain, and accelerates the breakdown and degradation of soil organic matter due to aeration enhancing microbial activity. Plowing also accelerates decomposition and reduces the amount of organic matter in the soil by disrupting soil structure, exposing more organic matter to oxidization and microbial breakdown (Kumar et al., 2017). Contrary to intuition, this makes plowing counterproductive for reducing runoff and promoting infiltration of rainfall into the soil to water crops. For the pass of a plow collapses void spaces like worm burrows and root casts, breaking up the natural drains that allow water to move down into the soil. Pulverizing the ground surface into a powder-like texture promotes crusting when it rains, increasing runoff over bare ground left vulnerable to erosion. So, more water runs off carrying away more soil, fertilizer, and pesticides to places where they can become problems. Tillage also disrupts soil life. The most obvious effect is on worms. A 2018 review of long-term farming trials found losses of between half and all worm biomass, with an average loss of more than 80% (Blakemore, 2018). Plowing can be a natural disaster for other soil life as well. For example, it chops up and disrupts the root-like hyphae of mycorrhizal fungi. The effects on bacteria are less clear-cut. While disturbed bacterial communities can bounce back more rapidly when boosted by a meal of freshly exposed soil organic matter, this burst of decomposition can reduce organic matter levels over time. For example, a 20-year comparison of soil quality under conventional cultivation and no-till farming in New Zealand found that microbial populations were two to three times higher in unplowed fields (Ross et al., 2002). The study also found that while conventionally cultivated fields had virtually no worms, no-till fields had close to the number of worms found in permanent pasture. Tillage also reduces the overall number of different species—the species richness—of soil fungi and bacteria (Anderson, Beare, Buckley, & Lear, 2017).

222

Rethinking Food and Agriculture

In addition, tillage also affects the timing of nitrogen delivery to crops. Enhanced microbial activity right after plowing makes nitrogen available to crops well before the period of rapid crop growth, before crops need it most or can take most of it up (Francis & Knight, 1993). So, some of the nitrogen liberated from tillage-degraded soil organic matter simply runs off and is lost from the fields without helping to grow a crop. And because they are soluble by design, nitrogen applied as chemical fertilizers is also susceptible to running off a farm before crops take it up. Less than half of the nitrogen applied as fertilizer to cereal crops is actually taken up by the plants (Raun & Johnson, 1999). Where does it go? Microbes that break down soil organic matter consume some of it. Some runs off into wells, streams, and lakes. There’s substantial room to improve the efficiency of fertilizer use given that global production of synthetic nitrogen increased about 10-fold since 1961. It is important to do so. Soil acidification resulting from long-term applications of nitrogen fertilizers is a growing global problem on conventionally farmed land (Tian & Niu, 2015). Certain soil-dwelling bacteria oxidize ammonia (NH3) to nitrite ( NO2 − ), and others then proceed to oxidize the nitrite to nitrate ( NO3 − ). Because each of these steps produces hydrogen ions (H+) that acidify the soil, soil pH gradually decreases with prolonged use of nitrogen fertilizers (especially ammonia). A 2015 review of global soil acidification found that soil pH decreased in proportion to the amount of nitrogen fertilizer added to soil (Tian & Niu, 2015). In other words, the more nitrogen fertilizer a soil receives, the more acidic it becomes. Soil acidification is increasingly problematic in the world’s grassland soils and when pH decreases below 5.5 it can restrict plant uptake of critical nutrients like calcium and magnesium (Tian & Niu, 2015). In places, the resulting soil acidification is depleting mineral nutrients like calcium, magnesium, and potassium, and increasing the solubility of other mineral elements that produce toxic effects at high concentrations (like aluminum, manganese, and iron) (Tian & Niu, 2015). The effects of synthetic nitrogen fertilizers on soil pH have led to the common practice of applying lime (calcium carbonate) to conventionally farmed fields to mitigate soil acidification. What does this all do to soil life? Synthetic nitrogen fertilizers can change soil community structure in ways that can have complicated effects on microbial biomass—some studies report increases, some find decreases. Yet, continued application of synthetic nitrogen fertilizers that decrease soil pH can have large impacts on soil life. If soil pH drops below 4.8 the soil becomes acidic enough that aluminum becomes soluble, retarding root growth, producing toxic effects on crops, and affecting nutrient uptake. Low pH can also affect microbial activity, decreasing or even curtailing the activity of the nitrogen-fixing bacteria that work in association with legumes. Acidification of agricultural soils also impacts bacterial communities that influence the availability of exchangeable cations that are important plant nutrients (Tian & Niu, 2015). The most important affects may be on changing the specific microbes present in the soil. This matters because plants depend to varying degrees on certain microbes—or communities of microbes—to help acquire several major elements and most mineral micronutrients. Synthetic nitrogen fertilization also stimulates soil microbes to break down and deplete soil organic matter. Khan, Mulvaney, Ellsworth, and Boast (2007) and Mulvaney,

Soil health and the revolutionary potential of Conservation Agriculture 223

Khan, and Ellsworth (2009) reviewed long-term field trials and concluded that reliance on synthetic nitrogen fertilizer uniformly led to a serious loss of nitrogen from the soil. Even applying more nitrogen than needed to account for incorporation into crops and crop residues led to major losses in soil nitrogen. Numerous studies using fertilizer tagged with nitrogen-15 as a tracer report that corn, wheat, and rice all take up more nitrogen from the soil than they actually take up from the fertilizer (Mulvaney et al., 2009). Hence, it appears that liberal applications of nitrogen fertilizer increase the pace of organic matter decay, drawing down soil nitrogen and in effect propping up crop production by mining native soil fertility. So, many of the major detrimental effects of conventional agriculture on soil health arise from overreliance on tillage and synthetic nitrogen fertilizers. Of particular importance is how the combination of habitual tillage and excessive nitrogen fertilizer use reduces soil organic matter and disrupts communities of soil life that build it. Relying on both practices together degraded the native fertility of agricultural soils, reducing their ability to produce high yields without sustained inputs. Fortunately, regenerative farming practices that reduce agricultural reliance on these two conventional practices could potentially reverse the historical pattern and rebuild soil organic matter.

11.4 Building soil health through Conservation Agriculture An important thing that doesn’t show up in existing global assessments of soil degradation is the loss of the soil life and soil organic matter that drive nutrient cycling in soils. Yet, in recent decades, growing recognition of the agricultural importance of soil life—of soil ecology—on soil fertility arose from a greater understanding of the roles that microbes in the rhizosphere play in influencing nutrient acquisition, chemical signaling, and plant defense (Montgomery & Biklé, 2016). This new understanding stresses the importance of soil organic matter as food for microbes that procure minerals and produce metabolites beneficial to the health of crops. It also stresses the role of soil life in building and maintaining soil health. The concept of health applies to living things. Thus the concept of soil health incorporates the role of soil life and not just the physical and chemical properties and quality of the mineral soil. The US Department of Agriculture (USDA) Natural Resources Conservation Service defines soil health as “the continued capacity of soil to function as a vital living ecosystem that sustains plants, animals, and humans.” Just as for human health, there are a number of ways to measure soil health and no single metric is adequate for all purposes. And even though both human and soil health may be hard to define crisply, they are readily recognized by their absence. Regenerative agricultural practices that build soil health offer a means to reverse soil degradation both rapidly and profitably (LaCanne & Lundgren, 2018; Montgomery, 2017). Examples from farms large and small in both developed and developing countries show that regenerative farming practices based on the principles

224

Rethinking Food and Agriculture

of Conservation Agriculture can reduce inputs and expenditures for diesel, fertilizer, and pesticides and yet still harvest comparable yields after a short, multiyear transition period (Montgomery, 2017). Hence, a key attraction of soil health building practices is the promise of a more profitable farm. Barriers to adopting such practices include force of habit and resistance to new ideas, lack of knowledge of how to implement new practices, and perceived economic risk during the transition. Yet, some benefits of rebuilding healthy, fertile soil are clear. According to a 2018 UN report the economic benefits of land restoration average 10 times the costs (IPBES, 2018). Rebuilding fertile soils is also one of the most promising ways to address hunger and malnutrition in Africa (Sanchez, 2002; Sanchez & Swaminathan, 2005), where the costs of inaction to reverse land degradation are typically three times the cost of addressing the problem (IPBES, 2018). Recent field trials in Ethiopia showed that smallholder farms with higher soil organic matter grew wheat crops with greater zinc and protein content, an important result given wheat’s role as a staple food crop and the widespread deficiencies of zinc and protein across the human population (Wood & Baudronc, 2018). Rebuilding the fertility of degraded soils can be done in a variety of ways. For example, soil depletion presents a major impediment to agricultural production in parts of Africa where decades of cropping removed nutrients from the soil without adequate replenishment with manure or mineral fertilizers (Sanchez, 2002). In this case, a combination of cropping practices that introduce nitrogen-fixing and biomass-­ accumulating crops and apply rock phosphates can form the foundation for replenishing soil fertility (Sanchez & Swaminathan, 2005). Strategies for building soil fertility as a consequence of agricultural practices include agroforestry and holistic grazing with frequent livestock movement and long recovery times. Conservation Agriculture, however, is much more than a practice. It is a whole system, a broader new paradigm that can include annual and/or perennial species, Integrated Pest Management, livestock management, and organic practices (whether certified or not). The defining principles of Conservation Agriculture involve using practices that minimally disturb the soil (no-till), keep the ground covered with living plants (cover crops), and plant diverse rotations and associations. This set of principles aims to promote building soil health through cultivating beneficial soil life (Kassam et  al., 2013). In 2015/16, Conservation Agriculture was adopted on more than 180 million hectares of cropland globally (12.5% of global cropland), and since 2008/09 it has been expanding at an annual rate of over 10 million hectares (Kassam, Friedrich, & Derpsch, 2019). Practices based on adopting all three of these principles build soil organic matter and improve soil health. Reviews of the effects of no-till farming on soil organic matter report increases in the amount of organic matter in topsoil, but mixed results for the full soil profile (Haddaway et al., 2017; Powlson et al., 2014). While these comparisons treat no-till as a standalone practice, the efficacy of no-till farming for increasing soil organic matter appears to depend on integration with other practices, particularly cover cropping and crop rotations (Montgomery, 2017). For example, a study in southern Brazil documented that 25 years of conventional tillage decreased soil organic matter to less than 20% of the amount in native soils. But two decades after switching

Soil health and the revolutionary potential of Conservation Agriculture 225

to high-intensity, no-till cropping using cover crops and a diverse rotation the soil organic matter recovered to more than 90% of the organic matter content of the native soil (de Oliveira Ferreira et  al., 2016). Montgomery (2017) reported similar results in South Dakota, Saskatchewan, Ghana, and Ohio where regenerative farms that adopted practices based on all three principles of Conservation Agriculture restored soil organic matter to levels close to those found in native soils in their respective regions. Regenerative practices that focus on soil building bring other benefits, notably comparable, if not higher, yields with fewer inputs. For example, Pretty et al. (2006) assessed low-input, resource-conserving agricultural practices in 57 countries in Latin America, Africa, and Asia, and evaluated 286 development projects that employed cover crops for nitrogen fixation and erosion control, applied pesticides only when crop diversity and rotations were not effective for pest management, and integrated livestock into farming systems. They found that for a wide variety of systems and crops, yields increased an average of 79%. For projects that had data on pesticide use, yields grew by 42%, while pesticide use declined by 71%. Practices that improved soil and crop health allowed effective pest control with minimal pesticide use. In other words, adopting more diversified, low-input farming methods offers an effective, demonstrated alternative for subsistence farmers for whom conventional input-­intensive methods can prove prohibitively expensive. And a 12-year study by a research team from Cornell University and the USDA showed that, compared to conventional tillage, no-till practices resulted in higher soil organic matter levels, greater microbial respiration (and thus biomass), greater levels of plant-available zinc in the soil, and higher infiltration rates (and therefore less runoff), thereby keeping more water in the soil where it can help crops grow (Nunes et  al., 2018). Introducing a cover crop further increased each of these positive effects, as well as the amount of plant-available iron in the soil. Coupling no-till systems with cover crops and diverse crop rotations enhances beneficial effects on soil biology through increases in soil organic matter, microbial biomass, and enzyme activity (Kinoshita, Schindelbeck, & van Es, 2017; Sharma, Singh, & Singh, 2013). Experiments at the University of Kentucky showed that crop rotations can greatly affect the community composition of mycorrhizal fungi, with greater mycorrhizal richness and diversity in rotated crops than found in continuously cultivated monocultures (Hendrix, Guo, & An, 1995). Soil microbial community composition rapidly shifted from one crop to the next in ways that could promote either mutualistic or pathogenic interactions with different crops. This means that while growing the same thing after itself time and again invites pests and pathogens to the agricultural party, but growing a diversity of plants recruits a diversity of microbial life—most of which compete with or consume pests or pathogens. After decades of trying to eradicate microbes from farm fields and sanitize them out of our lives the beneficial effects of microbiomes in maintaining the health of plants and people are increasingly recognized in both agriculture and medicine (Montgomery & Biklé, 2016). For over a century we viewed the microbial world primarily as agents of death and disease. To be sure, recognition of particular microbes as responsible for particular diseases, like tuberculosis and polio, led to effective ways to combat infectious pathogens. But we now know germ theory doesn’t cover the full

226

Rethinking Food and Agriculture

range of our relationships with the microbial world. Practices based on the principles of Conservation Agriculture appear to work to build healthy, fertile soils because they cultivate beneficial life in the soil (Montgomery, 2017). Mycorrhizal fungi are a good place to start in considering what a healthy root microbiome does for its plant host, because they mine trace minerals out of rocks, delivering them right to a plant’s doorstep, and in some cases right inside a plant’s green body. Soil life consumes and solubilizes mineral elements plants need but that remain locked up in soil particles unless biological assistance liberates them. What do the plants provide in return? They exude sugars and amino acids out of their roots to feed the soil life that promotes plant growth and health. This, in turn, produces more organic matter and exudates. And the tips of roots shed border cells that form ­polysaccharide-rich mucus, adding more organic matter to the rhizosphere. This promotes colonization by certain bacterial species in the soil around the growing root tip, which, in turn, helps suppress pathogenic fungi that infect root apices through competition that denies them the opportunity to occupy that space. Just as in the soil, members of our gut microbiome not only help keep their pathogenic brethren at bay, they also make compounds and molecules our bodies cannot make and yet we need for health (Montgomery & Biklé, 2016). Why do such relationships work across the microbial, plant, and animal kingdoms? They’re mutually beneficial. Conservation Agriculture takes a page from nature’s playbook in relying on practices that cultivate beneficial soil life, putting trillions of them to work for the farmer instead of making the farmer fight an uphill battle. Soil health and human health are connected on several levels. At a societal level, agricultural prosperity over the long run depends on the health of the soil (Montgomery, 2007a). At the farm scale, healthy soil benefits farm income, and reduced need for and reliance on agrochemicals would benefit farmer and farm worker health. For individuals it is worth asking to what degree the mineral and phytochemical density of our food reflects agronomic practices. Rethinking how we approach the health of the soil is a potentially revolutionary development.

11.5 The fifth revolution Rebuilding healthy, fertile soils as a consequence of intensive agricultural practices is one of the critically important challenges that humanity faces in the 21st century. Yet, practices based on two of the three core principles of Conservation Agriculture— cover crops and crop rotations or associations—are based on traditional practices adopted long ago in regions around the world. Coupling them with new technologies and methods can enable farmers to rebuild the fertility of the land as a consequence of intensive agriculture. Widespread adoption of practices based on the principles of Conservation Agriculture holds the potential to open a new chapter in humanity’s agricultural history (Montgomery, 2017). So far, in my view, humanity has pulled off four major agricultural revolutions. The first involved the break with hunting and gathering to planting and cultivating crops

Soil health and the revolutionary potential of Conservation Agriculture 227

followed by adoption of animal labor and the plow. The second arrived at different times around the world as farmers developed what became traditional practices to bolster soil fertility—rotating crops, planting legumes, and allowing livestock to manure fields. The third involved mechanization and industrialization as fossil fuels powered machinery that displaced the need for animal labor, and chemical fertilizers displaced the need for on-farm supplies of manure. The fourth came when the Green Revolution and subsequent biotechnology explosion reengineered the foundation of modern conventional agriculture. Now, the potential for a fifth agricultural revolution—one based on soil health— lies rooted in how we see the soil. Perhaps this should not be too surprising. Our understanding of the soil has changed dramatically since the dawn of agriculture, and could do so again. For most of the postglacial world, people have seen soil as something to be worked, an arena where human labor strove to reshape and compete against nature. With the advent of the Renaissance, soil seemed a decipherable mystery that could be understood through the application of reason. As natural philosophers began to contemplate its secrets, Leonardo da Vinci famously noted that, “We know more about the stars overhead than the soil underfoot.” While his words still ring true today, the principles of Conservation Agriculture bring into sharper focus a sound recipe for working with rather than against soil life—for cultivating instead of combating the hidden half of nature. Through the experience of centuries, farmers learned to see the combination of crop rotations, planting legumes, and animal husbandry as central to improving land and building fertile soil. Until, that is, the 19th century discovery of how chemical fertilizers could boost crop yields on even seriously degraded land led to seeing soil as a medium in which to root crops spoon-fed an agrochemical diet. In concert with the mechanization that reshaped agriculture in the 20th century this made soil the least expensive—and least valued—input in industrial crop production. Over the past century this perspective seriously degraded civilization’s agricultural foundation of healthy, fertile soil. Conservation Agriculture does not present an either/or choice between modern technology and time-tested traditions. It offers a way to update and combine traditional wisdom with new agronomic science and technology to change how we treat soil. Its principles offer flexible, adaptable guidelines for restoring soil health and ensuring that individual farmers can make a living without degrading our collective ability to feed the future. Additional societal benefits of regenerative farming practices that build soil health include pulling carbon from the atmosphere and putting it back into the soil, greater drought resilience, and a lower environmental footprint from reduced agrochemical use (Montgomery, 2017). Recent discoveries about the role of soil ecology in nutrient availability and cycling are changing views of soil fertility as we begin to accept that soil biology matters as much as soil chemistry and physics. Recognizing the foundational ecological role of soil life—and soil health—we can see the societal importance of agricultural soils rich in organic matter and the value of reinvesting in civilization’s most fundamental—and neglected—resource. To do this on a global scale we need a new system of

228

Rethinking Food and Agriculture

farming, an agricultural system that rebuilds and improves soil health. The reduced fuel, fertilizer, and pesticide use that would accompany such a transition would not only make economic sense for farmers but would greatly reduce agriculture’s environmental footprint.

References Anderson, C., Beare, M., Buckley, H. L., & Lear, G. (2017). Bacterial and fungal communities respond differently to varying tillage depth in agricultural soils. PeerJ, 5, e3930. Bai, Z. G., Dent, D. L., Olsson, L., & Schaepman, M. E. (2008). Proxy global assessment of land degradation. Soil Use and Management, 24, 223–234. Baumhardt, R. L., Stewart, B. A., & Sainju, U. M. (2015). North American soil degradation: Processes, practices, and mitigating strategies. Sustainability, 7, 2936–2960. Blakemore, R. J. (2018). Critical decline of earthworms from organic origins under intensive, humic SOM-depleting agriculture. Soil Systems, 2, 33. Brink, R. A., Densmore, J. W., & Hill, G. A. (1977). Soil deterioration and the growing world demand for food. Science, 197, 625–630. Campbell, J. E., Lobell, D. B., Genova, R. C., & Filed, C. B. (2008). The global potential of bioenergy on abandoned agriculture lands. Environmental Science and Technology, 42, 5791–5794. de Oliveira Ferreira, A., et al. (2016). Can no-till grain production restore soil organic carbon to levels natural grass in a subtropical Oxisol? Agriculture, Ecosystems and Environment, 229, 13–20. Dregne, H. E., & Chou, N. T. (1992). Global desertification dimensions and costs. In Degradation & restoration of arid lands (pp. 73–92). Lubbock, TX: Texas Tech. University. FAO. (2015). In L. Montanarella, et al. (Eds.), Status of the world’s soil resources. Rome: Food and Agricultural Organization of the United Nations. Francis, G. S., & Knight, T. L. (1993). Long-term effects of conventional and no-tillage on selected soil properties and crop yields in Canterbury, New Zealand. Soil and Tillage Research, 26, 193–210. Gibbs, H. K., & Salmon, J. M. (2015). Mapping the world’s degraded lands. Applied Geography, 57, 12–21. Haddaway, N. R., et al. (2017). How does tillage intensity affect soil organic carbon? A systematic review. Environmental Evidence, 6, 30. Hendrix, J. W., Guo, B. Z., & An, Z.-Q. (1995). Divergence of mycorhizal fungal communities in crop production systems. In H. P. Collins, G. P. Robertson, & M. J. Klug (Eds.), The significance and regulation of soil biodiversity (pp. 131–140). Dordrecht, Netherlands: Kluwer Academic Publishers. IPBES. (2018). In R. Scholes, et al. (Eds.), Summary for policymakers of the thematic assessment report on land degradation and restoration of the intergovernmental science-policy platform on biodiversity and ecosystem services. Bonn, Germany: IPBES Secretariat. Kassam, A., Basch, G., Friedrich, T., Shaxson, F., Goddard, T., Amado, T., … Mkomwa, S. (2013). Sustainable soil management is more than what and how crops are grown. In R. Lal, & R. A. Stewart (Eds.), Principles of soil management in agro-ecosystems. 2013. Advances in soil science CRC Press. Kassam, A., Friedrich, T., & Derpsch, R. (2019). Global spread of conservation agriculture. International Journal of Environmental Studies, 76, 29–51.

Soil health and the revolutionary potential of Conservation Agriculture 229

Khan, S. A., Mulvaney, R. L., Ellsworth, T. R., & Boast, C. W. (2007). The myth of nitrogen fertilization for soil carbon sequestration. Journal of Environmental Quality, 36, 1821–1832. Kinoshita, R., Schindelbeck, R. R., & van Es, R. M. (2017). Quantitative soil profile-scale assessment of the sustainability of long-term maize residue and tillage management. Soil Tillage Research, 174, 34–44. Kumar, A., et al. (2017). Effects of maize roots on aggregate stability and enzyme activities in soil. Geoderma, 306, 50–57. LaCanne, C. E., & Lundgren, J. G. (2018). Regenerative agriculture: Merging farming and natural resource conservation profitably. PeerJ, 6, e4428. Meade, R. H. (1982). Sources, sinks, and storage of river sediment in the Atlantic drainage of the United States. Journal of Geology, 90, 235–252. Montgomery, D. R. (2007a). Dirt: The erosion of civilizations. Berkeley, CA: University of California Press. Montgomery, D. R. (2007b). Soil erosion and agricultural sustainability. Proceedings of the National Academy of Sciences, 104, 13268–13272. Montgomery, D. R. (2017). Growing a revolution: Bringing our soil back to life. New York: W. W. Norton. Montgomery, D. R., & Biklé, A. B. (2016). The hidden half of nature: The microbial roots of life and health. New York: W. W. Norton. Mulvaney, R. L., Khan, S. A., & Ellsworth, T. R. (2009). Synthetic nitrogen fertilizers deplete soil nitrogen: A global dilemma for sustainable cereal production. Journal of Environmental Quality, 38, 2295–2314. Nunes, M. R., et al. (2018). No-till and cropping system diversification improve soil health and crop yield. Geoderma, 328, 30–43. Oldeman, L. R. (1994). The global extent of soil degradation. In D. J. Greenland, & I. Szabolcs (Eds.), Soil resilience and sustainable land use (pp. 99–119). Wallingford: CAB International. Pimentel, D., et al. (1995). Environmental and economic costs of soil erosion and conservation benefits. Science, 267, 1117–1123. Powlson, D. S., et al. (2014). Limited potential of no-till agriculture for climate change mitigation. Nature Climate Change, 4, 678–683. Pretty, J. N., et al. (2006). Resource-conserving agriculture increases yields in developing countries. Environmental Science & Technology, 40, 1114–1119. Raun, W. R., & Johnson, G. V. (1999). Improving nitrogen use efficiency for cereal production. Agronomy Journal, 91, 357–363. Ross, C., et  al. (2002). Soil quality under long-term cropping by no-tillage compared with conventional cultivation and permanent pasture in the Manawatu. In L. D. Currie, & P. Loganathan (Eds.), Dairy farm soil management (pp. 119–126). Palmerston North, New Zealand: Fertilizer and Lime Research Centre, Massey University. Occasional report no. 15. Sanchez, P. A. (2002). Soil fertility and hunger in Africa. Science, 295, 2019–2020. Sanchez, P. A., & Swaminathan, M. S. (2005). Hunger in Africa: The link between unhealthy people and unhealthy soils. The Lancet, 365, 442–444. Sharma, P., Singh, G., & Singh, R. P. (2013). Conservation tillage and optimal water supply enhances microbial enzyme (glucosidase, urease and phosphatase) activities in fields under wheat cultivation during various nitrogen management practices. Archives of Agronomy and Soil Science, 59, 911–928. Tian, D., & Niu, S. (2015). A global analysis of soil acidification caused by nitrogen addition. Environmental Research Letters, 10, 024019. Wood, S. A., & Baudronc, F. (2018). Soil organic matter underlies crop nutritional quality and productivity in smallholder agriculture. Agriculture, Ecosystems and Environment, 266, 100–108.

This page intentionally left blank

Climate change adaptability and mitigation with Conservation Agriculture

12

Emilio J. Gonzalez-Sancheza,b,c, Oscar Veroz-Gonzalezc, Manuel Moreno-Garciad, Manuel R. Gomez-Arizac, Rafaela Ordoñez-Fernandezc,d, Paula Trivino-Tarradasa, Amir Kassame, Jesús A. Gil-Ribesa,c, Gottlieb Baschb,f, and Rosa Carbonell-Bojollod a Department of Rural Engineering, School of Agricultural and Forestry Engineering, University of Cordoba, Cordoba, Spain, bEuropean Conservation Agriculture Federation (ECAF), Brussels, Belgium, cAsociación Española Agricultura de Conservación, Suelos Vivos (AEAC.SV), Cordoba, Spain, dArea of Ecological Production and Natural Resources, IFAPA Centro Alameda del Obispo, Cordoba, Spain, eUniversity of Reading, Reading, United Kingdom, fInstituto Ciências Agrárias e Ambientais Mediterrânicas, Universidade de Évora, Évora, Portugal

12.1 Introduction The study of the climate is a complex field of investigation and in constant evolution. Since the climate is influenced by a great number of factors, it is not a static system and therefore it is difficult to determine its effects with precision. In any case, climate change is occurring more rapidly than initially predicted (IPCC, 2007). Global warming due to human causes has its origins back in the 19th century. Although agriculture at that time was probably not a sector that contributed much to the phenomenon, current climate change models are based on what has subsequently contributed to this global warming and to climate change. Currently, agriculture accounts for 12% of the total global greenhouse gas (GHG) emissions (IPCC, 2019). From the very beginning of agriculture, soil tillage was developed as a means to facilitate the destruction of the prevailing vegetation and the control of weeds, as well as promote the establishment and growth of crops. These practices became more important with the use of tools pulled by animals and, when tractors became widely available in agriculture by the mid-20th century, tillage became one of the most important agricultural field tasks. Although the first plows were very primitive, over time these implements were progressively improved. The Romans, in particular, improved the plow to such an extent that the key elements of the ancient design are still part of common plows. Cultural and traditional knowledge, which is required for a good harvest tillage, was permanently imprinted in the minds of farmers for centuries. In fact, tillage has some benefits as it allows a proper incorporation of amendments and manure, promotes rapid nitrification of all organic matter present in the soil (with consequent short-term Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00012-8 © 2021 Elsevier Inc. All rights reserved.

232

Rethinking Food and Agriculture

positive and long-term negative effects), serves to decompact soils trampled by cattle, controls weeds present on the surface, and buries their seeds, among other aspects. Thus it is unquestionable that the plow played a fundamental role in agriculture. With the increased availability of tractors, moldboard plows were gradually adopted by farmers. These plows were able to mobilize a greater volume of soil, resulting in increased depth of tillage operations, while decreasing working time compared to animal-pulled plows. The availability of modern technical means, together with the perception that it was essential to plow to obtain good yields, soon led many farmers to practice excessive tillage. This has ended up causing agrienvironmental problems such as soil compaction, erosion, organic matter decline, and biodiversity loss, among others, exceeding any benefits. Even so, cultivation of the soil was passed down from generation to generation until it became part of the culture of farmers. However, tillage is one of the reasons why conventional agriculture is a net emitter of GHGs. The contribution of the agricultural sector to climate change is gaining more and more visibility leading to increasing interest in learning how agriculture can mitigate GHG emissions. Even if agriculture were not the only productive sector affected by climate change, the impacts on it would definitely have negative effects on food security and social welfare. Therefore it is essential to identify indicators that monitor the dynamics of changes and trends, and to assess the reduction of emissions achievable through better agricultural practices. According to the European Environment Agency, global warming has already altered the duration of the growing season in some areas. The periods of flowering and harvest of cereals are already several days ahead. It is foreseeable that these changes may continue to occur in many regions (EEA, 2016). A change is needed.

12.2 Climate change and agriculture: Why do we need to change the agricultural paradigm? Global GHG emissions were estimated to be 9 (± 4.5) Gt CO2 eq. in 2010 (IPCC, 2014), with approximately 21.2%–24% (10.3–12 Gt CO2 eq.) of emissions originating from soils in agricultural, forestry, and other land use (IPCC, 2014; Tubiello et al., 2015). Annual non-CO2 GHG emissions (primarily CH4 and N2O) from agriculture were estimated to be 5.2–5.8 Gt CO2 eq. year− 1 in 2010 (FAOSTAT, 2019; Tubiello et al., 2015), with approximately 4.3–5.5 Gt CO2 eq. year− 1 attributable to land use and land use change activities (IPCC, 2014). There is a link between soils and climate change. Fuel burning by agricultural machinery is often regarded as the main source of CO2 emissions in the primary sector, neglecting the CO2 emissions derived from agricultural land caused by the “burning” of organic residues left after harvest and soil organic carbon losses caused by intensive plow-based tillage, which is still considered “normal” and “good agricultural practice.” Intensification of agricultural production is an important factor influencing GHG emissions, particularly the relationship between intensive tillage and soil carbon loss (Reicosky & Archer, 2007).

Climate change adaptability and mitigation with CA

233

Agricultural soils are able to mitigate and adapt to climate change (GonzálezSánchez et al., 2017; Lal et al., 2011). The key to doing this is in the way that soil organic carbon is managed. For climate change mitigation, reducing carbon oxidation processes by diminishing the intensity of tillage is needed to increase soil organic matter (FAO, 2019; González-Sánchez, Ordóñez-Fernández, Carbonell-Bojollo, VerozGonzález, & Gil-Ribes, 2012; Nelson et al., 2009). Indeed, many authors agree that soil disturbance by tillage is one of the main causes of organic carbon losses in the soil (Balesdent, Mariotti, & Boisgontier, 1990; Olson, Lang, & Ebelhar, 2005; Six et al., 2004). Reicosky (2011) argues that intensive agriculture has contributed to the loss of between 30% and 50% of soil organic carbon in the last 2 decades of the 20th century. Kinsella (1995) estimates that, in only the first 10 years of tillage agriculture, 30% of the original organic matter in the soil was lost. In Europe, there are several estimations of carbon loss in agricultural soils. Vleeshouwers and Verhagen (2002) estimated an average loss of 78 Tg of carbon per year in the European Union. Ordóñez-Fernández, González-Fernández, Giráldez-Cervera, and Perea-Torres (2007) observed in Spain that 10 years of continuous tillage caused a decrease of 18% in organic matter content in the first 20 cm of a vertisol. Conventional tillage-based agriculture has driven numerous well-known environmental problems (Table  12.1). It is not only natural resources that suffer from ­tillage-based agriculture. The conventional system has also reached a plateau in average yields. According to Brisson et al. (2010), recent decades have witnessed a decline in the growth trend of cereal yields in many European countries, where access to the newest technologies is not a barrier. Effects in genetic changes through breeding, agronomy, and climate are investigated as possible causes. However, if farmers have the best technology at hand, why is there yield stagnation in many countries in Europe? In our view, soil degradation due to tillage is likely to be one of the major causes of this suboptimum yield ceiling and productivity (Brisson et al., 2010; Kassam et al., 2013). The Paris Agreement (2015) at COP21 by the United Nations Framework Convention on Climate Change builds upon previous COPs and for the first time brings all nations into a common cause to undertake ambitious efforts to combat climate change and adapt to its effects, with enhanced support to assist developing countries to do so. As such, it charts a new course in the global climate effort. Some world leaders, like the President of the European Commission, have stated the need for action even beyond the Paris Agreement (Von der Leyen, 2019), but the business-as-usual situation is unlikely to deliver this. So, there is a need for shifting the conventional tillage-based agricultural model to a new one. According to recent scientific studies, Conservation Agriculture (CA) is a promising sustainable alternative.

12.3 Conservation Agriculture: A sustainable farming system that mitigates and adapts to climate change CA is a holistic approach to sustainability, which comprises the practical application of three interlinked principles, namely: no or minimum mechanical soil disturbance, biomass mulch soil cover, and crop species diversification, in conjunction with other

Table 12.1  Agrienvironmental problems in and its relation to agricultural practices. Agrienvironmental problems

Crop type Herbaceous

Woody

Compaction

CO2 emissions

Decrease in biodiversity

Pollution of surface water

Pollution by pesticides

+ + ++++ +++++ ++

++ ++ ++++ +++++ ++

− − ++++ +++++ ++

– + +++ +++++ ++

− + ++++ +++++ ++

− + ++++ +++++ ++

+++

+++

+++

+++

+++

+++

+++

+++++

++++

+++++

+++++

+++++

+++++

+++++

Agricultural practices

Erosion/ Desertification

Organic matter

CT MT NT NT + CC Timid CC 30% Average CC 60% Total CC 90%

− − ++++ +++++ ++

CC, cover crop; CT, conventional tillage; MT, minimum tillage; NT, no-tillage. CC 30%: cover crop in between tree rows, in 30% of the surface without trees. CC 60%: Id. Id. 60%; CC 90%: Id. Id. 90%. Grading of environmental effect: +, timidly positive; +++++, very positive. – indifferent or negative. Adapted from Gonzalez-Sanchez, E. J., Veroz-Gonzalez, O., Blanco-Roldan, G. L., Marquez-Garcia, F., Carbonell-Bojollo, R. (2014). A renewed view of conservation agriculture and its evolution over the last decade in Spain. Soil and Tillage Research, 146(PB), 204–212.

Climate change adaptability and mitigation with CA

235

complementary good agricultural practices of integrated crop and production management (Kassam, Friedrich, & Derpsch, 2019). ●





Continuous no or minimum soil disturbance. In practice this means no-tillage. At least 30% of the soil must be covered after sowing to effectively protect it against erosion. However, it is recommended to leave more than 60% of the soil covered to have almost complete control over the soil degradation process. Permanent soil mulch cover. In other words, this means maintaining crop biomass and stubble in annual crops and planting or preserving groundcovers between rows of trees in woody crops. In this way, soil organic matter and water infiltration into the soil increases, some weeds are inhibited, and water evaporation from the soil is limited. Crop diversification through rotations and associations. In this way, pests and diseases are better controlled by breaking cycles that are maintained in monocultures, in addition to including crops that can improve the natural fertility of the soil and biodiversity.

CA is one of the most studied agrosciences in the world. In 2015/16, CA was practiced globally on about 180 M ha of cropland, corresponding to about 12.5% of total global cropland. In 2008/09, the spread of CA was reported to be about 106 M ha. This change constitutes an increase of some 69% globally since 2008/09. In 2015/16, CA adoption was reported in 78 countries, an increase in adoption by 42 more countries since 2008/09. The average annual rate of global expansion of CA cropland area since 2008/2009 has been 10.5 M ha. Adoption has been greatest in South and North America, followed by Australia and New Zealand, Asia, Russia and Ukraine, Europe, and Africa (Kassam et al., 2019). The principles of CA have been recognized by different initiatives. LIFE + ClimAgri (www.climagri.eu) project’s best management practices (BMP) are presented in Table 12.2, where CA principles are comprised in numbers BMPs 1, 2, and 3. CA mitigates climate change and also favors the adaptation of agricultural ecosystems to the negative effects of climate change by increasing crop resilience, efficiency, and productivity (González-Sánchez et  al., 2017; Kassam et  al., 2013; Lal, 2010). Sometimes controversial results can be found in the literature that are attributed to CA (Giller, Witter, Corbeels, & Tittonell, 2009; Gowing & Palmer, 2008; Pittelkow et al., 2014; Powlson, Whitmore, & Goulding, 2011, Powlson et al., 2014; Sumberg & Thompson, 2012). However, a closer look at these results often reveals that some of the key interlinked CA principles were not applied, thus the results were not based on CA systems (Derpsch et al., 2014; Reicosky, 2015). Indeed, according to Derpsch et al. (2014), there is a broad lack of understanding of what CA systems research means.

12.3.1 Conservation Agriculture: Mitigating climate change CA has a double effect on the reduction of GHG concentration in the atmosphere. On the one hand, the changes introduced by CA increase the carbon in the soil, reducing its emissions into the atmosphere (Fig. 12.1). On the other hand, the drastic reduction of tillage along with ceasing the mechanical alteration of the soil lead to a fall in CO2 emissions because of the reduction of the mineralization processes of soil organic matter as well as from energy saving and lower fuel consumption. Through CA, farmers can both sequester CO2 in the form of soil organic carbon and organic matter, and reduce the emissions of CO2 to the atmosphere (Fig. 12.2).

Table 12.2  LIFE + ClimAgri project best management practices and major contributions to either climate change mitigation and/or adaptation. Best management practice

Mitigation

Adaptation

Increased carbon sequestration in soil due to more biomass returned Reduction of CO2 emissions due to lower energy consumption and lower emissions from the ground Increased carbon sequestration in soil due to more root biomass

Increase in soil resilience and function

1

Permanent soil cover

2

No or minimum soil disturbance

3

Crop diversification

4

Optimized use of inputs

Reduction of CO2 emissions due to lower energy consumption

Increase in input use efficiency and factor productivity

5

Appropriate use of agrochemicals

Reduction of CO2 emissions due to lower energy consumption

Increase in agrochemical use efficiency and productivity

6

Use of advanced technology (decision-making aid systems, precision agriculture, etc.)

Reduction of CO2 emissions due to lower energy consumption

Increase in energy and labor use efficiency and productivity

Increase in soil resilience and function

Increase in soil resilience and function

7

Implementation of optimum and deficit irrigation strategies

Reduction of CO2 emissions due to lower energy consumption

Improved water use efficiency and productivity

8

Joint consideration of optimized agricultural, technical, and financial practices to improve irrigation water management

Reduction of CO2 emissions due to lower energy consumption

Improved water use efficiency Reduced risk for crops from temperature and water stress conditions

9

Implementation of multifunctional margins and retention structures

Increased carbon sequestration in soil

Increased resilience of the agricultural ecosystem

10

Measures for the promotion of biodiversity

Reduction in agrochemical application

Increased resilience of the agricultural ecosystem

Source: Adapted from www.climagri.eu.

238

Rethinking Food and Agriculture

Fig. 12.1  Barley in Conservation Agriculture (Rabanales University of Cordoba farm, Spain).

Fig. 12.2  Mitigating climate change mechanisms through Conservation Agriculture. Source: Own elaboration.

Climate change adaptability and mitigation with CA

239

By leaving crop biomass on the soil surface, CA induces an organic matter dynamic closer to natural ecosystems than conventional tillage-based agriculture can. On the other hand, the less soil is disturbed through tillage, the more it is able to absorb and store carbon. This carbon would previously have been fixed in the plant biomass, including roots, due to photosynthesis synthesizing more organic matter, which in the long run increases soil productive capacity, and at the same time decreases CO2 emissions. Many studies and metaanalyses demonstrate that with CA it is possible to sequester significant amounts of soil carbon per hectare per year with annual and woody crops, compared to tillage-based systems. The estimation for EU-28 countries of potential soil organic carbon sequestration through the adoption of CA in annual crops when compared to conventional tillage-based systems would be 189 M t ha− 1 year− 1 (González-Sánchez et al., 2017). Similarly, the estimate of annual carbon sequestration potential in African agricultural soils through CA amounts to 145 M t of C per year, that is 533 M t of CO2 per year. This figure represents about 95 times the current sequestration figure (Gonzalez-Sanchez et al., 2019). Based on the potential of CO2 sequestration in Europe of 189 M t ha− 1 year− 1, the average yearly CO2 sequestration that can be achieved by the implementation of CA is 1.82 t ha− 1 year− 1. This amount is derived by dividing the CO2 sequestration in Europe by the potential area that can be managed under CA, which is 103 M ha. To put these figures in context, some facts and figures are provided (González-Sánchez et al., 2017): ●











Just 4 ha under CA would negate the average annual emissions of a European citizen. Adoption of CA across Europe would sequester the CO2 emitted by 18 million households, or the emissions from electricity generation for 25 million households. Carbon sequestration due to the adoption of CA across Europe would be equivalent to emissions saving obtained by the installation of over 43,000 wind turbines. Implementation of CA in Europe would reduce emissions equal to the closure of 50 coalfired power plants. If all European farmland was converted to CA, it would reduce atmospheric carbon by as much as planting 65 million ha of forest. For every hectare converted to CA in Europe the emissions of a return flight from London to Athens are removed from the atmosphere.

Besides the soil carbon sink effect, the adoption of CA supposes a drastic reduction of tillage operations. The LIFE + Agricarbon project (www.agricarbon.eu) demonstrated the notable savings in respect of energy utilization of CA due to higher efficiency and optimized use of inputs. After four seasons, energy savings were estimated at 13.8% for wheat, 21.6% for sunflowers, and 24.4% for leguminous plants. These savings mean lower CO2 emissions, corresponding to 166 kg ha− 1 for wheat, 64 kg ha− 1 for sunflowers, and 107 kg ha− 1 for leguminous plants. Carbon dioxide emissions derived from the mechanical action on the soil are directly related to the stability of its aggregates. Under natural conditions, organic matter is encapsulated inside the aggregates, and it is not accessible to the attack of

240

Rethinking Food and Agriculture

­ icroorganisms present in the soil. The less stable an aggregate, the lower its resism tance to alteration processes that may cause its breakage and therefore the organic matter of its interior may be more easily accessible to microorganisms, favoring the processes of mineralization and CO2 generation as a by-product that would be emitted into the atmosphere. The biological carbon-based inert cement that encapsulates and binds soil particles and organic matter together is called glomalin. Glomalin is produced by mycorrhizae, which form a vast network in the soil, connecting individual plants with each other through their root systems. This symbiotic relationship has positive impact on water and nutrient availability and on plant resilience. Under CA, mycorrhizae are encouraged to flourish, whereas under tillage-based agriculture, they are destroyed and dysfunctional. The adoption of CA’s interlinked principles has allowed not only greater control of soil erosion but also a decrease in organic matter losses and CO2 emissions as a result of nonsoil disturbance. The nonalteration of the soil promoted by conservation practices improves its structure, increasing the stability of the aggregates against the processes of disaggregation. This allows greater protection of soil organic matter against the attacks of the edaphic microfauna, and the maintenance of the CO2 “trapped” in the porous space of the soil, resulting from the mineralization processes of organic matter. Therefore the avoidance of tillage reduces and slows the decomposition of crop residues, storing the atmospheric CO2 (fixed in the structure of the crop and returned to the ground in the form of crop residue) in the soil. In this way, the soil is able to store atmospheric CO2, thus helping to mitigate the GHG emissions generated by other activities. In a study carried out in the United States (Reicosky & Archer, 2007), the shortterm effects on CO2 emissions of two soil management systems were evaluated (Fig. 12.3). The investigation found higher emissions in both the short and medium term from conventionally tilled plots in comparison with no-tillage plots. Compared

Fig. 12.3  Accumulated CO2 emissions (g m− 2) 5 h after tillage. Source: Adapted from Reicosky, D.C. (1997). Tillage-induced methods CO2 emissions from soil. Nutrient Cycling in Agroecosystems, 49, 273–285.

Climate change adaptability and mitigation with CA

241

to those from the no-till system, emissions were 3.8 times higher in tilling processes where tillage was more superficial (10 cm) and 10.3 times higher in the case of deeper tillage.

12.3.2 Conservation Agriculture: Adapting to climate change It is not only important to adopt strategies to mitigate phenomena that contribute to climate change, but also to adopt practices that increase the resilience of agricultural ecosystems to the consequences of climate change, and that favor the adaptation of crops to the new climatic scenarios predicted by the atmospheric circulation models. The term “adaptation” refers to all adjustments that need to be made in a system to respond to actual or anticipated changes resulting from climate change, and taking advantage of the opportunities given by the new climatic scenarios. According to Lal (2010), CA is a good strategy not only to mitigate climate change, but also to adapt agricultural ecosystems to their effects, by increasing crop resilience facing climatic variations. In CA systems, erosion is reduced, the quality and fertility of the soil is improved, and the runoff water loss is reduced, allowing the crop to have more water in dry periods (Table 12.3). All this increases the responsiveness of CA systems to changes in climate, meaning crops under CA systems have a much better capacity for adaptation. The adoption and development of CA practices lead to a number of benefits in the water supply system within the agricultural ecosystems, such as greater availability of water for the crop and improvement of its quality (Fig. 12.4). The adaptation advantages offered by CA will be particularly beneficial in ecosystems that experience a decrease in availability of water resources, and/ or in those regions where the phenomena of runoff are increased due to the increase of extreme precipitation events. On the basis of studies published by the European Environment Agency (EEA, 2012), a reduction in precipitation in the Mediterranean and Continental regions is expected. Therefore there will be an increased demand for water resources in agriculture, which will make some areas more vulnerable due to the lack of water. On the other hand, an increase in extreme precipitation events in the Atlantic regions is expected, which will affect water quality and facilitate erosion processes. Concerning water balance of the soil-cropping system, CA systems improve the uptake, conservation, and better use of available water in the soil by the crops, thanks to the fact that these systems favor infiltration, reduce runoff, increase water-holding capacity, and reduce evaporation. All of this is achieved due to the maintenance of soil mulch cover with crop biomass, the effects of which are described in Table 12.4. According to López-Garrido, Madejón, Murillo, and Moreno (2010), in fields under CA systems, the volumetric space in the first 20 cm is 20%–30% higher than in soils under conventional tillage. Muriel et al. (2005) concluded that CA techniques not only allow a greater retention of water in the soil profile, especially in the first 30 cm of depth, but also slow down the water discharge rate, which has a positive impact on the development of spring/summer crops, where the limiting factor of production is the availability of water.

242

Rethinking Food and Agriculture

Table 12.3  Possible actions to increase resilience of agrarian ecosystems and agricultural techniques helping to adapt to climate change. Components

Actions to increase resilience

Suggested techniques/systems

Water

Increased infiltration Reduced runoff Optimization of water use Improvement of water balance

Soil

Reduced runoff Increase in organic carbon (organic matter) Improvement of structure Increased soil fertility Increase of the epigeal fauna Improvement of conditions for the habitability of steppe birds Enhancement of pollinating species

Conservation Agriculturea Deficit irrigation Precision farming Improvement of irrigation infrastructures and pipelines Use of irrigation monitoring systems Implantation of green filters, protective bands, and vegetation in the margins of the plot (multifunctional margins) Conservation Agriculturea High flotation tires Soil health cards

Biodiversity

Crops

Increased resistance to drought Escape from water stress Reduction of weeds Reduced incidence of pests and diseases

Conservation Agriculturea Use of integrated pest management Implantation of green filters, protective bands, and vegetation in the margins of the plot (multifunctional margins) Conservation Agriculturea Use of varieties resistant to drought Advancement of planting date Use of native varieties Crop cycle variation Use of integrated pest management

a

Conservation Agriculture refers to the three interlinked practices of the Conservation Agriculture system. Source: Own elaboration.

Finally, CA reduces water evaporation as it prevents the direct incidence of radiation on moist soil and reduces the turbulent transfer of vapor into the atmosphere. As a result, crops in dry lands can better withstand difficult conditions. Thus by keeping the soil unaltered and covered by crop residues, CA causes a decrease in soil water evaporation during periods of high temperatures, and this means that the soil stays wetter during the summer and early spring. Moret, Arrúe, López, and Gracia (2006) observed, during three periods of long fallow (16–18  months), that soil under an intensive tillage system with a moldboard plow had lost 14 times more water by evaporation than in the no-tillage system in the 24 h after the first tillage operations.

Climate change adaptability and mitigation with CA

243

Fig. 12.4  Conservation Agriculture processes related to water benefits. Source: Own elaboration. Table 12.4  Effects of permanent soil covers on edaphic moisture. Effects

Direct causes

Indirect causes

Increase in infiltration— reduction of runoff

Greater retention of rainwater in permanent soil covers

Through the increase of organic matter, disintegration resistance and structure are improved Increased soil fauna (earthworms), which generate galleries and pores, favoring the circulation of water

Protection of the soil against the impact of raindrops Reduction of evaporation

No direct incidence of radiation on wet soil Reduction of the turbulent transfer of water vapor to the atmosphere

Source: Own elaboration.

12.4 Conclusions The status quo of agriculture based on soil tillage is unacceptable from a climate point of view. To reverse agriculture’s field performance from that of a net GHG emitter to a GHG mitigator requires a new paradigm.

244

Rethinking Food and Agriculture

CA is a holistic agricultural system that is able to mitigate and adapt to climate change. The three interlinked principles of CA enable the system to deliver many benefits in terms of carbon sequestration and climate adaptation, especially with regards to soil, water, nutrient, and energy management.

Acknowledgments To the European Commission for cofinancing the project LIFE AGROMITIGA-LIFE17 CCM/ES/000140: Development of climate change mitigation strategies through carbon-smart agriculture.

References Balesdent, J., Mariotti, A., & Boisgontier, D. (1990). Effects on tillage on soil organic carbon mineralization estimated from 13C abundance in maize fields. Journal of Soil Science, 41, 584–596. Brisson, N., Gate, P., Gouache, D., Charmet, G., Oury, F. X., & Huard, F. (2010). Why are wheat yields stagnating in Europe? A comprehensive data analysis for France? Field Crops Research, 119(1), 201–212. https://doi.org/10.1016/j.fcr.2010.07.012. Derpsch, R., Franzluebbers, A. J., Duiker, S. W., Reicosky, D. C., Koeller, K., Friedrich, T., … Weiss, K. (2014). Why do we need to standardize no-tillage research? Soil and Tillage Research, 137, 16–22. EEA. (2012). Climate change, impacts and vulnerability in Europe 2012. https://www.eea.europa. eu/publications/climate-impacts-and-vulnerability-2012. (accessed 19 October 2019). EEA. (2016). Agriculture and climate change. European Environment Agency. https://www. eea.europa.eu/signals/signals-2015/articles/agriculture-and-climate-change. (accessed 19 October 2019). FAO. (2019). Conservation agriculture website. http://www.fao.org/conservation-agriculture/en/. (accessed 2 December 2019). FAOSTAT. (2019). FAOSTAT website. http://www.fao.org/faostat/en/#home. (accessed 20 November 2019). Giller, K. E., Witter, E., Corbeels, M., & Tittonell, P. (2009). Conservation agriculture and smallholder farming in Africa: The heretics’ view. Field Crops Research, 114(1), 23–34. https://doi.org/10.1016/j.fcr.2009.06.017. González-Sánchez, E. J., Moreno-García, M., Kassam, A., Holgado-Cabrera, A., TriviñoTarradas, P., Carbonell-Bojollo, R., … Basch, G. (2017). Conservation agriculture: Making climate change mitigation and adaptation real in Europe. European Conservation Agriculture Federation (ECAF). González-Sánchez, E. J., Ordóñez-Fernández, R., Carbonell-Bojollo, R., Veroz-González, O., & Gil-Ribes, J. A. (2012). Meta-analysis on atmospheric carbon capture in Spain through the use of conservation agriculture. Soil and Tillage Research, 122, 52–60. Gonzalez-Sanchez, E. J., Veroz-Gonzalez, O., Conway, G., García, M., Kassam, A., Mkomwa, S., … Carbonell, R. (2019). Meta-analysis on carbon sequestration through Conservation Agriculture in Africa. Soil and Tillage Research, 190, 22–30. Gowing, J. W., & Palmer, M. (2008). Sustainable agricultural development in sub-Saharan Africa: The case for a paradigm shift in land husbandry. Soil Use and Management, 24, 92–99.

Climate change adaptability and mitigation with CA

245

IPCC. (2007). Climate change 2007: Synthesis report. In R. K. Pachauri & A. Reisinger (Eds.), Contribution of Working Groups I, II and III to the fourth assessment report of the intergovernmental panel on climate change. Core Writing Team. Geneva, Switzerland: IPCC. IPCC. (2014). Summary for policymakers. In O. Edenhofer, R. Pichs-Madruga, Y. Sokona, E. Farahani, S. Kadner, K. Seyboth, … J. C. Minx (Eds.), Climate change 2014, mitigation of climate change, contribution of working group III to the fifth assessment report of the intergovernmental panel on climate change (pp. 1–30). Cambridge, UK and New York, NY, USA: Cambridge University Press. IPCC. (2019). Climate change and land. In IPCC special report on climate change, desertification, land degradation, sustainable land management, food security, and greenhouse gas fluxes in terrestrial ecosystem. Summary for policymakers. WG I, WG II, WG III. IPCC, Geneva, Switzerland. Kassam, A., Basch, G., Friedrich, T., Shaxson, F., Goddard, T., Amado, T., … Mkomwa, S. (2013). Sustainable soil management is more than what and how crops are grown. In R. Lal & R. A. Stewart (Eds.), Advances in soil science. Principles of soil management in agro-ecosystems (pp. 337–400). Boca Raton, FL: CRC Press. Kassam, A., Friedrich, T., & Derpsch, R. (2019). Global spread of conservation agriculture. International Journal of Environmental Studies, 76(1), 29–51. https://doi.org/10.1080/00 207233.2018.1494927. Kinsella, J. (1995). The effects of various tillage systems on soil compaction. In Soil and Water Conservation Society (Ed), Farming for a better environment: A white paper (pp. 15–17). Ankeny, IA: Soil and Water Conservation Society. Lal, R. (2010). Enhancing eco-efficiency in agro-ecosystems through soil carbon sequestration. Crop Science, 50, 120–131. Lal, R., Delgado, J. A., Groffman, P. M., Millar, N., Dell, C., & Rotz, A. (2011). Management to mitigate and adapt to climate change. Journal of Soil and Water Conservation, 66(4), 276–285. López-Garrido, S., Madejón, E., Murillo, J. M., & Moreno, F. (2010). Soil quality alteration by mouldboard ploughing in a commercial farm devoted to no-tillage under Mediterranean conditions. Agriculture, Ecosystems and Environment, 140, 182–190. Moret, D., Arrúe, J. L., López, M. V., & Gracia, R. (2006). Influence of fallowing practices on soil water and precipitation storage efficiency in semiarid Aragon (NE SPAIN). Agricultural Water Management, 82, 161–176. Muriel, J.L., Vanderlinden, K., Perea, F., Jiménez, J.A., García-Tejero, I., Pérez, J.J. (2005). Régimen hídrico en suelos arcillosos de campiña sometidos a distintos sistemas de manejo. En: Actas Congreso Internacional sobre Agricultura de Conservación: 537-542. AEAC.SV. Córdoba. Nelson, R. G., Hellwinckel, C. M., Brandt, C. C., West, T. O., Ugarte de la, T., & Marland, G. (2009). Energy uses and carbon dioxide emissions from cropland production in Unites States, 1990-2004. Journal of Environmental Quality, 38, 418–425. Olson, K. R., Lang, J. M., & Ebelhar, S. A. (2005). Soil organic carbon changes after 12 years of no tillage and tillage of Grantsburg soils in southern Illinois. Soil & Tillage Research, 81, 217–225. Ordóñez-Fernández, R., González-Fernández, P., Giráldez-Cervera, J. V., & Perea-Torres, F. (2007). Soil properties and crop yields after 21 years of direct drilling trials in southern Spain. Soil and Tillage Research, 94, 47–54. Paris Agreement. (2015). UNFCC. https://unfccc.int/process-and-meetings/the-paris-agreement/ the-paris-agreement. (accessed 19 November 2019).

246

Rethinking Food and Agriculture

Pittelkow, C., Xinqiang, L., Lindquist, B., van Groeningen, K. J., Lee, J., Lundy, M. E., … van Kessel, C. (2014). Productivity limits and potentials of the principles of conservation agriculture. Nature Letter, 517, 365–368. Powlson, D. S., Stirling, C. M., Jat, M. L., Gerard, B. G., Palm, C. A., Sanchez, P. A., & Cassman, K. G. (2014). Limited potential of no-till agriculture for climate change mitigation. Nature Climate Change, 4(8), 678–683. Powlson, D. S., Whitmore, A. P., & Goulding, K. W. T. (2011). Soil carbon sequestration to mitigate climate change: A critical re-examination to identify the true and the false. European Journal of Soil Science, 62(1), 42–55. Reicosky, D. C. (2011). Conservation agriculture: Global environmental benefits of soil carbon management. In Vol. 1. Fifth world congress on conservation agriculture (pp. 3–12). Canberra, ACT: ACIAR. Reicosky, D. C. (2015). Conservation tillage is not conservation agriculture. Journal of Soil and Water Conservation, 70(5), 103A–108A. Reicosky, D., & Archer, D. W. (2007). Moldboard plow tillage and short-term carbon dioxide release. Soil and Tillage Research, 94, 109–121. Six, J., Ogle, S. M., Breidt, F. J., Conant, R. T., Mosiers, A. R., & Paustian, K. (2004). The potential to mitigate global warming with no-tillage management is only realized when practiced in the long term. Global Change Biology, 10, 155–160. Sumberg, J., & Thompson, J. (Eds.), (2012). Contested agronomy: Agricultural research in a changing world. London: Routledge. Tubiello, F. N., Salvatore, M., Ferrara, A. F., House, J., Federici, S., Rossi, S., … Smith, P. (2015). The contribution of agriculture, forestry and other land use activities to global warming, 1990–2012. Global Change Biology, 21, 2655–2660. Vleeshouwers, L. M., & Verhagen, A. (2002). Carbon emission and sequestration by agricultural land use: A model study for Europe. Global Change Biology, 8, 519–530. Von der Leyen, U. (2019). Political guidelines for the next European Commission 2019–2024. A Union that strives for more. My agenda for Europe. https://ec.europa.eu/commission/sites/ beta-political/files/political-guidelines-next-commission_en.pdf. (accessed 19 November 2019).

Will gene-edited and other GM crops fail sustainable food systems?

13

Allison K Wilson The Bioscience Resource Project, Ithaca, NY, United States

13.1 Introduction Conventional agriculturea is a major driver of climate change (Foley et  al., 2005). Its intensive use of natural resources, synthetic fertilizers, and pesticides further degrades air, soil, and water quality and causes large-scale biodiversity loss (Foley et al., 2005; Horrigan, Lawrence, & Walker, 2002; Kremen & Merenlender, 2018; Pretty et al., 2000; Sánchez-Bayo & Wyckhuys, 2019; Tilman, 1998). Such assessments have driven wide agreement that conventional agriculture must become more sustainable. A transition to sustainable agriculture is also an essential component of sustainable food systems. However, there are very different views on how to improve the sustainability of agriculture while meeting food security goals (Godfray & Garnett, 2014; HoltGiménez & Altieri, 2013; Kremen & Merenlender, 2018; McMichael & Schneider, 2011; Mercer, Perales, & Wainwright, 2012; Perfecto & Vandermeer, 2010; Zaidi et al., 2019). The sustainability impact of biotech crops remains a key area of controversy. For good or ill, this impact is likely substantial. According to the most commonly cited source, in 2017 genetically modified (GM) crops were planted on 189.8 million ha in 24 countries (ISAAA, 2017b). The majority of GM crops grown commercially are either herbicide tolerant (HT crops) or they produce GM pesticides originating from the bacterium Bacillus thuringiensis, an insect pathogen (Bt crops). The main HT and/ or Bt commodity crops grown globally are soybean (94.1 million ha), maize (59.7 million ha), cotton (24.1 million ha), and canola (10.2 million ha) (ISAAA, 2017). The narrow view of sustainability sees the role of agriculture as providing food security. It further equates food security with high yields (Latham, this volume). To obtain its yields, conventional agriculture combines large-scale monoculture cropping systems with the extensive use of off-farm inputs (e.g., hybrid seeds, agrochemicals, water, and fuel). Labor is replaced with mechanization. Soil fertility and pests are managed with synthetic fertilizers, insecticides, fungicides, and herbicides. a

Conventional (also called industrial) agriculture includes input-heavy large-scale monoculture commodity cropping systems and confined animal feeding operations. b ISAAA (International Service for the Acquisition of Agri-biotech Applications) is funded by both government organizations and the biotech industry to promote the uptake of agricultural biotechnology products, including GM crops. Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00013-X © 2021 Elsevier Inc. All rights reserved.

248

Rethinking Food and Agriculture

Hybrid seeds must be purchased each year. Consequently, increased sustainability is framed as increased monocrop yield/acre and/or a concomitant decrease in the use offfarm inputs. Biotechnology, specifically the use of GM crops (also called genetically engineered or bioengineered, e.g., transgenic, cisgenic, RNAi, or gene-edited crops), is often seen as necessary to achieve these sustainability goals (Ammann, 2005; Dibden, Gibbs, & Cocklin, 2013; Godfray & Garnett, 2014; Pretty, 2001). In this paradigm it is argued that, by increasing yields, GM crops can also benefit smallholder farming and sustainable agricultural systems (Holt-Giménez & Altieri, 2013; Mercer et al., 2012; Shelton, Hossain, Paranjape, & Azad, 2018). The broad view of sustainability, in contrast, sees agriculture as intrinsically multifunctional, having diverse and interconnected environmental and social impacts (McIntyre et al., 2009). To be sustainable therefore agriculture must (1) provide food security and healthy diets and (2) support rural livelihoods and culture, while also reducing poverty and inequality and (3) increase biodiversity and environmental health, placing a strong emphasis on lessening climate change (Horrigan et al., 2002; Kremen & Merenlender, 2018; McMichael & Schneider, 2011). Within this paradigm, technologies and practices that support low-input small-scale agriculture, food and seed sovereignty, and local food systems are considered essential (Adhikari, 2014; Holt-Giménez & Altieri, 2013; Kremen & Merenlender, 2018; McIntyre et al., 2009). However, there are many different possible sustainable farming systems. These can include organic, agroecological, agroforestry, and traditional systems. What they have in common is the use of biodiversity to achieve multiple agricultural, environmental, and societal goals (Altieri, 1999; Thrupp, 2000; Wickson, Binimelis, & Herrero, 2016). For example, both soil health and pests are managed by increasing on-farm biodiversity via practices that include multispecies crop rotation, cover cropping, and intercropping, as well as push/pull, SRIc and no-till techniques, and the incorporation of livestock and trees (Anderson, 2015; Hailu, Niassy, Zeyaur, Ochatum, & Subramanian, 2018; Kremen & Miles, 2012; Midega, Pittchar, Pickett, Hailu, & Khan, 2018; Pretty, 2001; Thakur, Uphoff, & Stoop, 2016; Zhang, Postma, York, & Lynch, 2014). Such techniques can drastically reduce the need for off-farm inputs, with the goal of eliminating them entirely (Nicholls, Altieri, & Vazquez, 2016). Within this broad agroecological paradigm of sustainability, GM crops are often seen as incompatible (Adhikari, 2014; Altieri, 2005; Barker, 2014; Fischer, 2016; Garibaldi et al., 2017; Kesavan & Swaminathan, 2018; Kremen & Miles, 2012; McIntyre et al., 2009; Pengue, 2005; Schütte et al., 2017; Wickson et al., 2016). The availability of a wide diversity of crops and cultivars is an integral component of sustainable agriculture. Historically, farmer seed saving and selection created an enormous diversity of crop varieties with widely varying properties and adaptations (Villa, Maxted, Scholten, & Ford-Lloyd, 2005). The intention of plant breeding is to contribute to this diversity by producing crops with beneficial new trait combinations or characteristics. Conventional plant breeding achieves this via genetic crossing of sexually compatible plants, and the selection of offspring with the desired traits. c

SRI stands for system of rice intensification, a set of practices that can reduce the use of water, seeds, and other inputs, while increasing soil biodiversity and improving yields.

Will gene-edited and other GM crops fail sustainable food systems?249

Occasionally, conventional plant breeders use intentional mutagenesis, somaclonal mutation, or wide crosses to introduce novel traits (Wilson, Latham, & Steinbrecher, 2004). GM crop developers, in contrast, use a combination of lab-based techniques (e.g., recombinant DNA [rDNA] technology, tissue culture, and plant transformation [e.g., Agrobacterium infection or particle bombardment]) to introduce specific rDNA sequences, specifying novel traits, into crop plants (Barampuram & Zhang, 2011). GM techniques expand the range of traits available by conferring the ability to combine, alter, and transfer DNA from any organism (e.g., viruses, bacteria, fungi, mammals, nonfood plants) into a crop’s genome (Wickson et al., 2016). A suite of new GM techniques (nGMs), including a variety of “gene editing” systems, has been developed for use in plants (Eckerstorfer, Heissenberger, Reichenbecher, Steinbrecher, & Waßmann, 2019). In contrast to standard GM techniques, gene editing techniques can target DNA integration and/or other modifications to specific regions of the genome (Fichtner, Castellanos, & Ülker, 2014). The claimed benefits of GM and nGM technologies are their ability (1) to transfer or alter specific sequences of DNA and (2) to introduce novel DNA modifications and traits that cannot be introduced via conventional plant breeding.

13.2 Impacts of HT and Bt crops In theory, HTd crops and Bte crops were intended to promote sustainable agriculture by (1) reducing overall pesticide use and (2) substituting safer pesticides (e.g., plant-­ produced Bt toxins or glyphosate herbicides) for more harmful ones (Ammann, 2005; Andow, 2010; Koch et al., 2015). HT crops have also been claimed to improve sustainability by facilitating the uptake of no-till agriculture. In practice, however, the widespread use of Bt and HT crops has led to the problematic development of pest resistance, “superweeds,” and secondary pests (Benbrook, 2018; Bonny, 2016; Carrière et al., 2016; García et al., 2019; Gould, Brown, & Kuzma, 2018; Kilman, 2010; Kranthi, 2016; Mortensen, Egan, Maxwell, Ryan, & Smith, 2012; Stone & Flachs, 2018; Tabashnik & Carrière, 2017). In response to these problems, farmers increased both insecticide and herbicide use. Some also increased tillage and other mechanical methods of weed control (Bonny, 2016; Green, 2014). The seed industry response has been to add multiple Bt pesticide and/or HT traits (stacked and d

HT GM crops have one or more transgene(s) that specify a protein that either resists or detoxifies a specific herbicide. For example, the herbicide glyphosate targets an essential plant enzyme, 5-enolpyruvoylshikimate-3-phosphate synthase (epsps), to cause lethality. Roundup Ready (RR) glyphosate-tolerant GM crops have a transgene that specifies a glyphosate tolerant version of the epsps enzyme. e Bt crops have one or more recombinant cry transgenes (e.g., cry1Ab, cry3A) inserted into their genome, each transgene specifying a different GM Bt toxin (Latham, Love, & Hilbeck, 2017). Plant-produced Bt toxins are derived from the natural toxins produced by the insect gut pathogen Bacillus thuringiensis (Sanchis, 2011). Different classes of Bt toxins are considered lethal to different orders of insects, and are used to target different plant pests (Carrière, Fabrick, & Tabashnik, 2016). For example, B. thuringiensis Cry1 toxins are thought to specifically target Lepidoptera (e.g., corn borers and cotton bollworms), while Cry3 toxins are believed to specifically target Coleoptera (e.g., corn rootworm, Colorado potato beetle).

250

Rethinking Food and Agriculture

pyramided traitsf) to each variety, and to develop new plant-produced pesticides (e.g., VIP protein toxins and RNAi-based insecticides) (Bøhn & Lövei, 2017; Carrière et al., 2016; Chakroun, Banyuls, Bel, Escriche, & Ferré, 2016; Gould et al., 2018). In addition, the introduction of Bt and HT crops and their attendant pesticides has encouraged a variety of changes to farmer practice that themselves have had highly detrimental environmental impacts. First, the adoption of Bt and HT crops has undermined the use of integrated pest management (Gray, 2010) and sustainable techniques. They do so by substituting GM crop-produced and chemical pesticides for pesticide-free control measures that include tillage, short season crops, cover crops, crop rotation, and biological controls (Brainard, Haramoto, Williams, & Mirsky, 2013; Gutierrez, Ponti, Herren, Baumgärtner, & Kenmore, 2015; Kesavan & Malarvannan, 2010; LaCanne & Lundgren, 2018; Lang, Oehen, Ross, Bieri, & Steinbrich, 2015; Schütte et al., 2017; Stone & Flachs, 2018; Tooker, 2015). Second, growing Bt crops (a decision made before actual insect pressures are known) exposes the landscape, and consumers, to pesticide whether or not the targeted pest is a threat, and whether or not Bt toxins provide effective protection. Wide uptake of Bt crops is thus prophylactic pesticide use (Douglas & Tooker, 2015; Gray, 2010; Stone & Flachs, 2018; Tooker, 2015). Finally, HT crops have ­decreased biodiversity by encouraging simplified crop rotations and/or farming systems. They have also permitted unrestricted spraying of broad-spectrum herbicides throughout the growing season, further exacerbating biodiversity losses (Schreiner, 2009; Schütte et al., 2017). As another consequence, HT soybeans on the US market have a high level of glyphosate contamination (Bøhn et al., 2014). Thus Bt and HT traits have exacerbated and expanded the pesticide treadmill (Altieri, 2000; Binimelis, Pengue, & Monterroso, 2009; Douglas & Tooker, 2015; Mortensen et  al., 2012; Pengue, 2005; Stone & Flachs, 2018). The resulting “technology-­ facilitated pesticide treadmill” is described by Douglas and Tooker (2015): Neonicotinoid seed treatments may also have “tagged along” with other technologies that were attractive to farmers. They are usually one component of larger packages, that, for instance in maize, can include germplasm (i.e., crop variety), up to eight transgenes, and up to six or more different seed treatments (fungicides, nematicides, and insecticides).

These well-documented outcomes indicate the adoption of HT and Bt crops is leading to dramatic increases in pesticide use over time, including the use of pesticides known to be extremely toxic such as neonicotinoids, glufosinate, 2,4-d, and dicamba (Douglas & Tooker, 2015; Mortensen et al., 2012; Schütte et al., 2017; Tooker, 2015). Increased herbicide use with HT crops has been repeatedly demonstrated in the scientific literature (Schütte et  al., 2017). However, some authors claim that Bt crops can reduce pesticide use (e.g., Klümper & Qaim, 2014; Naranjo, 2009). Short-term f

Stacked resistance traits are defined as multiple transgenes specifying tolerance to more than one herbicide (e.g., glyphosate, dicamba, and 2,4-d) and/or to more than one target pest (e.g., corn root worm and corn borer). Pyramided traits have more than one transgene targeting the same pest (e.g., Cry1Ab + Vip3A to target Lepidoptera).

Will gene-edited and other GM crops fail sustainable food systems?251

s­ tudies (e.g., before resistance develops), or failure to quantify the amount of plant-­ produced Bt toxin(s) and/or seed coat insecticides, can account for these discrepancies (Benbrook, 2012; Douglas, 2016). When applied pesticides, GM crop-produced insecticides (e.g., Benbrook, 2012; Clark, Phillips, & Coats, 2005; Nguyen & Jehle, 2007; Saxena, Stewart, Altosaar, Shu, & Stotzky, 2004; U.S. Environmental Protection Agency, 2010; van der Hoeven, 2014), and seed coat pesticides are taken fully into account, both Bt and HT crops increase pesticide use in farming systems (e.g., Benbrook, 2012; Bøhn & Lövei, 2017; Bonny, 2016; Capellesso, Cazella, Schmitt Filho, Farley, & Martins, 2016; Douglas & Tooker, 2015; Heinemann, Massaro, Coray, AgapitoTenfen, & Wen, 2014; Kranthi, 2016; Perry, Ciliberto, Hennessy, & Moschini, 2016; Yang, Iles, Yan, & Jolliffe, 2005).

13.2.1 Toxicity of GM crop-associated pesticides Developers and US regulators of GM crops claim that Bt toxins and glyphosate are low-toxicity pesticides (Koch et  al., 2015; Williams, Kroes, & Munro, 2000). However, for Bt crops, there is an ever-growing body of evidence showing Bt toxins and Bt plants have harmful off-target effects, including toward mammals, beneficial insects, and aquatic invertebrates (Andreassen et al., 2015; Hilbeck & Schmidt, 2006; Latham et al., 2017; Paula et al., 2014; Venter & Bøhn, 2016). Many researchers have pointed out the need for further biosafety research, in particular in planta studies and research on the sublethal and long-term effects of exposure to Bt crops (e.g., Andow, 2010; Arpaia et al., 2017; Hilbeck & Otto, 2015; Latham et al., 2017; Sanchis, 2011; Wolfenbarger, Naranjo, Lundgren, Bitzer, & Watrud, 2008). Similar concerns apply to glyphosate-based herbicides. For example, glyphosate and/or its formulations affect the composition of soil and gut microbiota and have negative effects on earthworms, beneficial insects, and aquatic organisms (Schütte et al., 2017; Sharma, Jha, & Reddy, 2018). They are also linked to cancer and chronic kidney disease in humans (e.g., Jayasumana, Gunatilake, & Senanayake, 2014; McHenry, 2018; Myers et al., 2016). Due to the large body of evidence documenting their harmful off-target impacts, coupled with significant research gaps, there is no scientific consensus that Bt toxins and glyphosate-based herbicides are low-toxicity pesticides (Ardekani & Shirzad, 2019; Hilbeck et al., 2015; Krimsky, 2015).

13.3 Unintended traits in GM crops Regardless of the intended trait, GM technology is frequently acclaimed for its precision. In particular, the ability to introduce novel traits without the problem of “yield drag,” a problem that can complicate conventional plant breeding (Gepts, 2002). Yet, despite these claims, reports of unexpected and harmful unintended traits (UTs) in GM crops periodically surface in the media. For example, in 2012 various news outlets, including the Wall Street Journal, reported that the stalks of GM corn and soy were much tougher than those of conventional crops (Tita, 2012). The tougher GM stalks puncture tractor tires. This unexpected trait has both economic and environmental

252

Rethinking Food and Agriculture

costs, as farmers are forced to buy new and expensive reinforced tractor tires and/ or replace tires more frequently. Like most such reports, the “tough stubble” trait has not been fully followed up or acknowledged in the scientific literature. Yet, it raises a host of important and neglected questions: (1) How precise and predictable is GM technology? (2) Do developers and regulators prevent GM crops with harmful UTs from reaching the market? (3) Are UTs an underappreciated barrier to sustainability? A UT, sometimes called an unintended effect, is defined here as any significant difference, other than the intended GM trait, between a GM crop compared to a non-GM isogenic line. UTs thus include, for example, statistically significant differences in characters such as seed germination, weed suppression, pest resistance, drought tolerance, height, yield, and flowering time. UTs further include compositional differences in nutrients, toxins, and other biochemicals. Such UTs are often revealed by transcriptomic, proteomic, and metabolomic profiling studies (Cellini et  al., 2004). Several reviewers have collected examples of a wide variety of UTs recorded in the scientific literature (Cellini et al., 2004; Haslberger, 2003; Kuiper, Kok, & Engel, 2003; Nature Institute, 2019; Ricroch, Bergé, & Kuntz, 2011). Nevertheless, the number of documented examples is much greater than those already collected, precluding a comprehensive review.

13.3.1 Precision and predictability If GM technology was precise and predictable, biotechnologists would only need to create one GM plant, and this would be identical to the parent plant except for the intended new trait. However, biotechnologists instead produce many hundreds or even thousands of initial transformants.g For example, to identify a Roundup Ready (RR) wheat event for commercialization, Hu et  al. (2003) used either Agrobacterium or the gene gun to introduce rDNA into over 98,000 plant tissue fragments. At the same time they tested a number of different transgenes, since it was not clear which would be the most effective in wheat. From their initial populations they selected over 1300 glyphosate-tolerant plants for further development, discarding the rest. Subsequent rounds of selection assessed Roundup resistance and basic agronomic performance. After four generations of such selection, six suitable events remained. Finally, after 3 years of “large-scale field trials,” one event was selected for commercialization and submitted to the US Department of Agriculture (USDA) for approval. The petition for deregulation of this event was subsequently withdrawn. Research on large populations of initial transformants, created for the development of Bt rice (Shu et al., 2002), Bt, or blight-resistant potato (Davidson et al., 2004; Felcher, Douches, Kirk, Hammerschmidt, & Li, 2003), virus-resistant tobacco (Xu, Collins, Hunt, & Nielsen, 1999), and virus-resistant barley (Bregitzer, Halbert, & Lemaux, 1998), suggest similar problems for other GM crops and traits. Defects in basic agronomic traits such as yield, height, stem, and leaf morphology are frequent in regenerated GM plants, and many initial transformants exhibit multiple UTs. Even when only a few (from 2 to 22) traits are assessed, the proportion of initial t­ransformants g

A transformant is a cell or an organism, such as a plant, into which foreign DNA has been introduced.

Will gene-edited and other GM crops fail sustainable food systems?253

with UTs usually ranges between 20% and almost 100% (e.g., Bregitzer et al., 1998; Dale & McPartlan, 1992; Davidson et  al., 2004; Felcher et  al., 2003; Hoekema, Huisman, Molendijk, van den Elzen, & Cornelissen, 1989; Kumar, Rakow, & Downey, 1998; Shu et al., 2002; Vickers, Grof, Bonnett, Jackson, & Morgan, 2005). Later in GM crop development, diverse UTs are still frequently identified, even after multiple rounds of selection. To take rice as a sample crop, decreased yield, seed size, or vigor have all been reported for different Bt rice lines (e.g., Bashir et al., 2004; Chen, Snow, Wang, & Lu, 2006; Jiang et al., 2018; Shu et al., 2002; Tu et al., 2000; Wang et  al., 2012; Wei-xiang, Qing-fu, Hang, Xue-jun, & Wen-ming, 2004; Wu, Shu, Wang, Cui, & Xia, 2002), as have alterations to grain and straw quality (e.g., Bashir et al., 2004; Li et al., 2008; Wei-xiang et al., 2004; Wu et al., 2002). As well, height, yield, and developmental UTs have been documented for glufosinate-tolerant rice lines (Oard et al., 1996). Additional examples of UTs are collected in Tables 13.1 and 13.2 of this review. These, combined with examples collected by other reviewers, confirm that UTs are not limited to any particular GM technique, trait, or plant species (Cellini et al., 2004; Haslberger, 2003; Kuiper et al., 2003; Nature Institute, 2019; Ricroch et al., 2011).

13.3.2 UTs in commercial GM crops Commercial GM crops are considered “the best of the best” that GM plant breeding can offer. They undergo years of selection and development prior to “rigorous compositional, nutritional, and safety evaluations” and, in some cases, environmental risk assessments (Larkin & Harrigan, 2007). Finally, most GM crops currently grown have undergone some form of regulatory process before their commercial release or import was permitted (Davison, 2010; Pelletier, 2005). Nevertheless, Table 13.1 provides examples of some of the many UTs that have been identified in commercial GM crops. These UTs are documented in peer-reviewed papers and/or the petition submitted to USDA/APHIS (U.S. Department of Agriculture/Animal and Plant Health Inspection Service) to deregulate a particular crop or event in the United States. The examples given in Table 13.1 have been selected because they have implications for sustainability. Important conclusions can be drawn from Table 13.1. First, it is not difficult to find examples of commercial crops with UTs. In fact, many commercial GM crops have multiple UTs. UTs documented for Mon810 maize, for example, include numerous compositional differences (including increased lignin and the presence of an allergen), increased moisture content, and negative impacts on beneficial soil organisms. Second, UTs have different origins. Some, such as the increased lignin levels associated with cry1Ab Bt maize, are likely pleiotropic effects of the transgene (i.e., the trait itself), as they are seen with many independent cry1Ab events. Others, such as the loss of resistance to root knot nematode in Bt Paymaster cotton, are event specific and likely due to mutational damage at or near the transgene insertion site (Colyer et al., 2000). Third, many UTs commonly documented in commercial GM crops have had unequivocal negative impacts on sustainability. For example, high moisture content, as documented for MON810 and Bt11 maize, requires added energy during drying

Table 13.1  Examples of unintended traits in commercial HT, Bt, or virus-resistant GM crops. GM crop Bt maize

Transgene event (transgene namea,b) Event Mon810 (cry1Ab)

Intended trait Lepidopteran insect resistant (pesticidal)

Unintended traits compared to non-GM parent or isogenic line (references) ●













Bt maize

Event Bt11 (cry1Ab)

Lepidopteran insect resistant (pesticidal)









Bt maize

Event 176 (cry1Ab)

Lepidopteran insect resistant (pesticidal)







Increased lignin in stem (Flores, Saxena, & Stotzky, 2005; Poerschmann, Gathmann, Augustin, Langer, & Górecki, 2005; Saxena & Stotzky, 2001) Altered sugar content, osmolytes, branched amino acids, and proteins (including truncated proteins and the presence of an allergen) in kernels (Barros et al., 2010; Manetti et al., 2006; Zolla, Rinalducci, Antonioli, & Righetti, 2008) Decreased protozoan and nematode numbers, and drier rhizosphere soils (Griffiths et al., 2005; Höss et al., 2008) Increased aphid susceptibility (Faria, Wäckers, Pritchard, Barrett, & Turlings, 2007) and thrip numbers (Bourguet et al., 2002) Delay in seed and plant maturation (La Paz, Pla, Centeno, Vicient, & Puigdomènech, 2014) Higher moisture content in whole plant and grain at harvest (Ma & Subedi, 2005) Delayed decomposition of Mon810 maize plant residues (Flores et al., 2005; Stotzky, 2004) Increased lignin in stem (Flores et al., 2005; Saxena & Stotzky, 2001) Increased aphid susceptibility (Faria et al., 2007) Higher moisture content in the whole plant and grain at harvest (Ma & Subedi, 2005) Detrimental impacts on corn root colonization by beneficial mycorrhizal fungi (Castaldini et al., 2005) Increased lignin in stem (Poerschmann et al., 2005; Saxena & Stotzky, 2001) Increased aphid susceptibility (Faria et al., 2007) Detrimental impacts on corn root colonization by beneficial mycorrhizal fungi (Castaldini et al., 2005)

Roundup Ready maize

Event NK603 (cp4 epsps)

Resistant to glyphosate herbicide











Roundup Ready (RR1) soybean

Event 40–3-2 (cp4 epsps)

Genuity Roundup Ready 2 Yield (RR2Y) soybean Virusresistant squash

Event MON89788-1 (cp4 epsps)

Resistant to glyphosate herbicide

Event CZW-3 (three transgenes specifying coat proteins from CMV, ZYMV, and WMV2) Atlantic NewLeaf Clone 6 (cry3A)

Resistant to three viruses affecting cucurbits

Bt potato Atlantic NewLeaf

Resistant to glyphosate herbicide















Coleopteran insect resistant (pesticidal)



Decreased Υ-tocopherol and inositol; increases in potentially toxic polyamines, e.g., N-acetylcadaverine (2.9-fold), N-acetylputrescine (1.8-fold), putrescine (2.7-fold), and cadaverine (28fold) in kernel (Barros et al., 2010; Mesnage et al., 2016) Evidence of kidney and liver toxicity (Fagan, Traavik, & Bøhn, 2015) Unintended proteomic and plant defense-related phytohormone differences in leaves (Benevenuto et al., 2017) Statistically significant differences for ear height and days to 50% silking (USDA/APHIS petition 00-011-01p) Average maize yield for epsps glyphosate-tolerant GM maize varieties (such as NK603) decreased by 5.98 bushels per acre compared to conventional varieties (Shi, Chavas, & Lauer, 2013) Yield decrease of 7%–11% (Gordon, 2007; Nelson, Renner, & Hammerschmidt, 2002; Elmore et al., 2001; Benbrook, 1999; USDA/APHIS Petition P93–258-01) Note: grown extensively in major soy producing countries such as the United States, Brazil, and Argentina (e.g., over 93% of the US soybean crop is RR HT soy) (Oliveira & Hecht, 2016) 5% decrease in height (USDA/APHIS Petition 06-SB-167U; Horak et al., 2015) RR1 and RR2Y have similar yields (Mason, Walters, Galusha, Wilson, & Kmail, 2017)

Beta-carotene decreased to 1.5%, iron decreased to 87%, and fat to 50%, of control levels (Table V, USDA/APHIS Petition no. 95-352-01p) Vitamin A levels increased twofold and sodium increased fourfold, compared to control lines (Table V, USDA/APHIS Petition no. 95-352-01p) Other unintended traits include differences in floral traits and altered bee visits compared to an isogenic line (Prendeville & Pilson, 2009) Loss of resistance to golden nematode, a key trait present in its non-GM parent plant (Brodie, 2003; Brodie & Mai, 1989) Continued

Table 13.1 Continued GM crop

Transgene event (transgene namea,b)

Bt cotton Paymaster

Event 15560BG (cry1Ac)

Bollgard Bt cotton, INGARD Chinese Bt cotton

Event: N/A (cry1Ac)

Bt cotton

Event GK97 (N/A)

Roundup Ready oilseed rape

Brassica napus cv. Quest Event: N/A (cp4 epsps) Events: N/A Two cultivars: Falcon pat (pat) and cultivar ArtusLL (N/A)

HT winter rape

Event: N/A (cry1Ac)

Intended trait Lepidopteran insect resistant (pesticidal) Lepidopteran insect resistant (pesticidal) Lepidopteran insect resistant (pesticidal) Lepidopteran insect resistant (pesticidal) Resistant to glyphosate herbicide Resistant to glufosinate herbicide

Unintended traits compared to non-GM parent or isogenic line (references) ●













Loss of resistance to root knot nematode (Colyer, Kirkpatrick, Caldwell, & Vernon, 2000)

Several varieties showed increased susceptibility to Fusarium fungal disease (Kochman et  al., 2000) Two varieties had decreased resistance to Fusarium oxysporum fungal disease compared to ­controls (Li et al., 2009) Altered composition of root exudates (Li et al., 2009) Quantitative and qualitative differences in volatiles (Yan et al., 2004)

Altered diversity of root-endophytic bacteria (Siciliano & Germida, 1999; Siciliano, Theoret, De Freitas, Hucl, & Germida, 1998)

Altered flowering in both cultivars (Pierre et al., 2003)

The UTs listed have implications for sustainability, for example, via decreased yields or potential impacts on ecological interactions. Bt, Bacillus thuringiensis; CMV, cucumber mosaic cucumovirus; epsps, 5-enolpyruvoylshikimate-3-phosphate synthase; GM, genetically modified; HT, herbicide tolerant; WMV, watermelon mosaic ­potyvirus 2; ZYMV, zucchini yellow mosaic potyvirus. a “N/A” indicates no information was provided in the reference. b Many transgenic events also include additional marker genes, which are usually recombinant antibiotic or herbicide resistance genes. Associated marker genes are not noted in most references and have not been included in this table.

Will gene-edited and other GM crops fail sustainable food systems?257

Table 13.2  Examples of unintended traits (UTs) identified in crops having “complex” genetically modified traits intended to improve their agronomic performance, resistance to abiotic stress, or nutritional value. Crop

Intended trait

Potato Potato

Increased tuber dormancy Reduced browning

Potato

High amylose

Cotton

Increased salt tolerance Increased salt tolerance

Buckwheat

Rice and barley

Increased salt tolerance

Tomato

Increased provitamin A levels Novel flavonoids

Tomato Maize Rice

Increased lysine content Increased tryptophan levels

Unintended traits compared to non-GM parent or isogenic line Lower tuber yield and/or fewer tubers per plant (Marmiroli et al., 2000) Unintended alterations to glycoalkaloid and sesquiterpene levels (Matthews, Jones, Gans, Coates, & Smith, 2005) Large unintended increase in phosphorus; altered sucrose levels, yield, and growth (Hofvander, Andersson, Larsson, & Larsson, 2004) Decreased seed yield under normal growth conditions (Zhang et al., 2009) Unintended agronomic and compositional changes when grown under greenhouse conditions (Chen, Zhang, & Xu, 2008) Numerous unexpected metabolic changes in both rice and barley (Jacobs, Lunde, Bacic, Tester, & Roessner, 2007) Reduced lycopene (Römer et al., 2000) Decreased seed set, color changes, parthenocarpy, and enzymatic differences (Schijlen et al., 2006) Agronomic and metabolic UTs (Bicar et al., 2008) Agronomic and metabolic UTs (Wakasa et al., 2006)

(Ma & Subedi, 2005). UTs such as a loss of pest resistance or decreased yield potential undermine sustainability, particularly when farmers use more external inputs (e.g., fuel, pesticides, fertilizer, water) to maintain yields. Table  13.1 includes the large yield decreases documented for RR soybeans (7%–11% decrease), glyphosate HT maize (5.98 bushels/acre decrease), and corn root worm-protected Bt maize (12.22 bushels/acre decrease) (Gordon, 2007; Shi et al., 2013). Loss of pest or pathogen resistance documented in Table  13.1, includes increased aphid and/or thrip numbers on Bt maize varieties, loss of nematode resistance in Atlantic NewLeaf Bt potato and Paymaster Bt cotton, and decreased resistance to Fusarium fungal disease observed with various Bt cotton varieties (Bourguet et  al., 2002; Brodie, 2003; Colyer et  al., 2000; Faria et  al., 2007; Li et  al., 2009). Table  13.1 thus illustrates that commercial GM commodity crops, many grown on millions of acres worldwide, frequently have detrimental UTs, in addition to the "tough stalks" of HT and Bt crops, UTs that contribute to their large and negative environmental impacts.

258

Rethinking Food and Agriculture

Nevertheless, many biotechnologists tend to dismiss UTs as unimportant, even when they are identified in commercial varieties (e.g., Fox, Morrison-Saunders, & Katscherian, 2006; Larkin & Harrigan, 2007; Ricroch et al., 2011; Shepherd, McNicol, Razzo, Taylor, & Davies, 2006; Sidhu et al., 2000). This was the case with the yield loss UT documented for RR soybean. In their regulatory petition to the USDA, the developers of RR soybean claimed they would improve yields through further breeding, yet they were never able to do so (USDA/APHIS, Petition P93–258-01, 2020). Likewise, large and statistically significant compositional UTs in CZW3 squash were also dismissed, rather than interpreted as a red flag indicating a need for further risk assessment, or for rejection of that particular event or crop (USDA/APHIS Petition 95–352-01p). Yet, small and large compositional UTs, such as quantitative or qualitative alterations to metabolites, nutrients, or potential toxins, can negatively impact ecological interactions (Arpaia et  al., 2017; Li et  al., 2009; Mesnage et  al., 2016; Venter & Bøhn, 2016) and/or food or feed safety (Haslberger, 2003; Pelletier, 2005; Schubert, 2008), traits essential to sustainability. Statistically significant differences seen under one environment or in one field trial, but not others, are also routinely dismissed. However, transgenes may be prone to large transgene × environment interactions (Zeller, Kalinina, Brunner, Keller, & Schmid, 2010). Therefore UTs seen only under specific conditions should also be considered as a starting point for more risk assessment, as they could indicate a defect in environmental response or other significant problems (Agapito-Tenfen, Guerra, Wikmark, & Nodari, 2013). It is important to emphasize the examples in Table 13.1 are not exhaustive of UTs that have been documented in commercial GM crops. A number of factors further combine to make it likely that the publicly available data seriously underestimate the number of commercial lines with UTs and the number of UTs in each commercial line. Many important sustainability traits, such as increased outcrossing, seed dormancy or seedbank persistence, might never be assessed, despite their potential for negative impacts (e.g., Altieri, 2005; Bergelson, Purrington, & Wichmann, 1998; Linder, 1998; Linder & Schmitt, 1995). Other limitations for commercial GM crops include the lack of standardization for compositional studies, a lack of -omic analyses, the use of inadequate or inappropriate test conditions for many lab and field trials (e.g., inappropriate organisms or life stage tested for toxicity, inadequate distances between field plots) and a lack of studies on the long-term and sublethal effects of GM crops and products (Arpaia et al., 2017; Booij, 2014; Hilbeck, Meier, & Trtikova, 2012; Pelletier, 2005; Schubert, 2008). In addition, the proprietary nature of GM crops acts to restrict independent research (Waltz, 2009). Universities and funders also do not encourage research that could find harm from GM crops, and findings of harm are often heavily contested or even suppressed (Fagan et  al., 2015; Peekhaus, 2010; Waltz, 2009). Consequently, most commercialized GM crops have undergone little or no independent testing or risk assessment that could identify UTs (Diels, Cunha, Manaia, Sabugosa-Madeira, & Silva, 2011; Séralini, Mesnage, Defarge, & de Vendômois, 2014). In summary, UTs are frequent, if not ubiquitous, in all crops developed using standard GM techniques, including commercial GM varieties. These UTs frequently have a negative impact on sustainability, making GM crops an inappropriate choice for sustainable agriculture and food systems (Kesavan & Swaminathan, 2018).

Will gene-edited and other GM crops fail sustainable food systems?259

13.3.3 Standard GM techniques contribute to UTs UTs can arise from unintended effects of the transgene (or accompanying selectable marker genes). However, the techniques used to produce a GM crop can also give rise to UTs (Wilson, Latham, & Steinbrecher, 2006). Transgenes are introduced into a plant cell, usually via infection with Agrobacterium, or else a “gene gun” is used to bombard plant cells with DNA-coated particles. The transgene subsequently integrates into damaged regions of the plant genome, via the plant’s natural DNA repair mechanisms. The genomic location of transgene integration is therefore uncontrolled, and differs for each independent integration event. Modified plant cells are then regenerated back into whole plants via tissue culture. Regenerated plants with one or more transgenic events are selected (often with the aid of a cotransferred marker gene that specifies antibiotic or herbicide resistance) for further analysis. Transgene insertion thus inevitably disrupts the endogenous plant genome. Furthermore, Agrobacterium infection, particle bombardment, and tissue culture have all been shown to be highly mutagenic. Together they can produce many thousands of mutations. These mutations can be at, or linked to, the site of transgene integration, and also spread throughout the genome (Wilson et al., 2006). Such mutations include base pair changes, large and small DNA insertions and deletions, large-scale genome rearrangements, as well as unintended integration of bacterial chromosomal DNA, vector DNA, multiple transgenes, and transgene fragments. Thus, the mutagenicity of GM techniques contributes to the frequency and variety of UTs documented in GM crops (Wilson et al., 2006). In some cases UTs can be removed via genetic recombination (outcrossing or backcrossing). However, UTs genetically linked to the transgene insertion site will be difficult if not impossible to separate from the desired trait. UTs are an even greater problem for commercial crops that are propagated clonally or are otherwise difficult or impossible to cross. These include potato, banana, cassava, and most tree crops.

13.4 New GM traits and techniques The UTs described in previous sections were identified in GM plants engineered for a very limited number of traits: pest resistance, virus resistance, and/or herbicide tolerance. These are simple traits, specified by single transgenes whose novel products were not intended to alter normal plant functions, structures, or biochemical pathways. For complex traits with the potential to benefit sustainable agriculture, such as increased tolerance to drought, salt, heat, or flood, intentionally altered levels of specific nutrients, or increased yield, UTs are likely to be an even greater obstacle (e.g., Flowers & Yeo, 1995; Kollist et al., 2019). Table 13.2 lists some of the many documented examples of GM crops with such complex traits that also exhibit UTs.

13.4.1 Complex GM traits have a history of failure Despite receiving frequent and positive media attention, most attempts to develop commercial GM crops with complex GM traits have failed. At best they have lagged

260

Rethinking Food and Agriculture

far behind conventional plant breeding in producing viable products (e.g., Barker, 2014; Gilbert, 2016; McFadden, Smith, Wechsler, & Wallander, 2019; Stone & Glover, 2011, 2017). It is only recently that an extremely limited number of GM crops with complex traits have become commercially available (e.g., McFadden et al., 2019; Waltz, 2015). However, despite the high risk of UTs, these new GM crops have undergone even less independent research, risk assessment, and regulatory scrutiny than previous commercialized GM crops (Camacho, Van Deynze, Chi-Ham, & Bennett, 2014; Waltz, 2016, 2018). In fact, for “drought-tolerant” soybean HB4, the developers themselves admit they do not understand the mechanism of action behind their trait (Waltz, 2015).

13.4.2 Golden Rice case study Golden Rice is a widely cited example of a nutritionally enhanced GM crop (Bollinedi et al., 2014). Unusually, a significant amount of research on Golden Rice has been published in the scientific literature. Golden Rice is thus useful as a case study of a GM trait that could theoretically benefit sustainable agriculture. Golden Rice varieties contain two transgenes specifying enzymes in the β-carotene biosynthesis pathway (Bollinedi et al., 2014; Ye et al., 2000). In theory, the targeted production of these enzymes in the rice endosperm will increase grain levels of β-carotene (provitamin A) (Dubock, 2014). Since the first Golden Rice paper was published in 2000, public and private sector researchers have produced many iterations of Golden Rice, each one intended to further increase the levels of β-carotene in the rice grain (Bollinedi et al., 2014). Syngenta donated six of its GR2 events for public sector use (Bollinedi et al., 2014). The International Rice Research Institute (IRRI) has used these in breeding efforts targeted to countries deemed to have populations with high levels of vitamin A deficiency (Bollinedi et al., 2014; IRRI, 2019; Stone & Glover, 2017). Two Golden Rice events, GR2-R1 and GR2E, have been the subjects of the most research and development (e.g., Bollinedi et al., 2017, 2019; Paine et al., 2005; Schaub et al., 2017).

13.4.2.1 The GR2-R1 event causes agronomic defects, including dramatic yield loss For many years, the GR2-R1 event was the focus of Golden Rice breeding efforts (Bollinedi et al., 2017; Stone & Glover, 2017). GR2-R1 lines, however, gave consistently low yields (Dubock, 2014; Stone & Glover, 2017). In addition, Indian researchers documented other UTs in GR2-R1, including dwarfism, bushy stature, pale green leaves, root defects, late flowering, and low fertility (Bollinedi et al., 2017). At least two underlying defects contribute to the UTs observed in GR2-R1 rice. The first pertains to the introduced DNA itself. In GR2-R1 plants the enzymes specified by the transgenes are active in other tissues apart from the grain (Bollinedi et al., 2017). This indicates the failure of the GR2 transgene regulatory sequences to function as intended, at least in GR2-R1 (Paine et al., 2005). The second defect was discovered when the Indian researchers sequenced the site of GR2-R1 integration. In GR2-R1,

Will gene-edited and other GM crops fail sustainable food systems?261

the integrated transgene disrupts a native gene, called OsAux1, that specifies an auxin transport protein (Bollinedi et al., 2017). Auxins are plant hormones with vital functions in growth and behavior. Both the regulatory and the insertion-site defects are predicted to impact several additional plant hormones, all with important roles in plant growth and development. These include abscisic acid, gibberellin, and cytokinin. Indeed, the researchers found levels of these three hormones were altered in leaves, stems, and flowering parts of GR2-R1 rice, as compared to non-GM isogenic lines (Bollinedi et al., 2017). In light of its many UTs and inherent molecular defects, efforts to further develop GR2-R1 were abandoned.

13.4.2.2 GR2E: Low levels of β-carotene in grain As GR2-R1’s defects became clear, a second event, GR2E, was incorporated into IRRI’s Golden Rice breeding program (Dubock, 2014). However, the effectiveness of the Golden Rice trait to produce provitamin A varies widely between events (Bollinedi et al., 2019; Paine et al., 2005). Of Syngenta’s six GR2 events, GR2E has the lowest β-carotene levels (Bollinedi et al., 2014; Paine et al., 2005). While other, sometimes higher, measurements exist in the scientific and regulatory literature (Bollinedi et al., 2019; FSANZ, 2017; Paine et al., 2005; Schaub et al., 2017), the data submitted to regulators worldwide gave the β-carotene level of GR2E rice as only 3.5 μg/g when milled and 0.5–2.35 μg/g when unmilled (FDA, 2018a, 2018b).

13.4.2.3 Golden Rice: β-carotene degrades rapidly in storage Two different research groups have reported that β-carotene levels in GR2E rice grains decrease rapidly in storage (Bollinedi et al., 2019; Schaub et al., 2017). After 3 weeks of storage, Golden Rice GR2E retained only 60% of its original levels. After 10 weeks, only 13% remained (Schaub et al., 2017). The second paper reported similar results, this time for both GR2E and GR2-R1 (Bollinedi et al., 2019). For GR2-R1, rapid degradation was shown to occur in several different genetic backgrounds. Cooking was shown to further decrease β-carotene levels (Bollinedi et al., 2019). Together, these results suggest that rapid degradation of β-carotene during normal storage and cooking conditions is a general problem of Golden Rice varieties.

13.4.2.4 Golden Rice: Commercialization despite missing benefit and risk assessment? Agronomic and biosafety UTs in Golden Rice are likely to arise from two aspects of the trait. First, UTs can arise from unintended alterations to the many biosynthetic pathways that intersect with the β-carotene biosynthesis pathway, as demonstrated in GR2-R1. These intersecting pathways produce a wide variety of compounds in addition to plant hormones, including volatiles, other carotenoids, and unknown signaling molecules (DellaPenna & Pogson, 2006). Such UTs in the grain might impact nutrition, toxicity, seed dormancy, germination, and fertility, for example.

262

Rethinking Food and Agriculture

Second, the rapid breakdown of β-carotene in the grain raises questions about the level and biosafety of the breakdown products (Schaub et al., 2017). However, despite the high probability of UTs, there is currently a complete absence of -omic data or other applicable research for GR2E. Furthermore, key human efficacy and safety studies are still lacking for targeted populations (Schubert, 2008; Stone & Glover, 2017; Then & Bauer-Panskus, 2018). Yet, regulators in Australia, the United States, and Canada have accepted developer’s biosafety claims for GR2E (IRRI, 2019). The data on Golden Rice thus conflict with proponents’ claims that critics and overregulation are responsible for the ongoing failure of Golden Rice (Dubock, 2014; Lee & Krimsky, 2016; Stone & Glover, 2017). Instead, the available data suggest the commercialization of Golden Rice has been consistently hindered by technical difficulties inherent to GM plant breeding. Furthermore, the current leading candidate, GR2E, is unlikely to make a useful contribution to the stated humanitarian goal of helping to alleviate vitamin A deficiency in target populations. This is due in part to the low initial levels, and subsequent rapid degradation, of β-carotene in GR2E grains. Its commercialization would, however, introduce unnecessary agronomic and biosafety risks into the food system. Vacuum packaging of Golden Rice, which has been suggested to slow β-carotene degradation, would further undermine food system sustainability (Bollinedi et al., 2019).

13.4.2.5 Golden Rice: Illuminating the failures of GM plant breeding The development of Golden Rice, a complex GM trait, exemplifies many of the inherent technical challenges faced by all GM plant breeders. These include (1) the imprecision and mutagenic nature of the techniques used to introduce GM traits, (2) inadequate scientific understanding of the biological processes underlying the relationships between transgenes and genome structure and function, and (3) the limited scientific understanding of the relationships between genes and traits, and how these are impacted by developmental and/or environmental factors. These technical difficulties combine to make GM plant breeding prone to UTs and, ultimately, failure. The history of Golden Rice development suggests that the production of safe and robust crop varieties that successfully express complex GM traits, those most likely to be useful for sustainable agriculture, is likely to be even more problematic. Golden Rice further highlights the overall institutional failure of regulators to implement adequate risk assessment and regulation for GM crops. As many researchers have already noted, more stringent regulation is needed (1) to ensure GM traits and crops fulfill their stated purpose and also (2) to safeguard the food system and the environment (Fox et al., 2006; Freese & Schubert, 2004; Heinemann, Agapito-Tenfen, & Carman, 2013; Heinemann, Kurenbach, & Quist, 2011; Hilbeck et al., 2015; Hilbeck & Otto, 2015; Latham et al., 2017; Mandel, 2003; Modonesi & Gusmeroli, 2018; Pelletier, 2005, 2006; Schubert, 2008; Venter & Bøhn, 2016; Wilson et al., 2006).

Will gene-edited and other GM crops fail sustainable food systems?263

13.4.3 Are new GM techniques more precise? Biotechnologists now claim that a new generation of genome modification techniques (nGMs) is essential to sustainably increase food production (Stone, 2017; Zaidi et al., 2019). These nGMs include cisgenesis/intragenesish and RNAi.i For the specific risks arising from intragenic, cisgenic, and RNAi-based traits, the reader is referred to other reviews (Casacuberta et al., 2015; Eckerstorfer et al., 2019; Gelinsky & Hilbeck, 2018; Heinemann et al., 2013; Lundgren & Duan, 2013; Senthil-Kumar & Mysore, 2011). The most recently developed nGMs for plants, and by far the most discussed, include techniques that are referred to as “gene editing” (Agapito-Tenfen, Okoli, Bernstein, Wikmark, & Myhr, 2018; Casacuberta et  al., 2015; Eckerstorfer et  al., 2019; Hou, Atlihan, & Lu, 2014; Lusser & Davies, 2013). Gene editing is a disparate family of techniques that include oligonucleotide-directed mutagenesis (ODM),j and/ or the use of site-directed nucleases such as meganucleases, TALENs,k ZFN,l and CRISPR/Cas9m Of these, CRISPR/Cas9 is the most widely used (Fichtner et al., 2014; Lusser & Davies, 2013; Sauer et al., 2016). The claimed benefit of gene editing is that genome modifications can be precisely targeted to specific genomic locations. A very wide variety of genome modifications can be intentionally introduced via gene editing (Ahmad, Rahman, Mukhtar, Zafar, & Zhang, 2019; Eckerstorfer et al., 2019; Fichtner et al., 2014; Lusser & Davies, 2013; Puchta, 2017). These extend from single base-pair changes to large-scale insertions or deletions of DNA. Insertions could include transgenes, cisgenes, RNAi-based traits, regulatory sequences, or multiple transgenes. CRISPR/Cas9 can further be used to create multiple changes at the same time, in a single gene or at multiple sites in the genome. Called multiplexing, this technique can be used, for example, to mutate or knock out several different members of a gene family (Fichtner et al., 2014). A key question for sustainability is whether gene editing, which is claimed to be far more precise than standard GM, can introduce beneficial traits without the introduction of UTs. While the publicly available data for edited crop plants is still extremely

h

Cisgenic/intragenic traits utilize only DNA derived from the host plant or a cross-compatible plant. In this they differ from standard transgenic traits that routinely utilize DNA from distantly related organisms. i RNAi-based traits specify double-stranded RNA molecules that trigger RNA interference (RNAi) pathways. This disrupts the cellular processes that connect specific genes with the production of specific proteins (Heinemann et al., 2013). The target of the RNAi is specified by the nucleotide sequence of the RNAi molecule. j ODM is a generic term for a wide range of different methodologies that use synthetic oligonucleotides to introduce a specific mutation at a particular site in the plant genome (ACRE, 2011). The oligonucleotides used for ODM are homologous to the targeted endogenous plant sequences except for the site of the intended mutation. k TALENs stands for transcription activator-like effector nucleases. TALENs are engineered nucleases that cut DNA at specific target sequences. l ZFN stands for zinc finger nuclease. ZFNs are engineered DNA-binding proteins that cut DNA at specific target sequences. m CRISPR stands for clustered regularly interspaced short palindromic repeats. In the CRISPR/Cas system, the engineered CRISPR RNA acts as a “guide” RNA that combines with a protein, for example, the Cas9 nuclease, and targets it to a specific DNA sequence.

264

Rethinking Food and Agriculture

limited, there are reasons to expect that plant gene editing methods are also prone to introducing UTs. First, virtually all gene editing protocols utilize standard GM techniques, i.e., tissue culture and either Agrobacterium infection or the gene gun (Ahmad et al., 2019; Ding, Li, Chen, & Xie, 2016; Eckerstorfer et  al., 2019). These techniques serve to introduce the gene editing reagents, which can include DNA, RNA, protein, or oligonucleotides, into plant cells. For example, Agrobacterium infection can be used to introduce DNA that specifies the CRISPR RNA guide sequence and the Cas9 nuclease, either transiently or via DNA integration. However, as discussed previously, tissue culture, Agrobacterium infection, and the gene gun are highly mutagenic, able to introduce thousands of mutations throughout the genome (Wilson et al., 2006). Proposed alternative methods, such as direct uptake of DNA-free reagents into protoplasts, may be less mutagenic and thus less likely to introduce UTs. However, this assumption remains to be tested experimentally. Furthermore, such methods are currently not available for most crop species (Ding et al., 2016). Second, new evidence from both animals and plants indicates that gene editing itself can result in unintended mutations at or near the target site. These include the insertion of vector, bacterial, and other superfluous DNA, and the unintended introduction of large DNA deletions and rearrangements (Biswas et  al., 2020; Kosicki, Tomberg, & Bradley, 2018; Li et al., 2015; Norris et al., 2019; Ono et al., 2015). Third, new research from animals suggests that even precise and intended edits can cause frequent on-target mRNA misregulation (Sharpe & Cooper, 2017; Tuladhar et al., 2019). These include “exon skipping” and unintentionally altered RNA splicing. Both can produce new protein coding sequences with the potential to result in UTs. Fourth, it has been shown in both plants and animals that gene editing reagents can make cuts at unintended sites in the genome. These cuts can result in off-target edits and potentially UTs (Ahmad et al., 2019; Biswas et al., 2020; Fichtner et al., 2014; Jin et al., 2019). Fifth, gene editing is being applied to situations where researchers have little prior research to guide them. Thus some researchers suggest gene editing can be used for “fast tracking development of underutilized species or perhaps wild species into widely adapted options to help improve global food security” (Van Eck, 2018). Additionally, it is being adapted to target regions of the genome that, during conventional breeding, are usually protected from genomic change (Kawall, 2019). The results of multiplexing would also be difficult if not impossible to introduce via conventional plant breeding (Kawall, 2019). Such novel uses of gene editing will likely increase the already high likelihood of introducing UTs. These observations support the conclusion that plant gene editing outcomes are imprecise and unpredictable, and that, depending on the combination of techniques used, gene editing can be highly mutagenic. However, because gene editing is a new field of research, particularly for plants, there are still many knowledge gaps (Ahmad et al., 2019; Schindele, Wolter, & Puchta, 2018). Lacking are whole genome sequences for gene-edited crop plants and nonedited comparators. Also lacking are systematic analyses of UTs in gene-edited crops. The knowledge gaps are especially large with regard to the unintended effects of different types of gene editing techniques and different

Will gene-edited and other GM crops fail sustainable food systems?265

types of edits, particularly in crops being developed for commercial use. That UTs in gene-edited organisms will be frequent, however, is suggested by a recent news report. This described gene-edited animals with UTs that ranged from spotted fleece to big tongues, extra vertebrae, sickness, and death (Rana & Craymer, 2018). Nevertheless, most plant gene editing papers have not systematically tested for UTs. However, one group found that CRISPR-Cas9-generated promoter variants of the maize ARGOS8 gene, intended to improve maize grain yield under drought stress, exhibited UTs, including statistically significant differences in plant height, ear height, and grain moisture (Shi et al., 2017). Therefore while biotechnologists frequently claim, “CRISPR in agriculture should be best considered as simply a ‘new breeding method’ that can produce identical results to conventional methods in a much more predictable, faster and even cheaper manner” (Gao, 2018), it is clear that CRISPR and other gene editing techniques are more similar to GM than conventional plant breeding due to the use of plant transformation techniques to introduce DNA or other reagents, the use of tissue culture, and the potential for exogenous DNA insertion (Eckerstorfer et al., 2019; Kawall, 2019). Like standard GM, the number and type of UTs introduced using gene editing will depend, in part, on the new trait being introduced and, in part, on the unintended effects wrought on the genome by the techniques themselves. All benefits, hazards, and risks must therefore be assessed experimentally on a case-by-case basis for each independently derived nGM trait and crop (Biswas et al., 2020; Eckerstorfer et al., 2019; Gelinsky & Hilbeck, 2018; Hilbeck et al., 2015). This should include whole genome sequencing comparisons with an isogenic line and -omic analyses.

13.5 Sustainable agriculture and plant breeding As discussed, the introduction of GM crops has not made conventional agriculture more sustainable. However, the failures of GM plant breeding, and of GM agriculture more broadly, provide insight into the changes necessary for a transition to sustainable systems. An important insight comes from the experience of farmers who have turned to regenerative agriculture. These farmers replace GM crops and their high-input management systems with some combination of sustainable practices that increase biodiversity, decrease topsoil loss, and increase natural soil fertility (e.g., complex cover crops, intercropping, multiyear multicrop rotations, GM-free no-till agriculture, the reintroduction of livestock). Research suggests regenerative farming leads to increased financial and environmental sustainability (LaCanne & Lundgren, 2018). Financial benefits occur primarily through the lower cost of conventional seeds, and because improved soils and decreased pest pressures reduce the need for costly and polluting inputs, including synthetic pesticides and fertilizers. In other words, the introduction of new traits or cultivars is only one of the many components necessary to improve sustainability. Conversely, to support sustainable agriculture, plant breeders must develop traits specifically tailored to, and selected within, low-input sustainable systems (Murphy, Campbell, Lyon, & Jones, 2007; van Bueren et al., 2011).

266

Rethinking Food and Agriculture

To date, however, the vast majority of conventional crop varieties used within organic and other sustainable systems have been bred for and selected within conventional systems (Murphy et al., 2007; van Bueren et al., 2011). But varieties that give high yields or perform best in conventional systems do not always do best in organic systems (Murphy et  al., 2007; van Bueren et  al., 2011). This is because pests and pathogens of primary concern for resistance breeding differ between the two systems. Weeds also are differentially problematic, since herbicides are not used in organic systems (van Bueren et al., 2011). Additionally, non-GM cultivars developed for conventional systems can have extremely negative trade-offs for sustainability. For example, semidwarf cereal varieties, introduced to prevent lodging, have a variety of UTs, such as decreased mineral nutrition and protein content; decreased root size and depth; decreased disease resistance and nutrient use efficiency; and poorer weed suppression (Marles, 2017; van Bueren et al., 2011). These defects can be masked in high-input systems. Thus crops and traits bred specifically for sustainable systems could greatly benefit both the yields and the performance of sustainable systems.

13.5.1 Traits for sustainable systems Researchers have identified a number of characters and traits that are likely to be of general importance when breeding cultivars for organic agriculture and other sustainable systems (van Bueren et al., 2011). These include increased nutrient-use efficiency (vigorous root systems or root exudates that promote beneficial symbiosis with soil microbiota) or uptake (increase in fine roots); resistance to fungal and bacterial disease; insect resistance (e.g., changes in life history, gross morphology, physical characteristics, or metabolism to promote resistance; Carmona, Lajeunesse, & Johnson, 2011); improved ability to compete against weeds; improved tolerance to abiotic stressors; and quality improvements, including improved nutritional value. Modern conventional crop varieties, and, in particular, landraces or farmers varieties and their wild relatives, all provide valuable sources of variation when breeding for sustainable systems (Deb, 2014; Dwivedi et al., 2016; van Bueren et al., 2011). The usefulness and impact of a particular trait will depend on the specific cropping system and the crop species (van Bueren et  al., 2011). For example, within no-till systems, weeds are potential problems, as are deep planting depths and soil moisture (Joshi, Chand, Arun, Singh, & Ortiz, 2007). Consequently, useful traits for herbicide-­ free no-till systems include those leading to faster seed emergence (or other traits that increase competitiveness against weeds); faster (or in some cases slower) residue decomposition; the ability to germinate when deep seeded; and resistance to mechanical weeding (Joshi et al., 2007; van Bueren et al., 2011). Other traits with potential benefits include resistance to pests and pathogens that survive on crop residues and resistance to phytotoxic organic acids released by some residues (Joshi et al., 2007). The priorities for cultivars intended for use in polyculture systems (e.g., the ancient “three sisters” maize/bean/squash system of the Americas, cereal and legume systems in Africa or Asia, or covercrop polycultures) differ from those of no-till. Useful traits tend to promote complementarity rather than competition between the crops. For example, researchers found maize, squash, and beans grown in a polyculture have

Will gene-edited and other GM crops fail sustainable food systems?267

a niche complementarity-dependent yield advantage (as compared to being grown in monoculture), likely arising from differences in root nutrient foraging strategies (Zhang et al., 2014). Therefore traits that promote root or shoot plasticity could be explored for polyculture crops (Zhang et al., 2014). On the other hand, allelopathic traits, as suggested for no-till systems to promote weed suppression (van Bueren et al., 2011), might negatively impact polyculture symbioses. Natural ecosystems can suggest further traits. For example, prairies have inspired efforts to breed perennial grains (e.g., wheat) and sunflowers (Piper, 1993). A recent review suggests perennial grains, which require well-developed root systems, have lower input requirements, and can support multifunctional sustainable systems. For example, they can protect and improve soil on sloped land, when intercropped with annuals or perennials, and/or in grazing systems or long-term rotations (Ryan et al., 2018).

13.5.2 Sustainable breeding: Considerations and strategies For plant breeders to support the kinds of food and seed sovereignty goals embodied by agroecological farming systems, they need to be responsive to farmer and community needs. This in turn requires breeding strategies that are flexible and easily adapted to differences in environment, scale, sustainable practice, and markets. Flexibility and adaptability are essential given the large variety of low-input and sustainable systems possible and the widely varying needs of local food systems. To these can be added the growing uncertainty generated by climate change. Participatory plant breeding strategies, where breeders collaborate with farmers (and sometimes others, including consumers and traditional farmer-focused breeding companies) are a promising method to achieve these multiple outcomes (Ceccarelli & Grando, 2019; Cleveland, Daniela, & Smith, 2000; Mercer et  al., 2012; Murphy, Lammer, Lyon, Carter, & Jones, 2005; van Bueren et al., 2011). Involving farmers at the start of the breeding program broadens understanding and should better ensure varieties provide farmer-preferred traits. Holding variety trials in farmers’ fields, as well as in test plots, and involving farmers in the selection process are other components of participatory plant breeding. In addition to providing well-adapted local varieties, participatory plant breeding has multiple other benefits. These include decreasing costs while educating and empowering all parties. It also facilitates uptake of new varieties (Najeeb et al., 2018). Participatory plant breeding can also facilitate the optimization and use of suitable variety mixtures. Mixtures are a promising strategy to increase resilience, especially for self-pollinating or clonal crops. Different varieties planted in a single field can improve yield stability under variable biotic or abiotic stress conditions, or allow for continued adaptation to changing conditions (Cleveland et al., 2000; Murphy et al., 2005; Phillips & Wolfe, 2005). For example, East African farmers grow mixtures of many varieties of common bean that are resistant to different diseases at different sites. Some Andean farmers maintain a large variety of different potato cultivars via harvesting and planting bulk mixtures (Cleveland et al., 2000). Evolutionary plant breeding is another strategy that can provide genetically diverse and resilient crops for sustainable agriculture (Döring, Knapp, Kovacs, Murphy, & Wolfe, 2011; Murphy et al., 2005;

268

Rethinking Food and Agriculture

Phillips & Wolfe, 2005; Raggi et al., 2017). In this case breeders produce heterogeneous composite cross populations with high inherent genetic diversity, for example to biotic or abiotic stressors (Jackson, Kahler, Webster, & Allard, 1978). These populations are successively selected under low-input conditions in natural cropping systems, often within participatory plant breeding programs. Over time, breeder and/ or farmer selection can produce heterogeneous populations or pure lines well-adapted to variable low-input cropping systems. A final important consideration running through all of plant breeding is one of control, and thus power. Conventional and GM plant breeding currently encourage or require the yearly purchase of seeds by farmers. GM crops also have patent restrictions on research, seed saving, and use for breeding. Their development requires specialized knowledge, equipment, and reagents. These factors keep control in the hands of seed companies and large institutions. To support food and seed sovereignty, in addition to using participatory methods, sustainable plant breeders must prioritize varieties that facilitate farmer seed saving and adaptation. These should be unpatented and free from other restrictions, particularly on seed sharing, breeding, or research.

13.6 Conclusions: Obstacles and opportunities While in theory it might someday be possible to create a GM crop that meets the broad requirements of sustainable agriculture, in practice this seems highly unlikely to ever happen (Kesavan & Swaminathan, 2018; Wickson et al., 2016). Nevertheless, despite their numerous technical, ecological, and social failings (e.g., Benbrook, 2018; Fischer, 2016; Wickson et al., 2016), GM crops are commercially successful, dominating the market for specific commodity crops in a number of countries. This success has been attributed to various factors. Farmer surveys suggest HT crops can save time and provide more spraying flexibility, and Bt crops are considered “insurance” to reduce risk. Research also suggests farmers often have limited seed options, and they often are locked into technological treadmills, in part due to a loss of knowledge about alternatives, or a belief that a technology is inevitable (Pechlaner, 2010; Stone & Flachs, 2018). Ultimately, however, the commercial successes of GM crops are due to politics, rather than technical factors. Science and technology are not neutral (O'Brien, 1993). GM crops support conventional agriculture, which in turn supports a vast corporate agro-industrial complex (Lima, 2015). For economic and ideological reasons, the US government, nongovernmental organizations, universities, and academics work with agribusiness to promote the uptake of GM crops and technologies, tailoring public research, government regulation, and subsidies to promote their rapid acceptance and expansion, while suppressing unwelcome findings and locking out alternatives (Binimelis et al., 2009; Cáceres, 2015; Capellesso et al., 2016; Foscolo & Zimmerman, 2013; Harsh, 2014; Peekhaus, 2010; Pelletier, 2005, 2006; Robinson, Holland, Leloup, & Muilerman, 2013; Schnurr, 2013; Schnurr & Gore, 2015; Schreiner, 2009; Vanloqueren & Baret, 2009; Waltz, 2009). The mainstream media further support

Will gene-edited and other GM crops fail sustainable food systems?269

these efforts by consistently portraying GM crops and technology as promising and technically successful (Barker, 2014; Stone, 2017). In spite of these systemic biases, there are signs of change. Some commodity crop farmers in the United States have abandoned GM crops and substituted more sustainable regenerative practices (LaCanne & Lundgren, 2018). Some GM plant breeders have switched from GM to conventional and participatory plant breeding (Ceccarelli & Grando, 2019; Gilbert, 2016). Meanwhile, the acreage of land under organic production has increased (Paull, 2017). Numerous researchers from different disciplines have called for both agricultural practices and plant breeding to become more socially and ecologically sustainable and reject GM crops (Kesavan & Swaminathan, 2018; Kremen & Miles, 2012; McIntyre et al., 2009; van Bueren, Struik, van Eekeren, & Nuijten, 2018). These are all hopeful signs that the scientific and political momentum is building to end the transgenic treadmill and transition to the agroecological and regenerative practices needed to underpin a sustainable food system, one that can support healthy people on a healthy planet (Anderson & Rivera Ferre, 2020; Valenzuela, 2016).

References ACRE. (2011). Advice on a plant breeding technique involving oligo-directed mutagenesis: RTDS™. Submitted by ACRE, U.K for Cibus HT oilseed rape. PDF Accessed May 2019 http://www.ogtr.gov.au/internet/ogtr/publishing.nsf/ Content/8884A10B0BA5CF42CA2580B10016087D/$File/Cibus%20-%203.pdf. Adhikari, J. (2014). Seed sovereignty: Analysing the debate on hybrid seeds and GMOs and bringing about sustainability in agricultural development. Journal of Forest and Livelihood, 12(1), 33–46. Agapito-Tenfen, S. Z., Guerra, M. P., Wikmark, O. G., & Nodari, R. O. (2013). Comparative proteomic analysis of genetically modified maize grown under different agroecosystems conditions in Brazil. Proteome Science, 11(1), 46. Agapito-Tenfen, S. Z., Okoli, A. S., Bernstein, M. J., Wikmark, O. G., & Myhr, A. I. (2018). Revisiting risk governance of GM plants: The need to consider new and emerging gene-­ editing techniques. Frontiers in Plant Science, 9, 1874. Ahmad, N., Rahman, M. U., Mukhtar, Z., Zafar, Y., & Zhang, B. (2019). A critical look on CRISPR‐based genome editing in plants. Journal of Cellular Physiology, 235. Altieri, M. A. (1999). The ecological role of biodiversity in agroecosystems. In Invertebrate biodiversity as bioindicators of sustainable landscapes (pp. 19–31): Elsevier. Altieri, M. A. (2000). The ecological impacts of transgenic crops on agroecosystem health. Ecosystem Health, 6(1), 13–23. Altieri, M. A. (2005). The myth of coexistence: Why transgenic crops are not compatible with agroecologically based systems of production. Bulletin of Science, Technology & Society, 25(4), 361–371. Ammann, K. (2005). Effects of biotechnology on biodiversity: Herbicide-tolerant and insect-­ resistant GM crops. Trends in Biotechnology, 23(8), 388–394. Anderson, R. L. (2015). Integrating a complex rotation with no-till improves weed management in organic farming. A review. Agronomy for Sustainable Development, 35(3), 967–974.

270

Rethinking Food and Agriculture

Anderson, M. D., & Rivera Ferre, M. G. (2020). Unsustainable by design: Extractive narratives of ending hunger and regenerative alternatives. Current Opinion in Environmental Sustainability. in press. Andow, D. A. (2010). Bt Brinjal: The scope and adequacy of the GEAC environmental risk assessment. http://re.indiaenvironmentportal.org.in/files/Bt_Brinjal.pdf. Accessed April 2019. Andreassen, M., Rocca, E., Bøhn, T., Wikmark, O. G., van den Berg, J., Løvik, M., … Nygaard, U. C. (2015). Humoral and cellular immune responses in mice after airway administration of bacillus thuringiensis Cry1Ab and MON810 cry1Ab-transgenic maize. Food and Agricultural Immunology, 26(4), 521–537. Ardekani, A. M., & Shirzad, M. (2019). Genetically modified (GM) foods and the risk to human health and environment. Health Biotechnology and Biopharma, 3(2), 61–73. Arpaia, S., Birch, A. N. E., Kiss, J., Van Loon, J. J., Messéan, A., Nuti, M., … Tebbe, C. C. (2017). Assessing environmental impacts of genetically modified plants on non-target organisms: The relevance of in planta studies. Science of the Total Environment, 583, 123–132. Barampuram, S., & Zhang, Z. J. (2011). Recent advances in plant transformation. In Plant chromosome engineering (pp. 1–35). Totowa, NJ: Humana Press. Barker, D. (2014). Genetically engineered (GE) crops: A misguided strategy for the twenty-first century? Development, 57(2), 192–200. Barros, E., Lezar, S., Anttonen, M. J., Van Dijk, J. P., Röhlig, R. M., Kok, E. J., & Engel, K. H. (2010). Comparison of two GM maize varieties with a near‐isogenic non‐GM variety using transcriptomics, proteomics and metabolomics. Plant Biotechnology Journal, 8(4), 436–451. Bashir, K., Husnain, T., Fatima, T., Latif, Z., Mehdi, S. A., & Riazuddin, S. (2004). Field evaluation and risk assessment of transgenic indica basmati rice. Molecular Breeding, 13(4), 301–312. Benbrook, C. (1999). Evidence of the magnitude and consequences of the roundup ready soybean yield drag from university-based varietal trials in 1998. Vol. 1. Sandpoint, ID: Benbrook Consulting Services. Benbrook, C. M. (2012). Impacts of genetically engineered crops on pesticide use in the US-the first sixteen years. Environmental Sciences Europe, 24(1), 24. Benbrook, C. M. (2018). Why regulators lost track and control of pesticide risks: Lessons from the case of glyphosate-based herbicides and genetically engineered-crop technology. Current Environmental Health Reports, 5(3), 387–395. Benevenuto, R. F., Agapito-Tenfen, S. Z., Vilperte, V., Wikmark, O. G., van Rensburg, P. J., & Nodari, R. O. (2017). Molecular responses of genetically modified maize to abiotic stresses as determined through proteomic and metabolomic analyses. PLoS One, 12(2), e0173069. Bergelson, J., Purrington, C. B., & Wichmann, G. (1998). Promiscuity in transgenic plants. Nature, 395(6697), 25. Bicar, E. H., Woodman-Clikeman, W., Sangtong, V., Peterson, J. M., Yang, S. S., Lee, M., & Scott, M. P. (2008). Transgenic maize endosperm containing a milk protein has improved amino acid balance. Transgenic Research, 17(1), 59–71. Binimelis, R., Pengue, W., & Monterroso, I. (2009). “Transgenic treadmill”: Responses to the emergence and spread of glyphosate-resistant johnsongrass in Argentina. Geoforum, 40(4), 623–633. Biswas, S., Tian, J., Li, R., Chen, X., Luo, Z., Chen, M., … Shi, J. (2020). Investigation of CRISPR/Cas9-induced SD1 rice mutants highlights the importance of molecular characterization in plant molecular breeding. Journal of Genetics and Genomics. in press.

Will gene-edited and other GM crops fail sustainable food systems?271

Bøhn, T., Cuhra, M., Traavik, T., Sanden, M., Fagan, J., & Primicerio, R. (2014). Compositional differences in soybeans on the market: Glyphosate accumulates in roundup ready GM soybeans. Food Chemistry, 153, 207–215. Bøhn, T., & Lövei, G. L. (2017). Complex outcomes from insect and weed control with transgenic plants: Ecological surprises? Frontiers in Environmental Science, 5, 60. Bollinedi, H., Dhakane-Lad, J., Krishnan, S. G., Bhowmick, P. K., Prabhu, K. V., Singh, N. K., & Singh, A. K. (2019). Kinetics of β-carotene degradation under different storage conditions in transgenic Golden Rice® lines. Food Chemistry, 278, 773–779. Bollinedi, H., Gopala, K. S., Sundaram, R. M., Sudhakar, D., Prabhu, K. V., & Singh, N. K. (2014). Marker assisted biofortification of rice with pro-vitamin a using transgenic golden rice lines: Progress and prospects. Indian Journal of Genetics, 74(4), 624–630. Bollinedi, H., Prabhu, K. V., Singh, N. K., Mishra, S., Khurana, J. P., & Singh, A. K. (2017). Molecular and functional characterization of GR2-R1 event based backcross derived lines of golden rice in the genetic background of a mega rice variety Swarna. PLoS One, 12(1), e0169600. Bonny, S. (2016). Genetically modified herbicide-tolerant crops, weeds, and herbicides: Overview and impact. Environmental Management, 57(1), 31–48. Booij, C. J. H. (2014). Ecological and experimental constraints for field trials to study potential effects of transgenic Bt-crops on non-target insects and spiders (no. 592). Cogem. Bourguet, D., Chaufaux, J., Micoud, A., Delos, M., Naibo, B., Bombarde, F., … Pagliari, C. (2002). Ostrinia nubilalis parasitism and the field abundance of non-target insects in transgenic bacillus thuringiensis corn (Zea mays). Environmental Biosafety Research, 1(1), 49–60. Brainard, D. C., Haramoto, E., Williams, M. M., & Mirsky, S. (2013). Towards a no-till nospray future? Introduction to a symposium on nonchemical weed management for reduced-­ tillage cropping systems. Weed Technology, 27(1), 190–192. Bregitzer, P., Halbert, S. E., & Lemaux, P. G. (1998). Somaclonal variation in the progeny of transgenic barley. Theoretical and Applied Genetics, 96(3–4), 421–425. Brodie, B. B. (2003). The loss of expression of theH 1 gene inBt transgenic potatoes. American Journal of Potato Research, 80(2), 135–139. Brodie, B. B., & Mai, W. F. (1989). Control of the golden nematode in the United States. Annual Review of Phytopathology, 27(1), 443–461. Cáceres, D. M. (2015). Accumulation by dispossession and socio‐environmental conflicts caused by the expansion of agribusiness in a rgentina. Journal of Agrarian Change, 15(1), 116–147. Camacho, A., Van Deynze, A., Chi-Ham, C., & Bennett, A. B. (2014). Genetically engineered crops that fly under the US regulatory radar. Nature Biotechnology, 32(11), 1087. Capellesso, A. J., Cazella, A. A., Schmitt Filho, A. L., Farley, J., & Martins, D. A. (2016). Economic and environmental impacts of production intensification in agriculture: Comparing transgenic, conventional, and agroecological maize crops. Agroecology and Sustainable Food Systems, 40(3), 215–236. Carmona, D., Lajeunesse, M. J., & Johnson, M. T. (2011). Plant traits that predict resistance to herbivores. Functional Ecology, 25(2), 358–367. Carrière, Y., Fabrick, J. A., & Tabashnik, B. E. (2016). Can pyramids and seed mixtures delay resistance to Bt crops? Trends in Biotechnology, 34(4), 291–302. Casacuberta, J. M., Devos, Y., Du Jardin, P., Ramon, M., Vaucheret, H., & Nogue, F. (2015). Biotechnological uses of RNAi in plants: Risk assessment considerations. Trends in Biotechnology, 33(3), 145–147.

272

Rethinking Food and Agriculture

Castaldini, M., Turrini, A., Sbrana, C., Benedetti, A., Marchionni, M., Mocali, S., … Nuti, M. P. (2005). Impact of Bt corn on rhizospheric and soil eubacterial communities and on beneficial mycorrhizal symbiosis in experimental microcosms. Applied and Environmental Microbiology, 71(11), 6719–6729. Ceccarelli, S., & Grando, S. (2019). Participatory plant breeding: Who did it, who does it and where? Experimental Agriculture, 56, 1–11. Cellini, F., Chesson, A., Colquhoun, I., Constable, A., Davies, H. V., Engel, K. H., … Lehesranta, S. (2004). Unintended effects and their detection in genetically modified crops. Food and Chemical Toxicology, 42(7), 1089–1125. Chakroun, M., Banyuls, N., Bel, Y., Escriche, B., & Ferré, J. (2016). Bacterial vegetative insecticidal proteins (Vip) from entomopathogenic bacteria. Microbiology and Molecular Biology Reviews, 80(2), 329–350. Chen, L. H., Zhang, B., & Xu, Z. Q. (2008). Salt tolerance conferred by overexpression of Arabidopsis vacuolar Na+/H+ antiporter gene AtNHX1 in common buckwheat (Fagopyrum esculentum). Transgenic Research, 17(1), 121. Chen, L. Y., Snow, A. A., Wang, F., & Lu, B. R. (2006). Effects of insect‐resistance transgenes on fecundity in rice (Oryza sativa, Poaceae): A test for underlying costs. American Journal of Botany, 93(1), 94–101. Clark, B. W., Phillips, T. A., & Coats, J. R. (2005). Environmental fate and effects of Bacillus thuringiensis (Bt) proteins from transgenic crops: A review. Journal of Agricultural and Food Chemistry, 53(12), 4643–4653. Cleveland, D. A., Daniela, S., & Smith, S. E. (2000). A biological framework for understanding farmers’ plant breeding. Economic Botany, 54(3), 377–394. Colyer, P. D., Kirkpatrick, T. L., Caldwell, W. D., & Vernon, P. R. (2000). Root-knot nematode reproduction and root galling severity on related conventional and transgenic cotton cultivars. Journal of Cotton Science, 4(4), 232–236. Dale, P. J., & McPartlan, H. C. (1992). Field performance of transgenic potato plants compared with controls regenerated from tuber discs and shoot cuttings. Theoretical and Applied Genetics, 84(5–6), 585–591. Davidson, M. M., Takla, M. F. G., Jacobs, J. M. E., Butler, R. C., Wratten, S. D., & Conner, A. J. (2004). Transformation of potato (Solanum tuberosum) cultivars with a cry 1Ac9 gene confers resistance to potato tuber moth (Phthorimaea operculella). New Zealand Journal of Crop and Horticultural Science, 32(1), 39–50. Davison, J. (2010). GM plants: Science, politics and EC regulations. Plant Science, 178(2), 94–98. Deb, D. (2014). Folk rice varieties, traditional agricultural knowledge and food security (pp. 45–57). Penang, Malaysia: Multicultural Knowledge and the University. Multiversity/ Citizens International. DellaPenna, D., & Pogson, B. J. (2006). Vitamin synthesis in plants: Tocopherols and carotenoids. Annual Review of Plant Biology, 57, 711–738. Dibden, J., Gibbs, D., & Cocklin, C. (2013). Framing GM crops as a food security solution. Journal of Rural Studies, 29, 59–70. Diels, J., Cunha, M., Manaia, C., Sabugosa-Madeira, B., & Silva, M. (2011). Association of financial or professional conflict of interest to research outcomes on health risks or nutritional assessment studies of genetically modified products. Food Policy, 36(2), 197–203. Ding, Y., Li, H., Chen, L. L., & Xie, K. (2016). Recent advances in genome editing using CRISPR/Cas9. Frontiers in Plant Science, 7, 703. Döring, T. F., Knapp, S., Kovacs, G., Murphy, K., & Wolfe, M. S. (2011). Evolutionary plant breeding in cereals—Into a new era. Sustainability, 3(10), 1944–1971.

Will gene-edited and other GM crops fail sustainable food systems?273

Douglas, M. R. (2016). Ecological trade-offs associated with insecticide use, from Pennsylvania to Bangladesh. Dissertation The Pennsylvania State University The Graduate School Department of Entomology. Accessed Sept 2019 https://etda.libraries.psu.edu/files/ final_submissions/12394. Douglas, M. R., & Tooker, J. F. (2015). Large-scale deployment of seed treatments has driven rapid increase in use of neonicotinoid insecticides and preemptive pest management in US field crops. Environmental Science & Technology, 49(8), 5088–5097. Dubock, A. (2014). The present status of golden rice. Journal of Huazhong Agricultural University, 33(6), 69–84. Available at: https://sfiar.ch/fileadmin/user_upload/documents/ dubock_state_of_golden_rice_2014_10_.pdf. (Accessed June 2019). Dwivedi, S. L., Ceccarelli, S., Blair, M. W., Upadhyaya, H. D., Are, A. K., & Ortiz, R. (2016). Landrace germplasm for improving yield and abiotic stress adaptation. Trends in Plant Science, 21(1), 31–42. Eckerstorfer, M. F., Heissenberger, A., Reichenbecher, W., Steinbrecher, R. A., & Waßmann, F. (2019). An EU perspective on biosafety considerations for plants developed by genome editing and other new genetic modification techniques (nGMs). Frontiers in Bioengineering and Biotechnology, 7, 31. Elmore, R. W., Roeth, F. W., Nelson, L. A., Shapiro, C. A., Klein, R. N., Knezevic, S. Z., & Martin, A. (2001). Glyphosate-resistant soybean cultivar yields compared with sister lines. Agronomy Journal, 93(2), 408–412. Fagan, J., Traavik, T., & Bøhn, T. (2015). The Seralini affair: Degeneration of science to re-­ science? Environmental Sciences Europe, 27(1), 19. Faria, C. A., Wäckers, F. L., Pritchard, J., Barrett, D. A., & Turlings, T. C. (2007). High susceptibility of Bt maize to aphids enhances the performance of parasitoids of lepidopteran pests. PLoS One, 2(7), e600. FDA. (2018a). GR2E biotechnology notification file no. 000158 note to the file date: May 8, 2018. Available at: https://www.fda.gov/downloads/Food/IngredientsPackagingLabeling/ GEPlants/Submissions/ucm607450.pdf. (Accessed June 2018). FDA. (2018b). GR2E response letter RE: Biotechnology notification file no. BNF 000158. Available at: https://www.fda.gov/downloads/Food/IngredientsPackagingLabeling/GEPlants/ Submissions/ucm608797.pdf. (Accessed June 2018). Felcher, K. J., Douches, D. S., Kirk, W. W., Hammerschmidt, R., & Li, W. (2003). Expression of a fungal glucose oxidase gene in three potato cultivars with different susceptibility to late blight (Phytophthora infestans Mont. deBary). Journal of the American Society for Horticultural Science, 128(2), 238–245. Fichtner, F., Castellanos, R. U., & Ülker, B. (2014). Precision genetic modifications: A new era in molecular biology and crop improvement. Planta, 239(4), 921–939. Fischer, K. (2016). Why new crop technology is not scale-neutral—A critique of the expectations for a crop-based African Green revolution. Research Policy, 45(6), 1185–1194. Flores, S., Saxena, D., & Stotzky, G. (2005). Transgenic Bt plants decompose less in soil than non-Bt plants. Soil Biology and Biochemistry, 37(6), 1073–1082. Flowers, T. J., & Yeo, A. R. (1995). Breeding for salinity resistance in crop plants: Where next? Functional Plant Biology, 22(6), 875–884. Foley, J. A., DeFries, R., Asner, G. P., Barford, C., Bonan, G., Carpenter, S. R., … Helkowski, J. H. (2005). Global consequences of land use. Science, 309(5734), 570–574. Foscolo, J., & Zimmerman, M. (2013). Alternative growth: Forsaking the false economies of industrial agriculture. Fordham Environmental Law Review, 25, 316. Fox, S., Morrison-Saunders, A., & Katscherian, D. (2006). Biotechnology risk assessment in Australia: A molecular perspective. Environmental and Planning Law Journal, 23(3), 236–248.

274

Rethinking Food and Agriculture

Freese, W., & Schubert, D. (2004). Safety testing and regulation of genetically engineered foods. Biotechnology and Genetic Engineering Reviews, 21(1), 299–324. FSANZ. (2017). Supporting document 1. Safety assessment report (at approval)—Application A1138. Food derived from Provitamin A rice line GR2E. Available at: http://www.foodstandards.gov.au/code/applications/Documents/A1138_SD1_at approval.pdf. (Accessed June 2019). Gao, C. (2018). The future of CRISPR technologies in agriculture. Nature Reviews. Molecular Cell Biology, 19(5), 275–276. García, M. J., Palma-Bautista, C., Rojano-Delgado, A. M., Bracamonte, E., Portugal, J., Alcántara-de la Cruz, R., & De Prado, R. (2019). The triple amino acid substitution TAPIVS in the EPSPS gene confers high glyphosate resistance to the Superweed Amaranthus hybridus. International Journal of Molecular Sciences, 20(10), 2396. Garibaldi, L. A., Gemmill-Herren, B., D’Annolfo, R., Graeub, B. E., Cunningham, S. A., & Breeze, T. D. (2017). Farming approaches for greater biodiversity, livelihoods, and food security. Trends in Ecology & Evolution, 32(1), 68–80. Gelinsky, E., & Hilbeck, A. (2018). European Court of Justice ruling regarding new genetic engineering methods scientifically justified: A commentary on the biased reporting about the recent ruling. Environmental Sciences Europe, 30(1), 52. Gepts, P. (2002). A comparison between crop domestication, classical plant breeding, and genetic engineering. Crop Science, 42(6), 1780–1790. Gilbert, N. (2016). Frugal farming. Nature, 533(7603), 308–310. Godfray, H. C. J., & Garnett, T. (2014). Food security and sustainable intensification. Philosophical Transactions of the Royal Society, B: Biological Sciences, 369(1639), 20120273. Gordon, B. (2007). Manganese nutrition of glyphosate-resistant and conventional soybeans. Better Crops, 91(4), 12–13. Gould, F., Brown, Z. S., & Kuzma, J. (2018). Wicked evolution: Can we address the sociobiological dilemma of pesticide resistance? Science, 360(6390), 728–732. Gray, M. E. (2010). Relevance of traditional integrated pest management (IPM) strategies for commercial corn producers in a transgenic agroecosystem: A bygone era? Journal of Agricultural and Food Chemistry, 59(11), 5852–5858. Green, J. M. (2014). Current state of herbicides in herbicide‐resistant crops. Pest Management Science, 70(9), 1351–1357. Griffiths, B. S., Caul, S., Thompson, J., Birch, A. N. E., Scrimgeour, C., Andersen, M. N., … Krogh, P. H. (2005). A comparison of soil microbial community structure, protozoa and nematodes in field plots of conventional and genetically modified maize expressing the Bacillus thuringiensis Cry1Ab toxin. Plant and Soil, 275, 135–146. Gutierrez, A. P., Ponti, L., Herren, H. R., Baumgärtner, J., & Kenmore, P. E. (2015). Deconstructing Indian cotton: Weather, yields, and suicides. Environmental Sciences Europe, 27(1), 12. Hailu, G., Niassy, S., Zeyaur, K. R., Ochatum, N., & Subramanian, S. (2018). Maize–legume intercropping and push–pull for Management of Fall Armyworm, Stemborers, and Striga in Uganda. Agronomy Journal, 110. Harsh, M. (2014). Nongovernmental organizations and genetically modified crops in Kenya: Understanding influence within a techno-civil society. Geoforum, 53, 172–183. Haslberger, A. G. (2003). Codex guidelines for GM foods include the analysis of unintended effects. Nature Biotechnology, 21(7), 739. Heinemann, J. A., Agapito-Tenfen, S. Z., & Carman, J. A. (2013). A comparative evaluation of the regulation of GM crops or products containing dsRNA and suggested improvements to risk assessments. Environment International, 55, 43–55.

Will gene-edited and other GM crops fail sustainable food systems?275

Heinemann, J. A., Kurenbach, B., & Quist, D. (2011). Molecular profiling—A tool for addressing emerging gaps in the comparative risk assessment of GMOs. Environment International, 37(7), 1285–1293. Heinemann, J. A., Massaro, M., Coray, D. S., Agapito-Tenfen, S. Z., & Wen, J. D. (2014). Sustainability and innovation in staple crop production in the US Midwest. International Journal of Agricultural Sustainability, 12(1), 71–88. Hilbeck, A., Binimelis, R., Defarge, N., Steinbrecher, R., Székács, A., Wickson, F., … Novotny, E. (2015). No scientific consensus on GMO safety. Environmental Sciences Europe, 27(1), 4. Hilbeck, A., Meier, M., & Trtikova, M. (2012). Underlying reasons of the controversy over adverse effects of Bt toxins on lady beetle and lacewing larvae. Environmental Sciences Europe, 24(1), 9. Hilbeck, A., & Otto, M. (2015). Specificity and combinatorial effects of Bacillus thuringiensis Cry toxins in the context of GMO environmental risk assessment. Frontiers in Environmental Science, 3, 71. Hilbeck, A., & Schmidt, J. E. (2006). Another view on Bt-proteins-how specific are they and what else might they do. Biopesticides International, 2(1), 1–50. Hoekema, A., Huisman, M. J., Molendijk, L., van den Elzen, P. J., & Cornelissen, B. J. (1989). The genetic engineering of two commercial potato cultivars for resistance to potato virus X. Bio/Technology, 7(3), 273. Hofvander, P., Andersson, M., Larsson, C. T., & Larsson, H. (2004). Field performance and starch characteristics of high‐amylose potatoes obtained by antisense gene targeting of two branching enzymes. Plant Biotechnology Journal, 2(4), 311–320. Holt-Giménez, E., & Altieri, M. A. (2013). Agroecology, food sovereignty, and the new green revolution. Agroecology and Sustainable Food Systems, 37(1), 90–102. Horak, M. J., Rosenbaum, E. W., Kendrick, D. L., Sammons, B., Phillips, S. L., Nickson, T. E., … Perez, T. (2015). Plant characterization of roundup ready 2 yield® soybean, MON 89788, for use in ecological risk assessment. Transgenic Research, 24(2), 213–225. Horrigan, L., Lawrence, R. S., & Walker, P. (2002). How sustainable agriculture can address the environmental and human health harms of industrial agriculture. Environmental Health Perspectives, 110(5), 445–456. Höss, S., Arndt, M., Baumgarte, S., Tebbe, C. C., Nguyen, H. T., & Jehle, J. A. (2008). Effects of transgenic corn and Cry1Ab protein on the nematode, Caenorhabditis elegans. Ecotoxicology and Environmental Safety, 70(2), 334–340. Hou, H., Atlihan, N., & Lu, Z. X. (2014). New biotechnology enhances the application of cisgenesis in plant breeding. Frontiers in Plant Science, 5, 389. Hu, T., Metz, S., Chay, C., Zhou, H. P., Biest, N., Chen, G., … Fry, J. (2003). Agrobacteriummediated large-scale transformation of wheat (Triticum aestivum L.) using glyphosate selection. Plant Cell Reports, 21(10), 1010–1019. IRRI. (2019). https://www.irri.org/golden-rice. (Accessed 17 April 2019). ISAAA. (2017). Global status of commercialized biotech/GM crops: 2017. ISAAA Brief No. 53. Ithaca, NY: ISAAA. https://www.isaaa.org/resources/publications/pocketk/16/. (Accessed July 2019). Jackson, L. F., Kahler, A. L., Webster, R. K., & Allard, R. W. (1978). Conservation of scald resistance in barley composite cross populations. Phytopathology, 68, 645–650. Jacobs, A., Lunde, C., Bacic, A., Tester, M., & Roessner, U. (2007). The impact of constitutive heterologous expression of a moss Na+ transporter on the metabolomes of rice and barley. Metabolomics, 3(3), 307–317.

276

Rethinking Food and Agriculture

Jayasumana, C., Gunatilake, S., & Senanayake, P. (2014). Glyphosate, hard water and nephrotoxic metals: Are they the culprits behind the epidemic of chronic kidney disease of unknown etiology in Sri Lanka? International Journal of Environmental Research and Public Health, 11(2), 2125–2147. Jiang, Y., Ling, L., Zhang, L., Wang, K., Li, X., Cai, M., … Cao, C. (2018). Comparison of transgenic Bt rice and their non-Bt counterpart in yield and physiological response to drought stress. Field Crops Research, 217, 45–52. Jin, S., Zong, Y., Gao, Q., Zhu, Z., Wang, Y., Qin, P., … Gao, C. (2019). Cytosine, but not adenine, base editors induce genome-wide off-target mutations in rice. Science, 364, eaaw7166. Joshi, A. K., Chand, R., Arun, B., Singh, R. P., & Ortiz, R. (2007). Breeding crops for reduced-­ tillage management in the intensive, rice–wheat systems of South Asia. Euphytica, 153(1–2), 135–151. Kawall, K. (2019). New possibilities on the horizon: genome editing makes the whole genome accessible for changes. Frontiers in Plant Science, 10, 525. Kesavan, P. C., & Malarvannan, S. (2010). Green to evergreen revolution: Ecological and evolutionary perspectives in pest management. Current Science, 99(7), 00113891. Kesavan, P. C., & Swaminathan, M. S. (2018). Modern technologies for sustainable food and nutrition security. Current Science, 115(10), 1876–1883. Kilman, S. (2010). Superweed outbreak triggers arms race. Wall Street Journal. Klümper, W., & Qaim, M. (2014). A meta-analysis of the impacts of genetically modified crops. PLoS One, 9(11), e111629. Koch, M. S., Ward, J. M., Levine, S. L., Baum, J. A., Vicini, J. L., & Hammond, B. G. (2015). The food and environmental safety of Bt crops. Frontiers in Plant Science, 6, 283. Kochman, J., Moore, N. Y., Obst, N. R., O’Neill, W. T., Salmond, G., & Bentley, S. (2000). Management strategies for Fusarium wilt of cotton. In Cotton: Meeting the challenge: Conference program: 10th Australian cotton conference (pp. 443–453). Orange, NSW: Australian Cotton Grower’ Research Association. Kollist, H., Zandalinas, S. I., Sengupta, S., Nuhkat, M., Kangasjärvi, J., & Mittler, R. (2019). Rapid responses to abiotic stress: Priming the landscape for the signal transduction network. Trends in Plant Science, 24, 25–37. Kosicki, M., Tomberg, K., & Bradley, A. (2018). Repair of double-strand breaks induced by CRISPR–Cas9 leads to large deletions and complex rearrangements. Nature Biotechnology, 36(8), 765. Kranthi, K. (2016). Fertilizers gave high yields, Bt-only provided cover. In Cotton statistics and news. 2016–2017, no. 39, 27 December 2016. Kremen, C., & Merenlender, A. M. (2018). Landscapes that work for biodiversity and people. Science, 362(6412), eaau6020. Kremen, C., & Miles, A. (2012). Ecosystem services in biologically diversified versus conventional farming systems: Benefits, externalities, and trade-offs. Ecology and Society, 17(4). Krimsky, S. (2015). An illusory consensus behind GMO health assessment. Science, Technology & Human Values, 40(6), 883–914. Kuiper, H. A., Kok, E. J., & Engel, K. H. (2003). Exploitation of molecular profiling techniques for GM food safety assessment. Current Opinion in Biotechnology, 14(2), 238–243. Kumar, A., Rakow, G., & Downey, R. K. (1998). Isogenic analysis of glufosinate-ammonium tolerant and susceptible summer rape lines. Canadian Journal of Plant Science, 78(3), 401–408. La Paz, J. L., Pla, M., Centeno, E., Vicient, C. M., & Puigdomènech, P. (2014). The use of massive sequencing to detect differences between immature embryos of MON810 and a comparable non-GM maize variety. PLoS One, 9(6), e100895.

Will gene-edited and other GM crops fail sustainable food systems?277

LaCanne, C. E., & Lundgren, J. G. (2018). Regenerative agriculture: Merging farming and natural resource conservation profitably. PeerJ, 6, e4428. Lang, A., Oehen, B., Ross, J. H., Bieri, K., & Steinbrich, A. (2015). Potential exposure of butterflies in protected habitats by Bt maize cultivation: A case study in Switzerland. Biological Conservation, 192, 369–377. Larkin, P., & Harrigan, G. G. (2007). Opportunities and surprises in crops modified by transgenic technology: Metabolic engineering of benzylisoquinoline alkaloid, gossypol and lysine biosynthetic pathways. Metabolomics, 3(3), 371–382. Latham, J. R., Love, M., & Hilbeck, A. (2017). The distinct properties of natural and GM cry insecticidal proteins. Biotechnology and Genetic Engineering Reviews, 33(1), 62–96. Lee, H., & Krimsky, S. (2016). The arrested development of golden rice: The scientific and social challenges of a transgenic biofortified crop. International Journal of Social Science Studies, 4, 51. Li, X., Huang, K., Zhu, B., Liang, Z., Wei, L., & Luo, Y. (2008). Comparative physicochemical properties and structure of rice containing the sck+ cryIAc genes and its nontransgenic counterpart. Journal of Food Science, 73(1), S64–S69. Li, X. G., Liu, B., Heia, S., Liu, D. D., Han, Z. M., Zhou, K. X., … Zheng, Y. P. (2009). The effect of root exudates from two transgenic insect-resistant cotton lines on the growth of Fusarium oxysporum. Transgenic Research, 18(5), 757–767. Li, Z., Liu, Z. B., Xing, A., Moon, B. P., Koellhoffer, J. P., Huang, L., … Cigan, A. M. (2015). Cas9-guide RNA directed genome editing in soybean. Plant Physiology, 169(2), 960–970. Lima, T. (2015). Agricultural subsidies for non-farm interests: An analysis of the US Agroindustrial complex. Agrarian South: Journal of Political Economy, 4(1), 54–84. Linder, C. R. (1998). Potential persistence of transgenes: Seed performance of transgenic canola and wild× canola hybrids. Ecological Applications, 8(4), 1180–1195. Linder, C. R., & Schmitt, J. (1995). Potential persistence of escaped transgenes: Performance of transgenic, oil‐modified Brassica seeds and seedlings. Ecological Applications, 5(4), 1056–1068. Lundgren, J. G., & Duan, J. J. (2013). RNAi-based insecticidal crops: Potential effects on nontarget species. Bioscience, 63(8), 657–665. Lusser, M., & Davies, H. V. (2013). Comparative regulatory approaches for groups of new plant breeding techniques. New Biotechnology, 30(5), 437–446. Ma, B. L., & Subedi, K. D. (2005). Development, yield, grain moisture and nitrogen uptake of Bt corn hybrids and their conventional near-isolines. Field Crops Research, 93(2–3), 199–211. Mandel, G. N. (2003). Gaps, inexperience, inconsistencies, and overlaps: Crisis in the regulation of genetically modified plants and animals. William and Mary Law Review, 45, 2167. Manetti, C., Bianchetti, C., Casciani, L., Castro, C., Di Cocco, M. E., Miccheli, A., … Conti, F. (2006). A metabonomic study of transgenic maize (Zea mays) seeds revealed variations in osmolytes and branched amino acids. Journal of Experimental Botany, 57(11), 2613–2625. Marles, R. J. (2017). Mineral nutrient composition of vegetables, fruits and grains: The context of reports of apparent historical declines. Journal of Food Composition and Analysis, 56, 93–103. Marmiroli, N., Agrimonti, C., Visioli, G., Colauzzi, M., Guarda, G., & Zuppini, A. (2000). Silencing of G1-1 and A2-1 genes. Effects on general plant phenotype and on tuber dormancy in Solanum tuberosum L. Potato Research, 43(4), 313–323. Mason, S. C., Walters, C. G., Galusha, T. D., Wilson, R. K., & Kmail, Z. (2017). Planting saved roundup ready 1 soybean seed implications on yield and profit. Agronomy Journal, 109(2), 729–734.

278

Rethinking Food and Agriculture

Matthews, D., Jones, H., Gans, P., Coates, S., & Smith, L. M. (2005). Toxic secondary metabolite production in genetically modified potatoes in response to stress. Journal of Agricultural and Food Chemistry, 53(20), 7766–7776. McFadden, J., Smith, D., Wechsler, S., & Wallander, S. (2019). Development, adoption, and management of drought-tolerant corn in the United States. Economic Information Bulletin, No. 204. USDA, Economic Research Service. McHenry, L. B. (2018). The Monsanto papers: Poisoning the scientific well. The International Journal of Risk & Safety in Medicine, 29, 1–13. preprint. McIntyre, B. D., Herren, H. R., Wakhungu, J., & Watson, R. T. (2009). Agriculture at a crossroad: Global report. Washington: International Assessment of Agricultural Knowledge, Science & Technology for Development, IAASTD, Island Press. McMichael, P., & Schneider, M. (2011). Food security politics and the millennium development goals. Third World Quarterly, 32(1), 119–139. Mercer, K. L., Perales, H. R., & Wainwright, J. D. (2012). Climate change and the transgenic adaptation strategy: Smallholder livelihoods, climate justice, and maize landraces in Mexico. Global Environmental Change, 22(2), 495–504. Mesnage, R., Agapito-Tenfen, S. Z., Vilperte, V., Renney, G., Ward, M., Séralini, G. E., … Antoniou, M. N. (2016). An integrated multi-omics analysis of the NK603 roundup-­ tolerant GM maize reveals metabolism disturbances caused by the transformation process. Scientific Reports, 6, 37855. Midega, C. A., Pittchar, J. O., Pickett, J. A., Hailu, G. W., & Khan, Z. R. (2018). A climate-­ adapted push-pull system effectively controls fall armyworm, Spodoptera frugiperda (JE Smith), in maize in East Africa. Crop Protection, 105, 10–15. Modonesi, C., & Gusmeroli, F. (2018). Evidence-based science or science-based evidence? The GM crops between false myths and ecological systems. Organisms. Journal of Biological Sciences, 2(1), 63–72. Mortensen, D. A., Egan, J. F., Maxwell, B. D., Ryan, M. R., & Smith, R. G. (2012). Navigating a critical juncture for sustainable weed management. Bioscience, 62(1), 75–84. Murphy, K., Lammer, D., Lyon, S., Carter, B., & Jones, S. S. (2005). Breeding for organic and low-input farming systems: An evolutionary–participatory breeding method for inbred cereal grains. Renewable Agriculture and Food Systems, 20(1), 48–55. Murphy, K. M., Campbell, K. G., Lyon, S. R., & Jones, S. S. (2007). Evidence of varietal adaptation to organic farming systems. Field Crops Research, 102(3), 172–177. Myers, J. P., Antoniou, M. N., Blumberg, B., Carroll, L., Colborn, T., Everett, L. G., … Vandenberg, L. N. (2016). Concerns over use of glyphosate-based herbicides and risks associated with exposures: A consensus statement. Environmental Health, 15(1), 19. Najeeb, S., Sheikh, F. A., Parray, G. A., Shikari, A. B., Kashyp, S. C., Ganie, M. A., & Shah, A. B. (2018). Farmers’ participatory selection of new rice varieties to boost production under temperate agro-ecosystems. Journal of Integrative Agriculture, 17(6), 1307–1314. Naranjo, S. E. (2009). Impacts of Bt crops on non-target invertebrates and insecticide use patterns. Biocontrol News and Information, 30(2), 1R. Nature Institute. (2019). Online database of non-target effects. Accessed 1 April 2019 at: http:// natureinstitute.org/nontarget/. Nelson, K. A., Renner, K. A., & Hammerschmidt, R. (2002). Cultivar and herbicide selection affects soybean development and the incidence of Sclerotinia stem rot. Agronomy Journal, 94(6), 1270–1281. Nguyen, H. T., & Jehle, J. A. (2007). Quantitative analysis of the seasonal and tissue-­specific expression of Cry1Ab in transgenic maize Mon810. Journal of Plant Diseases and Protection, 114(2), 82–87.

Will gene-edited and other GM crops fail sustainable food systems?279

Nicholls, C. I., Altieri, M. A., & Vazquez, L. (2016). Agroecology: Principles for the conversion and redesign of farming systems. Journal of Ecosystem and Ecography, S5, 010. Norris, A. L., Lee, S. S., Greenlees, K. J., Tadesse, D. A., Miller, M. F., & Lombardi, H. (2019). Template plasmid integration in germline genome-edited cattle. BioRxiv, 715482. https:// doi.org/10.1101/715482. Oard, J. H., Linscombe, S. D., Braverman, M. P., Jodari, F., Blouin, D. C., Leech, M., … Christou, P. (1996). Development, field evaluation, and agronomic performance of transgenic herbicide resistant rice. Molecular Breeding, 2(4), 359–368. O’Brien, M. H. (1993). Being a scientist means taking sides. Bioscience, 43(10), 706–708. Oliveira, G., & Hecht, S. (2016). Sacred groves, sacrifice zones and soy production: globalization, intensification and neo-nature in South America. The Journal of Peasant Studies, 43(2), 251–285. https://doi.org/10.1080/03066150.2016.1146705. Ono, R., Ishii, M., Fujihara, Y., Kitazawa, M., Usami, T., Kaneko-Ishino, T., … Ishino, F. (2015). Double strand break repair by capture of retrotransposon sequences and reverse-transcribed spliced mRNA sequences in mouse zygotes. Scientific Reports, 5, 12281. Paine, J. A., Shipton, C. A., Chaggar, S., Howells, R. M., Kennedy, M. J., Vernon, G., … Drake, R. (2005). Improving the nutritional value of Golden Rice through increased pro-vitamin a content. Nature Biotechnology, 23(4), 482. Paula, D. P., Andow, D. A., Timbó, R. V., Sujii, E. R., Pires, C. S., & Fontes, E. M. (2014). Uptake and transfer of a Bt toxin by a Lepidoptera to its eggs and effects on its offspring. PLoS One, 9(4), e95422. Paull, J. (2017). Four new strategies to grow the organic agriculture sector. AGROFOR International Journal, 2(3), 61–70. Faculty of Agriculture, University of East Sarajevo. Pechlaner, G. (2010). Biotech on the farm: Mississippi agriculture in an age of proprietary biotechnologies. Anthropologica, 52(2), 291. Peekhaus, W. (2010). The neoliberal university and agricultural biotechnology: Reports from the field. Bulletin of Science, Technology & Society, 30(6), 415–429. Pelletier, D. L. (2005). Science, law, and politics in the Food and Drug Administration’s genetically engineered foods policy: FDA’s 1992 policy statement. Nutrition Reviews, 63(5), 171–181. Pelletier, D. L. (2006). FDA’s regulation of genetically engineered foods: Scientific, legal and political dimensions. Food Policy, 31(6), 570–591. Pengue, W. A. (2005). Transgenic crops in Argentina: The ecological and social debt. Bulletin of Science, Technology & Society, 25(4), 314–322. Perfecto, I., & Vandermeer, J. (2010). The agroecological matrix as alternative to the land-­ sparing/agriculture intensification model. Proceedings of the National Academy of Sciences of the United States of America, 107(13), 5786–5791. Perry, E. D., Ciliberto, F., Hennessy, D. A., & Moschini, G. (2016). Genetically engineered crops and pesticide use in US maize and soybeans. Science Advances, 2(8), e1600850. Phillips, S. L., & Wolfe, M. S. (2005). Evolutionary plant breeding for low input systems. The Journal of Agricultural Science, 143(4), 245–254. Pierre, J., Marsault, D., Genecque, E., Renard, M., Champolivier, J., & Pham‐Delègue, M. H. (2003). Effects of herbicide‐tolerant transgenic oilseed rape genotypes on honey bees and other pollinating insects under field condtions. Entomologia Experimentalis et Applicata, 108(3), 159–168. Piper, J. K. (1993). A grain agriculture fashioned in nature’s image: The work of The Land Institute. Great Plains Research, 3, 249–272. Poerschmann, J., Gathmann, A., Augustin, J., Langer, U., & Górecki, T. (2005). Molecular composition of leaves and stems of genetically modified Bt and near-isogenic non-Bt maize— Characterization of lignin patterns. Journal of Environmental Quality, 34(5), 1508–1518.

280

Rethinking Food and Agriculture

Prendeville, H. R., & Pilson, D. (2009). Transgenic virus resistance in cultivated squash affects pollinator behaviour. Journal of Applied Ecology, 46(5), 1088–1096. Pretty, J. (2001). The rapid emergence of genetic modification in world agriculture: Contested risks and benefits. Environmental Conservation, 28(3), 248–262. Pretty, J. N., Brett, C., Gee, D., Hine, R. E., Mason, C. F., Morison, J. I. L., … van der Bijl, G. (2000). An assessment of the total external costs of UK agriculture. Agricultural Systems, 65(2), 113–136. Puchta, H. (2017). Applying CRISPR/Cas for genome engineering in plants: The best is yet to come. Current Opinion in Plant Biology, 36, 1–8. Raggi, L., Ciancaleoni, S., Torricelli, R., Terzi, V., Ceccarelli, S., & Negri, V. (2017). Evolutionary breeding for sustainable agriculture: Selection and multi-environmental evaluation of barley populations and lines. Field Crops Research, 204, 76–88. Rana, P., & Craymer, L. (2018). Big tongues and extra vertebrae: The unintended consequences of animal gene editing. The Wall Street Journal. Accessed May 2019 https://www.wsj.com/ articles/deformities-alarm-scientists-racing-to-rewrite-animal-dna-11544808779. Ricroch, A. E., Bergé, J. B., & Kuntz, M. (2011). Evaluation of genetically engineered crops using transcriptomic, proteomic, and metabolomic profiling techniques. Plant Physiology, 155(4), 1752–1761. Robinson, C., Holland, N., Leloup, D., & Muilerman, H. (2013). Conflicts of interest at the European food safety authority erode public confidence. Journal of Epidemiology and Community Health, 67(9), 717–720. Römer, S., Fraser, P. D., Kiano, J. W., Shipton, C. A., Misawa, N., Schuch, W., & Bramley, P. M. (2000). Elevation of the provitamin a content of transgenic tomato plants. Nature Biotechnology, 18(6), 666. Ryan, M. R., Crews, T. E., Culman, S. W., DeHaan, L. R., Hayes, R. C., Jungers, J. M., & Bakker, M. G. (2018). Managing for multifunctionality in perennial grain crops. Bioscience, 68(4), 294–304. Sánchez-Bayo, F., & Wyckhuys, K. A. (2019). Worldwide decline of the entomofauna: A review of its drivers. Biological Conservation, 232, 8–27. Sanchis, V. (2011). From microbial sprays to insect-resistant transgenic plants: History of the biospesticide Bacillus thuringiensis. A review. Agronomy for Sustainable Development, 31(1), 217–231. Sauer, N. J., Narváez-Vásquez, J., Mozoruk, J., Miller, R. B., Warburg, Z. J., Woodward, M. J., … Walker, K. A. (2016). Oligonucleotide-mediated genome editing provides precision and function to engineered nucleases and antibiotics in plants. Plant Physiology, 170(4), 1917–1928. Saxena, D., Stewart, C. N., Altosaar, I., Shu, Q., & Stotzky, G. (2004). Larvicidal Cry proteins from Bacillus thuringiensis are released in root exudates of transgenic B. thuringiensis corn, potato, and rice but not of B. thuringiensis canola, cotton, and tobacco. Plant Physiology and Biochemistry, 42(5), 383–387. Saxena, D., & Stotzky, G. (2001). Bt corn has a higher lignin content than non‐Bt corn. American Journal of Botany, 88(9), 1704–1706. Schaub, P., Wüst, F., Koschmieder, J., Yu, Q., Virk, P., Tohme, J., & Beyer, P. (2017). Nonenzymatic β-carotene degradation in provitamin A-biofortified crop plants. Journal of Agricultural and Food Chemistry, 65(31), 6588–6598. Schijlen, E., Ric de Vos, C. H., Jonker, H., Van Den Broeck, H., Molthoff, J., Van Tunen, A., … Bovy, A. (2006). Pathway engineering for healthy phytochemicals leading to the production of novel flavonoids in tomato fruit. Plant Biotechnology Journal, 4(4), 433–444.

Will gene-edited and other GM crops fail sustainable food systems?281

Schindele, P., Wolter, F., & Puchta, H. (2018). Transforming plant biology and breeding with CRISPR/Cas9, Cas12 and Cas13. FEBS Letters, 592(12), 1954–1967. Schnurr, M. A. (2013). Biotechnology and bio-hegemony in Uganda: Unraveling the social relations underpinning the promotion of genetically modified crops into new African markets. Journal of Peasant Studies, 40(4), 639–658. Schnurr, M. A., & Gore, C. (2015). Getting to ‘yes’: Governing genetically modified crops in Uganda. Journal of International Development, 27(1), 55–72. Schreiner, Z. R. (2009). Frankenfuel: Genetically modified corn, ethanol, and crop diversity. Energy Law Journal, 30, 169. Schubert, D. R. (2008). The problem with nutritionally enhanced plants. Journal of Medicinal Food, 11(4), 601–605. Schütte, G., Eckerstorfer, M., Rastelli, V., Reichenbecher, W., Restrepo-Vassalli, S., RuohonenLehto, M., … Mertens, M. (2017). Herbicide resistance and biodiversity: Agronomic and environmental aspects of genetically modified herbicide-resistant plants. Environmental Sciences Europe, 29(1), 5. Senthil-Kumar, M., & Mysore, K. S. (2011). Caveat of RNAi in plants: The off-target effect. In RNAi and plant gene function analysis (pp. 13–25). Humana Press. Séralini, G. E., Mesnage, R., Defarge, N., & de Vendômois, J. S. (2014). Conflicts of interests, confidentiality and censorship in health risk assessment: The example of an herbicide and a GMO. Environmental Sciences Europe, 26(1), 13. Sharma, A., Jha, P., & Reddy, G. V. (2018). Multidimensional relationships of herbicides with insect-crop food webs. Science of the Total Environment, 643, 1522–1532. Sharpe, J. J., & Cooper, T. A. (2017). Unexpected consequences: Exon skipping caused by CRISPR-generated mutations. Genome Biology, 18(1), 109. Shelton, A. M., Hossain, M., Paranjape, V., & Azad, A. K. (2018). Bt eggplant project in Bangladesh: History, present status, and future direction. Frontiers in Bioengineering and Biotechnology, 6, 106. Shepherd, L. V., McNicol, J. W., Razzo, R., Taylor, M. A., & Davies, H. V. (2006). Assessing the potential for unintended effects in genetically modified potatoes perturbed in metabolic and developmental processes. Targeted analysis of key nutrients and anti-nutrients. Transgenic Research, 15(4), 409–425. Shi, G., Chavas, J. P., & Lauer, J. (2013). Commercialized transgenic traits, maize productivity and yield risk. Nature Biotechnology, 31(2), 111. Shi, J., Gao, H., Wang, H., Lafitte, H. R., Archibald, R. L., Yang, M., … Habben, J. E. (2017). ARGOS 8 variants generated by CRISPR‐Cas9 improve maize grain yield under field drought stress conditions. Plant Biotechnology Journal, 15(2), 207–216. Shu, Q. Y., Cui, H. R., Ye, G. Y., Wu, D. X., Xia, Y. W., Gao, M. W., & Altosaar, I. (2002). Agronomic and morphological characterization of agrobacterium-transformed Bt rice plants. Euphytica, 127(3), 345–352. Siciliano, S. D., & Germida, J. J. (1999). Taxonomic diversity of bacteria associated with the roots of field-grown transgenic Brassica napus cv. Quest, compared to the non-transgenic B. napus cv. Excel and B. rapa cv. Parkland. FEMS Microbiology Ecology, 29(3), 263–272. Siciliano, S. D., Theoret, C. M., De Freitas, J. R., Hucl, P. J., & Germida, J. J. (1998). Differences in the microbial communities associated with the roots of different cultivars of canola and wheat. Canadian Journal of Microbiology, 44(9), 844–851. Sidhu, R. S., Hammond, B. G., Fuchs, R. L., Mutz, J. N., Holden, L. R., George, B., & Olson, T. (2000). Glyphosate-tolerant corn: The composition and feeding value of grain from glyphosate-tolerant corn is equivalent to that of conventional corn (Zea mays L.). Journal of Agricultural and Food Chemistry, 48(6), 2305–2312.

282

Rethinking Food and Agriculture

Stone, G. D. (2017). Dreading CRISPR: GMOs, honest brokers, and Mertonian transgressions. Geographical Review, 107(4), 584. Stone, G. D., & Flachs, A. (2018). The ox fall down: Path-breaking and technology treadmills in Indian cotton agriculture. The Journal of Peasant Studies, 45(7), 1272–1296. Stone, G. D., & Glover, D. (2011). Genetically modified crops and the ‘food crisis’: Discourse and material impacts. Development in Practice, 21(4–5), 509–516. Stone, G. D., & Glover, D. (2017). Disembedding grain: Golden rice, the green revolution, and heirloom seeds in the Philippines. Agriculture and Human Values, 34(1), 87–102. Stotzky, G. (2004). Persistence and biological activity in soil of the insecticidal proteins from bacillus thuringiensis, especially from transgenic plants. Plant and Soil, 266(1–2), 77–89. Tabashnik, B. E., & Carrière, Y. (2017). Surge in insect resistance to transgenic crops and prospects for sustainability. Nature Biotechnology, 35(10), 926. Thakur, A. K., Uphoff, N. T., & Stoop, W. A. (2016). Scientific underpinnings of the system of rice intensification (SRI): What is known so far? In Vol. 135. Advances in agronomy (pp. 147–179). Academic Press. Then, & Bauer-Panskus. (2018). Testbiotech on risk assessment for Golden Rice by FSANZ February 2018. Available at: https://www.testbiotech.org/en/publikationen. (Accessed June 2019). Thrupp, L. A. (2000). Linking agricultural biodiversity and food security: The valuable role of agrobiodiversity for sustainable agriculture. International Affairs, 76(2), 265–281. Tilman, D. (1998). The greening of the green revolution. Nature, 396(6708), 211. Tita, B. (2012). Genetically modified tires: Tire makers learn to cope with bioengineered crops. The Wall Street Journal. updated July31, 2012. Accessed 21 April 2019 https://www.wsj. com/articles/SB10000872396390444097904577539012713121388. Tooker, J. F. (2015). Sustainability of genetically engineered, insect-resistant crops: A view from the fringe. NABC Report 27 (p. 197). Tu, J., Zhang, G., Datta, K., Xu, C., He, Y., Zhang, Q., … Datta, S. K. (2000). Field performance of transgenic elite commercial hybrid rice expressing Bacillus thuringiensis δ-endotoxin. Nature Biotechnology, 18(10), 1101. Tuladhar, R., Yeu, Y., Piazza, J. T., Tan, Z., Clemenceau, J. R., Wu, X., … Hwang, T. H. (2019). CRISPR/Cas9-based mutagenesis frequently provokes on-target mRNA misregulation. bioRxiv, 583138. U.S. Environmental Protection Agency. (2010). Biopesticides registration action document— Cry1Ab and, Cry1F Bacillus thuringiensis (Bt) corn plant-incorporated protectants. Accessed Sept 2019 at: http://www3.epa.gov/pesticides/chem_search/reg_actions/pip/ cry1f-cry1ab-brad.pdf. USDA/APHIS. (2020). Petitions for determination of nonregulated status. Obtained from: https://www.aphis.usda.gov/aphis/ourfocus/biotechnology/permits-notifications-petitions/ petitions/petition-status. (Accessed 16 June 2019). Valenzuela, H. (2016). Agroecology: A global paradigm to challenge mainstream industrial agriculture. Horticulturae, 2(1), 2. van Bueren, E. L., Jones, S. S., Tamm, L., Murphy, K. M., Myers, J. R., Leifert, C., & Messmer, M. M. (2011). The need to breed crop varieties suitable for organic farming, using wheat, tomato and broccoli as examples: A review. NJAS-Wageningen Journal of Life Sciences, 58(3–4), 193–205. van Bueren, E. T. L., Struik, P. C., van Eekeren, N., & Nuijten, E. (2018). Towards resilience through systems-based plant breeding. A review. Agronomy for Sustainable Development, 38(5), 42.

Will gene-edited and other GM crops fail sustainable food systems?283

van der Hoeven, M. (2014). Bacillus thuringiensis toxins: Their mode of action and the potential interaction between them. Commissioned and published by: COGEM (The Netherlands Commission on Genetic Modification)-(CGM 2014–02). Van Eck, J. (2018). Genome editing and plant transformation of solanaceous food crops. Current Opinion in Biotechnology, 49, 35–41. Vanloqueren, G., & Baret, P. V. (2009). How agricultural research systems shape a technological regime that develops genetic engineering but locks out agroecological innovations. Research Policy, 38(6), 971–983. Venter, H. J., & Bøhn, T. (2016). Interactions between Bt crops and aquatic ecosystems: A review. Environmental Toxicology and Chemistry, 35(12), 2891–2902. Vickers, J. E., Grof, C. P. L., Bonnett, G. D., Jackson, P. A., & Morgan, T. E. (2005). Effects of tissue culture, biolistic transformation, and introduction of PPO and SPS gene constructs on performance of sugarcane clones in the field. Australian Journal of Agricultural Research, 56(1), 57–68. Villa, T. C. C., Maxted, N., Scholten, M., & Ford-Lloyd, B. (2005). Defining and identifying crop landraces. Plant Genetic Resources, 3(3), 373–384. Wakasa, K., Hasegawa, H., Nemoto, H., Matsuda, F., Miyazawa, H., Tozawa, Y., … Miyagawa, H. (2006). High-level tryptophan accumulation in seeds of transgenic rice and its limited effects on agronomic traits and seed metabolite profile. Journal of Experimental Botany, 57(12), 3069–3078. Waltz, E. (2009). Under wraps. Nature Biotechnology, 27(10), 880–882. Waltz, E. (2015). First stress-tolerant soybean gets go-ahead in Argentina. Nature Biotechnology, 33(7), 682–683. Waltz, E. (2016). CRISPR-edited crops free to enter market, skip regulation. Nature Biotechnology, 34, 582. Waltz, E. (2018). With a free pass, CRISPR-edited plants reach market in record time. Nature Biotechnology, 36, 6–7. Wang, F., Ye, C., Zhu, L., Nie, L., Cui, K., Peng, S., … Huang, J. (2012). Yield differences between Bt transgenic rice lines and their non-Bt counterparts, and its possible mechanism. Field Crops Research, 126, 8–15. Wei-xiang, W., Qing-fu, Y., Hang, M., Xue-jun, D., & Wen-ming, J. (2004). Bt-transgenic rice straw affects the culturable microbiota and dehydrogenase and phosphatase activities in a flooded paddy soil. Soil Biology and Biochemistry, 36(2), 289–295. Wickson, F., Binimelis, R., & Herrero, A. (2016). Should organic agriculture maintain its opposition to GM? New techniques writing the same old story. Sustainability, 8(11), 1105. Williams, G. M., Kroes, R., & Munro, I. C. (2000). Safety evaluation and risk assessment of the herbicide roundup and its active ingredient, glyphosate, for humans. Regulatory Toxicology and Pharmacology, 31(2), 117–165. Wilson, A., Latham, J., & Steinbrecher, R. (2004). Genome scrambling–myth or reality. In Transformation-induced mutations in transgenic crop plants. Brighton, UK: EcoNexus. Wilson, A. K., Latham, J. R., & Steinbrecher, R. A. (2006). Transformation-induced mutations in transgenic plants: Analysis and biosafety implications. Biotechnology and Genetic Engineering Reviews, 23(1), 209–238. Wolfenbarger, L. L., Naranjo, S. E., Lundgren, J. G., Bitzer, R. J., & Watrud, L. S. (2008). Bt crop effects on functional guilds of non-target arthropods: A meta-analysis. PLoS One, 3(5), e2118. Wu, D. X., Shu, Q. Y., Wang, Z. H., Cui, H. R., & Xia, Y. W. (2002). Quality variations in transgenic rice with a synthetic cry1Ab gene from Bacillus thuringiensis. Plant Breeding, 121(3), 198–202.

284

Rethinking Food and Agriculture

Xu, D., Collins, G. B., Hunt, A. G., & Nielsen, M. T. (1999). Agronomic performance of transgenic burley tobaccos expressing TVMV or AMV coat protein genes with and without virus challenges. Crop Science, 39(4), 1195–1202. Yan, F., Bengtsson, M., Anderson, P., Ansebo, L., Xu, C., & Witzgall, P. (2004). Antennal response of cotton bollworm (Heliocoverpa armigera) to volatiles in transgenic Bt cotton. Journal of Applied Entomology, 128(5), 354–357. Yang, P., Iles, M., Yan, S., & Jolliffe, F. (2005). Farmers’ knowledge, perceptions and practices in transgenic Bt cotton in small producer systems in Northern China. Crop Protection, 24(3), 229–239. Ye, X., Al-Babili, S., Klöti, A., Zhang, J., Lucca, P., Beyer, P., & Potrykus, I. (2000). Engineering the provitamin A (β-carotene) biosynthetic pathway into (carotenoid-free) rice endosperm. Science, 287(5451), 303–305. Zaidi, S. S. E. A., Vanderschuren, H., Qaim, M., Mahfouz, M. M., Kohli, A., Mansoor, S., & Tester, M. (2019). New plant breeding technologies for food security. Science, 363(6434), 1390–1391. Zeller, S. L., Kalinina, O., Brunner, S., Keller, B., & Schmid, B. (2010). Transgene× environment interactions in genetically modified wheat. PLoS One, 5(7), e11405. Zhang, C., Postma, J. A., York, L. M., & Lynch, J. P. (2014). Root foraging elicits niche complementarity-­dependent yield advantage in the ancient ‘three sisters’(maize/bean/ squash) polyculture. Annals of Botany, 114(8), 1719–1733. Zhang, H., Dong, H., Li, W., Sun, Y., Chen, S., & Kong, X. (2009). Increased glycine betaine synthesis and salinity tolerance in AhCMO transgenic cotton lines. Molecular Breeding, 23(2), 289–298. Zolla, L., Rinalducci, S., Antonioli, P., & Righetti, P. G. (2008). Proteomics as a complementary tool for identifying unintended side effects occurring in transgenic maize seeds as a result of genetic modifications. Journal of Proteome Research, 7(5), 1850–1861.

Sustaining agricultural biodiversity and heterogeneous seeds

14

Patrick Mulvany Centre for Agroecology, Water and Resilience, Coventry University, Coventry, United Kingdom

14.1 Agricultural biodiversity is intentionally heterogeneous The ways in which peasants and other small-scale food providers interact with other living beings, and the ecosystems and environment in which they live, are at the root of how they perceive, conserve, use and dynamically manage biodiversity. FAO (2020)

Agricultural biodiversity underpins the food system and also provides fiber, feed, fuel, pharmaceuticals, and materials for shelter; it is a vital subset of biodiversity that includes the seeds, breeds, and ecosystems within which food and other goods are grown and harvested by people. Agricultural biodiversity evolved through human interaction and coevolution with other living beings in the biosphere. Intentionally, and through innovative interactions, agricultural biodiversity, developed by women and men peasant farmers and gardeners, herders and pastoralists, fishers, forest dwellers, and other small-scale food providers, is very heterogeneous. That is to say agricultural biodiversity is rich in inherent diversity both between and within species and in the ecosystems that support production. Over millennia, through these interactions and the dynamic management of thousands of species of plants and animals, a very wide range of agricultural biodiversity evolved. Through the spread and exchange of seeds, plants, and livestock breeds, and associated knowledge, within and between communities, countries and continents, agricultural biodiversity was enhanced in, and adapted to, local communities’ needs and their “managed” ecosystems across the world. The dissemination of agricultural biodiversity was accelerated in the past 500 years, through the mass movement of peoples across the world, caused especially by colonization, migration, and slavery. As they moved, they took with them their plants and animals and adapted them to local productive environments and biodiverse ecosystems across all continents. In every part of the world, people’s knowledge and skills developed the high levels of agricultural biodiversity needed to sustain the heterogeneous seeds, breeds, and resilient ­agroecological production systems essential for an equitable food system that can confront the climate, biodiversity, and nutrition crises. Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00014-1 © 2021 Elsevier Inc. All rights reserved.

286

Rethinking Food and Agriculture

14.1.1 What agricultural biodiversity includes? Agricultural biodiversity encompasses the variety and variability of animals, plants and micro-organisms which are necessary to sustain key functions of the agroecosystem, its structure and processes for, and in support of, food production and food security. FAO (1998) All food systems depend on biodiversity; how this biodiversity is nurtured and managed determines the security of future food supplies and the sustainability of productive ecosystems. Vandermeer, Lawrence, Symstad, and Hobbie (2002, pp. 221–236)

Agricultural biodiversity includes the variety and variability of animals, plants, and microorganisms that are used directly or indirectly for food and agriculture, including crops, livestock, forestry, and fisheries. It also includes the diversity of associated nonharvested species that directly support production above and below ground and in waters (e.g., soil and water microorganisms, predators, pollinators), and those in the wider environment that support agroecosystems (agricultural, pastoral, forest, and aquatic) as well as the diversity of the agroecosystems themselves. Most agricultural biodiversity has evolved through the nurturing or deliberate domestication of useful varieties and breeds of uncultivated and untamed “wild” species over the past 10–20 millennia, since the “dawn of agriculture” (Fowler & Mooney, 1990). The origin of some species commonly used for food today predate this. Perennial species of plants, such as walnuts and Brazil nuts, for example, became established long before human agricultural activity, though, subsequently, humans have influenced the cultivars that are currently grown through selection and grafting. Agricultural biodiversity evolved with people in most of the world’s terrestrial ecosystems that they manage: “depending on definition, as much as 90% of the terrestrial surface of the earth is estimated to be maintained in some sort of managed state, usually forestry or agriculture and this very abundance should be sufficient to place managed ecosystems at the centre of the analysis of biodiversity and ecosystem function” (Vandermeer et al., 2002). Worldwide, flowering plant diversity is estimated to be around 248,000 species (Judd, 2002). Within these there are between 30,000 and 50,000 edible plant species of which about 7000 are cultivated, and, for these, millions of varieties have been bred by farmers. Between 35 and 40 animal species have been domesticated and around 7600 distinct breeds of livestock have been selected and developed by herders, pastoralists, and other livestock keepers. In tandem with this, many fodder and other livestock feed species have been developed for their diets. Forest dwellers have selected, developed, and nurtured thousands of tree and other species, including many subspecies and varieties, some of which support agroecosystem functions (Fielder, Landis, & Wratten, 2008, pp. 254–271). Many species of edible fungi, insects, and other invertebrates are also harvested by food providers and they cultivate some of them

Sustaining agricultural biodiversity and heterogeneous seeds 287

for consumption. It is estimated that fishers harvest nearly 2000 aquatic species or species groups and they nurture and use many other aquatic plant species for consumption and to sustain the fishery. Agricultural biodiversity, thus, includes not only the diversity and heterogeneity of plant species, varieties, livestock, and aquatic breeds directly used by humans, but also fodder crops and other feeds consumed by livestock and aquatic organisms. Agricultural biodiversity also encompasses the “associated biodiversity” of the species of animals, plants, and microorganisms at the genetic, species, and ecosystem levels that sustain structures, functions, and processes in and around production systems. This associated biodiversity comprises “the myriad components of biodiversity that support food and agricultural production by providing services such as pollination, pest control, soil formation and maintenance, carbon sequestration, purification and regulation of water supplies, reduction of disaster threats, and the provision of habitat for other beneficial species” (FAO, 2019). Of special significance among these are the uncultivated and “wild” relatives of species of social and economic importance that provide a genetic reservoir of diverse genes for the species (Hunter, Heywood, & Maxted, 2011). And, beyond species diversity per se, and the diversity of agroecosystems, which have spatial, temporal, and scale dimensions (FAO, 1998), agricultural biodiversity also includes biocultural and spiritual elements (Pimbert, 2006). Hence, the term is most commonly described in English as agricultural biodiversity, rather than the more reductionist term agrobiodiversity, to emphasize the cultural dimension of this vital subset of biodiversity. The scope of agricultural biodiversity is very broad—as the Food and Agricultural Organization (FAO) has repeated in its landmark assessment of the State of the World Biodiversity for Food and Agriculture, “it is difficult to establish definite boundaries” to agricultural biodiversity (FAO, 2019)—and it is recognized as hugely important to society (CBD, 2008). It is highly diverse, complex, and dynamic, being determined by a matrix of human factors and ecological feedback loops, in addition to underlying natural conditions (FAO, 1997b). By its nature, agricultural biodiversity is heterogeneous, a characteristic intentionally enhanced through human’s dynamic management of varieties, breeds, species, and ecosystems.

14.2 Distinctive features of agricultural biodiversity The distinctive features of agricultural biodiversity enable small-scale food providers to dynamically manage locally biodiverse and agroecological production systems that produce food of desired qualities and taste for local nutritional, social, and economic needs, and to do so in ways that increase the reliability of production and minimize risks. Within these production systems they have developed the agricultural biodiversity of their heterogeneous crops and livestock suited to varied environments. For ­example, crop varieties and domestic animal breeds were developed that could

288

Rethinking Food and Agriculture

withstand drought or floods, thrive on poor or rich soils, and resist pests and diseases. Likewise, а rich aquatic biodiversity, nurtured by local fishers, sustains local fish ­populations in inland waters and coastal fisheries. This dynamic management of agricultural biodiversity by biodiversity-conserving, small-scale food providers underpins their agroecological practices in the farms, gardens, rangelands, and inland and coastal waters from which they produce food primarily for localized peasant food webs. This contrasts dramatically with the reducing biodiversity and poor resilience of industrial food chains. As compared with wider biodiversity “the special nature of agricultural biodiversity, its distinctive features, and problems needing distinctive solutions” was recognized by the Convention on Biological Diversity (CBD) in its early years (CBD, 1994), confirming the need for globally agreed management and governance measures that are distinct from those governing the preservation of wider biodiversity through, for example, Protected Areas governance measures. These measures recognize that, for example: people have a key role in sustaining and dynamically managing agricultural biodiversity; many components of agricultural biodiversity, such as the diversity of the seeds of food crops, would not exist without this dynamic management and associated indigenous and local knowledge and culture; intraspecific diversity and intravarietal diversity within productive ecosystems are at least as important as diversity between species; access to, and control over, components of agricultural biodiversity can be restricted by legal and technological means thereby potentially limiting its availability; and that many food production systems are based on species introduced from other communities, countries, and continents, creating great interdependence on agricultural biodiversity across the world.

Small-scale food providers’ perspectives on the importance of agricultural biodiversity Biodiversity has as a fundamental base the recognition of human diversity, the acceptance that we are different and that every people and each individual has the freedom to think and to be. Seen in this way, biodiversity is not only flora, fauna, earth, water and ecosystems; it is also cultures, systems of production, human and economic relations, forms of government; in essence it is freedom (Via Campesina, 2000). ●



Agricultural biodiversity is enhanced by and supports agroecological production: it is an essential pillar of peasant strategies for survival and autonomy through reducing costs and risks. Peasant agroecological production systems have cultivated, sustained, and developed millions of heterogeneous varieties of crops and trees, breeds of livestock, and diverse aquatic organisms over millennia, and throughout the world. These nurture healthy populations of pollinators, pest predators, soil, and aquatic organisms above and below ground and in waters. Agricultural biodiversity is intertwined with our knowledge: it is more than the diversity of genetic resources, species, and ecosystems. It essentially includes the knowledge that led to its development and for its use. This knowledge is embedded in a dynamic web of relations between human beings and nature, continuously responding to new problems and finding new solutions.

Sustaining agricultural biodiversity and heterogeneous seeds 289









It makes the environment more resilient: environments rich in agricultural biodiversity are ecologically resilient to climate change and other threats, and also deliver other benefits to the environment and people, such as improved soil water retention, less soil erosion, increased soil biodiversity, improved pollination, as well as clean air and water. It improves the health of people and the planet: agricultural biodiversity is essential to human survival and health; when this biodiversity is diminished, disequilibrium results, which threatens health, both of humans and of nature. It is the product of their knowledge and collective rights: the successful enhancement of agricultural biodiversity—through exchanges within and between small-scale producers, countries, and continents—depends on peasants’ knowledge and collective rights of access to and control over territories, waters, seeds, and biodiversity. Its enhancement requires the freedom for peasants to choose the social system, the agrarian system, and culture that value it in a holistic sense, in the face of economic “values” imposed by a “free” market that destroys peasants’ seeds, biodiversity, and associated cultural freedoms. Adapted from IPC (2016).

14.3 Threats to agricultural biodiversity The negative impact on agricultural biodiversity from industrial agriculture, intensive livestock production, and large-scale mechanized fisheries is summarized in many international assessments that have been completed this century, from the International Assessment of Agricultural Knowledge, Science and Technology for Development (IAASTD) to FAO’s report on the State of the World’s Biodiversity for Food and Agriculture (FAO, 2019; IAASTD, 2009). Through land use change, destructive and unsustainable management of ecosystems and “downstream” pollution of more biodiverse production systems, industrial monocultures are the main cause of the loss of biodiversity. Industrial production and the value chains it supplies with their associated transport, processing, storage, and retailing systems are also principal drivers of climate change, through associated greenhouse gas emissions, causing further stresses for biodiversity (GRAIN & La Via Campesina, 2014). The loss of biodiversity in agroecosystems brings devastating consequences … simplified, human constructed agroecosystems may be unable to maintain their structure … [with] the accelerated loss of resilience and diversity and the erosion, salinization or decline in the fertility of soils. Egziabher (2002)

The impact of these industrial production systems on agricultural biodiversity is even more marked. Over the past 70 years, in the wake of 20th century global conflicts and subsequent instability in food supplies, international efforts to increase production of staple grains and commodities have been further intensified. These efforts include the consolidation, intensification, and simplification of peasant systems (Green Revolution ­technologies) with the expansion of genetically uniform monocultures displacing production using biodiverse peasant varieties. In addition, with the global

290

Rethinking Food and Agriculture

spread of the ­industrialization of agriculture and livestock production, including largescale land use changes from forests to plantations and cropping, and the expansion of industrial livestock production and large-scale fisheries, agricultural biodiversity has been further eroded. Causes of loss of agricultural biodiversity The social movements of smaller-scale food providers have summarized their views on the principal causes of the loss of agricultural biodiversity: The industrial model of production and consumption is rapidly eroding rural societies that manage agricultural biodiversity. The industrial model of production also displaces peasant varieties and breeds through using genetically uniform, and, increasingly, genetically modified, monocultures of crops, livestock, and fish, while locking up diversity in gene banks. Land grabs and ocean/water grabs extend the area under this model of production. Intensive use of pesticides, herbicides, and chemical fertilizers further reduces agricultural biodiversity and ecosystem functions. Climate change, exacerbated by this model, is putting new pressures on the local diversity of crops and livestock as weather patterns change, and new pests and diseases proliferate. Consequent disaster relief efforts distributing inappropriate, often industrial, seeds and livestock breeds undermine local agricultural biodiversity. Industrial research systems for this model devalue and erode peasant and indigenous knowledge, local research capacities, and the multitude of local innovation systems that foster agricultural biodiversity. Monopolies, favored by this model, control industrial seed, agrochemical, and industrial commodity markets and value chains, and this jeopardizes freedom for peasants to control, access, and use agricultural biodiversity. Intellectual property rights (IPRs) (sometimes dubbed “industrial” property rights by peasant organizations because they defend the interests of industry) and other laws that protect seed monopolies stimulate the widespread use of industrial varieties and can also criminalize peasant producers who develop, use, share, exchange, and sell their heterogeneous seeds. ●















Adapted from IPC (2016).

The industrial model of production favors the use of crop varieties, livestock, and aquatic breeds that respond to high applications of agrochemicals and intensive feeds, and the simplification of ecosystems. Its impacts in rural territories across the world include not only the rapid spread of monocultures but also massive increases in the use of associated pesticides and herbicides, resource consolidation, and the exodus of producers. The industrial model of production also produces excessive waste; it is built on an economy of surplus production of commodities rather than on the basis of ecological sustainability and realizing food sovereignty and the Right to Food. This model underlies the dysfunctionality of our food system (Mulvany & Ensor, 2011; van der Ploeg, 2009). An alternative view, as Eric Holt-Giménez has often pointed out, is that the system is not dysfunctional, but rather functioning as intended; it was never meant to provide good food for all, or preserve the environment, but to serve vested interests (Holt-Giménez, 2017).

Sustaining agricultural biodiversity and heterogeneous seeds 291

14.4 Few crops “Feed the World”? While industrial production systems per se undermine agricultural biodiversity for all the reasons cited previously, and more, a further determining factor is that, for complex reasons, probably related to power and control, human civilizations have selected relatively few crops for most of their food. By focusing on these few species, it can skew the discourse on sustaining agricultural biodiversity to the exclusion of concern about the diversity of the very many other species that are important for a sustainable food system. Estimates of the numbers of “how many crops feed the world” often build on the work of the Prescott-Allens (Prescott‐Allen & Prescott‐Allen, 1990). Their conclusion was that most food in the global food system comes from 82 species commodities as they describe them (i.e., species as well as genera, such as the cabbage family Brassicaceae) consisting in total of 103 species. These, they calculate, contribute 90% of national per capita supplies of food plants. Subsequently, Colin Khoury and colleagues concluded, based on the analysis of FAO data in 2014, that the number is slightly fewer, i.e., 94 species (Khoury et al., 2014). In answer to his own question, “So how many crops feed the world anyway?,” Guarino (2014), in the Agricultural Biodiversity Weblog, summarizes the state of knowledge and more or less concurs with the estimate of Khoury but adds Quinoa—increasing the number to 95 species— and reflects that precision about the exact numbers is not particularly important. The number is indicative of the limited number of species that dominate the global (especially industrial) food system.

Industrial Food Chains use few crop species and few varieties >7000

other species used in local food webs are under-reported?

4 species Maize, Rice, Wheat and Potatoes provide 60% industrial food

4species

+

83species +

95 species provide 90% industrial food

8species 12 species provide 75% industrial food

Significant Under-reporting of the Agricultural Biodiversity in Local Food Webs that provide most people with food?

Fig. 14.1  Industrial food chains use few crop species.

292

Rethinking Food and Agriculture

Fig. 14.1 shows the number of crops used directly for food, in part or wholly, which dominate the industrial food system. It is based on an interpretation of the data recorded by FAO for 2018. There seems to be general consensus that only three species—maize/corn, rice, and wheat— provide about half the world’s plant-derived calories (Tutwiler, Bailey, Attwood, Remans, & Ramirez, 2017), and, with potatoes, the proportion increases to about 60% of the formally recorded food system. The choice of the next eight species that, with maize/corn, rice, wheat, and potatoes, make up 75% of the industrial food supply, varies, depending on the author. They may include plantains and bananas, sweet potatoes, yams, cassava, and other roots, cereals such as barley, oats, rye/ triticale, sorghum/millets, chickpeas, and other pulses. The species in the additional 83 crops that provide 90% of food include some that might be suitable substitutes in the foregoing list of eight species, e.g., several vegetables and fruits that have many varieties that are important locally, and a few of these varieties are also globally dominant in the industrial food system such as tomatoes, mangoes, and apples. Also, the list includes many vegetable species, such as those in the Brassica (cabbage family) and Allium (onion/garlic family) genera, common fruits, and many legumes and pulses. In addition to these 95 species there are some 7000 species known to be used by people for their food. The prevalence and importance of these is probably underreported. What’s deliberately missing in this figure are industrial commodity crops, such as oil palm and other oil crops, soya, sugar cane, and sugar beet, which are traded as comestible commodities mainly for food processing and livestock feed, although on a smaller scale they may be used directly in local food webs.

Historically, human civilizations have selected relatively few species for their main foods. It is interesting to note that the four crops (maize/corn, rice, wheat, and potatoes), which make up 60% of the industrial food system, originated in regions that gave rise to significant empires; the Olmec, Aztecs, and Mayans developed their empires in MesoAmerica (today’s Mexico and Central America), the region where maize/corn was selectively developed from the grass Teosinte; rice developed in SE Asia where dynasties ruled for thousands of years; the Babylonian empires of the “fertile crescent” in West Asia arose in the region that developed wheat and other cereals; and in the Andean region, home to the Inca empire, potatoes were selected as an important food crop from the many roots and tubers that developed in the heterogeneous ecosystems on the slopes of the Andes. The most important aspect of the quantification of the number of food crops that “feed the world” to this discussion about agricultural biodiversity is the recognition that, in reality, humans regularly grow and harvest food from more than 7000 species, and that there are up to another 70,000 species known to have edible parts (Tutwiler et al., 2017). It is therefore plausible to suggest that there is underreporting of the wider number of species known to be used for food by people across the world. The few dominant species in the industrial food system, whose seed is increasingly controlled by a few corporations and whose production, trade, processing, and retail are captured in official records, bias attention in official statistics to these crops and, hence, in a self-­ reinforcing process also create a bias in food system presentation and planning. Though a limited number of species are used in the industrial food system, the security of the world’s food supplies was attained through the use of a multiplicity

Sustaining agricultural biodiversity and heterogeneous seeds 293

of crop varieties and livestock breeds. Large areas of single species can be relatively resilient and maintain productivity if there were significant varietal diversity within the production system. Millions of diverse varieties of the commonest crops have been developed by farmers over millennia as they selected seeds for various traits over many generations and when they adapted crops to new environments. For example, more than 100,000 varieties of rice are known to have been developed in India. In the in  vitro collections of the Peru-headquartered International Potato Centre there are more than 4000 varieties of potatoes and in the Global Seed Vault in Svalbard, a million accessions of diverse seed varieties are stored (Asdal & Guarino, 2018). While having the backup of diversity ex situ in gene banks provides some insurance, it is the in situ intraspecific and intravarietal diversity on-farm within a field that allows crops to evolve (Mulvany, 2017). This can compensate somewhat for limited species diversity in production systems. Tolerance of the different varieties of a single crop species within a field to varying weather patterns, rain/drought, heat/frost, changing soil conditions, diseases, insect attacks, and so on, can produce compensatory growth by one variety in the event that another variety suffers. Such has been the experience of peasant farmers in their biodiverse plots of, for example, potatoes. In Peru, peasant farmers in the Andes grow potato mixtures of up to 50 varieties in a “Challo” that, overall, can resist variable weather conditions and always provide consistent production from season to season. Now, it is this biodiversity—the number of varieties of each species in current production systems—that is fast disappearing. Assessments conclude that for some crops, more than 90% of varieties are no longer being grown regularly on-farm and the number of livestock breeds is also diminishing (FAO, 1997a, 2007). The single greatest cause of the loss of varietal diversity is the result of their substitution, on a very large scale, by relatively few industrial varieties (FAO, 1997a). This rapid erosion of the intraspecific variety within each species, i.e., the small number of varieties that dominate production, is of equal if not greater concern than the limited number of species grown. Genetically uniform monocultures of single varieties are easily susceptible to the effects of climate stresses and pests and diseases. Were this erosion of agricultural biodiversity to continue and production of the few industrial crops that dominate the industrial food system to become dependent on a limited number of genetically uniform varieties, the results could mirror, on a global scale, the impacts of the Great Irish Famine of the 1840s. This was caused, in part, because of failure of the staple potato crop; only two varieties of potato were being planted in the whole country at that time. They turned out to be susceptible to late blight and widespread crop failure resulted. So, it is not only important to recognize that few crops are dominant in the industrial food system but, more significantly, it is the diminishing number of varieties of crops and breeds of livestock that are currently grown and raised in fields and pastures that should be of greatest concern. However, the exact number is neither realistically knowable nor particularly relevant to the wider discourse about the imperative to increase the agricultural biodiversity of many species in all industrial agricultural and food production systems. Extensive monocultures of genetically uniform crops are damaging to agricultural biodiversity, whatever species is grown.

294

Rethinking Food and Agriculture

14.5 International governance of agricultural biodiversity and seeds The governance of agricultural biodiversity has great implications for the availability and use of the heterogeneous seeds needed for responsible agriculture. A number of international institutions and agreements govern laws, regulations, norms, commercial contracts, and technologies that impact on the use, ownership, sale and access to, and benefits derived from seeds (Tansey & Rajotte, 2008). Although the governance should cover all types of seeds and agricultural biodiversity, it overwhelmingly favors proprietary industrial seeds of relatively few species of plants. The international governance architecture for agricultural biodiversity and seeds is summarized in Fig. 14.2 and is dominated by four institutions: ●







The FAO, its Commission on Genetic Resources for Food and Agriculture (CGRFA), and International Treaty on Plant Genetic Resources for Food and Agriculture—commonly known as the International Seed Treaty; The CBD and its protocols; The World Intellectual Property Organization (WIPO) with its hosted convention—the Union for the Protection of New Varieties of Plants (UPOV in its French acronym); and The World Trade Organization (WTO) with its Trade-Related aspects of Intellectual Property Rights (TRIPs).

Internaonal Governance of Agricultural Biodiversity & Seeds affecng laws, regulaons, norms, commercial contracts and technologies

Other Instuons that support Industrial Seeds • IPPC /Seed health • CODEX Alimentarius /Food safety • OECD /Seed standards

UN UN CBD FAO / CGRFA /Nagoya & /Internaonal Seed Treaty Agricultural Cartagena IT PGRFA Biodiversity Protocols 3 2 MLS UN WIPO Patents/PVP /UPOV

SEEDS 1 4 TRIPS 27.3(b) /TM/GI SPS

ABS

WTO

Peasant Seed Systems are “collateral damage” in the internaonal governance of seeds?

Norms that could support Peasant Seeds Peasants’ Rights 1. Farmers’ Rights IT Arcle 9 / Rights of Peasants UNDROP 2. Indigenous Peoples Rights CBD Arcle 8j and UNDRIP Other Approaches 3. Sustainable Use of heterogeneous seeds, Agroecology /Food Sovereignty 4. Sui Generis PVP

Fig. 14.2  International governance of agricultural biodiversity and seeds.

Sustaining agricultural biodiversity and heterogeneous seeds 295

CBD—the Convention on Biological Diversity and its protocols: Cartagena Protocol on Biosafety (CPB), which governs the transboundary movement of GMOs (actually termed Living Modified Organisms in the text of the protocol), and the Nagoya Protocol (NP) on Access and Benefit Sharing (ABS). FAO—Food and Agriculture Organization of the United Nations, including the International Treaty on Plant Genetic Resources for Food and Agriculture (IT PGRFA)—commonly known as the International Seed Treaty—which includes, in Article 9, legally binding Farmers’ Rights; and the Commission on Genetic Resources for Food and Agriculture (CGRFA), which prepares regular reports and Global Action Plans on the state of the world’s genetic resources for food and agriculture (e.g., FAO, 1997a) and the landmark report on the State of the World’s Biodiversity for Food and Agriculture (FAO, 2019). WIPO—World Intellectual Property Organization and its hosted convention—the Union for the Protection of New Varieties of Plants (UPOV in its French acronym), which provides a TRIPs-compatible sui generis mechanism for providing intellectual property protection for commercial plant varieties. In its 1991 version (UPOV 91) the protection was strengthened and extended to include commercial ornamental flower varieties (required to access, for example, the Amsterdam flower market). WTO—the World Trade Organization with its agreement on TRIPs, which includes specific clauses, e.g., 27.3(b), on mandatory PVP and international governance oversight of trade marks and Geographical Indications that can be applied to goods from a specific territory. The WTO also has an Agreement on the Application of Sanitary and Phytosanitary Measures, which governs, inter alia, the health and cleanliness of traded seeds. ABS—the CBD mechanism, enshrined in the NP for ABS from the use of species and varieties/ breeds that are under national sovereignty. MLS—the Multi-Lateral System for access and benefit sharing, under the auspices of the International Seed Treaty (IT PGRFA), that covers, as listed in Annex 1 to the Treaty, seeds of about 64 species of crops, most of which have genes from multicountry origins. This allows for the efficient exchange of seeds across national jurisdictions, an opportunity used mainly by CGIAR institutions. IPPC—the International Plant Protection Convention governs standards on the phytosanitary requirements, including certification, for the international movement of seeds. CODEX Alimentarius—has agreed guidelines for assessing the safety of GM plant materials, which should be mutually compatible and supportive of the CPB text on biosafety and environmental protection, and which could be used to facilitate the introduction of GM seeds to third countries. OECD—Organization for Economic Co-operation and Development’s Seed Schemes provide internationally recognized varietal certification of seed moving in international trade. These cover 204 agricultural and vegetable species in eight groups of species.

The seed technologies prioritized by this governance architecture include hybridization, genetically modified organisms (GMOs), new-GM technologies (e.g., synthetic biology, genome editing, and gene drives, among others, which are often referred to as GM 2.0), Digital Sequence Information that facilitates the mining of the genetic codes of organisms, and the so-called fourth agricultural revolution (sometimes called AgTech 4.0) that embraces these seed technologies as well as making use of Big Datadriven automation, robots, drones, and so forth using satellite and 5G mobile network communications.

296

Rethinking Food and Agriculture

The devastating effect of this governance architecture and its norms, driven by pressures to prioritize trade and industry interests over environmentally and socially framed governance of food, farming, and the environment, is support for the development, protection (through the application of, among other instruments, Plant Variety Protection (PVP) and Plant Breeders’ Rights (PBRs)), and widespread use of industrial seeds of relatively few varieties of very few species of crops. Although millions of peasant farmers use, collectively manage, and develop agricultural biodiversity, including peasant seeds, on a practical, daily basis in their biodiverse, agroecological production systems, which provide food for the majority of people in the world, the international governance architecture undermines rather than supports them. Agricultural biodiversity and biodiverse peasant seed systems are, it could be argued, the “collateral damage” of this governance architecture. However, there are niches within the structure that can mitigate this impact to some extent. In some of the instruments cited and in the various intersections between institutions (illustrated in Fig. 14.2), there are niches that can be exploited for the governance of agricultural biodiversity, heterogeneous seeds, and peasants’ rights, necessary for developing biodiverse agroecology in the framework of food sovereignty. Even FAO, governed by ministries of agriculture, livestock, forestry and fisheries, is starting to recognize the need to break open these heavily defended ministerial or sectoral “silos” and also make links with ministries of environment, nutrition, and so on, in its determination to address the existential crises in food and agriculture. In FAO’s landmark assessment of the State of the World’s Biodiversity of Food and Agriculture, the common threats to agricultural biodiversity across all sectors of the industrial food system are identified and there are also many references to biodiversity-friendly practices across ecosystems that could be prioritized. The intersection of CBD and FAO governance provides coordination for the governance of agricultural biodiversity by both institutions, facilitated by the adoption of Decision III/11 in 1996 by the Conference of the Parties to the CBD, which defines the relationship between biodiversity and agriculture in its Annex 1, and the subsequent joint program of work on agricultural biodiversity with FAO, based, in part, on Annex 3 of the Decision (CBD, 1996). This Programme of Work prioritizes projects concerning soil biodiversity, pollinators, and nutrition. Niches within this governance structure (numbered 1–4 in Fig. 14.2) also allow for the establishment of norms that could support the development, protection, and use of peasant seeds. 1. The legal recognition and implementation of Farmers’ Rights in Article 9 of the International Seed Treaty, though these are subject to national laws. However, international recognition is provided through the UN’s Human Rights Council the UN Declaration on the Rights of Peasants (UNDROP) and other people working in rural areas, adopted in 2018. This recognizes, internationally, peasants’ collective rights to the use of peasant seeds and agricultural biodiversity in Articles 19 and 20. 2. In the CBD’s Article 8j, the rights of indigenous and local communities, and their innovations, skills, and practices, are recognized. This is further supported by the UN’s Declaration on the Rights of Indigenous Peoples (UNDRIP), which recognizes, in Article 31, ­indigenous peoples’ right to maintain, control, protect, and develop their seeds and their right to

Sustaining agricultural biodiversity and heterogeneous seeds 297

­ aintain, control, protect, and develop their intellectual property over associated traditional m knowledge about their seeds. 3. Article 6 of the International Seed Treaty covers the sustainable use of seeds and other planting material. This could provide for the development and use of heterogeneous peasant seeds in agroecological systems that support food sovereignty. 4. WTO/TRIPs Article 27.3(b) has an option for a country to adopt a sui generis form of Plant Variety Protection that could recognize and protect peasant seeds. However, States, in implementing these Articles, usually opt for measures that will favor industrial seed development and use, including the adoption of UPOV 91.

While these niches can be exploited beneficially for sustaining agricultural biodiversity, it is not enough, in the face of the overwhelming bias toward supporting biodiversity-reducing practices, to respond to the catastrophic climate, biodiversity, and nutrition emergencies. Governments may express their concern and agree to UN Decisions in favor of sustaining agricultural biodiversity. For example, there are eight Decisions of the CBD on agricultural biodiversity. These Decisions, the 1996 Landmark Decision III/11 and IV/6, V/5, VI/5, VII/3, VIII/23, IX/1, X/34, include some 156 Operational Elements of which 82 are currently “Active” but are mostly not implemented. Countries, for example, Australia, may note the “need for a workable regime of governance for agricultural biodiversity” yet they “have been particularly concerned to maintain and extend the neoliberal project of market rule” and the dominance of trade concerns (Lockie, 2009). A radical change in governance priorities is urgently required to secure nutritious food supplies for future generations in ways that can restore biodiversity and mitigate climate change. There’s a need to turn the international governance architecture on its head and give priority to sustaining peasants’ heterogeneous seed systems and agricultural biodiversity developed in the framework of food sovereignty.

14.6 Peasant seeds: More numerous and heterogeneous Peasants’ seeds are a heritage of peoples in the service of humanity. Seed is life, the basis of global food production, essential for peasants to produce healthy and culturally appropriate food and crucial for consumers and citizens who seek to find healthy and diversified food. Seed is part of peasant culture and is our heritage, allowing us to resist, maintain our ancestral wisdom and defend our peasant identity. Via Campesina: Global Campaign “Peasant seeds a heritage of peoples in the service of humanity, a way to promote Food Sovereignty” (Via Campesina, 2018).

Peasant seeds are prolific in number. Estimates are that the number of varieties of peasant seeds is some 20-fold those of registered industrial seeds. Peasants have developed around 2,100,000 biodiverse varieties as compared with some 100,000 registered varieties of uniform seeds (etcGroup, 2017). Peasant seeds are also inherently biodiverse—heterogeneous in their intravarietal diversity. They have co-evolved with smaller-scale food providers over millennia, in specific locations within biodiverse agroecological systems, which also have a wide

298

Rethinking Food and Agriculture

range of agricultural biodiversity above and below ground. Through their careful selection of crop cultivars from heterogeneous populations of varieties and species, using cultural norms, peasants enhance not only the agricultural biodiversity of their seeds but also the biocultural diversity of peasant production. These seeds are resilient and adaptable, producing crops that are richer in micronutrients and are more ­nutrient-dense (Fanzo, Hunter, Borelli, & Mattei, 2013). Peasant seeds are commonly varieties that reproduce through open pollination, which facilitates adaptation of a variety to changing local needs and conditions. Careful selection of seeds for subsequent crops embeds evolving diversity. With plants that reproduce vegetatively, the careful selection and subsequent use of a wide variety of seed stock or clones taken from one harvest for the next planting will keep the diversity alive. For some vegetatively propagated plants, diversity may be enhanced through exchanges and from biodiverse collections of crop varieties. For example, for the “False Banana” Enset, which is an important staple crop grown in African highlands, diversity is enhanced by the selection of new clones from sacred groves in which Enset plants and their wild relatives are protected through local customs for this purpose. Peasant seed systems function through collective efforts that, by sharing and exchanging seeds and associated knowledge within and between communities, enhance the diversity of seeds planted. Thus the importance of collective, as opposed to individual, rights to these seeds. The recognition of the importance of collective rights by the international community, through the process facilitated by the UN Human Rights Council to develop the UNDROP, is an important step in securing peasant seed systems. While UPOV and related laws and norms recognize the IPRs of natural and legal persons, UNDROP recognizes new rights that are framed as collective human rights (Claeys, 2019). Peasant seed systems, adapted to territorial diversity, are governed by peasants themselves, respecting their collective rights. Many networks, campaigns, and processes, organized by peasant organizations and social movements and supporting civil society organizations (CSOs), promote peasant seed systems and the on-farm conservation of their varieties (Mulvany, 2008). In most countries there are seed and agricultural biodiversity networks, some also organized regionally, continentally, and internationally, which raise the profile of peasant seed systems and help defend them from the devastation caused by industrial agriculture (Shiva, Anilkumar, & Singh, 2019). But translating this support into robust governance systems respected by States has remained elusive. In some cases, CSO support has undermined the integrity of these systems by trying to fit them into the dominant governance architecture outlined earlier to reduce the opportunity for States to criminalize peasants for using, saving, exchanging, and selling their farm-saved seeds. Or, through recognizing that peasants and their seed systems are harmed by existing seed laws, CSOs argue, in good faith, that seeds should not be owned and controlled by the state or indeed any natural or legal person, and should therefore be liberated from such laws. The danger of this approach is that this could open peasant seed systems to unchecked biopiracy and to contamination from industrial seeds, including GM seeds, which would reduce the heterogeneity of biodiverse peasant seeds. A governance system that can support

Sustaining agricultural biodiversity and heterogeneous seeds 299

heterogeneous peasant seed systems, organized separately from the systems used to protect and promote homogeneous industrial seeds, is urgently required. As illustrated in Fig.  14.3, the number of plant varieties, including many ornamental plants, registered in the “formal” seed system is many fewer than the numerous varieties of peasant seeds, mostly of food crops. The “formal” sector varieties are protected by PBRs, PVP, and similar mechanisms, which have satisfied the varietal purity criteria set by UPOV of being novel, distinct, stable, and uniform—in UPOV 1991, or homogeneous—in UPOV 1978 (Helfer, 2004). This “formal” seed system is a product of, and supported by, the governance architecture outlined in the earlier section. In contrast, peasant seed systems, with their numerous biodiverse varieties, prioritize heterogeneity, adaptability, and inherent diversity in their varieties and populations, governed by collective peasant rights. The two systems are distinct and require separate governance arrangements. However, there are a number of initiatives to find common ground between them. For example, in FAO’s Voluntary Guide for National Seed Policy Formulation (FAO, 2015), “formal” and “informal” seed systems are described, inviting governments to consider including both in the same seed policies. Unequal power relationships would, however, be reflected in the implementation of such policies at the cost of peasant seed systems. At the interface between the two systems there are alternatively named seed systems that can be promoted by CSOs, researchers, and government agencies. These may tend to serve the interests of the formal or peasant seed systems depending how they are organized, with whom and with what purpose.

Peasant Seed Systems: more numerous and heterogeneous seeds 0.1m* DUS vars

“Formal” Seed System PBRs PVP

2.1m* biodiverse varieties

Peasant Seed Systems Collective Peasant Rights

• • • • • •

ALTERNATIVE SEED SYSTEMS e.g. Farmer-Managed Seed Systems (FMSS) Community Seed Registers Community Seed Banks (CSB) Heritage Seed Collections Open Source Seed Systems (OSSS) Seeds as “Commons”

ALTERNATIVE PLANT BREEDING e.g. • Participatory Varietal Selection (PVS) • Participatory Plant Breeding (PPB) • Evolutionary Plant Breeding (EPB)

* Source: etcGroup (2017) Who Will Feed Us?

Fig. 14.3  Peasant seed systems: more numerous and heterogeneous seeds.

300

Rethinking Food and Agriculture

These alternatives may well provide biodiverse seeds, saved on-farm, for local food production, for example, in farmer-managed seed systems (AFSA & GRAIN, 2018). But they might simply copy industrial seed systems by providing seeds for chemically dependent monocultures—with the only difference being that the seeds are “owned” by the farmer. For example, the farmer-managed seed systems may focus on the production of quality declared seed (FAO, 2006). This seed is often designed for monocultural production by small-scale or commercial farmers and does not necessarily have heterogeneity embedded in the seeds. In some programs, seeds are stored and shared in “community seed banks” of various types as described in “A Typology of Community Seed Banks” (Lewis & Mulvany, 1997). Other alternative approaches may focus on recording varieties and making seeds available, for example, through heritage seed collections assembled by organizations that can be shared with, and multiplied by, their membership (Evans, 2007); or through “open source” seed systems, which try to fit farmers’ seeds into a formalized registration system that allows free access to the seed (Kloppenburg, 2016). There is also the broader concept of seeds as “commons,” which would not permit the private ownership of seeds, making seeds more freely available (Frison, 2016). In any of these approaches, seed collections, community registers, and varietal registration must also be accompanied by recording the knowledge and biocultural heritage embedded in the seeds if they are to serve their purpose (Argumedo & Pimbert, 2006). Unless carefully designed with peasant farmers, who practice biodiverse peasant agroecology in the framework of food sovereignty, they may not enhance agricultural biodiversity and the intravarietal biodiversity of seeds required for agroecology and food sovereignty. These approaches may address the limitations of the formal seed systems in terms of “ownership” and the “availability” of what is often a wide range of varieties of seeds, but unless explicitly included in the approaches, may not tackle the need for the seeds themselves to be heterogeneous and have adaptability embedded in their genes. With the proposition of seed “improvement,” a number of alternative participatory plant breeding methods have been developed by scientists working with peasant farmers. These methods include: participatory varietal selection (Li-Bird, 2017), participatory plant breeding (Ceccarelli, 2015; Halewood, Deupmann, Sthapit, Vernooy, & Ceccarelli, 2007), and evolutionary plant breeding of “populations” of varieties, as developed by, for example, CENESTA in Iran (Rahmanian et al., 2016). These cover a spectrum of methods from developing what are essentially monocultural seeds adapted to local environments along a transition toward having adaptable “populations” of seed that are heterogeneous with the diversity built into the seeds. As the godfather of participatory plant breeding, Dr. Melaku Worede has often said, it is the function of scientists to enhance existing varieties, not replace them: “The foundation for regenerating traditional farming systems is to work with communities to revive their former seed diversity and related knowledge. In most communities in Africa it is the women who have the greatest knowledge about seed breeding and biodiversity. Once this foundation is laid, scientists can add a little support to further enhance genetic and crop diversity and thereby productivity and climate change resilience” (Gaia Foundation, 2015).

Sustaining agricultural biodiversity and heterogeneous seeds 301

To help clarify how a range of approaches used to increase the heterogeneity within and among crops can be more precisely described, Wolfe and Ceccarelli (2019) summarized the transition from the breeding of monocultural varieties to static, then dynamic, mixtures, to the evolutionary breeding of populations. They have attempted to cover the range of populations of cereals that use a variety of different names, such as old varieties, landraces, ancient grains, heritage varieties, heirloom varieties, and varietal mixtures being brought back into cultivation, providing greater clarity about what each represents in terms of their inherent biodiversity. In summary, these alternative plant breeding programs are useful for peasant seed systems if the seeds that are enhanced by such programs are heterogeneous with high levels of intravarietal diversity. Only in this way would they be appropriate for biodiverse agroecology and food sovereignty.

14.7 Prioritizing biodiverse and heterogeneous peasant agroecology An increase and strengthening of AKST toward agroecological sciences will contribute to addressing environmental issues while maintaining and increasing productivity. Formal, traditional and community-based AKST [Agricultural Knowledge, Science and Technology] needs to respond to increasing pressures on natural resources, such as reduced availability and worsening quality of water, degraded soils and landscapes, loss of biodiversity and agroecosystem function, degradation and loss of forest cover and degraded marine and inshore fisheries. Agricultural strategies will also need to include limiting emission of greenhouse gases and adapting to humaninduced climate change and increased variability. International Assessment of Agricultural Knowledge, Science and Technology for Development (IAASTD) Finding #7 (IAASTD, 2008) Peasant agroecology is synonymous with dignity and is based on peasant knowledge, peasant knowhow, and peasant life skills. Peasant agroecology has life itself at its core. It is a way of living based on solidarity, on practices that are respected, shared and transmitted, and on common values and principles, most notably human rights. By virtue of its philosophical, social, environmental, and economic dimensions, peasant agroecology incorporates all forms of ecological, organic and fairlytraded agriculture. It is the key to safeguarding human life and the planet, today and tomorrow. Nyéléni (2017)

Peasant seed systems underpin heterogeneous peasant agroecology developed in the framework of food sovereignty. As declared by social movements at the Nyéléni Forum for Agroecology in 2015, “the production practices of agroecology (such as intercropping, traditional fishing and mobile pastoralism, integrating crops, trees, ­livestock and fish, manuring, compost, local seeds and animal breeds, etc.) are based on ecological principles like building life in the soil, recycling nutrients, and the dynamic management of biodiversity” (Nyéléni, 2015). This was further defined by

302

Rethinking Food and Agriculture

West African farmers in the Nyéléni Peasant Agroecology Manifesto in 2017. This manifesto importantly defined “peasant agroecology,” in the context of human rights principles, as being based on peasant knowledge, practices, and skills. A key pillar in this manifesto is to place cultural value upon and safeguard agricultural biodiversity, peasant seeds, and local breeds by fostering biodiversity, including the agricultural biodiversity of crops and farm animals, recognizing that these are the mainstay of present and future life (Nyéléni, 2017). Peasant agroecology can deliver multiple benefits in ways that will mitigate the climate, biodiversity, and nutrition crises. Given the increasing requirement for nutritious foods for a growing population, there are significant advantages for improving productivity through transitions toward, or enhancements of existing, biodiverse, agroecological production systems underpinned by heterogeneous peasant seeds, livestock breeds, and wider agricultural biodiversity. Not only can productivity, measured in terms of food produced or people fed per unit area, be greater than in monocultural and agrochemical-dependent industrialized production systems (Badgley et al., 2007), but also there are benefits in terms of reduced soil loss, fewer greenhouse gas emissions, and less agrochemical use as well as greater adaptive capacity and resilience and enhanced ecosystem functioning in these biodiverse and heterogeneous systems. Fig.  14.4 attempts to illustrate these advantages of transitioning toward peasant agroecology, underpinned by heterogeneous seeds, and also to make the case that conversion from industrial systems toward the more sustainable agroecological systems must be carefully managed to avoid catastrophic decreases in production.

Heterogeneous Peasant Seeds and Breeds and wider Agricultural Biodiversity underpin Peasant Agroecology

C

Food Provision/Nutrition per unit Land / Water A = Industrial, simplified, high external input production + monocultural (inc.GM&GM2.0) industrial seeds and breeds B = initial productivity levels

C =Agroecological, biodiverse, low external input production + heterogeneous peasant seeds and breeds

A

B X

Low----HETEROGENEITY / AGRICULTURAL BIODIVERSITY-----High Low---------------ADAPTIVE CAPACITY / RESILIENCE -------------- High Low----------------------------FOOD SOVEREIGNTY------------------------High Low-----------------PEOPLE’S / LOCAL KNOWLEDGE-----------------High High----------------------------------SOIL LOSS---------------------------------Low High------------GHG EMISSIONS / AGROCHEMICAL USE------------Low High----------CORPORATE CONTROL / $$ INVESTMENTS----------Low

Fig. 14.4  Heterogeneous peasant seeds, livestock breeds, and wider agricultural biodiversity underpin agroecology.

Sustaining agricultural biodiversity and heterogeneous seeds 303

Point B represents a productivity level of food/nutrition per unit of land and/or water, which, through human endeavor, research, extension and investment, society or corporate interests would wish to increase. Point A represents levels of production using approaches that require more fossil fuels and industrial monocultural seeds suited to high levels of inputs of agrochemicals and genetic technologies. These approaches result in the simplification of production systems, often with significant soil loss, reducing biodiversity, adaptive capacity, and resilience. These systems depend on commercial and proprietary industrial technologies (including GM seeds and livestock, pesticides, and fertilizers) that are patentable and are controlled by agribusiness corporations. These systems replace labor with capital-intensive equipment and inputs. Point C represents levels of production and productivity per unit area and/or per unit of water that, especially at smaller scales, can be higher than those achieved by the high input, carbon-­ intensive practices represented by point A. Point X is on the dotted line that shows productivity declining rapidly if external inputs were to be suddenly removed from the industrial system to a level below that in the unchanged production system represented by Point B. The dotted line also indicates that it could take a long time for the system to recover from the high-input, biodiversity-, and soil-eroding technologies used to boost the productivity of industrial production before the ecosystems can once again build the more biodiverse and more productive systems represented by Point C.

The technologies used to achieve the higher yields illustrated in Fig. 14.4 by Point C incorporate more agricultural biodiversity in more complex, resilient, and heterogeneous agroecological systems that can have lower, zero, or negative carbon costs and use nonappropriable technologies—technologies that cannot be privatized and which provide maximum benefit to local food providers who develop and use the technologies. These systems underpin agroecology and support the realization of food sovereignty. These more appropriate technologies, including the knowledge and skills used in the enhancement of heterogeneous seeds and seed populations, require more people and local knowledge for their development and implementation. There is, however, no such thing as an a priori “optimum” level and mixture of agricultural biodiversity in these biodiverse agroecological systems; the desirable configuration is determined by prevailing local, natural, and—equally importantly—socioeconomic and cultural circumstances (Cromwell, Cooper, & Mulvany, 2001). Given the origins and nature of agricultural biodiversity, it can be argued, however, that the levels needed for optimally functioning ecosystem processes is the degree of inter- and intraspecific, and inter- and intravarietal, heterogeneity existing within biodiverse agroecosystems at all scales of production, in gardens, fields, pastures, ponds, forests, landscapes, watersheds, and coastal ecosystems. There is abundant local and indigenous knowledge and innovation on how to boost production from Point B to Point C using biodiverse, heterogeneous, agroecological, resilient, low-carbon production systems with significant adaptive capacity. At ­present, this is far from the reality. In excess of 95% of resources are spent on the simplification of production systems using high levels of external inputs in ­monocultures of industrial seeds, illustrated by the shift from point B to point A, rather than on the transformation to agroecology (Pimbert & Moeller, 2018).

304

Rethinking Food and Agriculture

However, there is an equally urgent need to transform industrial production systems at all scales—systems that are low in biodiversity, increase soil loss, and have poor adaptive capacity (IDDRI, 2016)—toward biodiverse agroecology (the dashed line in Fig. 14.4 from Point A to Point C). This needs to be done without catastrophically losing productivity if external inputs, such as industrial seeds, chemical fertilizers, herbicides, insecticides, and fungicides, were suddenly to be removed. Such a shock could cut productivity to levels below the initial production system, Point B, illustrated in the diagram by Point X on the dotted line. Recovery from this point could take a long time because of the need to rebuild fertility, biodiversity, and ecosystem health to achieve the productivity gains in the more biodiverse systems represented by Point C. The urgent societal and scientific challenge must therefore be to remove the barriers and “lock-ins” that prevent this transformation—in terms of both the enhancement of degraded small-scale production using peasant seeds and the shift from industrial monocultural production using industrial seeds toward biodiverse agroecology using heterogeneous seeds—being the priority for research, development, and practice. There are, however, severe impediments to this transformation, as identified by IPES-Food in their publication “From Uniformity to Diversity.” They identified some eight “lock-ins” that prevent the transformation of industrial monocultures to more biodiverse and sustainable food systems: path dependency on investments that require large-scale industrial production; export orientation; expectation of “cheap” food; compartmentalized thinking (in silos); short-term planning; “feed the world” narratives; perverse measures of success; and concentration of power in food and agriculture (IPES-Food, 2016). Thus in the context of the climate, biodiversity, and nutritional crises, scientific research and innovation should be focused on the conversion of these degraded and simplified industrial production systems to more ecologically complex, biodiverse, and heterogeneous systems, building on the findings of many studies and international assessments. There is a further urgent challenge in this conversion process. How is it possible to prevent the commodification of, and corporate control over, the resources (seeds/biodiversity, soils, water, carbon, etc.) and systems (tenure, markets, etc.)? These processes would undermine people’s collective rights to their seeds and the commons needed for the full realization of productivity and food sovereignty in the biodiverse and heterogeneous agroecological systems represented by Point C. There seems to be significant resistance by governments to the acceptance of the imperative of challenging corporate pressures that argue for keeping and strengthening the biodiversity-eroding systems described earlier. This presents a considerable political challenge. One possibility is to continue “business as usual” operated by similar powerful actors and corporate interests, which defend industrial systems and the proprietary research and innovation that prop them up. Alternatively, disinvest from these and heavily regulate these systems in favor of supporting radical changes in research, development, and production priorities toward more agroecological, biodiverse, heterogeneous, and resilient local food provision systems. The inescapable truth should prevail: that the emergencies the world currently faces can, in part, be resolved if the mainstream biodiverse agroecological food system with its heterogeneous seeds, served by small-scale food providers in localized food webs, replaces the damaging international food chains of the industrial food system.

Sustaining agricultural biodiversity and heterogeneous seeds 305

14.8 Three coalitions contesting control over agricultural biodiversity and seed systems The reason why society has not embraced the transformation of food systems toward the more productive, more biodiverse, and resilient agroecological systems underpinned by heterogeneous seeds is because of significantly unequal power between the coalitions that control agricultural biodiversity and seed systems. In most parts of the world, it is possible to identify three coalitions of interests, which are contesting control over agricultural biodiversity and seed systems. Although other terms may be used to describe them, these could be called the coalitions of: Ecological Food Provision; Green Revolution; and Agribusiness (Coupe, Ensor, & Mulvany, 2011). In Annex 14.1 relevant attributes of each coalition are listed. The Ecological Food Provision coalition is the coalition that would best sustain agricultural biodiversity and heterogeneous peasant seeds. It is rooted in social and political movements of women and men farmers and other small-scale food producers, who are the providers of food for the majority of people in the world. These food providers do this mainly by using farm-saved peasant seed to produce crops in ecologically resilient and biodiverse agroecological production systems that use few if any agrochemicals. This coalition has growing support from civil society, including citizens in urban areas. It has a powerful discourse on peasant agroecology, agricultural biodiversity, and locally adapted biodiverse peasant seeds of a wide range of crops, developed in the framework of food sovereignty, mainly for localized food webs. This coalition robustly defends farmers’ and peasants’ rights to rural territories, water, agricultural biodiversity, seeds, livestock breeds, food, and livelihoods. At the same time, this coalition has a powerful critique on the activities of the other policy coalitions, which are perceived as hastening the destruction of ecosystemic natural resources, agricultural biodiversity, and peasant seeds/breeds, with negative impacts on climate change and the biosphere and increasing the exclusion of farmers from biodiverse food production and, ultimately, their territories, allowing for the expansion of agribusiness interests. The proposition of this coalition is that all food systems, including those defended by the other coalitions, should strengthen or transition to this more resilient, sustainable, and equitable approach to food provision, as recommended by many UN and civil society processes over the past decade. The Green Revolution coalition has a negative impact on agricultural biodiversity and peasant seed systems. It is given legitimacy by many international institutions and donors, including philanthrocapitalists, which influence the governance of agricultural biodiversity and seed systems. It prioritizes investment and profit from the exploitation of land, resources, and labor, purportedly to realize food security, but ultimately to reduce peasant control over food provision. This coalition emphasizes the packaging of science and technology, including homogeneous seeds, fertilizers, and pesticides, for delivery to small-scale farmers. The approaches favored may be presented not only as “Green Revolution,” but also as, for example, “sustainable intensification,” “climate resilient technologies,” and “climate smart agriculture,” using technologies and systems for the so-called benefit of smallholder farmers.

306

Rethinking Food and Agriculture

But these approaches are mainly focused on the production of monocultures of staple, often genetically uniform, crops. This coalition includes a growing number of alliances through which public, private, and voluntary sector energies are marshaled, for example, the Alliance for a Green Revolution in Africa, often co-opting the language of the ecological small-scale food provision coalition, e.g., agroecology, in promotion of its cause. This coalition can facilitate further colonization of peasant production systems, their resources, and territories for the ultimate benefit of the state and/or agribusiness corporations. The Agribusiness coalition, represented by powerful corporations, aggressively promotes capital-intensive proprietary technologies and economies of scale in industrial commodity monocultures. It has devastating negative impacts on agricultural biodiversity and biodiverse seeds. Its economic and political power allows it to capture resources and can hasten the exit of smaller-scale farmers from their territories into agricultural labor or urban centers. Although the fortunes of this coalition have fluctuated across the world over the past 70 years, in the past decade, with the increase in “land and water grabs,” more resources are now under the control of this coalition. In this coalition there is increasing concentration of power in fewer corporations, most notably in the seed industry, supported by internationally recognized, restrictive legislation, and in the linked agrochemical, machinery, and food industries. It is supported by well-funded private and public research systems that produce proprietary technologies and processes suited for capital-intensive industrial production systems, which require less labor and use much less biodiversity. What is clear from this analysis is that to sustain agricultural biodiversity and biodiverse peasant seed systems, emphasis needs to be put on supporting the ecological food provision coalition and finding ways to reduce the power and control exerted by the other coalitions. The central defining policy of the ecological food provision coalition is food sovereignty, which provides the framework for peasant agroecology and heterogeneous peasant seeds. Launched by Via Campesina in 1996, this framework is now recognized by peasants across the world. The initial stimulus for launching the food sovereignty framework was the threat of the WTO’s control over agriculture and the legal requirement on its members not to discriminate between “like products” based on their method, location, or scale of production. Biodiversity-destroying industrial production, or biodiversity-enhancing peasant production, of a crop produces, in WTO terms, the same product. This was one of the reasons for launching the food sovereignty framework and why Via Campesina still campaigns to keep agriculture out of the WTO. In subsequent years the food sovereignty framework was developed to embrace wider dimensions of the food system, including the Right to Food, agroecological production, and control over resources and the governance of seeds. After the Forum for Food Sovereignty in 2002, held in Rome in parallel with FAO’s “World Food Summit—5 years later,” the food sovereignty movement was further enlarged and met at the landmark Nyéléni 2007: Forum for Food Sovereignty, held in Sélingué, Mali, in 2007. Five hundred representatives of peasants, pastoralists, fishers, forest dwellers, and other smaller-scale food providers met to identify what

Sustaining agricultural biodiversity and heterogeneous seeds 307

would support food sovereignty, what barriers existed to its realization, and how these could be overcome. The Nyéléni Declaration (Nyéléni, 2007a), an outcome of the forum, defined food sovereignty and, in the Synthesis Report of the forum (Nyéléni, 2007b), a table summarized six pillars of food sovereignty, identifying what it is for and what it is against (Table 14.1). Food sovereignty is strongly contested by dominant power, but it is widely supported by peasant organizations and by progressive academics and civil society organizations. For example, it formed the basis of a People’s Food Policy in England (People’s Food Policy, 2017). The term food sovereignty has even been incorporated in the legislation of some countries, including Bolivia, Egypt, Mali, Nepal, Senegal, and Venezuela, though full implementation of the policies embedded in the food sovereignty framework have yet to be realized in any nation-state. Food sovereignty is the framework that enables the guardians of agricultural biodiversity, the peasant women and men who provide the food for most people in the world, to ensure there is inherent heterogeneity embedded within the seeds, breeds, and agroecological systems that they nurture.

14.9 Heterogeneous seed and agricultural biodiversity— The basis for responsible agriculture and food systems As has been summarized earlier, the heterogeneity of seeds, livestock breeds, aquatic organisms, and other agricultural biodiversity, at genetic, species, and ecosystems levels within agroecological food production systems, is important for resilient, sustainable, and biodiverse food production. Enhancing heterogeneous seeds and livestock breeds on-farm, in gardens, and within fields, pastures, waters, and wherever else food is produced is vital for people and the planet. These seeds underpin and depend upon agroecology and food sovereignty. These heterogeneous seeds and agricultural biodiversity have provided not only food for the world over millennia but also peasants’ biodiverse agroecology, based on such seeds and a broad range of agricultural biodiversity, still provides food for most people in ways that protect biodiversity and mitigate climate change. Yet, the prevalence of heterogeneous seeds and agricultural biodiversity within production systems is seriously under threat from the rapid expansion of monocultural, chemically-­dependent industrial production systems using homogeneous seeds of relatively few crops and varieties and few livestock breeds. As the previous section summarizes, these industrial production systems are promoted by the Agribusiness coalition, which aggressively captures financial, territorial, and productive resources in pursuit of its ends. The interests of the Agribusiness coalition, and also the coalition that supports the Green Revolution, are supported by an inequitable governance architecture. This ­architecture discriminates against peasants working in the framework of food

308

Rethinking Food and Agriculture

Table 14.1  Food sovereignty framework. Nyéléni (2007a, 2007b): Forum for Food Sovereignty Definition of Food Sovereignty—from the Declaration of Nyéléni (Nyéléni, 2007a) Food sovereignty is the right of peoples to healthy and culturally appropriate food produced through ecologically sound and sustainable methods, and their right to define their own food and agriculture systems. It puts the aspirations and needs of those who produce, distribute and consume food at the heart of food systems and policies rather than the demands of markets and corporations. It defends the interests and inclusion of the next generation. It offers a strategy to resist and dismantle the current corporate trade and food regime, and directions for food, farming, pastoral and fisheries systems determined by local producers and users. Food sovereignty prioritises local and national economies and markets and empowers peasant and family farmer-driven agriculture, artisanal fishing, pastoralist-led grazing, and food production, distribution and consumption based on environmental, social and economic sustainability. Food sovereignty promotes transparent trade that guarantees just incomes to all peoples as well as the rights of consumers to control their food and nutrition. It ensures that the rights to use and manage lands, territories, waters, seeds, livestock and biodiversity are in the hands of those of us who produce food. Food sovereignty implies new social relations free of oppression and inequality between men and women, peoples, racial groups, social and economic classes and generations Six Principles of Food Sovereignty From the Synthesis Report (Nyéléni, 2007b) Food Sovereignty: 1.

Focuses on Food for People:

2.

Values Food Providers:

3.

Localises Food Systems:

is FOR

is AGAINST

Food sovereignty puts the right to sufficient, healthy, and culturally appropriate food for all individuals, peoples, and communities, including those who are hungry, under occupation, in conflict zones and marginalised, at the centre of food, agriculture, livestock and fisheries policies; Food sovereignty values and supports the contributions, and respects the rights, of women and men, peasants and small scale family farmers, pastoralists, artisanal fisherfolk, forest dwellers, indigenous peoples, and agricultural and fisheries workers, including migrants, who cultivate, grow, harvest and, process food; Food sovereignty brings food providers and consumers closer together; puts providers and consumers at the centre of decision-making on food issues; protects food providers from the dumping of food and food aid in local markets; protects consumers from poor quality and unhealthy food, inappropriate food aid and food tainted with genetically modified organisms;

and rejects the proposition that food is just another commodity or component for international agri-business

and rejects those policies, actions and programmes that undervalue them, threaten their livelihoods and eliminate them

and rejects governance structures, agreements and practices that depend on and promote unsustainable and inequitable international trade and give power to remote and unaccountable corporations

Sustaining agricultural biodiversity and heterogeneous seeds 309

Table 14.1  Food sovereignty framework—cont’d Nyéléni (2007a, 2007b): Forum for Food Sovereignty Definition of Food Sovereignty—from the Declaration of Nyéléni (Nyéléni, 2007a) 4.

Puts Control Locally:

5.

Builds Knowledge and Skills:

6.

Works with Nature:

Food sovereignty places control over territory, land, grazing, water, seeds, livestock and fish populations on local food providers and respects their rights. They can use and share them in socially and environmentally sustainable ways which conserve diversity; it recognises that local territories often cross geopolitical borders and ensures the right of local communities to inhabit and use their territories; it promotes positive interaction between food providers in different regions and territories and from different sectors that helps resolve internal conflicts or conflicts with local and national authorities; Food sovereignty builds on the skills and local knowledge of food providers and their local organisations that conserve, develop and manage localised food production and harvesting systems, developing appropriate research systems to support this and passing on this wisdom to future generations; Food sovereignty uses the contributions of nature in diverse, low external input agroecological production and harvesting methods that maximise the contribution of ecosystems and improve resilience and adaptation, especially in the face of climate change; it seeks to “heal the planet so that the planet may heal us”;

and rejects the privatisation of natural resources through laws, commercial contracts and intellectual property rights regimes

and rejects technologies that undermine, threaten or contaminate these, e.g. genetic engineering

and rejects methods that harm beneficial ecosystem functions, that depend on energy intensive monocultures and livestock factories, destructive fishing practices and other industrialised production methods, which damage the environment and contribute to global warming

These six principles are interlinked and inseparable: in implementing the food sovereignty policy framework all should be applied.

310

Rethinking Food and Agriculture

s­ overeignty, who are represented in the Ecological Food Provision coalition. It might even criminalize them for pursuing their goal of extending biodiverse agroecology, based on heterogeneous seeds and agricultural biodiversity. As suggested earlier, “Peasant Seed Systems are ‘collateral damage’ in the international governance of seeds.” The expansion of industrial production, driven by powerful corporations in the Agribusiness coalition, supported by Green Revolution actors, and at the expense of those in the more responsible Ecological Food Provision coalition, is fueling the existential climate, biodiversity, and nutrition crises. However, as illustrated previously, food systems based on agroecology, especially as practiced by peasants in the framework of food sovereignty, can mitigate these crises. But key to the effectiveness of this approach is that heterogeneity is embedded in all the seeds, breeds, and agroecosystems used for food production. This heterogeneity needs to be expressed: ●





within the plant varieties, breeds of livestock, and aquatic organisms being cultivated, raised or harvested, and within the subspecies of “associated biodiversity” that provide essential ecosystem functions and support production (intravarietal and intervarietal biodiversity); in the biodiversity of species of crops, livestock, and associated species, above and below ground, such as soil organisms, pollinators, and predators and microbiota in soils, rumen, and waters (intraspecific and interspecific biodiversity); and in the biodiversity of the surrounding agroecosystems as well, from within fields and pastures and out to landscape and watershed levels (ecosystem biodiversity).

It is therefore imperative not only to maintain and enhance the diversity of crop species and varieties on-farm, as peasant agroecology and many alternative seed systems achieve, but it is also essential to increase their heterogeneity. It is necessary to enhance the intervarietal and intravarietal diversity of crop varieties and “populations” wherever food is grown. In other words, enhancing heterogeneity of the agricultural biodiversity within and between crops and their varieties in productive and biodiverse ecosystems is imperative for responsible agriculture. These biodiverse seeds will improve the resilience of production and secure future food (Global Alliance for the Future of Food, 2016). Beyond having heterogeneous seeds and breeds as the basis of production, enhancing the number and heterogeneity of associated species above and below ground is also essential for the sustainability of agroecosystems. For example, more heterogeneous populations of soil microbiota in organically managed, ­biodiversity-rich soils, improves soil health and hence the productivity of plants and the diversity of rumen microorganisms (Lupatini, Korthals, de Hollander, Janssens, & Kuramae, 2017). Such heterogeneity of agricultural biodiversity can be seen in biodiverse agroecological systems developed in the framework of food sovereignty and is achieved through the dynamic management of agricultural biodiversity and peasant seeds and breeds, especially by small-scale food providers (Mulvany, 2017). Increasing heterogeneity in production systems can be officially sanctioned, despite the hostile environment of the dominant governance architecture. The need

Sustaining agricultural biodiversity and heterogeneous seeds 311

for heterogeneity in seeds is now recognized by the European Union as an important dimension of sustainable and equitable food systems. These seeds are not uniform varieties and they are adapted to the needs of low-input agriculture in local or regional agroecological conditions. Plant breeders, cooperating with peasant and small-scale, often organic, farmers, have developed varietal mixtures. These include a barley mixture for brewing local beers developed in the former German Democratic Republic (East Germany) by Martin Wolfe and colleagues (described in Douthwaite, 1995). While Germany, once unified, restricted the use of such varietal mixtures, some 30 years later the European Union now recognizes that these are important. The European Union accepts that there are limitations of the term “plant variety” (taken from an IPR regime) in relation to the needs and efforts of organic farmers and plant breeders. The European Union has therefore allowed organic heterogeneous seeds and heterogeneous “populations” of single species to be marketed without complying with the requirements for registration and certification needed to register a uniform variety (Arche Noah, 2018). An example of such material, now licensed for sale in the European Union to organic farmers, is a heterogeneous wheat population initially developed through pioneering work by Martin Wolfe (Wolfe, Hinchsliffe, Clarke, Jones, & Haigh, 2006) and further developed as the “ORC-Wakelyns Wheat Population” (Organic Research Centre, 2016). In other regions, work by Ceccarelli with farmers and a nongovernmental organization (NGO), CENESTA, in Iran has introduced plantings of more than 1000 varieties of wheat from which local farmers have selected the population mixtures best suited to local ecosystems (Ceccarelli et al., 2010; CENESTA, 2013). Increasing the embedded heterogeneity within the food we eat could easily be promoted. Urban citizens are recognizing the need for increasing the biodiversity in their diets (McCarthy, 2018). If this were realized, it could significantly improve agricultural biodiversity within the production systems where the food is grown. This could be a driver for enhanced heterogeneity, resilience, and nutrient-density of food systems. To encourage this, and with due homage to Michael Pollan, one could say to these citizens “eat more biodiversity, mainly heterogeneous plants, but not too much!” Increasing the spread of knowledge of what needs to be done could be easily achieved if pushed by public-interest bodies and governments. But despite the urgency for systemic transformation of the dominant and damaging, monocultural, industrial food system, there are many barriers to achieving this—not least the power of corporations. Their “feed the world” narrative needs to be exposed for the self-­ serving offer that it is. Governments individually, and collectively in the UN and with support from civil society, need to exercise their responsibilities to strictly regulate these corporations and their activities rather than allowing corporations to co-opt them (Agroecology Now, 2020). To achieve the required changes, two approaches could be useful: 1. New governance priorities in place of, or separate from, those that currently exist, which would put primacy on heterogeneous seeds, breeds, and agroecosystems, would be an effective way for governments to demonstrate an intention to encourage and enforce changes. Priorities for this governance would include respecting the rights of peasants, and putting

312

Rethinking Food and Agriculture

control of heterogeneous seeds, livestock breeds, and other agricultural biodiversity back in their hands. 2. A parallel approach, which is more in the control of civil society and social movements, is to ensure that in all relevant programs and processes promoted and practiced by them, they are prioritizing the enhanced heterogeneity of seeds, breeds, and agroecosystems within production systems. As noted earlier, there are many approaches that, while espousing local determination of, and control over, localized food systems, may not explicitly embed heterogeneity in the systems they promote. This can undermine peasant agroecology and food sovereignty, and may result in the erosion of heterogeneous agricultural biodiversity and hence the resilience and adaptability of the systems. In research, development, and outreach programs designed to improve agroecological production systems, there is a need to explicitly include goals of enhancing heterogeneous seeds, livestock breeds, and agricultural biodiversity.

And how could one know if these approaches are really working in practice. In the context of developing “effective integrated approaches” to the sustainable use of agricultural biodiversity, the FAO assessment (FAO, 2019) calls for much greater attention to be placed on researching “the functional roles of various components of biodiversity for food and agriculture in key ecosystem processes within production systems and in wider landscapes or seascapes.” It also calls for better surveying, monitoring, and the development of appropriate indicators. Perhaps, in this context, what is needed is a new indicator for sustainability of food systems—an indicator that would show whether or not governance, policies, and practices are effective in sustaining agricultural biodiversity; an indicator based on measuring heterogeneity. Such an indicator should be designed to help all actors in the food system assess the heterogeneity of seeds and breeds, and other agricultural biodiversity above and below ground, within all food production systems. The indicator would not only measure the degree of heterogeneity of agricultural biodiversity within production systems at all levels—from seeds and breeds to ecosystems—but also show how this is changing. Widespread use of such an indicator could help highlight and reverse the devastating declines in agricultural biodiversity in many, mainly industrial, production systems, while supporting peasants’ biodiverse agroecology. From peasants, their social movements and civil society organizations to agricultural policy planners, enhancing the heterogeneity of agricultural biodiversity within all land, aquatic, or marine production systems should become a norm. An indicator of heterogeneity of agricultural biodiversity within production systems should become a dominant metric of responsible agriculture and food systems.

Annex 14.1  Three coalitions contesting control over agricultural biodiversity and seed systems Ecological Food Provision, Green Revolution and Agribusiness Coalitions

1.

Main characteristics of each coalition— based on the six pillars of food sovereignty summarized in the Nyéléni Synthesis Report (Nyéléni, 2007b) What is the priority output for the system?

Ecological Food Provision (in the framework of food sovereignty)

Green Revolution

Agribusiness—high external input industrial commodity production

Locally controlled, ecologically and socially sustainable and resilient food systems that fulfill the Right to Food and right of access to productive resources. Produce for subsistence and localized markets Emphasis on farmers’/peasants’ biodiverse, heterogeneous seeds for multivariety/species cropping integrated with livestock production Low-carbon, high-biodiversity production system

Seeks to increase single species productivity within existing framework of dominant power relations. Coalition encourages intensification through use of industrial seeds and agrochemicals in smallholder production, often for extractive markets. Emphasis on replacing locally adapted and resilient seeds with input-dependent “improved” hybrids and other industrial homogeneous seeds, including GMOs Reduced biodiversity and increased carbon footprint

Capital intensive monoculture production of commodities, dependent on agrochemicals, mechanization, and irrigation. Outputs traded to supply industrial processors and, often transnational, agribusinesses. Emphasis on homogeneous industrial seeds and breeds. Biodiversity erosion and high GHG emissions

Food for people, especially through localized food webs. It puts the right to sufficient, healthy, and culturally appropriate food for all individuals, peoples, and communities (including those who are hungry, under occupation, in conflict zones, and marginalized) at the center of food, agriculture, livestock, and fisheries policies and practices

Production of food mainly for market chains

Commodity production, usually monocultures, primarily for, and prices set by, external (national/regional) markets and international agribusiness linked through value chains

Continued

Annex 14.1  Three coalitions contesting control over agricultural biodiversity and seed systems—cont’d Ecological Food Provision, Green Revolution and Agribusiness Coalitions Ecological Food Provision (in the framework of food sovereignty) 2.

Which actors are valued?

3.

Is the food system delivering locally?

4.

Who controls the food system?

Green Revolution

Agribusiness—high external input industrial commodity production

Women and men, peasants and small-scale family farmers, pastoralists, artisanal fishers, forest dwellers, Indigenous Peoples, and agricultural and fisheries workers, including migrants, who cultivate, grow, harvest, and process food for (especially local) consumption Yes. Designed to bring small-scale food providers and consumers closer together in ensuring that good quality food is locally accessible

External actors—extension workers, scientists from formal public and private institutions and companies

Very capital intensive Low labor requirement Machine dependent Local livelihoods only from labor or contract farming

Not a priority other than to ensure production ultimately provides for food security

Local small-scale food providers and consumers and their organizations are at the center of decision-making on food issues. This attempts to place control over territory, land, grazing, water, seeds, livestock, and fish populations in the hands of local food providers and respects their rights

Defends dominant IPR systems and use-restriction technologies as “stimuli for innovation.” (Such benefits can only be realized by formal sector research and development.) Availability of external inputs but not control over them is a priority

No—based on demands and prices determined by unsustainable and inequitable international trade, often subsidized. Encourages “dumping” of products in others’ markets. Depends on governance, agreements, and practices that are determined remotely. Power ultimately in the hands of remote and unaccountable corporations Facilitates the privatization of natural resources (land, water, soil, biodiversity) and agricultural biodiversity through laws, commercial contracts, and IPR regimes and use-restriction technologies and processes

5.

Whose knowledge and skills are developed?

6.

Does it “work with nature?”

Builds on the skills and local knowledge of small-scale food providers and their local organizations that conserve, develop, and manage localized food production and harvesting systems, developing appropriate research and innovation systems to support this and passing on this wisdom to future generations Uses the contributions of nature in diverse, low external input, biodiverse, and carbon sequestering agroecological production and harvesting methods that maximize the contributions of ecosystems and improve resilience and adaptation, especially in the face of climate change

Accepts that the system is “knowledge intensive” and will need inclusion and “scaling up” of local and traditional knowledge, but does not call for local determination of priorities. Promotes proprietary, sometimes GM, seeds

To some extent—but seeks coexistence between industrial production and an ecosystem approach

The actors in formal R&D and extension—mainly private sector corporations with public sector work increasingly dependent on them. Promotes technologies that undermine, threaten, or contaminate local production systems, e.g., genetic engineering. Rejects local knowledge and innovations (including peasants’ heterogeneous seeds and breeds) No—uses methods that harm beneficial ecosystem functions that depend on energy-intensive monocultures and livestock factories, destructive fishing practices, and other industrialized production methods, which damage the environment and contribute significantly to global warming

Impacts on: Continued

Annex 14.1  Three coalitions contesting control over agricultural biodiversity and seed systems—cont’d Ecological Food Provision, Green Revolution and Agribusiness Coalitions Ecological Food Provision (in the framework of food sovereignty) Agricultural biodiversity

Seed systems

Ecosystem functions

Sustains and develops a wide range of agricultural biodiversity. Seeks improved animal/plant/soil interactions; improves adaptive capacity of production Develops heterogeneous local peasant seeds, adapted to local ecosystems, and protected through collective rights, in biodiverse agroecological production systems Maximizes effectiveness of ecosystem functions in biodiverse and heterogeneous agroecosystems

Green Revolution

Agribusiness—high external input industrial commodity production

Selectively builds upon existing agricultural biodiversity

Main cause of erosion of agricultural biodiversity

Can include use of biodiversityeroding “new seeds” and monocultural production

Uses proprietary industrial homogeneous seeds, protected through international and regional agreements, and national seed laws and norms, which can outlaw peasant seeds Replaces, where possible, natural ecosystem functions with external inputs in chemically dependent, homogeneous monocultures, with downstream impacts on other agroecosystems and wider biodiversity

Recognizes importance of ecosystem functions but can undermine these and the heterogeneity of agroecosystems and seeds through promotion of agrochemicals and GM seeds

Adapted from Coupe, S., Ensor, J., Mulvany, P.M. (2011). Food security, poverty reduction, climate change: placing Trócaire’s livelihoods work in context. Internal report, Practical Action. (This contributed to a 2012 Trócaire discussion paper, available at: https://www.trocaire.org/sites/default/files/resources/policy/food-security-poverty-reduction.pdf.)

Sustaining agricultural biodiversity and heterogeneous seeds 317

References AFSA, & GRAIN. (2018). The real seed producers: Small-scale farmers save, use, share and enhance the seed diversity of the crops that feed Africa. Alliance for Food Sovereignty in Africa, GRAIN. https://www.grain.org/en/article/6035-the-real-seeds-producers-smallscale-farmers-save-use-share-and-enhance-the-seed-diversity-of-the-crops-that-feed-africa. Agroecology Now. (2020). Prioritising agroecology and challenging corporate technology-led approaches in the 2021 UN food systems summit and beyond. Centre for Agroecology, Water and Resilience (CAWR), Coventry University. https://www.agroecologynow. com/prioritising-agroecology-and-challenging-corporate-technology-led-approaches-in-the-2021-un-food-systems-summit-and-beyond/. Arche Noah. (2018). Seeds in the new EU organic regulation. Information note January 2018 https://www.arche-noah.at/files/briefing_seeds_in_new_the_eu_organic_regulation_january_2018.pdf. Argumedo, A., & Pimbert, M. (2006). Protecting indigenous knowledge against biopiracy in the Andes. London: IIED. http://pubs.iied.org/14531IIED/. Asdal, Å., & Guarino, L. (2018). The Svalbard global seed vault: 10  years—1 million samples. Biopreservation and Biobanking, 16(5). https://www.liebertpub.com/doi/10.1089/ bio.2018.0025. Badgley, C., Moghtader, J., Quintero, E., Zakem, E., Chappell, M. J., Avilés-Vázquez, K., … Perfecto, I. (2007). Organic agriculture and the global food supply. Renewable Agriculture and Food Systems, 22(02), 86–108. https://doi.org/10.1017/s1742170507001640. CBD. (1994). CBD/COP decision II/15. Agricultural biodiversity. https://www.cbd.int/decision/ cop/?id=7088. CBD. (1996). Conservation and sustainable use of agricultural biological diversity. CBD COP decision III/11. Montreal, Canada: Convention on Biodiversity. https://www.cbd.int/ decision/cop/?id=7107. CBD. (2008). Importance of agricultural biodiversity. https://www.cbd.int/agro/importance.shtml. Ceccarelli, S. (2015). Seeds of future. https://www.researchgate.net/publication/278029868_ Seeds_of_Future. Ceccarelli, S., Grando, S., Maatougui, M., Michael, M., Slash, M., Haghparast, R., … Nachit, M. (2010). Plant breeding and climate changes. Journal of Agricultural Science, 148(6). Cambridge, UK https://www.cambridge.org/core/journals/journal-of-agricultural-science/ article/plant-breeding-and-climate-changes/55D9CF73F5D71708596DA4C1A18887EF. CENESTA. (2013). Evolutionary plant breeding: A method to adapt crops to climate changes, increase on-farm biodiversity and support sustainable livelihoods. Iran: CENESTA. https://www.cenesta.org/wp-content/uploads/2016/06/publication-cenesta-evolutionary-plant-breeding-2013-en.pdf. Claeys, P. (2019). The rise of new rights for peasants. From reliance on NGO intermediaries to direct representation. In Transnational Legal Theory. https://doi.org/10.1080/20414005.2 018.1563444. Coupe, S., Ensor, J., & Mulvany, P. M. (2011). Food security, poverty reduction, climate change: placing Trócaire’s livelihoods work in context. Internal report, Practical Action. Cromwell, E., Cooper, H. D., & Mulvany, P. M. (2001). Agriculture, biodiversity and livelihoods: Issues and entry points. In I. Koziell, & J. Saunders (Eds.), Living off biodiversity International Institute for Environment and Development. https://pubs.iied.org/ pdfs/7823IIED.pdf.

318

Rethinking Food and Agriculture

Douthwaite, R. (1995). Short circuit: Strengthening local economies for security in an unstable world. The Lilliput Press Ltd. https://www.feasta.org/documents/shortcircuit/Short_Circuit.pdf. Egziabher, T. B. G. (2002). The human individual and community in the conservation and sustainable use of biological resources. DEFRA, UK: Darwin Lecture. https://nyeleni.org/ IMG/pdf/Tewolde_Darwin_Lecture2002.pdf. etcGroup. (2017). Who will feed us?. https://www.etcgroup.org/whowillfeedus. Evans, P. (2007). Seed sense. London: The Guardian. https://www.theguardian.com/society/2007/jul/18/conservation.food. Fanzo, J., Hunter, D., Borelli, T., & Mattei, F. (Eds.). (2013). Diversifying food and diets: Using agricultural biodiversity to improve nutrition and health Routledge. https://www. bioversityinternational.org/fileadmin/user_upload/online_library/publications/pdfs/ Diversifying_food_and_diets_1688_01.pdf. FAO. (1997a). First report of the state of the world’s plant genetic resources for food and agriculture. http://www.fao.org/3/a-w7324e.pdf. FAO. (1997b). Technical workshop on farming systems approaches for the sustainable use and conservation of agricultural biodiversity and agro-ecosystems. Rome: FAO. https://www. cbd.int/doc/meetings/sbstta/sbstta-03/information/sbstta-03-inf-10-en.pdf. FAO. (1998). Sustaining agricultural biodiversity and agroecosystem functions: Opportunities, incentives and approaches for the conservation and sustainable use of agricultural biodiversity in agro-ecosystems and production systems. In Report of an international technical workshop organised jointly by the Food and Agriculture Organisation of the United Nations and the Secretariat of the Convention on Biological Diversity, prepared by Wino Aarnink (FAO), Sally Bunning (CBD), Linda Collette (FAO) and Patrick Mulvany (ITDG). www.cbd.int/doc/case-studies/agr/cs-agr-1998-12-rpt.pdf. FAO. (2006). Quality declared seed. Rome: FAO. http://www.fao.org/3/a-a0503e.pdf. FAO. (2007). The state of the world’s animal genetic resources for food and agriculture. Rome: FAO. http://www.fao.org/docrep/010/a1250e/a1250e00.htm. FAO. (2015). Voluntary guide for national seed policy formulation. Rome: FAO. http://www. fao.org/3/a-i4916e.pdf. FAO. (2019). The state of the world’s biodiversity for food and agriculture. In J. Bélanger, & D. Pilling (Eds.), FAO commission on genetic resources for food and agriculture assessments. www.fao.org/3/CA3129EN/CA3129EN.pdf. FAO. (2020). Biodiversity for food and agriculture: The perspectives of small‐scale food providers. Thematic study for FAO’s 2019 report on the state of the world’s biodiversity for food and agriculture. forthcoming. Fielder, A. K., Landis, D. A., & Wratten, S. D. (2008). Maximizing ecosystem services from conservation biological control: The role of habitat management. Biological Control, 45, 254–271. Fowler, C., & Mooney, P. (1990). The threatened gene: Food, politics and the loss of genetic diversity. Cambridge: Lutterworth Press. Frison, C. (2016). Redesigning the global seed commons. Law and policy for agricultural biodiversity and food security. Routledge. Gaia Foundation. (2015). Enhancing seed diversity. https://www.gaiafoundation.org/what-we-do/ food-seed-and-climate-change-resilience/dr-melaku-worede-seed-apprenticeship-trainings/. Global Alliance for the Future of Food. (2016). The future of food: Seeds of resilience—A compendium of perspectives on agricultural biodiversity from around the world. https://futureoffood. org/wp-content/uploads/2016/09/Future_of_Food_Seeds_of_Resilience_Report.pdf. GRAIN, & La Via Campesina. (2014). Food sovereignty: Five steps to cool the planet and feed its people. www.grain.org/article/entries/5102-food-sovereignty-five-steps-to-cool-the-planetand-feed-its-people.

Sustaining agricultural biodiversity and heterogeneous seeds 319

Guarino, L. (2014). So how many crops feed the world anyway? Agricultural biodiversity weblog. https://agro.biodiver.se/2014/03/so-how-many-crops-feed-the-world-anyway/. Halewood, M., Deupmann, P., Sthapit, B., Vernooy, R., & Ceccarelli, S. (2007). Participatory plant breeding to promote farmers' rights. Rome, Italy: Bioversity International. https:// www.bioversityinternational.org/fileadmin/_migrated/uploads/tx_news/Participatory_ plant_breeding_to_promote_Farmers__Rights_1254.pdf. Helfer, L. R. (2004). Intellectual property rights in plant varieties. International legal regimes and policy options for national governments. Rome: FAO. http://www.fao.org/3/a-y5714e.pdf. Holt-Giménez, E. (2017). A Foodie’s guide to capitalism: Understanding the political economy of what we eat. New York: Monthly Review Press and Food First Books. Hunter, D., Heywood, V. H., & Maxted, N. (2011). Crop wild relatives: A manual of in situ conservation. Reviewed in Plant Genetic Resources, 9(3), 486. https://doi.org/10.1017/ S1479262111000669. IAASTD. (2008). Global summary for decision makers of the global report of the international assessment of agricultural knowledge, science and technology for development. https://www.globalagriculture.org/fileadmin/files/weltagrarbericht/IAASTDBerichte/ GlobalSDM.pdf. IAASTD. (2009). Agriculture at a crossroads. International assessment of agricultural knowledge, science and technology for development. Island Press. www.fao.org/fileadmin/templates/est/Investment/Agriculture_at_a_Crossroads_Global_Report_IAASTD.pdf. IDDRI. (2016). Ten years for agroecology—Understanding agroecology in Europe. Paris: Institut du Développement Durable et des Relations Internationales (IDDRI). https://www. iddri.org/en/project/ten-years-agroecology-europe. Report at: http://www.efncp.org/download/Plaquette_AE-initiatives_VF_June_light.pdf. IPC (International Planning Committee for Food Sovereignty). (2016). Peasants give life to biodiversity. Agricultural Biodiversity Working Group of the IPC for Food Sovereignty. Illustrated Brochure http://www.foodsovereignty.org/wp-content/uploads/2016/10/ ENGLISH_spreads_lowRes_.pdf. IPES-Food. (2016). From uniformity to diversity: A paradigm shift from industrial agriculture to diversified agroecological systems. Report 02. International Panel of Experts on Sustainable Food systems. http://www.ipes-food.org/_img/upload/files/UniformityToDiversity_FULL.pdf. Judd, W. S. (2002). Plant systematics: A phylogenetic approach. Sinauer Associates. Khoury, C. K., Bjorkman, A. D., Dempewolf, H., Ramirez-Villegas, J., Guarino, L., Jarvis, A., … Struik, P. C. (2014). Increasing homogeneity in global food supplies. Proceedings of the National Academy of Sciences of the United States of America, 111(11), 4001–4006. https://doi.org/10.1073/pnas.1313490111. Kloppenburg, J. (2016). OSSI and the global seed system (p. 1). OSSI Newsletter. https://osseeds.org/ossi-and-the-global-seed-system/. Lewis,V., & Mulvany, P. M. (1997). A typology of community seed banks. NRI/ITDG. www.ukabc.org/ communityseedbanks.pdf. cited in Vernooy, R. et al. (2018) ‘Community seed banks: Origins, evolution and prospects’. Bioversity International https://www.bioversityinternational. org/e-library/publications/detail/community-seed-banks-origins-evolution-and-prospects/. Li-Bird. (2017). Participatory varietal selection: An approach to help farmers to evaluate crop varieties. Nepal: Li-Bird. http://libird.org/app/publication/view.aspx?record_id=228. Lockie, S. (2009). Agricultural biodiversity and neoliberal regimes of agri-environmental governance in australia. https://journals.sagepub.com/doi/abs/10.1177/0011392108101590. Lupatini, M., Korthals, G. W., de Hollander, M., Janssens, T. K. S., & Kuramae, E. E. (2017). Soil microbiome is more heterogeneous in organic than in conventional farming system. Frontiers of Microbiology. https://doi.org/10.3389/fmicb.2016.02064.

320

Rethinking Food and Agriculture

McCarthy, K. (2018). Eating biodiversity. London, UK: Royal Society for the Encouragement of Arts, Manufactures and Commerce. https://www.thersa.org/discover/ publications-and-articles/rsa-blogs/2018/12/eating-biodiversity. Mulvany, P. M. (2008). On-farm conservation of agricultural biodiversity: Needs, challenges and opportunities. In Agriculture for development Tropical Agriculture Association. No. 1, Spring 2008 https://taa.org.uk/wp-content/uploads/2018/10/A4DMarch2008.pdf. Mulvany, P. M. (2017). Agricultural biodiversity is sustained in the framework of food sovereignty. Biodiversity, 18(2–3), 84–91. https://www.tandfonline.com/doi/abs/10.1080/1488 8386.2017.1366872. Mulvany, P. M., & Ensor, J. (2011). Changing a dysfunctional food system: Towards ecological food provision in the framework of food sovereignty. Food Chain, 1(1), 34–51. Nyéléni. (2007a). Declaration of Nyéléni, Nyéléni 2007: Forum for food sovereignty, Sélingué, Mali. www.nyeleni.org/IMG/pdf/DeclNyeleni-en.pdf. Nyéléni. (2007b). Synthesis report. Nyéléni 2007: Forum for food sovereignty, Sélingué, Mali. www.nyeleni.org/IMG/pdf/31Mar2007NyeleniSynthesisReport-en.pdf. Nyéléni. (2015). Report of the Nyéléni forum for agroecology. https://ag-transition.org/wp-content/uploads/2015/10/NYELENI-2015-ENGLISH-FINAL-WEB.pdf. Nyéléni. (2017). Nyéléni Peasant agroecology manifesto. https://africaconvergence.net/IMG/ pdf/nyeleni_manifesto_of_peasant_agroecology_-_en.pdf. Organic Research Centre. (2016). Populations: Diversity in plant breeding. https://www. agricology.co.uk/sites/default/files/OWP_NOCC18.pdf. People’s Food Policy. (2017). A people’s food policy: Transforming our food system. https:// www.peoplesfoodpolicy.org/. Pimbert, M. (2006). Transforming knowledge and ways of knowing for food sovereignty and bio-cultural diversity. IIED. http://pubs.iied.org/pdfs/G01098.pdf. Pimbert, M. P., & Moeller, N. I. (2018). Absent agroecology aid: On UK agricultural development assistance since 2010. Sustainability, 10(2), 505. https://doi.org/10.3390/su10020505. Prescott‐Allen, R., & Prescott‐Allen, C. (1990). How many plants feed the world? Conservation Biology, 4, 365–374. https://doi.org/10.1111/j.1523-1739.1990.tb00310.x. Rahmanian, M., Salimi, M., Razavi, K., Haghparast, R., Ceccarelli, R., & Razmkhah, A. (2016). Evolutionary populations: Living gene banks in farmers’ fields in Iran. ILEIA. https://www. ileia.org/2016/04/16/evolutionary-populations-living-gene-banks-farmers-fields-iran/. Shiva, V., Anilkumar, P., & Singh, N. R. (2019). Seeds of sustenance & freedom vs seeds of suicide & surveillance. Dehra Dun, India: Research Foundation for Science, Technology and Ecology. https://seedfreedom.info/wp-content/uploads/2019/09/Seeds-of-Sustenance.pdf. Tansey, G., & Rajotte, T. (Eds.). (2008). The future control of food: A guide to international negotiations and rules on intellectual property. Earthscan: Biodiversity and Food Security, ISBN:9781844074297. https://www.idrc.ca/en/book/ future-control-food-guide-international-negotiations-and-rules-intellectual-property. Tutwiler, M. A., Bailey, A., Attwood, S., Remans, R., & Ramirez, M. (2017). Agricultural biodiversity and food system sustainability. In Mainstreaming agrobiodiversity in sustainable food systems: Scientific foundations for an agrobiodiversity index Bioversity International. https://www.bioversityinternational.org/mainstreaming-agrobiodiversity/. van der Ploeg, J. D. (2009). The new peasantries: Struggles for autonomy and sustainability in an era of empire and globalization. Earthscan. http://www.jandouwevanderploeg.com/EN/ doc/Struggles_for_Autonomy.pdf. Vandermeer, J. H., Lawrence, D., Symstad, A., & Hobbie, S. E. (2002). Effect of biodiversity on ecosystem functioning in managed ecosystems. In M. Loreau, S. Naeem, & P. Inchausti

Sustaining agricultural biodiversity and heterogeneous seeds 321

(Eds.), Biodiversity and ecosystem functioning: Synthesis and perspectives (pp. 221–236). Oxford University Press. Via Campesina. (2000). III International Conference. https://viacampesina.org/en/ biodiversity-and-genetic-resources/. Via Campesina. (2018). Peasant seeds a heritage of peoples in the service of humanity. https:// viacampesina.org/en/16-october-la-via-campesina-relaunches-global-campaign-forseeds-a-heritage-of-peoples-in-the-service-of-humanity/. Wolfe, M. S., Hinchsliffe, K. E., Clarke, S. M., Jones, H., & Haigh, Z. (2006). What will organic farming deliver? In Aspects of applied biology: Vol. 79. Evolutionary breeding of healthy wheat: From plot to farm Elm Farm Research Centre. https://orgprints.org/10177/1/ Evolutionary_breeding_of_healthy_wheat_-_from_plot_to_farm.pdf. Wolfe, M. S., & Ceccarelli, S. (2019). The increased use of diversity in cereal cropping requires more descriptive precision. In Short communication Wiley. https://doi.org/10.1002/ jsfa.9906.

This page intentionally left blank

Healthy diets as a guide to responsible food systems

15

Shireen Kassama,b, David Jenkinsc,d,e,f,g, Doug Bristorh, and Zahra Kassami,j a King’s College Hospital, London, United Kingdom, bWinchester University, Winchester, United Kingdom, cDepartment of Nutritional Sciences, Faculty of Medicine, University of Toronto, Toronto, ON, Canada, dDepartment of Medicine, Faculty of Medicine, University of Toronto, Toronto, ON, Canada, eClinical Nutrition and Risk Factor Modification Center, St. Michael’s Hospital, Toronto, ON, Canada, fLi Ka Shing Knowledge Institute, St. Michael’s Hospital, Toronto, ON, Canada, gDivision of Endocrinology and Metabolism, St. Michael’s Hospital, Toronto, ON, Canada, hIndependent Researcher, Cambridge, United Kingdom, i Stronach Regional Cancer Centre and the Princess Margaret Cancer Centre, Toronto, ON, Canada, jDepartment of Radiation Oncology, University of Toronto, Toronto, ON, Canada

15.1 Introduction During the last century, our diet, particularly in middle- and high-income countries, has transitioned from one centered around whole foods to one consisting of highly processed and animal-based foods, high in refined carbohydrates, added fat, sugar, and salt, often chosen for reasons such as convenience, ready availability, high palatability, and low cost and effective marketing campaigns (Rauber et al., 2018). This modern diet pattern, now typical in Western, high-income countries, is often termed the Western diet or the standard American diet. The consequence in many countries has been the exponential rise in obesity and diet-related diseases, such that our diet is one of the leading causes of death and disability globally (Gakidou et  al., 2017; Murray et al., 2018; Steel et al., 2018). Malnutrition, as defined by the 2018 Global Nutrition Report, is a global issue affecting every country, age group, and wealth bracket and is responsible for more illnesses and deaths than any other cause (Collaborators, 2018). In the report, malnutrition is a general term covering illnesses caused by insufficient food, including stunting and wasting in children under five, micronutrient deficiency, adult underweight, as well as illnesses related to overweight and obesity in all age groups. These different forms of malnutrition coexist within countries and people, meaning children can be affected by stunting and later by overweight and obesity. Micronutrient deficiencies can exist in those affected by food insecurity and underweight and those who are overweight due to overconsumption of nutrient-depleted foods. It is clear from the report that regardless of the type of malnutrition, country of residence, age group, and wealth bracket, people are eating too many refined grains and sugary foods and drinks and not enough whole plant foods, i.e., fruits, vegetables, whole grains, and legumes. Changes in dietary patterns have resulted in a worldwide epidemic of noncommunicable diseases (NCDs) (Fig.  15.1). According to the World Health Organization, Rethinking Food and Agriculture. https://doi.org/10.1016/B978-0-12-816410-5.00015-3 © 2021 Elsevier Inc. All rights reserved.

324

Rethinking Food and Agriculture

Fig. 15.1  Global death rate per 100,000 population between ages 5 and 74 years. IHD, ischemic heart disease; NCDs, noncommunicable diseases. Data from ghdx.healthdata.org.

NCDs cause 71% of deaths globally every year (WHO, 2016). The four main NCDs are cardiovascular disease (heart disease and stroke), cancer, chronic respiratory disease, and diabetes. The main risk factors leading to NCDs are high blood pressure, overweight/obesity, high blood sugars, and high blood cholesterol (Forouzanfar et al., 2017; Gakidou et al., 2017; Vos et al., 2017). These are ultimately all related to poor diet, high salt intake, and lack of physical activity, as well as tobacco smoking and

Healthy diets as a guide to responsible food systems325

excessive alcohol consumption. It was estimated that worldwide in 2010, around 9.5 million deaths were caused by dietary risk factors, including diets low in fruits, vegetables, nuts, seeds, and whole grains and high in salt and red and processed meat (Lim et al., 2012). The 2019 report from the Global Burden of Disease study group estimates that 11 million deaths a year are caused by unhealthy diets, which represents 20% of all deaths (GBD Diet Collaborators, 2019). So, where have we gone wrong? There are multiple factors involved in our food choices, including: public and government policy; industry influence over policy development, marketing of food, and scientific studies; food availability and accessibility; media representation of scientific data; social and cultural norms; and individual education, knowledge, and skill in nutrition and food preparation (Mozaffarian, Angell, Lang, & Rivera, 2018). National and international nutrition guidelines are not readily translated into individual behaviors for the reasons just mentioned. In addition, guidelines are often tempered for fear of industry pushback. Industry has repeatedly undermined evidence-based guidelines by attacking and denying the science, funding their own studies, and paying their own scientists to produce data that supports their narrative (Kearns, Schmidt, & Glantz, 2016). Medical professionals receive very little education in nutrition during their training and hence lack the knowledge and skills necessary to provide dietary counseling to patients (Katz, 2018). All of these factors either contribute to or are exacerbated by a food system that is increasingly providing consumers with cheap, processed, “food-like” products with low nutrient content rather than healthy, nutritious, whole foods that readily meet nutritional requirements. At the same time, this food system leaves nearly a billion people chronically hungry and another two billion suffering from overweight and obesity (Collaborators, 2018; Patel, 2012). Such a food system cannot be considered responsible by any measure. At the very least, responsible food systems must provide healthy and nutritious food for all consumers. The continued growth of the world’s population to an estimated 10 billion by the year 2050 is only going to worsen the problem, especially if more countries adopt a Western-style diet pattern. As such, the aim of this chapter is to show how healthy diets could serve as a guide for responsible food systems. To do so this chapter will review data on the impact of diet on health and illness and discuss how current dietary patterns are fueling an epidemic of chronic disease. It will discuss the role and influence of government and industry on our current diet choices, summarize the impact of food production on environmental health, and provide evidence supporting the need for a global shift to a predominantly plant-based dietary pattern for both optimal human and planetary health.

15.2 Current disease trends The exponential rise in chronic disease, termed NCDs, does not seem to be slowing. This trend is affecting all countries of the world with a higher rate of rise in low- and middle-income countries as they move from rural to urban dwelling, forego their traditional diets, and adopt a more Western diet pattern (India State-Level Disease Burden Initiative CVD Collaborators, 2018; Yang et al., 2013). There remains a high rate of

326

Rethinking Food and Agriculture

tobacco smoking in low- and middle-income countries, as the tobacco industry continues to push their products in countries that are yet to regulate advertising, sales, and smoking in public places (Reitsma et al., 2017). Cardiovascular disease is the commonest cause of death with 17.9 million deaths worldwide every year (WHO, 2016). It is estimated that 70%–80% of cases are preventable through the adoption of healthy lifestyle choices, including a healthy diet, regular physical activity, not smoking tobacco, and drinking alcohol in moderation or not at all. Risk factors leading to cardiovascular disease include high blood pressure, diabetes, and high blood cholesterol, all related to lifestyle factors. High blood pressure now affects more than 1 in 3 adults worldwide and causes 19% of deaths worldwide (Forouzanfar et al., 2017). In addition, diets high in sodium, predominantly from processed and convenience foods, are contributing considerably to the burden of high blood pressure globally (Strazzullo, D’Elia, Kandala, & Cappuccio, 2009). Diabetes, which in over 90% of cases is type 2, increases the risk of cardiovascular disease but in itself is a major cause of death and disability. In 2014, 422 million adults were living with diabetes and the global prevalence has nearly doubled since 1980, rising from 4.7% to 8.5% in the adult population (WHO, 2017a). Again, most cases of type 2 diabetes are caused by lifestyle factors, predominantly a diet that is promoting obesity and causing the accumulation of fat in muscle and liver cells, resulting in insulin resistance even in those who are of normal weight (Jacob et al., 1999). Cancer is the second commonest cause of death worldwide, with 9 million deaths annually (WHO, 2016). Lifestyle factors account for more than 40% of cases (WCRF, 2018). A Western diet pattern increases cancer risk through promotion of overweight and obesity, which is second only to tobacco smoking as the leading cause of cancer. In addition, the Western diet pattern contributes to cancer by being high in red and processed meat and ultraprocessed foods and low in fiber and phytonutrients (Li et al., 2018). Overweight and obesity is itself a chronic disease but also increases the risk of those diseases just mentioned. Prevalence of overweight and obesity is not only an issue for adults but is increasing among children (WHO, 2018). In 2016, more than 1.9 billion adults were overweight and 650 million were obese. This equates to 39% and 13% of adults being overweight and obese, respectively. Forty-one million children under the age of 5 and over 340 million children and adolescents aged 5–19 years were overweight or obese in 2016. Dietary patterns high in calorie-dense, nutrient-deficient foods leading to the overconsumption of calories is the main cause, with reduced levels of physical activity contributing to a lesser degree (Juul & Hemmingsson, 2015). Dementia has been termed by some as “type 3 diabetes” as it shares the same risk factors as diabetes and other NCDs. In 2016, 43.8 million people globally were living with dementia, a rise from 20.2 million in 1990 (Nichols et  al., 2018). The Lancet commission on dementia suggests that over a third of cases are preventable through the adoption of healthy lifestyle practices (Livingston et al., 2017). Overall, the rising burden of NCDs is contributing to a plateau in life expectancy and in some countries a rise in mortality rates (Dicker et  al., 2018). The current solution to the rising problem of NCDs has been an overemphasis on medical treatments with insufficient resources dedicated to prevention. Treatments rely heavily on

Healthy diets as a guide to responsible food systems327

p­ harmaceutical drugs and medical procedures, rather than lifestyle approaches. Half of adults are now taking at least one prescribed medication and one in four are on three medications. Yet, prescription medications and medical procedures are a leading cause of death (Gotzsche, 2014, 2016; Makary & Daniel, 2016). This rapid rise in NCDs and pharmaceutical approaches to treatment disadvantages those in low-­ income countries and from low-income households as they are more likely to have limited access to health services, inability to pay for health-related costs, and more likely to be exposed to harmful lifestyle products such as tobacco, alcohol, and unhealthy foods (WHO, 2016). Globally, the economic cost of NCDs is unsustainable even for high-income countries (Lozano et al., 2013). A report by the World Economic Forum in 2011 estimated that over the subsequent 20 years, NCDs in the United States could cost more than $30 trillion (World Economic Forum, 2011). In the United Kingdom, it is estimated that the annual cost of treating diabetes alone could reach £16.9 billion by 2035, which would represent 17% of the entire healthcare budget (Diabetes UK, 2012). Worldwide, it is estimated that dementia care will cost $2 trillion by 2030 (Alzheimers Disease International, 2015). The global outsourcing of clinical study design and funding to pharmaceutical companies is perpetuating the rising cost of treating NCDs. Little resource is being dedicated to research into lifestyle approaches, including prevention and reversal (Batt, 2016). Rather, the current medical and pharmaceutical model invariably means that once individuals are started on medications, they are required life-long. This is because medications generally treat symptoms (e.g., high blood glucose and blood pressure) rather than the underlying cause.

15.3 Diet patterns associated with health and longevity We have known for over a century that diet and lifestyle factors influence patterns of disease. Evidence comes from observational studies, large prospective cohort studies, and from human intervention studies. Regions of the world where people have the highest life expectancy and live in good health have been termed the Blue Zones (Buettner, 2012; Buettner & Skemp, 2016; Poulain et al., 2004). There are five Blue Zone regions: Loma Linda in California, Ikaria in Greece, the island of Sardinia, Okinawa in Japan, and the Nicoya Peninsula in Costa Rica. These regions share nine lifestyle factors in common. With regards to diet, they consume a predominantly whole food plant-based diet consisting of minimally processed fruits, vegetables, legumes, whole grains, nuts, and seeds. Animal-derived foods are eaten infrequently, with meat may be eaten five times per month (Table 15.1). A Blue Zone of particular interest is Loma Linda in California, home to a large population of Adventists. They pursue a healthy lifestyle and live 7–10 years longer than their fellow Americans. In general, they don’t smoke tobacco and limit alcohol and caffeine consumption. Their diet pattern, however, differs and falls into four main groups: omnivorous, pesco-vegetarian (excluding meat but eating fish, eggs, and dairy), ovo-lacto vegetarian (excluding fish and meat), and vegan (excluding all animal-­derived foods). All groups have a healthier diet than the typical American,

Table 15.1  Proportion of daily food consumption by food groups for traditional Blue Zone diets. Blue Zone regions Food groups

Okinawa 1949

Costa Rica 1960s

Loma Linda 1991

Crete (Greece) 1960

Montegiorgio (Italy) 1960

Average (excl. Okinawa)

Potatoesa Grains Oil Dairyb Meat and legumesc Sugar and honey Fruit and vegetables

66% 15%