FOOD AND AGRICULTURAL BIOTECHNOLOGY IN ETHICAL PERSPECTIVE. [3 ed.] 9783030612139, 3030612139

463 41 4MB

English Pages [428] Year 2021

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

FOOD AND AGRICULTURAL BIOTECHNOLOGY IN ETHICAL PERSPECTIVE. [3 ed.]
 9783030612139, 3030612139

Table of contents :
Acknowledgements
Introduction
Who Should Read this Book?
The Overall Plan
The Revised Edition
Some Disclosures
References
Contents
1 Biotechnology in the Context of Agriculture and Food: An Overview
1.1 Introduction
1.2 The Puzzle of Heredity
1.3 Altering the Genome
1.4 Genetic Engineering and Biotechnology
1.5 Recent Developments
1.6 What’s in a Name?
1.7 The Controversy in Ethical Perspective
1.8 Conclusion: Beyond Risk and Back Again
References
2 The Presumptive Case for Food Biotechnology
2.1 Technological Ethics: A Précis
2.2 Ethics and Risk
2.3 The Risk-Based Approach
2.4 The Logic of the Presumptive Case
2.5 The Social Dimension of the Presumptive Case
2.6 Making the Case for Biotechnology Badly
2.6.1 The Modernist Fallacy
2.6.2 The Naturalistic Fallacy
2.6.3 The Argument from Ignorance
2.6.4 The Argument from Hunger
2.7 Conclusion
References
3 Biotechnology, Policy and the Problem of Unintended Consequences: The Case of rBST
3.1 What is rBST? Why Does it Matter?
3.2 Biotechnology Policy and Philosophy
3.3 rBGH: Assessing Unwanted Consequences
3.3.1 Food Safety
3.3.2 Animal Welfare
3.3.3 Environmental Impact
3.3.4 Social Consequences
3.4 Ethical Disputes, Governance and Consensus Politics
3.5 Social Consequences Redux
3.6 Learning from rBST
References
4 Food Safety and the Ethics of Consent
4.1 The Ethics and Political Theory of Food Safety Regulation
4.2 Safety Criteria and Biotechnology
4.3 Ethical Gaps in Food Safety Governance
4.3.1 Bad Actors
4.3.2 Collateral Consequences
4.3.3 Social Uncertainty
4.4 The Philosophy of Food Safety
4.4.1 Classification
4.4.2 Purification
4.4.3 Optimization
4.5 Classification and Purification Versus Risk-Based Optimization
4.6 Food Safety and Ethics
4.7 Food Labels and Consent
4.8 Food Safety Risks in Ethical Perspective
References
5 Animal Health and Welfare
5.1 Animal Biotechnology and Food
5.2 Harm to Animals and the Risk-Based Approach
5.3 Drugs and Animal Feeds from Biotechnology
5.4 Engineered Animals
5.5 The Moral Status of Animals
5.6 Rollin’s Consensus Morality
5.7 Animal Telos, Animal Integrity and Objections to Genetically Engineered Animals
5.8 Against Changing Telos or Species Integrity of Food Animals
5.9 Animal Biotechnology and Moral Obligation
References
6 Ethics and Environmental Risk Assessment
6.1 The Environmental Debate
6.2 Environmental Risk: An Expected Value Approach
6.3 Expected Value and the Consequentialist Framework
6.4 Understanding Hazards and Harm
6.4.1 Acquired Pest Resistance
6.4.2 Weediness
6.4.3 Genetic Diversity
6.5 Probability, Precaution and the Quantification of Exposure
6.6 Exposure Quantification: Further Issues
6.7 Philosophy of Science and Risk Communication
6.8 Conclusion
References
7 Environmental Impact and Environmental Values
7.1 Environmental Impacts from Agrifood Biotechnology
7.2 Environmental Hazards and Environmental Values
7.3 Environmental Philosophy For and Against Expected Value
7.4 The Conceptual Landscape of Environmental Ethics
7.5 Environmental Ethics in Consumption
7.6 Environmental Ethics in Production
7.7 Ethics and Environmental Responsibility for Food Biotechnology
References
8 Social Impact and the Technology Treadmill
8.1 Technology, Politics and the Prediction of Social Change
8.2 The Social Consequences of Food Biotechnology
8.3 Theories of Justice
8.3.1 Utilitarianism and Utilitarian Theories of Justice
8.3.2 Justice and Rights
8.3.3 Justice and Virtue
8.4 Structural Injustice and Structural Criticisms
8.5 Social Consequences for Small and Family Farms
8.5.1 Family Farms: Utilitarian Arguments
8.5.2 Family Farms: Rights and Fairness
8.5.3 Family Farms and Moral Virtue
8.6 Conclusion
References
9 Can Agrifood Biotechnology Help the Poor?
9.1 The Ethics of Agricultural Research
9.2 Ethics and Agricultural Development
9.3 Social Consequences in Peasant Agriculture
9.4 World Feeders and Ethical Consumers
9.5 Social Consequences for the Conduct of Science
9.5.1 The Scientific Purity Argument
9.5.2 The Social Function Argument
9.5.3 The Public Trust Argument
9.6 Conclusion
References
10 Conceptions of Property and the Biotechnology Debate
10.1 Property Rights in Genetic Information: The Context
10.2 Ownership in Genetics: A Contested Subject
10.3 The Theory of Property
10.4 Instrumental Conceptions of Property
10.4.1 Libertarian Theory
10.4.2 Utilitarian Theory
10.5 Ontological Conceptions of Property
10.5.1 Natural Law Theory
10.5.2 The Labor Theory of Property
10.6 Linking Instrumental and Ontological Theories of Property
10.7 Property and Agrifood Biotechnology: Ontological Approaches
10.7.1 Ruling Out Ownership of Human Genes
10.7.2 Rivalry and Excludability
10.7.3 Do Scientists Own Their Labor?
10.8 Property and Agrifood Biotechnology: Utilitarian Approaches
10.9 Property and Agrifood Biotechnology: Libertarian Approaches
10.10 Against Property Rights in Food Biotechnology
10.11 Ownership and Commodities
10.12 Conclusion
References
11 Religiously Metaphysical Arguments Against Agrifood Biotechnology
11.1 Intrinsically Wrong
11.2 Situating Religious and Metaphysical Claims
11.3 Metaphysical and Religious Arguments: An Exhibitive Introduction
11.4 Analyzing the Religiously Metaphysical Case Against Genetic Engineering
11.5 Religious Statements on Genetic Technology Through 2005
11.6 The Religious Case Against Property Rights in Genes: 1985–1995
11.7 Academic Theology and Genetic Engineering
11.8 The Ethical Implications of Religious Views
References
12 Communication, Education and the Problem of Trust
12.1 Public Opinion and the Ethics of Gene Technology
12.2 The Problem of Trust
12.3 Communication and Public Understanding of Science
12.4 The Ethics of Science Communication
12.5 The Problem of Risk
12.6 Human Action, Risk, and Responsibility
12.7 Equivocation Problems and False Authority
12.8 Moral Reductionism and Political Exclusion
12.9 Conclusion
References
13 Gene Editing, Synthetic Biology and the Next Generation of Agrifood Biotechnology: Some Ethical Issues
13.1 Synthetic Biology: The Old New Thing
13.2 Gene Editing and the Next Generation of Biotechnology
13.3 The Products of Gene Editing in Food and Agriculture
13.4 Alternative Proteins
13.5 Novel Agricultural Ecosystems
13.6 Gene Drives
13.7 Conclusion: Should We Worry?
References
14 Biotechnology, Controversy and the Philosophy of Technology
14.1 The Global Controversy: What Role for Philosophy?
14.2 Philosophy of Technology
14.3 Risk Assessment and Its Enemies
14.4 Environmental Pragmatism: A Prolegomena
14.5 Technological Pragmatism: Ihde
14.6 Technological Pragmatism: Postphenomenology Expanded
14.7 Gene Technology as a Moral Apocalypse
14.8 Conclusion: Technological Pragmatism and World-Feeding Ideology
References
Index

Citation preview

The International Library of Environmental, Agricultural and Food Ethics 32

Paul B. Thompson

Food and Agricultural Biotechnology in Ethical Perspective Third Edition

The International Library of Environmental, Agricultural and Food Ethics Volume 32

Series Editors Michiel Korthals, Wageningen University, Wageningen, The Netherlands Paul B. Thompson, Michigan State University, East Lansing, USA

The ethics of food and agriculture is confronted with enormous challenges. Scientific developments in the food sciences promise to be dramatic; the concept of life sciences, that comprises the integral connection between the biological sciences, the medical sciences and the agricultural sciences, got a broad start with the genetic revolution. In the mean time, society, i.e., consumers, producers, farmers, policymakers, etc, raised lots of intriguing questions about the implications and presuppositions of this revolution, taking into account not only scientific developments, but societal as well. If so many things with respect to food and our food diet will change, will our food still be safe? Will it be produced under animal friendly conditions of husbandry and what will our definition of animal welfare be under these conditions? Will food production be sustainable and environmentally healthy? Will production consider the interest of the worst off and the small farmers? How will globalisation and liberalization of markets influence local and regional food production and consumption patterns? How will all these developments influence the rural areas and what values and policies are ethically sound? All these questions raise fundamental and broad ethical issues and require enormous ethical theorizing to be approached fruitfully. Ethical reflection on criteria of animal welfare, sustainability, liveability of the rural areas, biotechnology, policies and all the interconnections is inevitable. Library of Environmental, Agricultural and Food Ethics contributes to a sound, pluralistic and argumentative food and agricultural ethics. It brings together the most important and relevant voices in the field; by providing a platform for theoretical and practical contributors with respect to research and education on all levels.

More information about this series at http://www.springer.com/series/6215

Paul B. Thompson

Food and Agricultural Biotechnology in Ethical Perspective Third Edition

123

Paul B. Thompson Department of Philosophy Michigan State University East Lansing, MI, USA

ISSN 1570-3010 ISSN 2215-1737 (electronic) The International Library of Environmental, Agricultural and Food Ethics ISBN 978-3-030-61213-9 ISBN 978-3-030-61214-6 (eBook) https://doi.org/10.1007/978-3-030-61214-6 1st edition: © Chapman & Hall 1997 2nd edition: © Springer 2007 3rd edition: © Springer Nature Switzerland AG 2020 This work is subject to copyright. All rights are reserved by the Publisher, whether the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and transmission or information storage and retrieval, electronic adaptation, computer software, or by similar or dissimilar methodology now known or hereafter developed. The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication does not imply, even in the absence of a specific statement, that such names are exempt from the relevant protective laws and regulations and therefore free for general use. The publisher, the authors and the editors are safe to assume that the advice and information in this book are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the editors give a warranty, expressed or implied, with respect to the material contained herein or for any errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. This Springer imprint is published by the registered company Springer Nature Switzerland AG The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland

This book is dedicated to my mother, Joan Banks Thompson (1932–2008), with enduring respect for her appreciation of the influence of genes and a character building life environment.

Acknowledgements

For the first edition: I would like to thank the US National Science Foundation project SBR-9602968 for supporting research for this book. I would also like to thank Institute for Biosciences and Technology (IBT) and College of Liberal Arts at Texas A&M University for long term support of my research on biotechnology. The support of Charles Arntzen, Fuller Bazer and Benjamin Crouch was of singular importance. Rose Gilliver, Colette Holden and Marilyn Grant at Chapman and Hall, and Daralyn Wallace and Johanna White at Texas A&M gave assistance in preparing and editing the manuscript. The final stages immediately preceding publication were supported by the Joyce and Edward E. Brewer Chair in Applied Ethics at Purdue University. For the second edition: The US National Science Foundation supported work through a new award, SES-0403847. W. K. Kellogg Foundation and Michigan State University supported me through endowment funds for the W. K. Kellogg Chair in Agricultural, Food and Community Ethics. I would also like to thank the applied philosophy group at Wageningen University & Research Institute in the Netherlands, who hosted my sabbatical in 2004, when most of the revisions for the second edition were done. Bill Hannah assisted in preparing the index for the second edition. Fritz Schmul at Springer managed the publication of the revised manuscript in the Library of Environmental and Agricultural Ethics. My administrative assistant Julie Eckinger oversaw communication and management of materials. For the third edition: US Department of Agriculture supported my research through my Hatch Project MICL02620—Social and Ethical Issues in Advanced Biotechnology. Rebecca Grumet and Ned Walker, both of Michigan State, have offered encouragement and advice. Michigan State University’s W. K. Kellogg Chair in Agricultural, Food and Community Ethics continues to fund travel that supports my research. I would also like to acknowledge North Carolina State University’s Genetic Engineering and Society Center for including me in their NSF

vii

viii

Acknowledgements

funded workshop on gene editing, with special gratitude to Fred Gould and Jennifer Kuzma. Per Sandin, Kirill Thompson, Catherine Kendig and Jeffery Burkhardt offered valuable comments on the manuscript. My key contact at Springer for the third edition is Floor Oosting, ably supported by Christopher Wilby. Julie Eckinger continues be a vital aid in manuscript production and proofreading.

Introduction

Abstract: This introduction serves as a guide to potential readers by describing the structure and organization of the main body of this book. The concept of risk frames the main ethical issues associated with the development and utilization of agricultural and food biotechnologies. The book uses concepts and methods of risk analysis to address ethical issues in the identification of hazards from gene technology, the estimation of the likelihood that hazards will materialize, and the strategies for managing risk. Standard regulatory and scientific approaches to risk analysis underestimate the significance of philosophical value judgments in specifying risks. In this respect, the book complements existing scientific risk analysis of gene technology in the food system. However, risk analysis is less capable of addressing issues relating to the social environment (e.g. property rights), deeper philosophical and religious values or respectful communication and interaction with the non-science public. As such, the book further complements a risk-based approach by discussing its limitations. The introduction also identifies the audience for this new edition and outlines the main changes from previous editions. Keywords: risk assessment, genetic engineering, ethics, philosophy of biology, philosophy of technology Genetic engineering—the direct manipulation of an organism’s genetic code—is a signature technology for the transition years marking the end of the twentieth century and the beginning of a new millennium. In an era that saw the spread of personal computers, the arrival of smart phones, the development of the Internet, the robotic revolution of manufacturing and impressive changes in human medicine, there was arguably no technology that sparked a more intense and sustained public controversy than the application of genetic engineering in agriculture and the food system. A new round of technical innovations for gene editing has reinvigorated the controversy as humanity enters the third decade of the twenty-first century. This book provides a roadmap to ethical issues in the use of gene technology within the food system, and traces the debate back to a time before any products were available.

ix

x

Introduction

The first edition of Food Biotechnology in Ethical Perspective appeared in 1997. The first significant release of genetically engineered crops occurred in the same year. The next 5 years saw a series of snafus, controversies and imbroglios that continue to influence public’s reception of these crops, now widely referenced as GMOs, an acronym for “genetically modified organism.” These events precipitated revisions to the book that I began in 2003. Both the controversies and my revisions continued right up until the publication of the second edition in 2007. Although the basic philosophical argument and organizational structure were unchanged, the second edition reflected “lessons learned” from episodes including toxicologist Arpad Pusztai’s precautionary food safety concerns (see Ewan and Pusztai 1999), the recall of StarLink maize in 2000 (Taylor and Tick N.D.), notorious crop failures (Hilder and Boulter, 1999), so-called Terminator seed, (Herring 2006), African rejection of U.S. GMO maize as food aid, (Zerbe 2004), food animal cloning (Rollin 1997) and especially the widespread rejection of GMOs throughout the European Union, (Gaskell, Thompson and Allum 2002). Although the ethical controversy over GMOs abated only slightly, very little occurred after 2007 that presented the occasion for a third edition. That changed with the advent of gene editing (Wright, Nuñez, and Doudna 2016) as well as proposals for gene drives (Champer, Buchman and Akbar 2016). As such, this new edition includes new material on the ethical implications of these developments. The first six sections of this introductory chapter do exactly what one expects from an introduction. These sections give a broad overview of the approach and subject matter of the main text, explain and explain the rationale for producing a revised edition. The main body of the book is an attempt to provide a balanced and informative guide to ethical issues, rather than a normative argument for or against the technology. The introduction closes with an update of one feature from the original 1997 edition that was unusual for its time: a disclosure of financial and professional interests associated with the author’s research on agricultural genetic engineering.

Who Should Read this Book? The first edition of this book was addressed to food and agricultural scientists, and to research administrators involved in the oversight of what was already a controversial activity. It was included in a series of scientific titles offered by the now defunct science publisher Chapman and Hall. I had been publishing a newsletter in my capacity as Director of the Center for Biotechnology Policy and Ethics at Texas A&M University. In an era when the Internet was in its infancy, we mailed the newsletter to approximately 500 scientists, many of whom had attended a meeting of the (also now defunct) National Agricultural Biotechnology Council (NABC). The NABC itself was a consortium of U.S. and Canadian universities with major programs in agricultural research. It had been formed in the 1990s under the leadership of Ralph Hardy (1934–2016), a biochemist who led the DuPont Corporation’s transition to molecular technology before assuming a position at

Introduction

xi

Cornell University’s Boyce Thompson Institute in 1986. NABC was both a forum for discussion of social and regulatory issues associated with non-medical applications of biotechnology, and a vehicle for helping its member institutions understand the new social environment in which the products of biotechnology would be deployed, (Segelken 2016). NABC included an initiative that encouraged biotechnology researchers to become both more knowledgeable about the ethical criticisms being leveled against their research and to become more competent in the tools of ethical analysis, (Thompson 1998). The first edition was intended to serve a similar purpose, while the second edition updated the first by including brief discussions of controversial episodes. In retrospect, I was naïve about the science audience’s willingness to shoulder the burdens of a book-length treatment. Most of the readers who have cited either the first or the second edition are social scientists or philosophers. I eventually realized that scientists needed a more concise statement of the argument, which appeared in a scientific journal at almost the same time as the second edition, (see Thompson and Hannah 2008). This shift in my own understanding of the audience for the book has guided my approach to making revisions for the third edition. This third (and assuredly final) revision should be read as an extended case study in the risk-based philosophy of technology. I have made additions and clarifications to explain the risk-based approach to my expected audience of philosophers and science scholars, while the specific focus of my analysis—recombinant DNA methods for the transformation of plants and food animals—functions more as an illustrative application of the approach than as a comprehensive overview of food agricultural biotechnologies circa 2020. The book should be of special interest to the new generation in philosophy of science taking an interest in the role that values play in science, as well as those who are examining hitherto neglected areas, such as agricultural science. At the same time, the new edition includes a background discussion of applied agricultural genetics. Philosophers and social scientists who are well versed in medical genomics and molecular biology might not appreciate some basics in the way that these sciences are applied in the context of food and agricultural research. Moreover, as a philosopher of biology David Hull (1935–2010) argued, even highly educated people tend to imagine (falsely) that vertebrate reproduction models the hereditary transmission of genetic traits for all species. For Hull, a focus on plants opens one to a more sophisticated understanding of evolution as a process (Hull 2001). For me, it is crucial for appreciating both the risks of agricultural biotechnology, and for seeing how many people misjudge them. In addition, the new edition includes a chapter discussing how the risk-based philosophy of technological ethics might be applied to gene editing techniques that have been developed since the appearance of the second edition. The book does thus provide some substantive discussion of the philosophical issues that arise in applied molecular biology within the domains of food and agriculture. However, it does not pretend to be a thorough update on philosophical debates in food and agriculture. In short, the third edition stands as a contribution to science and technology studies with relevance to all practitioners of this field, as well as to the new

xii

Introduction

generation of philosophers who have taken up food ethics, (see Plakias 2019). The entire book should be of particular interest to philosophers of science now undertaking studies in inductive or epistemic risk, (see Elliott and Richards 2019). It also remains of value to practitioners in the agricultural and food sciences, as I have retained much of the original text that was develop to acquaint non-specialists with key concepts in ethics and the philosophy of technology. Ironically, it may be least useful for readers specializing in contemporary debates over non-medical applications of applied genetics, as I have made no attempt to provide a complete discussion of debates over genetically engineered insects or the impact of climate change on thinking about the ethics of food and agricultural biotechnology. For these readers, the value of the third edition will consist in providing a background on the history and trajectory of debates over non-medical biotechnology from the ASILOMAR conference in 1977 through to the publication of the second edition in 2007. It should counteract the tendency for scholars to write as if this 30-year period of intense scrutiny and contestation never existed. In service to the book’s historical ambitions, I include the birth and death years following the first mention of deceased persons, but not for persons who were still living at the time of the third edition’s publication. I have added a final chapter that situates my career interests in agriculture and technological risk within the philosophy of technology more generally.

The Overall Plan The basic theoretical orientation of risk analysis provides the organizational structure for the book. I take risk to be a phenomenon that human beings experience in the normal course of living. As occurrences within the stream of experience, risks are acts undertaken by oneself or others. Risks differ from the quotidian series of actions that constitute the life of a person, a group or a community because they are marked by a greater-than-normal potential for harm. They are exceptional in the sense of calling for reflective consideration. The concept of risk functions both to mark an action as requiring some form of rational evaluation or moral response, and to structure the reflective procedure that fulfills this requirement. In business, politics and (more recently) environmental planning, this procedure specifies the harm or evil that might ensue, evaluates its chance of actually occurring and decides what to do about it. This procedure has been formalized within the scientific risk assessment. In risk assessment, risks are understood as a composition of hazard (e.g. the bad thing that might happen) and exposure (the likelihood that the hazard will materialize at any given time and place). The identification of potential hazards and the assessment of exposure are primarily empirical matters, though they are not free of critical value judgements. Ethics comes in most prominently through risk management, where decision makers apply normative principles in choosing what to do about risks. One option is to do nothing; simply accept the risk. More typically in

Introduction

xiii

the cases of direct relevance to the GMO debate, government authorities apply principles that range from de minimus (e.g. take the least feasible risk) to some form of risk-benefit weighting. The centralization of decision-making in regulatory agencies means that ethical issues also arise in connection with the authority on which regulators exercise judgment, and the opportunity of individual citizens to apply values that are at odds with decisions taken in regulatory bodies. These issues are often classified under the heading risk communication in the jargon of the risk sciences. The application of scientific methods to the hazard-exposure-decision schema is a two-edged sword. On the one hand, it yields both clarity and insight into the dangers of contemporary life. In areas of life where hazards are uncontroversial (such as loss of life) and data on the likelihood of their occurrence are plentiful, the scientific quantification of hazard and exposure promotes planning and effective pursuit of both personal and social goals. Even when there is disagreement about the seriousness of a hazard (such as losing one’s job or psychological stress), the framework structures communication in a manner that promotes mutual understanding and democratic debate. On the other hand, some aspects of risk are less amenable to science. Viewed as actions, risks may call for courage or caution. They structure social relationships in terms of responsibility, dependency and self-reliance. These aspects of risk are less easily analyzed within the hazard-exposure-decision model. More problematically, specialists use the model to repress the influence that these ethically important aspects of risk might have in the policy arena. Significantly, the hazard-exposure-decision model seems to apply equally to exceptional and ordinary acts. People do, I think, generally acknowledge that virtually everything they do can function in a sequence leading to a bad outcome, but our concept of risk loses critical functionality if it is applied too broadly. For example, people do die in their sleep, yet we do not think of sleeping as a risky activity. In particular, we do not pause before bedtime to evaluate the chance of dying while asleep in order to decide about how we should proceed. Yet if one follows the logic of scientific risk assessment programmatically, it seems we should indeed make such an assessment: as risk assessors are fond of saying, “there is no zero risk.” In short, the concept of risk functions categorically to allocate cognitive effort. If the hazard-exposure-decision model is allowed to disrupt these functions, it is contributing to irrational (and possibly immoral) conduct, rather than the reverse. The book attempts to navigate these tensions. It follows the schema of risk analysis through six chapters at the center of the book (e.g. Chaps. 4 through 9) focused on the identification and management of risk-based ethical issues. Chapter 4 takes up food safety. Chapter 5 addresses harm to animals. Chapters 6 and 7 are focused on environmental risks and the ethics of the human relationship to the natural world. Potential impacts on the socioeconomic organization of the food system (including its research capacity) are discussed in Chap. 8, and distributional issues, including shifts in power relations that affect the poor, are taken up in Chap. 9. These six chapters are preceded by three chapters that situate these risk-oriented

xiv

Introduction

chapters within the general context of technological ethics and the specific developments that led up to direct genetic modification of the plants and animals consumed for food and fiber. Chapter 1 provides background on genetics and gene technology, while Chap. 2 describes the risk-based approach in more detail and discusses the burdens of proof that an ethical analysis of technology must meet. Chapter 3 is a case study of the controversy surrounding bovine somatotropin, one of the first products to move through global regulatory systems. Readers with a special interest in the acceptability of using genetic tools to affect an animal’s capacity to suffer will find that work I have published elsewhere covers themes that are not covered in this volume, (Thompson 2008, 2010). Chapters 10 through 12 discuss issues that fit less cleanly into the risk assessment framework. Many ethical issues take up patents in gene technology, the subject of Chap. 10. Chapter 11 covers so-called “intrinsic” arguments, often grounded in religious faith that there might be something inherently wrong with genetic engineering, irrespective of its risks and benefits. Chapter 12 pulls the risk-based themes together with debates over intellectual property and intrinsic arguments by viewing all of them through the lens of risk communication. It is here that the limitations of a risk-based approach become most relevant. Chapter 12 also contains the most prescriptive sections in the book, where recommendations on process and engagement become explicit. This was the final chapter in the previous two editions, but the new edition contains two new chapters. Chapter 13 updates the book’s overall plan with a discussion of gene editing. Chapter 14 situates the extended case study of agricultural biotechnology within ongoing discussions in the philosophy of applied technoscience. While limitations of an overly scientized version of the risk-based approach are discussed in nearly every chapter, it is in chapters on religious and metaphysical objections, communication and the philosophy of technology (e.g. Chaps. 11,12 and 14) that these problems are most squarely addressed. Brief words on terminology are in order, though a more philosophical discussion follows in Chap. 1. For many technical readers, the term food biotechnology would signal tools and techniques used in food processing but not in agricultural production. In fact, most of the specific applications discussed in the book are in agriculture. To acknowledge this, I frequently use the neologism agrifood to connote the systemic integration of production and processing in the industrial food system. I also use this term to acknowledge that cotton, tobacco and oilseeds are among the important products of agrifood biotechnology, even if some of us would not include them as food. I have modified the title of the new edition to communicate the book’s focus on agriculture more clearly. The GMO acronym is controversial in some quarters. I do not shy away from it despite the obvious fact that in some sense all the crops and livestock species we use as food have been genetically modified when compared to undomesticated species. In deference to still unresolved questions of common usage, I do not use this acronym in discussing the next generation of products being produced through gene editing. Finally, the word ethics covers an array of values-based topics that arise in connection with agrifood biotechnology. I use the term, as most

Introduction

xv

philosophers do, to signal the significance of interpretation and philosophical argument in addressing these topics. My usage does not imply particular ethical theories or a judgment about the conduct or character of any individual group. Food ethics is a philosophical inquiry conducted in response to the question “How should we act?” with respect to the production, processing, distribution and consumption of food, including the research that supports these activities. As such, food ethics should not be understood simply as a form of shopping or consumption behavior in which people choose to buy what they presume to be more ethically produced foods. On my interpretation, food safety and environmental risk are appropriate topics for food ethics. We cannot address the risks of any technology without deploying judgements of good and bad. There is no sharp distinction between ethics and risk assessment.

The Revised Edition I have not undertaken the wholesale restructuring that would have been required by writing an entirely different book. There is new material in every chapter, but only Chaps. 1, 13 and 14 are totally new contributions. Many revisions to the book accommodate changes in academic publishing. While it was my expectation that readers of the first two editions would use a printed copy, electronic platforms dictate several modifications. First, libraries are understandably reluctant to repurchase material that they have already paid for with a journal subscription. Two chapters from the 2007 edition were reprinted material from publications that are widely available through university libraries, (see Thompson 1999 and Thompson and Hannah 2008). These are dropped from the current edition. Second, many readers of electronic editions prefer chapters that function as self-contained units, with chapter abstracts and a complete set of references. Each chapter in the new edition can be read independently. These formatting changes should make the new edition more useful to the current generation of scholars. The revised edition will provide scholars coming to the subject for the first time with a risk-based schema for approaching ethical issues in agricultural and food biotechnology. This can be obtained by focusing on Chaps. 1 and 2, where a general and philosophically oriented discussion of the approach is given, and then selectively choosing other chapters that are of interest. As noted already, the new edition will also be a useful backward-looking review of scholarship completed before 2007 for specialists in the ethics and policy dimensions of the field. I especially recommend attention to the U.S. debate over recombinant bovine somatotropin (rBST), not only introduced in Chap. 3, but also discussed in Chaps. 5 and 8. This debate rose to the highest levels of politics between 1985 and 1995, with significant action taken by the U.S. Senate and the Executive Office of the President. Although rBST was a drug produced in a modified microorganism, rather than a genetically altered plant or animal, the rBST debate influenced the politics of biotechnology in numerous ways. Latecomers to the biotechnology

xvi

Introduction

debates (and this would include most European analysts) have underestimated both the seriousness with which rBST was debated in the U.S. and Canada and the way in which it was resolved that affected the subsequent attitudes of environmental and food safety activists when crops were released in 1997. Abstracts for individual chapters will help readers determine which aspects of the book are of interest. All 14 chapters are updated with references to recent events and literature, but I have also retained older citations and episodes discussed in earlier editions. I hope that readers will see this as an effort to ground the biotechnology controversy in a more accurate historical context. I am dismayed by recent critics’ and scholars’ apparent lack of knowledge about or interest in earlier stages in the debate. This earlier literature demonstrates that the science community had many opportunities to undertake more extensive engagement with ethical issues, but failed to do so. At the same time, the book was never envisioned as a discussion of regulatory policies for agrifood biotechnology. Regulatory bodies (such as the U.S. Food and Drug Administration or the European Food Safety Administration) and their policies are mentioned only in reference to the ethical norms that administrative law is intended to support. As debates have turned more explicitly toward the instruments of regulatory policy, the ethical rationale for preferred policies to manage risk is submerged. This book brings the competing versions of that rationale to the fore. Even before the appearance of the second edition, philosophers such as Eric Millstone and science scholars such as Brian Wynne had begun to publish critiques of the regulatory approach for agricultural and food biotechnology. They emphasized uncertainties associated with the impact of recombinant gene modification, sometimes with a specific focus on the potential for unanticipated hazards from food consumption (Millstone, Brunner and Mayer 1999) and sometimes more vaguely, in terms that could apply equally to food consumption or environmental impact, (Wynne 2001). More recently, Millstone and Wynne have joined forces to critique regulatory procedures that, in their view, are unduly influenced by scientists with pecuniary interests in the approval of GMOs, (Hilbeck and coauthors 2020; see also Millstone 2007). Zahra Meghani also advances the procedural critique, attributing weaknesses in the regulation of biotechnology to the neoliberal proclivity for favoring market mechanisms of social governance (Meghani 2017). In addition to these debates that revolve around the adequacy of food safety regulation, there has been a growing call for broadening the scope of regulation. In some cases, broadening implies integrating information from non-scientists into a regulatory decision-making process. In this view, reliance on scientific expertise becomes contrary to the democratic foundations of a regulatory policy when it excludes considerations that are of importance to citizens and other parties affected by regulatory policy. In such cases, ethics is understood to be a platform that opens up the regulator’s universe of discourse to these otherwise excluded perspectives, (Kaiser 2009; Cotton 2014). Others argue that regulatory systems should be expanded to include socioeconomic impacts, (Mölders 2014). Bjorn Kåre Myskja and Anne Ingeborg Myhr renew this call with respect to what they call “non-safety assessments of genome-edited organisms” (Myskja and Myhr 2020). A different

Introduction

xvii

way to expand the regulatory framework involves a focus on which organizations should be subjected to oversight. Meghani’s concerns about the influence of corporate power are also reflected in the work of analysts that study the influence of private capital on the research agenda for agriculture (Rudy et al. 2014). Some philosophical issues in this regulatory-oriented literature are addressed in the book, while others are at best touched on in passing. First, the procedural criticism of food safety policy holds that regulation should be democratized, meaning that it should reflect a broader swath of citizen values, and that it should especially not be controlled by individuals or organizations with a financial interest in marketing products. This is an issue that transcends food safety and applies equally well to the socioeconomic concerns. Indeed, it is relevant to the entire range of technological innovations. The cited critiques (and there are many others) are made relevant to food safety through the suggestion that existing regulatory procedures do not provide adequate assurance that unknown hazards might be overlooked in a review process that focuses on the safety of the intended gene product. One cannot be sure that some metabolic process under the control of a plant or animal genome has not been altered inadvertently through the process of inserting a gene construct with a known effect. I believe that the framework outlined in Chap. 2 does take account of this possibility, and I have included a discussion of this potential as it pertains to gene editing in Chap. 13. What is more, the final chapter of my 2015 book From Field to Fork explains why I do not find the inadvertent alteration argument to be a persuasive basis for either disallowing gene technology, or for requiring significantly more restrictive regulatory oversight (see Thompson 2015, pp. 227–239). Yet I have emphasized the underlying ethical principles that support such concerns, and have not directly addressed the question of whether specific regulatory approaches do or should (given concerns that arise in political theory) match the ethical analysis. As to the procedural arguments, the book as a whole is replete with arguments for democratizing the governance of emerging technology, and it emphasizes the role that a risk-based approach can play in doing so. This includes socioeconomic concerns, but food safety ranks high among the general population’s interest in food policy (Knight and coauthors, 2008). Activists hoping to constrain corporate exploitation and control of gene technologies or simply trying to expand democratic control over innovation for philosophical reasons have good reason to link their arguments to food safety because fear is a proven motivator for political change. In my case, however, that would also be deeply insincere. As much as I would like to see greater participation in steering the direction of technological innovation, the considerations discussed in Chap. 13 and my 2015 book explain why food safety risk is among the least of my worries for food biotechnology. Chapter 4 discusses why I believe that effective labeling would serve as an ethically appropriate response to those who do not share my confidence. There might be interesting points to discuss with respect to the way that Millstone, Wynne or Meghani approach these questions and the presumptive case I outline in Chap. 2, but as none of these authors has addressed my argument, I am without guidance as to how I might respond. The ethical rationale for including socioeconomic impacts in a risk-based approach was and remains a central focus of the book. Yet it is not immediately

xviii

Introduction

obvious that these concerns should be reflected in government regulation. Part of the implied social logic of technological innovation is that increasing the efficiency of production practices is generally, if not inherently, beneficial to society. Nevertheless, technology is a concern for social justice when specific products affect the distribution of economic rewards (and penalties) throughout society, or when less tangible social goods such as social cohesion and social legitimacy are damaged. Such impacts have been widely associated with agricultural biotechnology in rural settings. Some of the arguments have a history that extends back to the origins of the industrial revolution; others exemplify social concerns uniquely characteristic of the late twentieth century. Chapters 8 and 9 subdivide socioeconomic impacts into the social risks that yield enhancing agricultural technology imposes on rural communities, and the risks of shifting power relations through globalization, the rise of international corporations and the transformation of national and regional sovereignties. Readers desiring a more concise statement of my views on them should consult more recent articles, (see especially Thompson, 2014). As already mentioned above, the book’s treatment of agrifood biotechnology concluded with a critique of science-based risk assessment in its first and second editions. This chapter, entitled “Communication, Education and the Problem of Trust,” remains the ethical heart of the book. It argues for more participatory decision processes that include contrasting perspectives, and not only at the point of regulatory decision-making, as argued by Meghani (2017) and other recent critics (Hilbeck and coauthors, 2020). It is at this point that the book pivots from the advantages of a risk-based approach to some of its pitfalls. I argue that the conceptualization of risk that has dominated thinking among molecular biologists and regulatory scientists poses cognitive barriers to an effective communication process. Elements of this argument are now reflected in science communication scholars who criticize “the deficit model”, an approach that presumes telling people what they don’t know about science will fix everything. Although the original 1997 book discussed an important paper by John Zimon that laid down basic ideas in the deficit model critique (Zimon 1993), I did not use the term “deficit model,” in either previous edition. Although my argument in Chap. 12 overlaps with many elements of work by communication scholars who build upon the deficit model critique, it focuses more specifically on grammatical and cognitive aspects of discourse on risk. In this respect, it connects science communication theory to work on risk perception initiated in the 1970s by Amos Tversky (1937–1996), Daniel Kahneman and Paul Slovic. My revisions to this chapter have been designed to make these connections more obvious. As such, this chapter will, I hope, be of some interest to science scholars who work in these fields. I became aware of Sven Ove Hannson’s parallel work on risk after most of the work on the second edition had been completed. Aside from remarks in Chap. 14, the new edition does not attempt to reconcile my own analysis with Hannson’s (see Hansson 2013). I have subsequently published articles that make the arguments in this book apart from the context of debates over agricultural and food biotechnology (see Thompson 2012; 2018).

Introduction

xix

Although new work in Chaps. 13 and 14 considers some of the more recent innovations in plant and animal transformation, Chapter 12’s review of expertise, communication and trust still remains the part of the book that its author regards as most philosophically significant. In short, the practices of risk assessment and communication can be thought of as character defining, and the integrity with which they are carried out determines the extent to which the conglomeration of private firms, universities, public labs and regulatory agencies that control biotechnology is worthy of our trust. They may know what they are doing, but they also have ethical obligations to be more responsive to reasonable concerns. Science is not institutionally organized to fulfill these additional responsibilities, and that deficiency may be the most ethically damning fact about agrifood biotechnology. Chapter 14 is a brief essay reflecting on the relationship between these themes and the philosophy of technology.

Some Disclosures I have been accused of having been co-opted by financial rewards, sometimes rather rudely, by those whose view on biotechnology is less favorable than my own. In a book on ethics, it is reasonable to go well beyond what would normally be required in disclosing my financial and professional interests with respect to food biotechnology, and so I ask my readers’ forbearance in recounting some details. I have no personal investments in biotechnology firms. In 1997, I wrote that my research on biotechnology was a relatively small part of my total research portfolio, though it represented a larger percentage by 2007. Relatively, little of my work after 2007 addressed biotechnology. Research in philosophy is not expensive when compared to laboratory research, and the funding I received to conduct studies of biotechnology over the last 30 years is not large when compared to most university scientists. Research by philosophy professors is, in the vast majority of cases, done without external grant support of any kind. Humanities researchers seek funding to release them from teaching obligations. My research has required travel either to allow me opportunities to interact with scientists or activists in a non-structured fashion, or to sponsor workshops and symposia at which others are brought to me. My grants have also supported graduate students. A summary of the way that my research on agricultural biotechnology has been funded begins with approximately $150,000 I received from Texas A&M University’s Institute for Biosciences and Technology (IBT) over a period of 8 years from 1990 to 1997. These funds operated Center for Biotechnology Policy and Ethics (later renamed Center for Science and Technology Policy and Ethics and now defunct). They covered the cost of operating an office (secretary, telephone, copy machine), purchasing research materials (computers and published materials, in my case) and travel for A&M faculty to attend meetings or for bringing outside speakers to College Station. They also paid for the publication of a bimonthly

xx

Introduction

newsletter that reviewed topics of general relevance to bioethics and to science in society, not just food biotechnology. I did not draw salary from IBT funds. The second major source of support is my basic academic salary, paid by Texas A&M from 1981 until the summer of 1997, and by Purdue University from 1997 to 2003. I am now employed by Michigan State University. Virtually all of the time I spent working on biotechnology was covered by my base salary, rather than grants or contracts. It has been what university researchers call “unfunded research”. Since joining the Michigan State faculty in 2003, part of my salary has come from so-called Hatch funds. These are monies transferred to the university from U.S. Department of Agriculture (USDA). Recipients of Hatch funds are required to develop a project and to submit reports to USDA, and these Federal dollars have relieved me of some teaching responsibilities. My early Hatch project supported the revision of this book for a second edition. After this, my Hatch projects have not involved biotechnology until 2019, when I started a new project in light of changes wrought by gene editing. This project, MICL02620—Social and Ethical Issues in Advanced Biotechnology, is supporting the work I am doing on this revision. I received a small grant from the U.S. National Science Foundation (NSF) (Project SBR-9602968) in early 1997 when I was still working at Texas A&M University. But after the manuscript for the original book was largely completed, the total value of the grant was $49,787, major parts of which went to support the work of two colleagues at Texas A&M. I did derive some support from the grant for summer salary when I was editing the page proofs for the original book. These three sources, IBT, Texas A&M and NSF, provided the support that I needed to do my research for the first edition of Food Biotechnology in Ethical Perspective. With my move to Michigan State, my research turned toward sustainability, animal welfare and food nanotechnology, and with the exception of a brief collaboration on synthetic biology funded by Sloan Foundation, I have not received additional grant funds for biotechnology work since then. Sloan Foundation grant went to J. Craig Venter Institute, giving me additional opportunity for interaction with leading scientists in the field (though not with Dr. Venter himself). I realized a few weeks of salary from the project, which also covered my travel expenses for participating in project activities. I have received external grants throughout my career for non-biotechnology related work including grants for teaching projects from USDA, for work on environmental risk from the State of Texas, for work on ethics and development from Rockefeller Foundation and for other non-biotechnology projects from the NSF, including the large grant in 2004 for work on nanotechnology. In fact, my NSF biotechnology grant total of approximately $50,000 represents about 1% of my career total for external grants on all projects. At the time I wrote the first edition, I reported that I had also been supported in the form of travel and small honoraria to deliver lectures on ethics and food biotechnology, mostly at other universities. At that time, I had spoken on biotechnology at many American universities, and in England, Jamaica, Israel, Egypt and Thailand, and I could add a few other countries—Japan, China, Mexico —to the list now. Then as now, I addressed many meetings where I (or my own

Introduction

xxi

university) covered my expenses, but those who invited me to speak frequently covered my travel expenses. I reported in the first edition that I had received as much as $500 to make such presentations, above expenses, and had probably made somewhere in the neighborhood of five or six thousand dollars over the course of 10 years. In the interval since the book was published, I have received more opportunities to be remunerated for my work through similar kinds of contract arrangements. Canadian Biotechnology Advisory Committee, a quasi-government group, commissioned me to write a white paper that was the first draft of a paper that I wrote and published with a former student (Thompson and Hannah 2008). I served on a number of advisory boards. Some, such as Advisory Committee on Biotechnology of the Board of Agriculture and Natural Resources and National Research Council are unpaid, but I did advise two private firms, one of which is involved in biotechnology. The biotechnology company I advised was a start-up firm that paid me $2000 for comments on the potential ethical pitfalls in a business plans they were drafting. They are no longer in business. I signed a non-disclosure agreement with them and have not reported on any of their plans in this book or elsewhere. It does not violate the terms of that agreement to say that as far as I know, the product they were investigating has never even made it to regulatory review, much less appeared on the market. It would be difficult to make a precise estimate of how much I have made on speaking, writing and other projects related to biotechnology since I drafted the summary for the 2007 edition. I would estimate my lifetime total between $30,000 and $60,000, spread out over more than 30 years, (that is, between 1000 and 2000 a year, when averaged over my career). However, because this subject is so sensitive to many readers, I will provide a few details. I have already listed my $2000 windfall from a start-up firm. As the manuscript for the second edition was under review around 2006, I received my first and only invitation to consult with Monsanto, for which I was paid a flat fee of $1000 for a one-day meeting with top executives (they also took me to dinner). There were six to ten other outside consultants invited that day. When I was asked, what could they have done better, my advice was for them to encourage research on ethical issues, even when unfavorable to a planned product, because producing a robust public record of debate would provide more clarity about the context of developing products for agriculture. I do not believe that the company (now a part of Bayer) followed up on that advice. This brings my career total of private biotechnology industry supported income to $3000. I was also paid $2000 by Center for Ethics and Toxics to write a chapter for Engineering the Farm, a book that most would regard as hostile to agrifood biotechnology. By far the largest single source of income was a contract with the Commission on Environmental Cooperation, an inter-governmental agency for the United States, Canada and Mexico, which supports a number of research and consensus seeking activities involving the environment and trade. If memory does not fail me, I received $6000 for my contributions to a report on Mexican maize contamination, a report that was leaked to the public by Greenpeace, one of the leading anti-biotechnology civil society groups. In short, my income from

xxii

Introduction

anti-biotechnology groups exceeds my income from the biotechnology industry. Royalties from the book itself have been insignificant. The other source of income has come from universities or other organizations that at least attempt to take a non-partisan stance. The original Canadian paper I mentioned above paid something in the neighborhood of $1500, and a series of relationships with genome research centers in Canada ensued. I received comparable honoraria for my service on Science and Industry Advisory Committee for Genome Canada, which ended in 2008. There were also unpaid advisory roles for the U.S. National Research Council and the Genetics and Society Program at North Carolina State University, among others. I have received travel reimbursements (but not speaking or consulting fees) for participating in meetings of Biotechnology Industry Organization (BIO) (an industry trade group) and a workshop organized by DuPont. A few of the paid opportunities continued after the publication of the second edition. I received honoraria in the range of $500 a year for service on the advisory board of International Barcode of Life at the University of Guelph, for example. After a long hiatus, I have started to receive invitations again in the wake of debates on gene editing. In 2019, I received honoraria of about $1000 each for two lectures that presented some of the work contained in Chap. 13, maintaining the $2000 per year average noted above. These payments came from the Thomas Foley Center at Washington State University and from Canadian Society for Environmental Philosophy. I am sure that the vast majority of my non-salary income has come from universities, professional societies or other non-profit organizations. This could be considered a lot of speaking and a lot of travel (sometimes to attractive places, but more typically to dismal airport hotels). I would not dispute that it is a significant amount of money. Relatively few academic philosophers earn as much as $10,000 from philosophical employment beyond their teaching salaries over a professional lifetime, though leading figures (especially in medical bioethics) earn much more than this. I can only repeat what I wrote originally in 1997: if one were in it for the money, one would be wiser to take a critical stand. The summary just recounted suggests that anti-biotechnology groups have contributed more to my pocketbook than firms with a financial interest in selling something. Others may dispute my judgment. Indeed, an anonymous reviewer for the 2007 edition wrote “My impression is that there is money to be made in being a proponent and not taking a critical stand,” and notes that it may come in the form of travel and research contracts, rather than sales of books. I remain open to working with food industry firms, though I must insist that in my experience, this has not been a particularly lucrative research opportunity for philosophers. I regard much of the treatment I received as a result of my biotechnology work as unfair and even underhanded. I ran afoul of biotechnology’s critics more than once, and some of it can still be found by anyone who takes a bit of time to troll the web. I was accused of trying to murder Mexican peasants in order to steal their land when I spoke in Oaxaca. Websites still accuse me of advocating the blinding of chickens and of conspiring to promote the use of “Terminator” genes in the developing world. The latter accusation came about when a Purdue University staff writer posted a story intended to discuss the possibility of using genetically induced

Introduction

xxiii

seed sterility as a part of a strategy to limit the environmental risks from biotechnology (see Tally 2002). My experience with mainstream biotechnology companies is not, on the whole, positive, either. Prior to my involvement with them in 2005, Monsanto had placed another philosopher named Paul Thompson on its ethics advisory board, one who did not publish anything on agriculture or biotechnology until his 2013 book (which ignored my own work altogether). The timing of Monsanto’s invitation to the “other” Paul Thompson coincided with my appointment to the U.S. National Research Council committee that eventually produced the report Environmental Effects of Transgenic Crops (NRC 2002). It caused in an onslaught of attacks on my reputation from people who were accusing me of an industry connection that did not exist. Going further back, senior administrators who were in a position to know told me that one biotechnology company or another had pressured Texas A&M to close my program. Aside from 1 year when my budget for the Center was mysteriously cut from $40,000 to $2,000, Texas A&M stood firm in supporting my work. The life of an American college professor is an enviable one, and I feel very fortunate to have lived it. I have had a wonderful career and have been compensated better than the average philosophy professor for work that many would regard as being on the extreme margins of my discipline. Nevertheless, I believe it is not such a bad idea for philosophy professors to be more forthcoming in recounting what we have gained and not gained in pursuing a particular line of research.

References Champer, Jackson, Anna Buchman, and Omar S. Akbari. 2016. Cheating evolution: Engineering gene drives to manipulate the fate of wild populations. Nature Reviews Genetics 17(3): 146. Cotton, Matthew. 2014. Ethics and Technology Assessment: A Participatory Approach. Springer, Berlin, Heidelberg. Elliott, Kevin Christopher, and Ted Richards (eds.). 2017. Exploring inductive risk: Case studies of values in science. Oxford University Press, New York. Ewen, Stanley WB, and Arpad Pusztai. 1999. Effect of diets containing genetically modified potatoes expressing Galanthus nivalis lectin on rat small intestine. Lancet 354 (9187): 1353–4. doi:10.1016/S0140-6736(98)05860-7. PMID 1053386 Gaskell, George, Paul Thompson, and Nick Allum. 2002.Worlds Apart? Public Opinion in Europe and the USA. In Biotechnology: The Making of a Global Controversy, eds. M. W. Bauer and G. Gaskell, 351–375, Cambridge, UK: Cambridge University Press Hansson, Sven Ove. 2013. The ethics of risk: Ethical analysis in an uncertain world. New York: Palgrave Macmillan. Herring, Ronald J. 2007. Stealth seeds: Bioproperty, biosafety, biopolitics. The Journal of Development Studies 43:130– 157. Hilbeck, Angelika, Ilan H. Meyer, Brian, Wynne and Erik Millstone. 2020. GMO regulations and their interpretation: how EFSA’s guidance on risk assessments of GMOs is bound to fail. Environmental Sciences Europe Accessed June 24, 2020 at https://link.springer.com/content/ pdf/10.1186/s12302-020-00325-6.pdf. Hilder, Vaughan A., and Donald Boulter. 1999. Genetic engineering of crop plants for insect resistance–a critical review. Crop Protection, 18: 177–191.

xxiv

Introduction

Hull, David L. 2001. Science and selection: Essays on biological evolution and the philosophy of science. Cambridge, UK: Cambridge University Press. Kaiser, Matthias. 2009. Ethical aspects of livestock genetic engineering. In Genetic Engineering in Livestock: New Applications and Interdisciplinary Perspectives, eds. Engelhard, Margret, Kristin Hagen, and Matthias Boysen, 91–117, Dordrecht, NL: Springer. Knight, Andrew J., Michelle R. Worosz, Maria K. Lapinski, Toby A. Ten Eyck, Craig K. Harris, Leslie D. Bourquin, Thomas M. Dietz, Paul B. Thompson, and Ewen CD Todd. 2008. Consumer perceptions of the food safety system: implications for food safety educators and policy makers. Food Protection Trends 28: 391–406. Meghani, Zahra. 2017. Genetically engineered animals, drugs, and neoliberalism: The need for a new biotechnology regulatory policy framework. Journal of Agricultural and Environmental Ethics 30: 715–743. Millstone, Erik. 2007. Can food safety policy-making be both scientifically and democratically legitimated? If so, how? Journal of Agricultural and Environmental Ethics 20:483–508. Millstone, Erik, Eric Brunner, and Sue Mayer. 1999. Beyond ‘Substantial Equivalence’. Nature 401: 525–526. Mölders, Tanja. 2014. Multifunctional agricultural policies: pathways towards sustainable rural development? International Journal of Sociology of Agriculture & Food 21: 97–114. Myskja, Bjørn Kåre, and Anne Ingeborg Myhr. 2020. Non-safety assessments of genome-edited organisms: Should they be included in regulation? Science and Engineering Ethics https://doi. org/10.1007/s11948-00222-4. NRC (National Research Council). 2002. Environmental effects of transgenic plants: The scope and adequacy of regulation. Washington DC: National Academy Press. Plakias, Alexandra. 2019. Thinking through food: A philosophical introduction. Peterborough, ON: Broadview Press. Rollin, Bernard. 1997. Send in the clones... don’t bother, they’re here. Journal of Agricultural and Environmental Ethics 10: 25–40. Rudy, Alan P., and Dawn Coppin, Jason Konefal, Shaw, Byron T, T Ten Eyck, C J Harris and Lawrence Busch. 2014. Universities in the age of corporate science: The UC-berkeley-novartis controversy. Philadelphia: Temple University Press. Segelken, Ross A. 2016. Plant scientist Ralph Hardy, who led GMO foods debate, dies. Cornell Chronicle, Aug. 16. Accessed June 23, 2020 at https://news.cornell.edu/stories/2016/08/plantscientist-ralph-hardy-who-led-gmo-foods-debate-dies. Tally. S. 2002. Terminator tussle: Controversial technology needed, experts say. Purdue News. Accessed June 25, 2020 at https://www.purdue.edu/uns/html4ever/020418.Thompson. terminator.html. Taylor, Michael R., and Jody S. Tick. 2018. The star link case: Issues for the future. Washington, DC: Resources for the Future and Pew Initiative on Food and Biotechnology. Accessed Dec. 30 at https://web.archive.org/web/20130921055134/http://www.pewtrusts.org/uploadedFiles/ wwwpewtrustsorg/Reports/Food_and_Biotechnology/hhs_biotech_star_case.pdf . Thompson, Paul B. 2012a. Nature politics and the philosophy of agriculture. In ed. D. Kaplan, 214–232, Berkeley, CA: The Philosophy of Food University of California Press. Thompson, Paul B. 1998. Report of the NABC Ad-Hoc committee on ethics. Journal of Agricultural and Environmental Ethics 10: 105–125. Thompson, Paul B. 1999. Ethical issues in livestock cloning. Journal of Agricultural and Environmental Ethics 11: 197–217. Thompson, Paul B. 2012b. Ethics and risk communication. Science Communication 34: 618–641. Thompson, P. B., & Hannah, W. 2008. Food and agricultural biotechnology: A summary and analysis of ethical concerns. Advances in Biochemical Engineering and Biotechnology 111: 229–264. Thompson, Paul B. 2008. Current ethical issues in animal biotechnology. Reproduction, Fertility and Development 20: 67–73.

Introduction

xxv

Thompson, Paul B. 2010. Why using genetics to address welfare may not be a good idea. Poultry Science 89: 814–821. Thompson, Paul B. 2015. From field to fork: Food ethics for everyone. New York: Oxford U. Press. Thompson, Paul B. 2018. Communicating science-based information about risk: How ethics can help,” In Ethics and Practice in Science Communication, Priest, Susanna, Jean Goodwin, and Michael F. Dahlstrom, eds. 34–54, Chicago: University of Chicago Press. Wright, Addison V., James K. Nuñez, and Jennifer A. Doudna. 2016. Biology and applications of CRISPR systems: harnessing nature’s toolbox for genome engineering. Cell 164(1–2): 29–44. Wynne, Brian. 2001. Creating public alienation: Expert cultures of risk and ethics on GMOs. Science as Culture 10: 445–481. Zerbe, Noah, 2004. Feeding the famine? American food aid and the GMO debate in Southern Africa. Food Policy 29: 593–608. Zimon, Jan. 1992. Not knowing, needing to know, and wanting to know. In When Science Meets the Public, ed. Bruce Lewenstein, 13–20, Washington DC: American Association for the Advancement of Science.

Contents

1

Biotechnology in the Context of Agriculture and Food: An Overview . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.1 Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 1.2 The Puzzle of Heredity . . . . . . . . . . . . . . . . . . . . . 1.3 Altering the Genome . . . . . . . . . . . . . . . . . . . . . . . 1.4 Genetic Engineering and Biotechnology . . . . . . . . . 1.5 Recent Developments . . . . . . . . . . . . . . . . . . . . . . 1.6 What’s in a Name? . . . . . . . . . . . . . . . . . . . . . . . . 1.7 The Controversy in Ethical Perspective . . . . . . . . . 1.8 Conclusion: Beyond Risk and Back Again . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

1 1 3 6 10 11 14 18 20 22

2

The Presumptive Case for Food Biotechnology . . . . 2.1 Technological Ethics: A Précis . . . . . . . . . . . . 2.2 Ethics and Risk . . . . . . . . . . . . . . . . . . . . . . . 2.3 The Risk-Based Approach . . . . . . . . . . . . . . . . 2.4 The Logic of the Presumptive Case . . . . . . . . . 2.5 The Social Dimension of the Presumptive Case 2.6 Making the Case for Biotechnology Badly . . . . 2.6.1 The Modernist Fallacy . . . . . . . . . . . . 2.6.2 The Naturalistic Fallacy . . . . . . . . . . . 2.6.3 The Argument from Ignorance . . . . . . 2.6.4 The Argument from Hunger . . . . . . . . 2.7 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

25 26 28 30 32 36 37 38 41 42 44 48 49

3

Biotechnology, Policy and the Problem of Unintended Consequences: The Case of rBST . . . . . . . . . . . . . . . . . . . . . . . . . . 3.1 What is rBST? Why Does it Matter? . . . . . . . . . . . . . . . . . . . . 3.2 Biotechnology Policy and Philosophy . . . . . . . . . . . . . . . . . . .

53 55 57

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

xxvii

xxviii

Contents

3.3

rBGH: Assessing Unwanted Consequences . . 3.3.1 Food Safety . . . . . . . . . . . . . . . . . . 3.3.2 Animal Welfare . . . . . . . . . . . . . . . 3.3.3 Environmental Impact . . . . . . . . . . . 3.3.4 Social Consequences . . . . . . . . . . . 3.4 Ethical Disputes, Governance and Consensus 3.5 Social Consequences Redux . . . . . . . . . . . . 3.6 Learning from rBST . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

4

5

. . . . . . . . .

60 61 64 66 67 69 71 73 75

..........

79

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

81 83 86 86 88 89 90 91 92 93

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. 94 . 96 . 98 . 102 . 103

Animal Health and Welfare . . . . . . . . . . . . . . . . . . . . . . 5.1 Animal Biotechnology and Food . . . . . . . . . . . . . . . 5.2 Harm to Animals and the Risk-Based Approach . . . . 5.3 Drugs and Animal Feeds from Biotechnology . . . . . 5.4 Engineered Animals . . . . . . . . . . . . . . . . . . . . . . . . 5.5 The Moral Status of Animals . . . . . . . . . . . . . . . . . . 5.6 Rollin’s Consensus Morality . . . . . . . . . . . . . . . . . . 5.7 Animal Telos, Animal Integrity and Objections to Genetically Engineered Animals . . . . . . . . . . . . . 5.8 Against Changing Telos or Species Integrity of Food Animals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 5.9 Animal Biotechnology and Moral Obligation . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . . Politics . ...... . ...... . ...... .

Food Safety and the Ethics of Consent . . . . . . . . . . . . 4.1 The Ethics and Political Theory of Food Safety Regulation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.2 Safety Criteria and Biotechnology . . . . . . . . . . . . 4.3 Ethical Gaps in Food Safety Governance . . . . . . . 4.3.1 Bad Actors . . . . . . . . . . . . . . . . . . . . . . . 4.3.2 Collateral Consequences . . . . . . . . . . . . . 4.3.3 Social Uncertainty . . . . . . . . . . . . . . . . . 4.4 The Philosophy of Food Safety . . . . . . . . . . . . . . 4.4.1 Classification . . . . . . . . . . . . . . . . . . . . . 4.4.2 Purification . . . . . . . . . . . . . . . . . . . . . . 4.4.3 Optimization . . . . . . . . . . . . . . . . . . . . . 4.5 Classification and Purification Versus Risk-Based Optimization . . . . . . . . . . . . . . . . . . . . . . . . . . . . 4.6 Food Safety and Ethics . . . . . . . . . . . . . . . . . . . . 4.7 Food Labels and Consent . . . . . . . . . . . . . . . . . . 4.8 Food Safety Risks in Ethical Perspective . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

109 111 112 114 116 120 123

. . . . . . . . 125 . . . . . . . . 127 . . . . . . . . 131 . . . . . . . . 132

Contents

xxix

6

Ethics and Environmental Risk Assessment . . . . . . . . . . . . . . 6.1 The Environmental Debate . . . . . . . . . . . . . . . . . . . . . . . 6.2 Environmental Risk: An Expected Value Approach . . . . . 6.3 Expected Value and the Consequentialist Framework . . . . 6.4 Understanding Hazards and Harm . . . . . . . . . . . . . . . . . . 6.4.1 Acquired Pest Resistance . . . . . . . . . . . . . . . . . . 6.4.2 Weediness . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 6.4.3 Genetic Diversity . . . . . . . . . . . . . . . . . . . . . . . . 6.5 Probability, Precaution and the Quantification of Exposure 6.6 Exposure Quantification: Further Issues . . . . . . . . . . . . . . 6.7 Philosophy of Science and Risk Communication . . . . . . . 6.8 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

7

Environmental Impact and Environmental Values . . . . . . 7.1 Environmental Impacts from Agrifood Biotechnology . 7.2 Environmental Hazards and Environmental Values . . . 7.3 Environmental Philosophy For and Against Expected Value . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 7.4 The Conceptual Landscape of Environmental Ethics . . 7.5 Environmental Ethics in Consumption . . . . . . . . . . . . 7.6 Environmental Ethics in Production . . . . . . . . . . . . . . 7.7 Ethics and Environmental Responsibility for Food Biotechnology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

. . . . . . . . . . . . .

137 138 140 141 143 144 145 146 149 153 156 160 161

. . . . . . . 167 . . . . . . . 168 . . . . . . . 171 . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

173 177 178 181

. . . . . . . 186 . . . . . . . 189

8

Social Impact and the Technology Treadmill . . . . . . . . . . . . 8.1 Technology, Politics and the Prediction of Social Change 8.2 The Social Consequences of Food Biotechnology . . . . . . 8.3 Theories of Justice . . . . . . . . . . . . . . . . . . . . . . . . . . . . 8.3.1 Utilitarianism and Utilitarian Theories of Justice 8.3.2 Justice and Rights . . . . . . . . . . . . . . . . . . . . . . 8.3.3 Justice and Virtue . . . . . . . . . . . . . . . . . . . . . . . 8.4 Structural Injustice and Structural Criticisms . . . . . . . . . . 8.5 Social Consequences for Small and Family Farms . . . . . 8.5.1 Family Farms: Utilitarian Arguments . . . . . . . . . 8.5.2 Family Farms: Rights and Fairness . . . . . . . . . . 8.5.3 Family Farms and Moral Virtue . . . . . . . . . . . . 8.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

. . . . . . . . . . . . . .

193 194 196 201 201 203 205 207 211 211 212 214 216 217

9

Can 9.1 9.2 9.3

. . . .

. . . .

. . . .

. . . .

. . . .

223 224 227 231

Agrifood Biotechnology Help the Poor? . . . The Ethics of Agricultural Research . . . . . . . Ethics and Agricultural Development . . . . . . Social Consequences in Peasant Agriculture .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

xxx

Contents

9.4 9.5

World Feeders and Ethical Consumers . . . . . . . . Social Consequences for the Conduct of Science 9.5.1 The Scientific Purity Argument . . . . . . . 9.5.2 The Social Function Argument . . . . . . . 9.5.3 The Public Trust Argument . . . . . . . . . . 9.6 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

. . . . . . .

234 239 240 241 243 244 246

10 Conceptions of Property and the Biotechnology Debate . 10.1 Property Rights in Genetic Information: The Context 10.2 Ownership in Genetics: A Contested Subject . . . . . . 10.3 The Theory of Property . . . . . . . . . . . . . . . . . . . . . . 10.4 Instrumental Conceptions of Property . . . . . . . . . . . . 10.4.1 Libertarian Theory . . . . . . . . . . . . . . . . . . . 10.4.2 Utilitarian Theory . . . . . . . . . . . . . . . . . . . . 10.5 Ontological Conceptions of Property . . . . . . . . . . . . 10.5.1 Natural Law Theory . . . . . . . . . . . . . . . . . . 10.5.2 The Labor Theory of Property . . . . . . . . . . 10.6 Linking Instrumental and Ontological Theories of Property . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.7 Property and Agrifood Biotechnology: Ontological Approaches . . . . . . . . . . . . . . . . . . . . . 10.7.1 Ruling Out Ownership of Human Genes . . . 10.7.2 Rivalry and Excludability . . . . . . . . . . . . . . 10.7.3 Do Scientists Own Their Labor? . . . . . . . . . 10.8 Property and Agrifood Biotechnology: Utilitarian Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.9 Property and Agrifood Biotechnology: Libertarian Approaches . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10.10 Against Property Rights in Food Biotechnology . . . . 10.11 Ownership and Commodities . . . . . . . . . . . . . . . . . . 10.12 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

. . . . . . . . . .

251 253 257 260 262 263 263 264 264 266

. . . . . . .

. . . . . . .

. . . . . . . . 267 . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

268 269 271 272

. . . . . . . . 273 . . . . .

. . . . .

11 Religiously Metaphysical Arguments Against Agrifood Biotechnology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.1 Intrinsically Wrong . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.2 Situating Religious and Metaphysical Claims . . . . . . . . 11.3 Metaphysical and Religious Arguments: An Exhibitive Introduction . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.4 Analyzing the Religiously Metaphysical Case Against Genetic Engineering . . . . . . . . . . . . . . . . . . . . . . . . . . 11.5 Religious Statements on Genetic Technology Through 2005 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

. . . . .

275 277 280 282 283

. . . . . . 287 . . . . . . 288 . . . . . . 290 . . . . . . 292 . . . . . . 294 . . . . . . 298

Contents

xxxi

11.6 The Religious Case Against Property Rights in Genes: 1985–1995 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 11.7 Academic Theology and Genetic Engineering . . . . . . . 11.8 The Ethical Implications of Religious Views . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

. . . .

301 304 308 310

12 Communication, Education and the Problem of Trust . . . 12.1 Public Opinion and the Ethics of Gene Technology . . 12.2 The Problem of Trust . . . . . . . . . . . . . . . . . . . . . . . . 12.3 Communication and Public Understanding of Science . 12.4 The Ethics of Science Communication . . . . . . . . . . . . 12.5 The Problem of Risk . . . . . . . . . . . . . . . . . . . . . . . . . 12.6 Human Action, Risk, and Responsibility . . . . . . . . . . 12.7 Equivocation Problems and False Authority . . . . . . . . 12.8 Moral Reductionism and Political Exclusion . . . . . . . . 12.9 Conclusion . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

. . . . . . . . . . .

313 314 316 319 322 324 328 332 334 338 339

13 Gene Editing, Synthetic Biology and the Next Generation of Agrifood Biotechnology: Some Ethical Issues . . . . . . . . . . 13.1 Synthetic Biology: The Old New Thing . . . . . . . . . . . . . 13.2 Gene Editing and the Next Generation of Biotechnology . 13.3 The Products of Gene Editing in Food and Agriculture . . 13.4 Alternative Proteins . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.5 Novel Agricultural Ecosystems . . . . . . . . . . . . . . . . . . . 13.6 Gene Drives . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13.7 Conclusion: Should We Worry? . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

. . . . . . . . .

343 344 346 352 354 358 364 368 370

. . . . . . . .

. . . . . . . .

. . . . . . . .

375 376 378 381 386 389 391 394

14 Biotechnology, Controversy and the Philosophy of Technology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14.1 The Global Controversy: What Role for Philosophy? . . . . . 14.2 Philosophy of Technology . . . . . . . . . . . . . . . . . . . . . . . . . 14.3 Risk Assessment and Its Enemies . . . . . . . . . . . . . . . . . . . 14.4 Environmental Pragmatism: A Prolegomena . . . . . . . . . . . . 14.5 Technological Pragmatism: Ihde . . . . . . . . . . . . . . . . . . . . 14.6 Technological Pragmatism: Postphenomenology Expanded . 14.7 Gene Technology as a Moral Apocalypse . . . . . . . . . . . . . . 14.8 Conclusion: Technological Pragmatism and World-Feeding Ideology . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . References . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . .

. . . 396 . . . 398

Index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 401

Chapter 1

Biotechnology in the Context of Agriculture and Food: An Overview

1.1 Introduction Human beings have used tools to affect food since the prehistoric era. Most obviously, the use of sharpened stones for hunting and fire for both cooking and clearing landscapes are among the first know forms of technology. Even in this era, human activity affected the genetics of edible plants and animals, if only through the evolutionary pressure that hunting and gathering placed upon other species in the human environment. Direct genetic alteration coincides with the beginnings of agriculture. Humans chose plant specimens to cultivate giving preference to those with valuable traits such as ease of cultivation, non-shattering, and reliability of harvest. Taste may have also figured in the plants they favored. Humans have practiced trial and error forms of crop and animal breeding for millennia. They may not have known that they were making genetic changes in the foods that they were raising, but that was, in fact, what accounted for difference between wild types and the cultivated varieties that were preferred. In comparison, the era of biotechnology is only a few decades old. While people disagree about the precise meaning of biotechnology, the first techniques for altering plant and animal genomes using recombinant DNA became widely available to agricultural scientists in the 1980s. These techniques, and the plants and animals developed by using them, mark the most straightforward way to characterize food and agricultural biotechnology. Agricultural applications of biotechnology (also described as genetic engineering) were controversial from the very outset. One of the first experiments with an organism genetically engineered for agricultural use sparked debates about the oversight of environmental consequences. A bacterium modified to affect the formation of ice crystals on crops challenged the regulatory apparatus and sparked rancorous debate, (Gaertner and Kim 1987). Rachel Schurman and William Munro demonstrate the complexity of the controversy in their social history of the conflict as it developed from this dispute over the environmental risks of a single experiment through widespread opposition to gene technology that emerged in the first decade © Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_1

1

2

1 Biotechnology in the Context of Agriculture …

of the 21st century. Critics with concerns about animal welfare, the environmental impact of agriculture, the growing political power of science and technology, the expansion of global capitalism and the potential for eugenic uses of genetic modification on human beings were able to find common cause in opposing the use of genetic engineering in foods. Companies developing these products were utterly insensitive to these concerns and mounted countermeasures to stifle and marginalize critics. As these controversies matured, it became clear that ethical issues were at their heart, (Schurman and Munro 2010). Some controversies in agricultural use of rDNA techniques have their roots in agriculture itself. Does a given farming method or cropping system involve systemic violations of human rights? Plantation production of cotton in the American South before the end of slavery certainly did, and one can argue that farmers who depend upon migrant labor continue to commit such violations to this day. Does an agricultural technology conserve or exploit the natural resource base on which it depends? The use of guano for fertilizing soils throughout the 19th century was unsustainable because the global supply of guano was limited. The farming methods that relied on soil supplements would have crashed (with substantial shortfalls in the food supply) were it not for the invention of the Haber-Bosch process, which used fossil fuels to produce synthetic fertilizer. Synthetic fertilizer has high energy costs and contributes to climate change, so it is not at all clear that this technology makes food production more sustainable. These justice and environmental questions are typical for agricultural ethics. They intersect with debates on the use of genetic engineering, but genetic engineering is in no way a decisive criterion for resolving them. In some instances, questions about the use of gene technology turn upon these deeper or more comprehensive questions in agricultural ethics. The range of ethical questions that arise in agriculture is wide, but the scope of this book is narrow. This chapter will discuss some of the basic techniques for modifying the traits of plants and animals consumed for food. This presupposes some background knowledge in applied biology and genetics, but the goal in this chapter is keep the needed background knowledge to a minimum. Understanding the biotechnology debate also requires us to understand some things about people. Why did agricultural scientists and seed companies think that using genetic tools to alter agricultural crops might be a good idea? Why are many people who do not work in this area suspicious? These latter questions spill back into broader questions about the ethics of food and agriculture, and no concise approach could possibly cover everything of relevance. This chapter will provide background knowledge by explaining how practitioners in the mainstream of plant and animal science make sense of what they are doing, and proposing some plausible hypotheses to explain why others see it differently.

1.2 The Puzzle of Heredity

3

1.2 The Puzzle of Heredity Genetics is a way to make sense of heredity. For many scientists, it is more than this, but surely everyone can agree that genetics explains the heritability of certain traits. We can all see how some physical characteristics are passed from one generation to another. The fact that facial characteristics, hair and eye color, body type and even certain dispositions and talents are common to parent and child was certainly known by most observant individuals in antiquity. The process or mechanism that accounts for this, however, has been explained in many different ways. It would not have been unusual for highly educated people in 1800 to have thought that hereditary traits were contained in a bodily fluid, and that sexual reproduction involved the transfer of this fluid from the male to a receptacle in the female body. Early 19th century views of reproduction held that this receptacle (e.g. the woman) was completely lacking in ability to affect the heritable characteristics of the fetus. On this view, women contribute nothing to heredity. Such views reflected sexist and racist prejudices in some cases, and presumptions about human exceptionalism, in others. They were contrary to the practical folk knowledge of any livestock farmer, who would have known that the female also has hereditary influence. By 1900, no one with a college education would have entertained such a notion, but only someone more advanced training in biology would have associated heredity with genes. Charles Darwin (1809–1882) developed his approach to the evolution of species without the benefit of this fundamental concept. Up to a point, both theoretical and practical knowledge can function with a rather broad understanding of heredity that associates it with some material or bodily “stuff” transferred from parent to progeny during sex. That “stuff” might have been referred to as seed, blood, seminal fluid (not necessarily limited to males), gemmules, or genes. People operating with any of these seemingly distinct mental pictures might have had a functionally equivalent understanding of how heredity works. Although quite limited from the standpoint of contemporary molecular biology, such an understanding would have allowed many crop and livestock producers to make substantial improvements to the genetic stock of their plants and animals. This simple but functional understanding of heredity is arguably not unlike that of many people who today conduct successful careers in everything from plumbing to the practice of law. Despite whatever their school biology teachers may have tried to tell them, they get by with an understanding of heredity that does not really depend on detailed genetic science. By 2000, one did not need to be well educated to associate heredity with something called genes, but this did not imply that one understands how inheritance works at the level of genes. If we start, then, from the idea that otherwise intelligent people just think of genes as the stuff that carries hereditary traits, we should not be surprised that proposing to alter genes would puzzle them. The feeling might be a bit like taking your automobile to the garage and having the service department suggest replacing the engine (which you understand no better than genes, gemmules or seminal fluids) with crystals, flubber or a new device that channels cosmic forces directly to the wheels. We tend to trust our mechanics, but we do bring a broadly formed set of implicit beliefs to our

4

1 Biotechnology in the Context of Agriculture …

understanding of what makes an automobile work. The analogy highlights the critical role of trust (a theme explored in Chap. 12). Many people will be satisfied to place matters entirely in the hands of technical experts, but others will qualify their trust when technicians appear to be operating in violation of their basic understandings. Yet making any sense of the agricultural sciences involves a more detailed and technically complex interpretation of what they think they are doing when they develop new breeds or varieties. The following account will probably test the patience of some readers, though not necessarily for the same reasons. The notion of a gene that an early 20th century biologist was working with took genes to be microscopically small substances that were so strongly associated with observable traits of plants and animals that they could be thought of as determining them completely. On this view, having the gene causes the individual to have the trait, or alternately an individual’s genes explain the presence or absence of characteristics such as eye color, stature and other physical features, including certain diseases. Every individual inherited some genes from the male parent and others from the female parent, and that individual’s body-type (or phenotype) reflects the mixture. Early genetics recognized that genes were passed on in pairs, one from the male and one from the female. Traits could be dominant or recessive: An individual with only one gene from a dominant trait would exhibit the trait, while recessive traits would only be observed in individuals who inherited the same gene from both parents. This understanding paved the way for developing mathematical models of inheritance that became widely used by plant and animal breeders throughout the first half of the 20th century. Biologists did not have a detailed model of deoxyribonucleic acid (DNA), the molecule that transmits heritable traits before the 1950s. Educated non-specialists may be able to recite facts about the helical structure of DNA or its discovery by Rosalind Franklin (1920–1958), Maurice Wilkins (1916–2004), Francis Crick (1915–2004) and James Watson (b. 1928). However, the helix itself is less significant for understanding biotechnology than the discovery that “genes” are, in fact, sequences of paired chemical bases. The DNA molecule is itself a very long string of the chemical bases cytosine (C), guanine (G), adenine (A) and thymine (T). These chemicals associate in pairs, one on each side of the double helix. The sequence of base pairs on any organism’s DNA molecule is the genome. This paved the way for a subtle but important shift in the way that biologists understood genes. Rather than directly causing a trait, DNA functions as a “code” that determines many of an organism’s life functions through transcription of RNA (ribonucleic acids). RNA is causally responsible for initiating the action of cellular machinery needed to carry out the metabolic functions of life. Both DNA and RNA are chemical chains of the simpler chemical bases just mentioned. While a more detailed account of these chemical structures would certainly be requisite for a college level understanding of molecular biology, the point here is simply to note that a description of chemical sequences and the work that they do replaces the notion of a gene as a discrete entity that passes from parent to child. Some sequences in DNA code for RNA, while others do not. Regulatory sequences play a role in determining when RNA will be transcribed, or identify which parts of the genome code for RNA. In

1.2 The Puzzle of Heredity

5

biology, the idea of a self-contained gene that directly affects heredity has been replaced by a more complex system. Nevertheless, biologists continue to use the word ‘gene’ as a reference point for segments of the genome that play a role in a wide variety of biological functions. The double helix is significant for explaining cellular division and growth or sexual reproduction—two processes that would rank high in any basic biology course. When cells divide in the normal process of growth, the two strands of the double helix split. The metaphor of opening a zipper is apt. A chemical process (polymerase chain reaction or PCR) then matches up each single strand, yielding two identical double helix DNA molecules. Each DNA molecule contains the instructions to build a new cell, so two cells with the same DNA (the same sequence of base pairs) exist where only one existed before. DNA also contains instructions (coding sequences) that determine a cell’s structure and function depending on its particular location in a growing organism. That is, bodily materials as different as bone and blood can be built in accordance with the same instruction manual (the same DNA molecule), so an individual’s body will have an exact (or nearly exact) copy of the DNA molecule that was formed when the male and female parts unite. That process of sexual reproduction does indeed produce a mixture of both parents’ genetic material, just as early 20th century geneticists thought. It does so when strands of DNA from one parent pair with strands from the other, producing an entirely new and unique genome. That is, DNA can recombine in ways that that introduce novelty in the sequence: the sequence of a child is not exactly like that of either parent, though it does contain genes (sequences) from both. Farmers have always tried to exert some control over this process in hopes of getting progeny that were similar to the most productive plants in their fields (or livestock in their pens). The progress in 20th century biology that leads to biotechnology consists in learning how to exploit the recombinant aspects of DNA to both gain more targeted control over heredity and to introduce sequences (e.g. genes) that would have been virtually impossible to occur through ordinary sexual reproduction. The means of doing that are now understood to be associated with managing which strands of DNA (which genes) are incorporated into the genome. There is one last point that is worth making about heredity. Most people know that eye color is a hereditary trait, and as such, we would not find it strange to discuss the genes for blue, brown or green with respect to eye color. Scientists will talk this way, too, but they will also talk about “the gene for eye color” in a way that is intended to include all of these color variants. When they do this, they are referring to a particular location on the DNA molecule where the base pairs that code for eye color are found. This locus on the genome will not vary significantly from one individual in a species to another. (There are exceptions, but there are exceptions to almost everything in this chapter). Hence, when scientists talk about “the human genome” they are not, in fact, talking about a specific DNA molecule at all (though 99.9% of any human being’s genome is just like any others). They are talking about the order in which sequences that code for different functions (eye color, baldness, cystic fibrosis) line up, and sequences found at a given point on the genome become “the gene” for specific traits, like eye color. The more scientifically

6

1 Biotechnology in the Context of Agriculture …

rigorous way to talk about the difference between one individual’s sequence that codes for blue eyes and another individual’s sequence that codes for green eyes is to say that they possess different alleles. Alleles are variants in the sequence of a gene that produce observably distinct traits in the phenotype. A population may contain many individuals with many different alleles, yet one could still say that they all have the same gene (e.g. the gene that controls eye color). For some readers all of this repeats very basic concepts that, they might presume, everyone knows, while for others it has already gone far too deeply into details without really giving them the understanding needed to make sense out of things they were not really all that interested in, in the first place. I have always argued that it is possible to participate meaningfully in the ethical debates over biotechnology without having detailed understanding of the underlying biochemistry and genetics. Yet it will be helpful to understand that when molecular biologists talk of altering genes, or even adding new genes that originated in other species, what they are talking about is making additions and deletions of base pairs in a genome that is itself composed of billions of base pairs. The notion of a gene has itself become context dependent, though the word is still used broadly to describe elements of the genome that play a role in the heritability or reproduction of specific traits and characteristics, (Snyder and Gerstein 2003; Rheinberger and Müller-Wille 2018).

1.3 Altering the Genome Although some scientists dream of inventing functional sequences of base pairs that have never existed before, when present day plant and animal breeders say that they are changing genes, what they are talking about is altering a plant or animal’s genome. What they produce is colloquially called a GMO. GMO is short for “genetically modified organism,” but it is not a term with a scientifically precise meaning. Given the working definition of biotechnology introduced above (e.g. the application of recombinant DNA in changing a genome), all products of biotechnology are GMOs. In practice, however, the term is not usually applied to genetically engineered organisms produced for medical purposes. In some regulatory systems, classification as a GMO (or the similar LMO, for “living modified organism”) determines whether developers of a plant variety or animal breed must apply for government permission to develop or market it, or in some instances even do research. Chapter 13 discusses how gene-editing technologies present challenges to such classification systems. More broadly, some object to the terminology of GMOs in all contexts: Aren’t all food crops and livestock species genetically modified, they ask? They might suggest that the acronym GE (for genetically engineered) is more accurate. (Jiang and coauthors 2018). There are philosophical issues lurking here: Who gets to say what an expression means? Without pretending to settle this question, I will occasionally use both GMO and GE equivalently, except in cases where a distinction is explicit. All of the genes that are currently incorporated into GMOs originated in some plant, animal or microbial species that evolved its genetic structure under natural

1.3 Altering the Genome

7

conditions.1 As defenders of biotechnology never tire of pointing out, changes in the genomic sequence of the plant and animals that we eat have been underway for a very long time, (see, for example, McHughen 2000). As farming mechanized throughout the 19th century, farmers came to value seed that would grow into plants that all ripened or fruited at the same time. Mechanical harvesters move through a field in one pass, and typically destroy plants in their wake. Late ripening individuals were thus lost. Even something as simple as uniformity in height could be valuable to farmers, as machinery tended to either damage or overlook plants that are unusually short or tall. Botanists such as Luther Burbank (1849–1926) used trial and error to create varieties of familiar plants that would reliably reproduce traits. Some traits were valued for their taste and cooking properties, but more typically, agronomic traits (including drought, frost or disease resistance) that increase yields are valued because of the way they perform in a farmer’s field. Nevertheless, the development of quantitative genetics had significant effects on the changes plant and animal breeders make. The scientists who rediscovered the work of Austrian monk Gregor Mendel (1822–1884) and then improved upon Burbank’s approach were working on plants or animals for agriculture. This is the history that boosters of biotechnology are pointing to when they assert that there is nothing new about GMOs. Even by the 1920 s, plant breeders had developed an astonishingly diverse array of fruits and vegetables using trial and error or the tools of first generation breeding. Cabbage, broccoli, cauliflower, kale, Brussels sprouts and collard greens are all members of the same species, Brassica oleracea, despite the differences in their taste and appearance. These are common foods whose current shape, taste and growing characteristics were developed over the last 200 years. While one might think that crossing a Brussels sprout with a kale plant is “crossing species boundaries,” that is just biologically false. Breeders can also accurately say that their ability to develop plants and animals with characteristics that people wanted was never constrained by species boundaries in any absolute sense. Farmers of earlier generations would have been quite familiar with the mule, a work animal obtained by cross breeding of two distinct species, Equus ferus caballus (the domestic horse) and Equus africanus asinus (the donkey). The genomes of these distinct species are sufficiently similar to produce viable offspring through sexual coupling of a mare and a jack, despite the difference in the number of chromosomes. A second example is the garden strawberry, which is a cross between tiny wild strawberries from Virginia (Fragaria virginiana) and the larger (but less tasty) coastal strawberry (Fragaria chiloensis) of Latin America. One might think that crossing species boundaries is a new thing in plant or animal breeding, but that is just biologically false. 1 Exceptions

involve mutations that arose in response to high levels human-imposed stress, such as when seeds are exposed to radiation or plants are grown in environments that have been manipulated to have unusual concentrations of specific chemicals. Still, even in mutation breeding, scientists do not design or specify the novelty in a sequence of base pairs (Bado and coauthors 2015). That kind of change remains the dream of synthetic biology, and there is debate as to when and under what conditions a scientist might be said to have achieved it, (Deplazes-Zemp et al. 2015).

8

1 Biotechnology in the Context of Agriculture …

Plant breeders can also point to the occasional success of mutation breeding, where they induce genetic changes by exposing seed to high doses of chemicals or radiation. According to one source, more than 2500 varieties of plants developed through mutation breeding have been released for use by farmers and gardeners. Many of these varieties have proven to have traits of great value, (Ahloowalia et al. 2004). In his book The Wizard and the Prophet, Charles Mann reports that M.S. Swaminathan used mutation breeding to develop wheat and rice varieties used widely throughout Asia (Mann 2017, pp. 415–416; 442). One might think that none of our familiar foods had been altered through human induced change in their genetic structure prior to the introduction of gene transfer methods in the 1980s and 1990s, but that is just biologically false. It is eminently reasonable to be skeptical of any approach that induces change in a plant genome under conditions that would be unlikely to occur in nature. Indeed, most of the individual seeds subjected to the stress of mutation breeding do not produce viable plants. Plant breeders have developed confidence in the procedure because their experience is quite different from that of most people. This seemingly unexceptional point is highly significant. For most of us, increasing the chance of producing non-viable offspring is problematic from the start. The reproduction of human beings (or indeed many animals) presumes norms for viability, growth and development. Plant breeders feel no moral qualms about causing fatal damage to the genetic structure of many (thousands, perhaps) plants in search of the one change that, by pure chance, develops a functional change in its genome. These useful variants are called sports. The unusable, “weird” or nonviable plants subjected to radioactive or chemical stress are just disposed of. In discussions of evolution, words like selection and choice are used in a metaphorical sense. In plant and animal breeding, a scientist makes a real choice, an intentional selection, of the specific individuals that they want to work with, and they discard the rest. Thus in mutation breeding (as indeed in all forms of breeding, including biotechnology) it is not as if every genetic change gets incorporated into foods we eat. Not only do plant breeders choose which change to retain, before releasing a cultivar for use, they will have crossbred it back with other standard varieties in order to combine the novel trait with other characteristics that farmers also want. The breeder will have observed many (generally at least 7) generations of the plant and seen that it behaves like a typical individual from the plant type of which it is a token. Conventional plant breeding also makes extensive use of tissue culture. Tissue culture is a technique for regenerating a whole organism from a very small cluster of cells that have differentiated to a particular cell type, such as a root or a leaf. It exploits a non-sexual reproductive pathway for (most) plants that is simply not available for humans and other vertebrate species. Tissue culture can begin with only a few cells from a plant, and through manipulating the in vitro medium in which cells are maintained, the breeder stimulates the cell mass to differentiate into all the different cell types (e.g. flowers, leaves, etc.). Tissue culture artificially replicates the divisions and cell specializations that normally take place when plants grow from a single seed. Tissue culture has been used since the late 19th century, but it is only in light of contemporary knowledge of how regulatory sequences govern the

1.3 Altering the Genome

9

process of cell differentiation that it has clearly come to be understood as a form of reproduction that involves the manipulation of genes, (Sathyanarayana and Verghese 2007). Embryo rescue is a particular form of tissue culture that has significantly expanded plant breeders’ ability to have success with interspecies crosses, even when the two species have significantly different genomes, (Bridgen 1994). Animal breeding is largely a different story. Beyond the use of quantitative genetics and the occasional cross-species hybrid, animal breeders have mainly continued to rely on combining sperm and egg to achieve novel combinations of genetic material. Prior to the biotechnology era, breeders used tissue culture to maintain and manipulate animal cell cultures in vitro, but not to the end of regenerating an entire individual. The primary technological advances for animal breeding have been artificial insemination and in vitro fertilization (IVF), with the implantation of fertilized eggs into the uterus of a female. Readers are probably aware of IVF from its use in human beings, where it was pioneered in the 1970 s. Cattle breeders now use IVF frequently. However, it would be misleading to understate the power of quantitative genetics in conventional breeding for livestock species. Indeed, ethical studies of animal production often note the way in which conventional breeding has led to the development of animals with increased susceptibility to disease or injury, or vulnerability to other welfare issues such as the large breasted turkey’s inability to perform normal acts of copulation (Gamborg and Sandøe 2002). This discussion of breeding falls short of the educational ambitions that any biology teacher would have in covering similar topics. It also may go more deeply into the gory details about the plants and animals we eat than many people want. Yet most agricultural scientists and many practicing farmers would have had some familiarity with this history of change in the methods for altering genomes when the gene transfer techniques of modern agricultural biotechnology began to be discussed in the 1980 s. They certainly did not think that the strawberries, corn or wheat varieties and breeds of poultry or beef cattle that appear on dinner tables were the same or even highly similar to what would have been at the time of Socrates (d. 399 BCE), Julius Caesar (100 BCE-44 BCE) or even George Washington (1732–1799). What is more, the routine utilization of these techniques for a half century (without apparent incident) had generated an expectation that the standards and practices in use by plant and animal breeders were generally acceptable. The presumption that past standards and practices were adequate needs to be subjected to two significant ethical qualifiers. First, as previously noted, animal breeding was already coming under significant criticism for unintended but unacceptable consequences for animal welfare, (Sandøe and coauthors 1999). Second, plant breeders were certainly aware that conventional breeding was not without hazards. Toxic tropane alkaloids, for example, are produced plants of the genus Solanum, which includes tomatoes and potatoes as well as deadly nightshade. Although it is safe to eat potatoes and tomatoes, the green parts of these plants contain toxins. The alteration of regulatory genes that govern whether tropane alkaloids are produced in the tuber or fruit can happen simply through interbreeding with related varieties (Oksman-Caldentey and Arroo 2000). It was not for nothing that people avoided tomatoes for decades based on their relationship to belladonna, (Smith 2013). The

10

1 Biotechnology in the Context of Agriculture …

qualifier here, however, is not to imply that procedures were inadequate. To my knowledge, we do not have an episode of toxic potatoes or tomatoes coming out of plant breeding programs in industry or the public sector ever reaching consumers. The point, rather, is that we must see that the standards and practices of conventional plant breeders include their knowledge of risks, and their willingness to look out for such hazards.

1.4 Genetic Engineering and Biotechnology How do genetic engineering, biotechnology and GMOs differ from other methods that breeders and agricultural scientists have used to alter plant and animal genomes? The short answer notes two main points of difference. First, as a process, biotechnology uses recombinant DNA (rDNA) to accomplish changes in the genome. These other methods are exploiting the fruits of recombination, and in each case, including very early forms of trial and error breeding, changes in the genome occurred. Yet the farmers and breeders did not change the sequence of chemical bases that make up a molecule of DNA by acting directly on strings of C, G, A or T. They did not even have to understand that change in the sequence of base pairs had anything to do with the changes they sought. Second, from a practical standpoint biotechnology enables much more control over which specific genetic constructs (strands of DNA) are being altered. All of the methods described above involve quite a bit of randomness about which genes are being introduced into a plant and animal genome. In fact, the only way to see what change had been made was to grow the seed out and see how it performed in response to the water, sunlight, temperature, soils and other features of a crop’s environment. Scientists who engage in this work describe the alterations they are now making with biotechnology as enabling a much higher degree of precision than could be achieved with older techniques. The ability to target which genes will be altered is the cumulative result of many specific discoveries and innovations. They include the ability to identify a locus on the genome, to identify the sequence, to distinguish coding and regulatory DNA and to insert new sequences into the genome of a microorganism, plant or animal. The earliest applications of gene transfer were done on prokaryotes: microorganisms such as bacteria where DNA is not contained in a cell nucleus or membrane. Prokaryotes can be modified simply by injecting DNA into the organism where it will be taken up into its genome. For example, the human gene for production of insulin was identified and then inserted into a bacterium. This enabled the production of high-grade insulin for diabetics, (Johnson 1983). Similar ideas launched the pharmaceutical biotechnology industry, including the animal drug rBST (discussed at some length in Chap. 3). Chymosin, an enzyme for cheese making traditionally harvested from slaughtered calves, is now produced in a genetically engineered bacterium, (Mohanty and coauthors 1999). Up to the present, the two most utilized techniques for getting DNA into plants are agrobacterium and the gene gun. Agrobacterium is a natural organism—a prokaryote—that infects many plant species, inserting bits of its own DNA into the

1.4 Genetic Engineering and Biotechnology

11

infected material. The bacterium itself can be modified to carry the new construct into the infected material. A normal (that is, uninfected) plant can be reconstituted from these modified cells using tissue culture. The gene gun literally blasts microscopically small particles coated with DNA into a cluster of plant cells, where some will rub off and recombine with DNA present in cell nuclei. Then tissue culture regenerates a whole plant, (Chrispeels and Sadava 2003). In early uses of these techniques, scientists needed some test that would determine whether the DNA had been successfully incorporated. The solution was to pair the target gene with an antibiotic marker gene. Plants that had successfully incorporated the desired construct would respond differentially when challenged by microorganisms. Concerns about the impact of this antibiotic were among the early points raised against agricultural biotechnology. As the tools and techniques for altering genomes have progressed, it is no longer necessary to use antibiotic markers, (Daniell et al. 2002). Perhaps more significantly, tools like agrobacterium and the gene gun allowed control over which genes were being inserted, but the insertion might take place virtually anywhere in the genome. Insertions that landed in an infelicitous location had the potential to disrupt essential genetic functions. Plant scientists’ experience with mutation breeding forms the backdrop for evaluating the risk posed by such disruptions. Non-functional plants or plants that exhibited undesired changes are not selected for further development. Only the transformations that scientists deemed successful would go forward. Nevertheless, the inability to target a specific location in the genome was an inconvenience, and it limited the applicability of biotechnology to uses that did not involve subtler changes in specific gene sequences.

1.5 Recent Developments During the last decade, a series of techniques that do allow breeders to target specific loci have been developed. The most widely publicized is CRISPR Cas9, which is associated with gene editing. Although the range of applicability for gene editing is still being explored, the consensus is that it will allow even more precision: a specific gene can be introduced at a specific location. What is more, many applications of gene editing may not involve gene transfer, at all. Instead, the modification will be to adjust gene regulation: the question of whether a gene is “turned on” (e.g. producing proteins that affect plant or animal metabolism) or off. This seemingly less intrusive modification might enable the production of foods with a different nutrient profile, or plants that fight infectious diseases more effectively, though still with their natural (e.g. preexisting) repertoire of responses. CRISPr Cas9 may be bringing technical capabilities well beyond the increase in precision. Using CRISPr to effect a gene change may turn out make genetic modification much easier to accomplish. This could put the capability for making genetic changes in plants and animals at the disposal of actors who have hitherto been unable to perform genetic engineering of eukaryotes. Although this is still a speculative possibility with respect to agricultural plants and animals, some are

12

1 Biotechnology in the Context of Agriculture …

already discussing the ethical significance of making genetic engineering “easy.” Biologists have long recognized that genetic engineering could be used for nefarious purposes, (see Goodfield 1977). Developments that make it easier to undertake gene transfer could make genetic engineering accessible to criminals or terrorists, (Kang and coauthors 2017; DiEuliis and Giordano 2018). If gene editing simplifies the process of gene transfer, it could also put the development of food and agricultural GMOs within the reach of well-intentioned groups who lack the experience and background of agricultural plant and animal breeders, (Cha 2013; Waltz 2016). CRISPr Cas9 is itself a genetic construct found in nature. Its function seems to involve repair of sequences in the genome that, when altered, could affect the viability of the organism. However, as a genetic construct, CRISPr Cas 9 can itself be incorporated into the microbe, plant or animal being modified. This means that a scientist can now incorporate a genetic change that will continue to make further changes after the organism has left the confines of the laboratory. This potential has considerably expanded the universe of imagined applications for modification of genomes, (Doudna and Charpentier 2014). It is, for example, one of the key features of gene editing that has breathed new life into the discussion of germline genetic modification in humans (see Bosley and coauthors 2015). The agricultural applications of gene editing are taken up in Chap. 13. Synthetic biology is also an expansion of the capabilities for gene technology. The term synthetic biology is used within both the scientific community and more widely for the manipulation of nucleic acids, the products of such manipulations and broad programs of scientific research and technological development. It is not a precise term, and some uses overlap with the term biotechnology itself. The expression may have originated with a remark made by Waclaw Szybalski in a 1974 discussion of transcription, the cellular process of conversion from DNA to RNA. Szybalski speculated on the emergence of a new kind of biology in which entirely novel sequences of the four bases from which DNA is composed are constructed and utilized both for control and regulation of transcription, and potentially for the creation of novel proteins (Szybalski 1974). Currently, gene transfer works with genetic constructs discovered and identified through their activity in living organisms. Just as synthetic chemistry allowed chemists to develop chemical constructs that are not found in nature, Szybalski’s conception of synthetic biology implies an ability to string base pairs together in ways that would produce utterly unknown substances or novel biological functions. A flurry of discussion and debate over synthetic biology occurred between 2010 and 2015. Two developments sparked this episode, though both were far short of the speculative capability that Szybalski described in 1974. First, a group of scientists working at the J. Craig Venter Institute developed what they called a synthetic genome. The chemical properties of guanine, cytosine, adenine and thymine (the four bases that make up a molecule of DNA) make it possible to assemble a string manually, and then to automate that process of stringing G, C, A and T together. When one knows the sequence one wants to produce, it is thus possible to put together strings of bases that replicate segments of DNA found in nature. However, until the Venter group developed techniques for assembling shorter strings into longer ones,

1.5 Recent Developments

13

it had been impossible to create a nucleotide sequence long enough to function as the genome for an organism. The Venter group’s accomplishments consist in solving a series of technical problems in connecting and authenticating shorter strands into longer functional units (Gibson and coauthors 2008, 2010). What the Venter group had done was to replicate the genome of mycoplasma genitalium (a relatively simple microorganism) by assembling strings of guanine, cytosine, adenine and thymine. They did not invent a novel gene, much less a genome. The second development in synthetic biology was the emergence of a research program intend to identify gene sequences with known functions, and to develop a method of assembling these “biological parts” in order to accomplish basic tasks in bioengineering. The BioBricks effort is among the better known programs of this sort (Shetty et al. 2008). When realized, such efforts would allow the application of design principles adapted from electrical and computer engineering to the development and modification of organisms (Endy 2005). The basic technical problem is that previous methods of copying the DNA produced non-standard versions of the sequence. Different versions of the gene could function in the new organism, but would couple with other segments of DNA, and be located in slightly different places in the overall sequence. This effectively makes genetic engineering into a “one shot” operation: the sequence produced by any effort at genetic engineering was, in its details, potentially unique. Creating a standardized set of sequences with known functions could make it possible to string together several genetic constructs and have confidence that they would function as expected. As with the Venter group’s effort, this was about connecting bits of DNA, rather than designing entirely new sequences. Both BioBricks and the synthetic genome were steps toward creating a platform technology in molecular biology, and this, rather than synthesis as such, is the larger significance of these developments so far. Much of the genetic engineering practiced in both commercial and academic laboratories can be best understood as a form of problem solving science. The research team organizes its work around a specific goal: efficient production of human insulin or development of Vitamin A enhanced rice, for example. Resources and activities are then organized in pursuit of this goal, and the technical capabilities, knowledge base and imagination of team members are deployed in service to it. In contrast, synthetic biology moves innovation into what Adrian Mackenzie calls ‘design space’. Once a research team has developed a functional system of biological working parts for producing a vaccine or a biologic, it will become possible to redesign the system for a much larger range of applications. In contrast, present day genetic engineers must, in a real sense, go ‘back to the drawing board’ with each new problem they hope to resolve (Mackenzie 2010). In summary, biotechnology comprises a large number of techniques for analyzing and modifying the sequence of base pairs in a molecule of DNA. These techniques exploit the recombinant characteristics of DNA, hence one can define biotechnology as any approach to genetic modification that utilizes rDNA. Agrifood biotechnology is simply the utilization of this tool set in the practice of genetic modification that has long occupied the breeders of plants and animals used in agriculture. These include foods, but they also include fiber crops (e.g. cotton) as well as tobacco and

14

1 Biotechnology in the Context of Agriculture …

crops intended for use as feed, as energy stocks or for other industrial purposes. Biotechnology differs from older techniques of plant and animal breeding both in its use of rDNA and in the specificity of the genes and gene process that can be targeted through technological means. With the advent of gene editing and synthetic biology, the tool kit is expanding, and so are the potential applications of biotechnology.

1.6 What’s in a Name? The preceding description of biotechnology is broad, but it would probably satisfy most scientists. It excludes mutation breeding and tissue culture (both discussed above), but it would not be unreasonable to include these techniques as forms of biotechnology. Both require relatively sophisticated equipment and training that would not have been found in the agricultural laboratories of many less industrialized countries as recently as 1990. Neither could be performed by the average farmer working alone. Hence, a research institute that develops a capacity for tissue culture might claim to be doing biotechnology as a way to assert that they are advancing in their technical capability. More substantively, although neither mutation breeding nor tissue culture involve the transfer of genes, both are known to affect the rate of mutations in plant genomes. This topic will be revisited in Chap. 13 in connection with off-target effects from gene editing. As already noted, use of the GMO acronym is controversial. Aren’t all of the foods we eat “genetically modified?” some plaintively ask. In fact, virtually all of the terminology introduced so far can be contested. For scientists, more precise and complex ideas supplant the concept originally intended by the word “gene”. Yet, scientists still find it handy to talk about genes, though it is not at all clear that they have the same conception of genes as an average person. Scientists are not trained in the humanities disciplines that study language and meaning. One of the primary venues in which philosophy can be helpful in controversies is in being attentive to the ways in which a choice of words might spark a misunderstanding. For an extreme example, take luciferase. This word was coined by Raphaël Dubois (1849–1929) as a term for enzymes that produce bioluminescence, the light emitting characteristic we associate with fireflies. To my knowledge, this particular coinage has never led to controversy. Nevertheless, those of us who interact with Biblical literalists or people whose worldviews reflect the predominance of theology over biology can speculate a bit. What if some new product—a food, a drug, a home cleaning compound—advertised itself as containing luciferase? Are there people out there who might presume that a satanic force was being corralled for commercial purposes? As farfetched as this might seem in some quarters, it illustrates how a person’s worldview and social context can trigger concerns that would be regarded as laughable by scientists whose own social context orients them to a specific (and quite different) set of associations. Two words that do occasion misunderstanding are mutation and, indeed, biotechnology, itself.

1.6 What’s in a Name?

15

In the context of tissue culture, mutation just means any change in the sequence that constitutes the genome. Although the recombination of a DNA molecule in cell division or sexual reproduction is said to replicate the sequence that was present in the original molecule, in fact, small changes can and do occur naturally. We could refer to these changes as errors in transcription or recombination, and, indeed, many biologists do refer to them in just that way. The error rate (or frequency of spontaneous change) cell growth and reproduction is quite low and seems not to make much difference to a plant or animal’s metabolic processes. Yet every molecular biologist will concede that there are some mutations that could have effects to which we should pay heed. At the cellular level, a mutation could affect how a cell functions, and could be lethal to that cell. The death of a single cell is not a big deal in the context of an organism (like a carrot, a pig or a human being) that has billions of cells. However, some mutations can affect the phenotype, and could be lethal to the affected individual. In the context of nature, this, too, is not a big deal. There are many individuals in a reproducing population. Nevertheless, heritable mutations are thought to play a role in evolution. They mostly have no effect, and in the event that the effect is harmful, it just leads to the loss of one among many. Nevertheless, in that exceedingly rare case where the effect is beneficial, it can give one organism in the population an advantage that spreads throughout the entire gene pool of a species. Therefore, in biology, mutations are not necessarily bad, and they can be exceedingly good (at least from the standpoint of species adaptation). In comic books, a mutant is a human being who starts to exhibit superhuman powers at puberty. In the movies, mutants have been infected by toxic waste that turns them into bloodravenous zombies. All of these meanings can be found in Wikipedia by searching on “mutant” “mutant (Marvel)” and “mutant (film)” respectively. All of them draw on understandings of genetics and evolution, though in the case of comic books and the film Mutant, the understanding is a kind of folk biology. This suggests that nonbiologists might be predisposed to worry when talk of mutations in the genome starts to be thrown around in a casual manner. What is more, these ordinary folk would not be without some support from the science community itself. Although the error rate from transcription or recombination is always small when compared to the number of base pairs in a genome, it can increase when these processes occur under stressful conditions. Here, “stress” means simply that the presence or absence of abiotic elements (such as heat or radioactive isotopes) or biotic elements (such as chemicals) affects the frequency with which spontaneous mutations occur. This is, in fact, the basic idea behind mutation breeding, where chemical or radioactive stress is applied to induce potentially useful changes in an organism’s genome. In fact, tissue culture also increases the rate of mutation. Those molecular biologists (and they are a minority) who have concerns about the safety of agrifood biotechnology base part of their concern on the fact that when the number of mutations increases, the probability that a harmful mutation will occur increases, as well (see Druker 2017). Hence, if a technique utilizing recombinant DNA then requires tissue culture to regenerate the plant, the higher mutation rate associated with tissue culture will necessarily place the transformation into a higher risk category.

16

1 Biotechnology in the Context of Agriculture …

This returns us to the meaning of the word ‘biotechnology.’ The conceptualization that will be utilized throughout this book stresses the use of recombinant DNA and excludes tissue culture. Yet if practical applications of recombinant DNA techniques also involve tissue culture, and if tissue culture increases the rate of spontaneous change in the sequence of base pairs, there is a potential for misleading statements and even dissemblance. For example, one scientific publication (Akcakaya and colleagues 2018) demonstrates a method for limiting the number of mutations after gene editing and cites this as evidence for the safety of gene editing. However, this is arguably misleading if it is taken to imply that gene edited foods are safe. 2 It is misleading because it implies that gene edited plants will have low mutation rates, but reconstitution of an agricultural plant (necessary after gene editing) will be achieved with tissue culture, a technique that itself is associated with an accelerated rate of mutations. On the other hand, someone who relies on the mutation rate in tissue culture to claim that gene editing (or any other specific tool of biotechnology) is the source of an increase in the rate of mutations (and hence risky) would also be encouraging a faulty inference. This is not a distinction that makes a difference when gene edited plants are compared to many foods that have been eaten safely for a century or more. In fact, the debate over biotechnology is full of exchanges that exhibit exactly this form of claim and counterclaim. The publication and subsequent retraction of a 2012 study by French scientist Gilles-Éric Séralini and his colleagues is a case in point. Séralini’s group had been conducting a number of toxicological studies on glyphosate, the chemical in the Monsanto weed killer Round-Up™. When applied to Round-Up Ready™ varieties of GMO cotton, soy and other crops, glyphosate provides a highly effective method of weed control with no effect (or perhaps we should say, no putative effect) on the herbicide resistant crop itself. Although Séralini’s lab had conducted studies on glyphosate having nothing to do with GMOs, the study in question involved some feeding trials of Round-Up Ready™ potatoes to laboratory rats. The controversy over the study’s validity notwithstanding, the toxicity allegedly observed in the study was due to the presence of glyphosate. It did not derive from the process of rDNA plant transformation that made these potatoes resistant to glyphosate, and the original article was careful not to make causal claims, in any case. Yet the study and the ensuing controversy over glyphosate is widely promoted among anti-biotechnology advocates as evidence against gene transfer. Séralini’s own lab was promoting their findings on a website with the heading “Séralini GMO” (Thompson 2015, pp. 234– 236). Here again the opportunities for miscommunication extend in both directions. Creating resistance to glyphosate’s lethal impact on plant growth is one of the most widely used applications of agrifood biotechnology. It has been deemed safe because tests of glyphosate in the 1970s determined that it degraded into harmless 2 In

fact, the Akcakaya paper cited makes no claims about gene editing for food crops. It is focused on gene editing in a biomedical context where cells will be grown in vitro, as in in vitro fertilization techniques that are widely practiced on humans. This is not to say that in vitro cell cultures do not carry their own risks, yet these risks would appear to have been deemed acceptable by both physicians and patients seeking treatments for infertility.

1.6 What’s in a Name?

17

compounds fairly rapidly. However, the development of crops resistant to glyphosate has certainly increased its use and has also changed the pattern of use. Farmers make repeated sprays throughout the growing season. As studies have challenged the initial findings on the safety and environmental persistence of glyphosate, it is certainly reasonable to question its use in connection with herbicide resistant GMOs. Yet it is also misleading to lay the moral responsibility for problems with glyphosate at the doorstep of biotechnology because glyphosate is used in many other applications that have nothing to do with biotechnology. In addition, many applications of biotechnology have nothing to do with glyphosate. Using the debate over glyphosate to tarnish the tools of biotechnology is a form of dissembling, and that is a problem in the ethics of agrifood biotechnology. The critics of GMOs are displaying a character flaw. Then when someone simply refuses to take the evidence of toxicity seriously because anti-biotechnology critics brought the issue forward, that too is an ethical issue. Here, the industry scientists are displaying a character flaw. This shows how the ethical problems depend as much on people as they do on the technology itself. In sum, the ethical issues in agrifood biotechnology concern the way that we talk about it, as well as its biochemical mechanisms and their toxicological or environmental impact. The mere fact that this talk has become so divisive is itself an ethical issue, and one that has had a profound impact on the politics of agriculture and the practice of agricultural science. One lesson for the science community is simply that as a group they should be much more sensitive to the language questions than they typically are. This theme will be sounded most prominently in Chap. 12, but readers should be alert to the way in which definitions and social practice shape opinion from the outset. One final observation again concerns the meaning of the word ‘biotechnology.’ The approach I have chosen emphasizes the technology, but remains open to the possibility that the technology might be socially institutionalized in many different ways. Yet if one consults the financial pages of any major newspaper, one will see that biotechnology is being defined less in terms of specific tools and techniques than in terms of those for-profit firms who are using recombinant technology to develop products for medicine and agriculture. In other words, biotechnology is a profit seeking activity performed by corporate firms. Notice that on this interpretation, plant or animal varieties developed by universities or other non-profit ventures would not count as biotechnology, because biotechnology is by definition linked to profit seeking enterprise. However, many public sector scientists still talk (and perhaps believe) that GMOs could emerge from their labs free of the profit-seeking motives we associate with the biotech industry. There is, on their interpretation (which is close to mine), no necessary tie between biotechnology and capitalism or the profit motive. This may be a naïve belief when socio-economic realities are added to the picture. Chapters 8 and 9 will examine these connections more closely.

18

1 Biotechnology in the Context of Agriculture …

1.7 The Controversy in Ethical Perspective The use of recombinant DNA to modify the genetic structure of plants, animals and microbes and the ability to clone adult cells from mammals jointly contributed to an international controversy that has several axes of contention. While theologians and philosophers have thus far focused primarily on applications in the field of human medical science, the broader public has arguably been equally (if not more) concerned with the use of these techniques in food and agriculture (see Lassen and coauthors 2002; Einseidel and coauthors 2002). This popular concern with biotechnology is both prudential and moral. There are worries that the technology may have unknown and unacceptable risks, but there is also apprehension about the ethics of this seemingly new and radical activity (Frewer and coauthors 1997; Midden and coauthors 2002; Schurman and Munro 2010). Furthermore, risks can be converted into moral concerns. As such, there is ample terrain for prima facie analysis of ethical issues associated with food and agricultural biotechnology. The goal throughout this book is to illuminate the normative basis for alternate judgments about the acceptability, advisability and justifiability of using biotechnology in the production of agricultural plants and animals. A normative basis for action and judgment stipulates ideals, values or standards that ought to be reflected in human conduct. It is distinguished from matters of fact that may also form a component of the basis for action or judgment in a particular case. On the one hand, ethics deals with almost universally recognized norms that are both implicit within everyday social interaction and explicitly articulated in public sources such as legal or professional codes of practice, religious texts, folktales, literature and philosophy. On the other hand, the ethical dimension of conduct and reflection is often characterized as inherently personal, introspective and inherently unsuited to public discourse. Given this range of interpretation, ethical concerns associated with food and agricultural biotechnology can be expected to comprise highly idiosyncratic personal reactions of individuals, identifiable traditions and values of particular social groups, and broadly shared social norms. One approach is to present the debate in terms of opposing pro and con arguments, as several studies by philosophers have done. Gregory Pence (2002) for example, emphasizes the way in which proponents of biotechnology emphasize humanitarian goals of ending hunger, while opponents see biotechnology as unnatural, a “mutant harvest.” Pence’s focus on the issue of whether biotechnology is natural was also the main organizing principle for an earlier study by Michael Reiss and Roger Straughan (1996) that covered medical as well as agricultural biotechnology. Gary Comstock (2000) also takes up the possibility that biotechnology might be unnatural, but emphasizes how he himself came to see the humanitarian rationale for biotechnology as overriding his own concerns about the social and environmental risks associated with transgenic crops and genetically engineered animal drugs. All these authors wind up on the “pro” side of the debate. This way of framing the debate in terms of benefit from increasing agricultural productivity, on the one hand, and risky technology, on the other, has also been the subject of a lengthy and careful study by Hugh Lacey

1.7 The Controversy in Ethical Perspective

19

(2005), who is less inclined toward the “pro” point of view. Lacey believes that the pro-biotech perspective is rooted in an ethical perspective that valorizes processes of control and predictability, while the anti-biotech perspective can be traced to skepticism about the viability and desirability of control. Perhaps the most aggressively positive philosophical appraisal of agrifood biotechnology came from R. Paul Thompson in 2013. Thompson’s treatment includes a detailed and reassuring discussion of the biological mechanisms for gene flow. This exercise in the philosophy of agricultural science is embedded within a larger consequentialist ethical framework. Thompson proposes a form of risk-benefit thinking to evaluate GMOs in ethical terms. He describes genetically engineered crops as an alternative to mainstream, industrial agriculture, emphasizing the contrast between crops that have been developed to resist glyphosate herbicides or to produce bacillus thuringiensis (a toxin approved for organic production) in plant tissues. GMOs win this competition on two counts. First, they impose a significantly reduced toxic burden on the environment, and second, they are more efficient at controlling insects and weeds than chemical insecticides or herbicides, leading to an increase in the usable yield per acre of agricultural commodities. This latter benefit is especially significant in light of the need to achieve adequate global yields, which are expected to require increases due both to population increase and the loss of arable land from climate change. Thompson then undertakes a comparison between genetically engineered crops and organic or “sustainable” agriculture. Although he concludes that this alternative may have unrealized promise, GMOs still win out in virtue of their proven ability to meet global food needs, (R. Paul Thompson 2011). A critical discussion of the contrasts between the two Paul Thompsons can be found in From Field to Fork, (Thompson 2015, pp. 239–252). My approach interprets controversy over agricultural biotechnology as an episode in several ongoing and overlapping social, political and ethical struggles over the appropriate guidance, (the ethics, that is) of food and food production. These struggles range over disputes about food safety, where the normative dimension (avoidance of mortality and morbidity) is virtually uncontested, to the accommodation of culturally or religiously based norms that define what is and is not considered to be food, irrespective of nutritive or health-related concerns. Because food consumption is both rich in symbolic or cultural significance and biologically necessary for human life, any technology for producing or preparing food has ethical ramifications of one kind or another. These include the way that the technology affects safety and access to food, as well as other questions of fairness and equity associated with the broad system for producing and distributing food. One should expect that any novel food technology such as biotechnology would raise such ethical issues. They are a classic instance of technological ethics. From this perspective, there are no deep philosophical differences in the kind of question being considered when one shifts from energy technologies, such as nuclear power, to information technology or geoengineering to biotechnology. Nevertheless, any superficial survey of the global controversy over food and agricultural biotechnology demonstrates how this technology has been subject to far more public debate and criticism than has been typical of food production, processing or

20

1 Biotechnology in the Context of Agriculture …

marketing technologies in recent years. As will become clear in Chaps. 4 through 9, much of the debate involves ethical matters that are risk-based. Any food technology involves risk, and similar issues could be raised about organic food. However, the use of biotechnology in the food system has raised two large issues that do not arise in connection with chemical, mechanical and other food technologies. First, the use of gene technologies has been accompanied by significant transformations in the system of property rights that govern the food system, causing disruption in both research and farming practice. Second, the suggestion that there is something unnatural about using the tools of recombinant DNA to alter agricultural plants and animals overlaps with questions that, as Reiss and Straughan argue, have been prominent in biomedical applications of genetic technology. To put it bluntly, the mere act of modifying genes this way has been opposed as contrary to nature in the sense of violating a religiously or metaphysically grounded moral principle. It is just intrinsically wrong, without regard to benefits and risks. These two questions complicate the way in which civil societies create institutions for the governance of risk because they introduce reasons for opposing GMOs that have very little to do with management of risks to human or animal health, the environment or to the socio-economic well-being of various actors in the food system. Thus, biotechnology gives rise to institutional concerns that relate to the nature of technical expertise, its use social decision-making and governance, and concomitant issues associated with public trust in science. Must one, as the approach of Pence, Comstock, R. Paul Thompson and (to a lesser degree) Lacey implies, be “for” or “against” agricultural biotechnology? My approach opposes the view that one must take sides. The analysis in this book situates agricultural biotechnology within broader ethical debates, and interprets pro and con arguments about agricultural biotechnology as being motivated by philosophical positions that the parties to these arguments have adopted with respect to these broader debates. The reader is thus invited to understand the controversy as, in fact, a conglomeration of multiple controversies, each having a history and logic of its own, and in some cases operating in spheres of social and political concern that might have been thought to have little relation to one another. While advocates of positions within any of these multiple controversies might have hoped to enroll allies in their respective fights by portraying agricultural biotechnology in stark pro and con terms, it is not clear that such a portrayal leads to a philosophically sophisticated, much less philosophically honest, understanding of the issues involved. As such, the analysis that follows proceeds along organizational principles that break the debate up into the three broad categories listed above, each of which can be subsequently broken into sub-categories of its own.

1.8 Conclusion: Beyond Risk and Back Again The risk assessment framework that is covered in Chaps. 4 through 9 represents an important way to approach ethical issues in agrifood biotechnology for at least three reasons. Most importantly, it yields both insight into many of the key issues that need

1.8 Conclusion: Beyond Risk and Back Again

21

to be considered and a framework for addressing them that allows for a systematic organization of the issues. The proof of this claim is in the pudding, of course. I hope that the book itself testifies to the fruitfulness of my general approach. Secondly, the scientific community is deeply invested in the risk assessment framework. Demonstrating lapses or errors in the way that risks are interpreted or analyzed is a way of speaking their language. Problems that can be identified through the risk assessment lens are problems that scientists should be willing to admit and address. I do not claim that alternative approaches are mistaken or unimportant, but a properly philosophical critique of the risk-based approach demands that we have actually understood it, and that we have attempted to take it as far as it can go. My analysis holds that there are key points of misalignment between the scientific community’s approach to the ethical issues in biotechnology and that of society as a whole. Debates over the social institution of intellectual property and the role of religious or metaphysical thinking, the topics of Chaps. 10 and 11, respectively provide an especially dramatic example of misalignment. However, the most crucial issues really do revolve around communication, procedure and the ways in which current regulatory, governance and social institutions require much more from scientists than they are willing or able to give, given the current way that science itself is institutionally structured. Science is organized through disciplines that function through university departments, public and private laboratories, publications, funding agencies and organizations created to support the activity of sciences (including the private sector, where scientists work to develop salable products). These are the institutions of science, but they are ill equipped and under-resourced for the social tasks of engagement with non-scientists, including many scholars from humanities or social science disciplines in the scientists’ own universities. If, as I believe, a properly expanded and carefully considered risk-based approach would address many of the issues that people really care about, it will nevertheless stumble and fail given our current lack of investment in the procedures for public engagement and expert review of ethical issues. The investments needed include restructuring within scientific institutions, but they will also require that non-scientists make an honest effort to see things as the scientists see them, if only to articulate any remaining concerns in a language they everyone can understand. Does this imply that the words of molecular biologists are rational and truthful and that what devotees of Pachamama or Papa Legba say is not? Giving that question its due would require an entirely different book, so I will simply assert that I think not. Even those who reject the presumptions or doctrines of applied biology would be wise to learn a bit of its language, if only because doing so would place them in a much stronger position to articulate concerns in a manner that would be compelling in public discourse. Whether they should be required to do this, either by political morality or coercion, is also an entirely different question. Therefore, I will simply assert allegiance to the view that they should not. Whether learning to appreciate the applied biologists’ perspective sublimates one’s identity, making one complicit in the repression of marginalized people, is yet another entirely different set of philosophical concerns. Here, I will say that I think not.

22

1 Biotechnology in the Context of Agriculture …

The book that I wrote in the 1990s was intended to enrich scientist’s own picture of their activity so as to engage some issues that were being neglected more adequately. I also hoped it would document the trace of ethically oriented debate over gene technology for future scholars. Writing in the 2020s, I believe that the book has been far more successful in the second task than the first. The book may be most helpful to those who do want to understand a broad range of social and ethical concerns in a manner that is consistent with a generalized version of the perspective that scientifically trained people bring to the moral and political evaluation of emerging technology. There are many reasons as well as non-rational motivations why a reader might want to understand this perspective, and I have undertaken revisions to accommodate as many of them as possible. A conception of technological democracy grounds the rationale for understanding the risk assessment perspective that I favor, and Chap. 2 explains it. Nonetheless, it would not be consistent with democratic values to insist that everyone has to think like a biologist, and the book’s structure and argument should not be read as asserting or implying such an absurdity.

References Akcakaya, P., M.L. Bobbin, J.A. Guo, J.M. Lopez, M.K. Clement, S.P. Garcia, M.D. Fellows, M.J. Porritt, M.A. Firth, A. Carreras, T. Baccega, F. Seeliger, M. Bjursell, S.Q. Tsai, N.T. Nguyen, R. Nitsch, L. Mayr, L. Pinello, M. Bohlooly-Y, M.J. Aryee, M. Maresca and J.K. Joung. 2018. In vivo CRISPR editing with no detectable genome-wide off-target mutations. Nature 561: 416–419. https://doi.org/10.1038/s41586-018-0500-9. Ahloowalia, B.S., M. Maluszynski, and K. Nichterlein. 2004. Global impact of mutation-derived varieties. Euphytica 135: 187–204. Bado, S., B.P. Forster, S. Nielen, A.M. Ali, P.J.L. Lagoda, B.J. Till, and M. Laimer. 2015. Plant mutation breeding: Current progress and future assessment. Plant Breeding Reviews 39: 23–88. Bosley, K.S., M. Botchan, A.L. Bredenoord, D. Carroll, R.A. Charo, E. Charpentier, R. Cohen, J. Corn, J. Doudna, G. Feng, and H.T. Greely. 2015. CRISPR germline engineering—the community speaks. Nature Biotechnology 33: 478–486. Bridgen, M.P. 1994. A review of plant embryo culture. HortScience 29: 1243–1246. Cha, A.E. 2013. Glowing plant project on Kickstarter sparks debate about regulation of DNA modification. The Washington Post, October 3, Accessed June 6, 2019 at https://www.washingto npost.com/national/health-science/glowing-plant-project-on-kickstarter-sparks-debate-aboutregulation-of-dna-modification/2013/10/03/e01db276-1c78-11e3-82ef-a059e54c49d0_story. html?noredirect=on&utm_term=.07c85fda555c. Chrispeels, M.J., and D.E. Sadava. 2003. Plants, Genes, and Crop Biotechnology, 2nd ed. Sudbury, MA: Jones & Bartlett. Comstock, G. 2000. Vexing Nature? On the Ethical Case against Agricultural Biotechnology. Boston: Kluwer Academic Publishers. Daniell, H., M.S. Khan, and L. Allison. 2002. Milestones in chloroplast genetic engineering: an environmentally friendly era in biotechnology. Trends in Plant Science 7: 84–91. Deplazes-Zemp, A., D. Gregorowius, and N. Biller-Andorno. 2015. Different understandings of life as an opportunity to enrich the debate about synthetic biology. NanoEthics 9: 179–188. DiEuliis, D., and J. Giordano. 2018. Gene editing using CRISPR/Cas9: Implications for dual-use and biosecurity. Protein and Cell 9: 239–240. Doudna, J.A., and E. Charpentier. 2014. The new frontier of genome engineering with CRISPRCas9. Science 346: 1258096.

References

23

Druker, S. (2017) The UK’s Royal Society: A case study in how the risks of GMOs have been systematically misrepresented. Independent Science News. Accessed Feb 17, 2019 at https://www.independentsciencenews.org/health/the-uks-royal-society-how-the-health-risksof-gmos-have-been-systematically-misrepresented/. Einseidel, E., A. Allansdottir, N. Allum, M.W. Bauer, A. Berthomier, A. Chatjouli, S. de Cheveigné, R. Downey, J.M. Gutterling, M. Kohring, M. Leonarz, F. Manzoli, A. Olofsson, A. Przestalski, T. Rusanen, F. Seifert, A. Stathopoulou, and W. Wagner. 2002. Brave new sheep—The clone named Dolly. In Biotechnology: The Making of a Global Controversy, ed. M.W. Bauer and G. Gaskell, 313–347. Cambridge, UK: Cambridge U Press. Endy, D. 2005. Foundations for engineering biology. Nature 438: 449–453. Frewer, L.J., C. Howard, and R. Shepherd. 1997. Public concerns in the United Kingdom about general and specific applications of genetic engineering: Risk, benefit and ethics. Science, Technology and Human Values 22: 98–124. Gaertner, F., and L. Kim. 1988. Current applied recombinant DNA projects. Trends in Ecology & Evolution 3: S4–S7. Gamborg, C., and P. Sandøe. 2002. Breeding and biotechnology in farm animals. In Key Issues in Bioethics: A Guide for Teachers, ed. R. Levinson and M.J. Reiss, 133–142. London: Routledge Falmer. Gibson, D.G., et al. 2008. Complete chemical synthesis, assembly, and cloning of a mycoplasma genitalium genome. Science 319: 1215–1220. Gibson, D.G., et al. 2010. Creation of a bacterial cell controlled by a chemically synthesized genome. Science 329: 52–56. Goodfield, J. 1977. Playing God: Genetic Engineering and the Manipulation of Life. New York: Random House. Jiang, K., B.N. Anderton, P.C. Ronald, and G.A. Barnett. 2018. Semantic network analysis reveals opposing online representations of the search term “GMO”. Global Challenges 2: 1700082. Johnson, I.S. 1983. Human insulin from recombinant DNA technology. Science 219: 632–637. Kang, X.J., C.I.N. Caparas, B.S. Soh, and Y. Fan. 2017. Addressing challenges in the clinical applications associated with CRISPR/Cas9 technology and ethical questions to prevent its misuse. Protein and Cell 8: 791–795. Lacey, H. 2005. Values and Objectivity: The Current Controversy about Transgenic Crops. Lanham, MA: Lexington Books. Lassen, J., A. Allansdottir, M. Liakopoulos, A.T. Mortensen, and A. Olofsson. 2002. Testing times— The reception of Roundup Ready soya in Europe. In Biotechnology: The Making of a Global Controversy, ed. M.W. Bauer and G. Gaskell, 279–312. Cambridge: Cambridge U. Press. Mackenzie, A. 2010. Design in synthetic biology. BioSocieties 5: 180–198. McHughen, A. 2000. Pandora’s Picnic Basket: The Potential and Hazards of Genetically Modified Foods. New York: Oxford University Press. Midden, C., D. Boy, E. Einseidel, B. Fjæstad, M. Liakopoulos, J.D. Miller, S. Öhman, and W. Wagner. 2002. The Structure of Public Perceptions. In Biotechnology: The Making of a Global Controversy, ed. M.W. Bauer and G. Gaskell, 203–223. Cambridge: Cambridge U Press. Mohanty, A.K., U.K. Mukhopadhyay, S. Grover, and V.K. Batish. 1999. Bovine chymosin: Production by rDNA technology and application in cheese manufacture. Biotechnology Advances 17: 205–217. Oksman-Caldentey, K.M., and R. Arroo. 2000. Regulation of tropane alkaloid metabolism in plants and plant cell cultures. In Metabolic Engineering of Plant Secondary Metabolism, ed. R. Verpoorte and A.W. Alfermann, 253–281. Dordrecht, NL: Springer. Pence, G.E. 2002. Designer Food: Mutant Harvest or Breadbasket of the World?. Lanham, MD: Rowman and Littlefield. Reiss, M.J., and R. Straughan. 1996. Improving Nature? The Science and Ethics of Genetic Engineering. Cambridge, UK: Cambridge U Press. Rheinberger, H., and S. Müller-Wille. 2018. The Gene: From Genetics to Postgenomics. Chicago: University of Chicago Press.

24

1 Biotechnology in the Context of Agriculture …

Sandøe, P., B.L. Nielsen, L.G. Christensen, and P. Sørensen. 1999. Staying good while playing god: The ethics of breeding farm animals. Animal Welfare 8: 313–328. Sathyanarayana, B.N., and D.B. Verghese. 2007. Plant Tissue Culture: Practices and New Experimental Protocols. New Delhi: IK International. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists and Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press. Shetty, P., D.Endy Reshma, and T.F. Knight Jr. 2008. Engineering biobrick vectors from biobrick parts. Journal of Biological Engineering 2: 5. https://doi.org/10.1186/1754-1611-2-5. Smith, K.A. 2013. Why the tomato was feared in Europe for more than 200 years. Smithsonian Magazine. Accessed 3 July 2017 at http://www.smithsonianmag.com/arts-culture/why-the-tom ato-was-feared-in-europe-for-more-than-200-years-863735/#dZrfv7GUTUh8Hgok.99. Snyder, M., and M. Gerstein. 2003. Defining genes in the genomics era. Science 300: 258–260. Szybalski, W. 1974. Discussion on in vivo and in vitro initiation of transcription. In Control of Gene Expression, ed. A. Kohn and A. Shatkay, 404–405. New York: Plenum Press. Thompson, P.B. 2015. From Field to Fork: Food Ethics for Everyone. New York: Oxford U. Press. Thompson, R.P. 2011. Agro-Technology: A Philosophical Introduction. Cambridge, UKn: Cambridge University Press. Waltz, E. 2016. CRISPR-edited crops free to enter market, skip regulation. Nature Biotechnology 34: 582.

Chapter 2

The Presumptive Case for Food Biotechnology

Abstract Hans Jonas’ principle of responsibility establishes a basic framework for evaluating novel technology in ethical terms. Risk assessment provides a further development of Jonas’s framework as it is applied to agrifood biotechnology. A riskbased approach consists in distinguishing four tasks for implementing technological ethics: hazard identification, exposure quantification, management and communication. The risk-based approach is effective when it operates against the background assumption that technologies passing risk-based tests are at least prima facie acceptable on ethical grounds. However, a complex of social institutions must be in place for this assumption to be valid. These institutions, combined with a risk-based assessment of the potential for unwanted consequences, constitute the presumptive case for agricultural and food biotechnology. This implies that innovators are not ethically required to demonstrate the case for their technology, and that the primary task of ethics is to focus on arguments against the technology. The chapter also discusses some logically and ethically problematic adaptations of the presumptive case. Keywords Risk assessment · Utilitarianism · Libertarianism · The principle of responsibility · Technological ethics · Logical fallacies · World hunger This chapter introduces the risk-based approach to addressing the ethical issues in food and agricultural biotechnology, and situates it within the philosophy of technology. As discussed in the Introduction, the risk-based approach understands risk as the conditional probability that harmful or deleterious events will occur, given the introduction of some new tool, product or technical process. In the parlance of risk assessment, risk is a function of hazard and exposure. Hazard comprises the harms or adverse and unwanted outcomes that might occur as a result of introducing a new technology. Exposure is the probability that hazards will materialize, given specifiable conditions of use. Risk assessment is a technically dense procedure that combines inductive reasoning, systems modeling, data collection and analysis in the attempt to profile the risks associated with a new product or process. This chapter will not delve into these technicalities, though some are discussed in later chapters. Rather, my claim here is that the hazard/exposure characterization provides a useful

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_2

25

26

2 The Presumptive Case for Food Biotechnology

starting point for understanding the ethical issues associated with agricultural and food biotechnology. The risk-based approach is ethically useful for three broad reasons. First, it conforms well to philosophical schemes that are already widely in use by ethicists and policy analysts. It is a specific development of rudimentary decision theory, and as such can be easily linked to normative theories in ethics and public choice. Second, the terminology of hazard and exposure brings consistency and conceptual clarity to ethical questions that would otherwise seem chaotic and unfocused. This virtue is especially significant in light of the scientific community’s level of comfort with the use of statistical methods to quantify correlations and assign probabilities to future events. Finally, the framework is particularly suitable to exposing ethically significant gaps and lacunae in existing governance procedures for products of gene technology. Having the framework in place proves helpful for identifying limitations of the framework, itself. Once a risk-based approach has been developed and applied, one can see how it threatens to introduce cognitive biases into the ethical evaluation of new technology. Although a risk-based approach is a valuable tool for organizing the ethical issues associated with agriculture and food production, it does not, in itself, produce a comprehensive accounting for every ethical issue. This chapter begins with a discussion of technological ethics, including its history and recent development. The risk-based approach is then described as a promising development in technological ethics, and I offer a policy-relevant version of the approach that I developed for the original 1997 edition of the book. My interpretation is compatible with regulatory risk assessments used for policy making by government and intergovernmental agencies, but it is not equivalent to them. Finally, the chapter closes by examining how fallacious applications of risk-based thinking have had pernicious impact on the biotechnology discourse.

2.1 Technological Ethics: A Précis The 20th century was a time of unsurpassed technological progress, but it was also a time in which humanity learned that technological changes bring unintended social and environmental consequences. Government agencies for managing unintended consequences appeared early in the century, but it was not until the 1970s that explicitly philosophical rationales and approaches were articulated. The German philosopher Hans Jonas (1903–1993) argued for a systematic method of anticipating and evaluating technology in ethical terms in a 1979 book whose English title is The Principle of Responsibility: The Search for Ethics in a Technological Age. Jonas understood that this would depart from traditional ethics in that technology has impacts that extend indefinitely in space and time. Traditional ethics had focused on relations between people who lived together in a common place, and who could expect to have repeated encounters with one another. Jonas argued that technological ethics must integrate science-based attempts to understand the systematic and temporally distant effects of technology with ethical concepts attuned to the fact

2.1 Technological Ethics: A Précis

27

that many of the affected parties will not be known to those who plan and execute a technological practice. Jonas called for what he called a principle of responsibility (Prinzip verantwortung) as a response to this situation, (Jonas 1984). Still today, Jonas represents a break from the dominant conceptualization of ethics in science and engineering. Much of contemporary bioethics is limited to standards for scientific integrity and the responsibilities that arise in connection with human subjects. Jonas called for an ethical inquiry into the purposes and general trajectory of technology. He noted that it would be necessary to re-conceptualize the impact of human projects on the natural world in moral terms, and he noted special concern for technological developments (such as atomic weapons) having the potential to extinguish “autonomous reason” from the universe. Yet in one respect Jonas’s approach was not radical. The implicit logic of the principle of responsibility accepts the basic legitimacy of technological innovation, and does not challenge the presumptive norms that support the discovery and implementation of new technical methods and products. These norms draw on two of the most venerable philosophical traditions of the industrial age: utilitarianism and libertarianism. The implicit logic of technology is utilitarian in that new technologies are presumed to offer new opportunities, new possibilities of action, to human beings. These new opportunities present alternatives to the status quo, and are evaluated according to whether the outcome of utilizing a new technology is expected to be an improvement on the current situation. Utilitarianism mandates that an actor should always choose the course of action that produces the best outcome. The specification of what counts as an improvement remains open in this unexceptional description of technology, and in practice, engineers, designers and entrepreneurs evaluate technological innovations in terms of workplace standards already in play at the time and place in which an innovation is made. These standards often reflect a firm’s need to economize on scarce or expensive inputs in the production process, resulting in time savings, increased production or, what has been most important under capitalism, an ability to sell the product for a lower price. Alternative “ways of making or doing” (e.g. technology) that do not economize in this way are simply not taken up, with the result that technological innovation comes to be closely, perhaps even inherently, associated with increasing efficiency in the production process. Translating localized workplace efficiencies into global, social efficiencies requires a process for ensuring that efficiencies have not been achieved simply by “externalizing” costs, that is, by imposing costs on other parties. Utilitarianism requires that all of these costs be included in calculating whether the innovation improves the ratio of positive to negative outcomes, (see Schmid 2004, for a concise review of economic analyses of technological innovation). The implicit logic of technology is libertarian in that new technical methods enable particular modes of human activity. The libertarian ethic holds that human beings should be maximally free of constraint, subject to the condition that their actions should not harm or constrain others. Innovators should be free to innovate and to use their innovations, subject, of course, to the limitation that they are not free to harm others. Although the specification of harm or constraint is left open here, the libertarian view establishes a key burden of proof for technological innovation. If

28

2 The Presumptive Case for Food Biotechnology

no one complains, there is no basis for constraining innovation. Importantly, there is no reason why the innovator has to have an argument in favor of the innovation. Those who would constrain the innovator are in the position of needing to show how someone who has not agreed to bear risk would be harmed. However, Jonas contests this way of understanding the libertarian ethic by saying that innovators must take this burden of proof upon themselves. The Prinzip verantwortung calls for scientists and engineers to make an active attempt to anticipate possible forms of harm. If they do this and find no harm, they act in conformity to the principle of responsibility and are liberty to proceed with their technological application. In addition, the expectation that technological innovations will improve workplace efficiencies provides a global argument that further supports them doing so. There is, in short, no real need for a “pro” argument for biotechnology or any other technology, at least not at the outset. There is only the need for a responsible effort to ascertain the unintended consequences of technical change.

2.2 Ethics and Risk Risk analysis is one of the main social responses to Jonas’s call for a Prinzip verantwortung. Risk analysis is an adaptation of social choice intended to help decision makers develop a more inclusive and forward-looking understanding of the potential for unanticipated or externalized costs. For present purposes, social choice is defined as the application of decision theoretic approaches (including ethical theories) to problems where society as a whole is viewed as the choice-making agent. Succinctly, society as a whole (or some relevant population) is treated as analogous to an individual evaluating a choice situation. This decision-maker surveys policy options—which can include technological innovations—with the goal of choosing how to go forward. The analytic content of the decision situation includes some description or demonstration of the value associated with each option. Social choice approaches value options in terms of their impact or significance for all members of the relevant social group. Utilitarian interpretations of social choice require a predictive assessment of each option’s impact on social welfare, an accounting commonly expressed in terms of costs and benefits, but alternative interpretations are available. Amartya Sen and Martha Nussbaum have specified an approach that assigns value to decision alternatives in terms of capabilities, for example (see Gaspar 1997). The social choice perspective is most obviously useful in policymaking contexts where a governmental decision is expected to transcend partisan interests, but it can also be used to comparatively evaluate future scenarios that have been generated by technical capabilities. Social choice models have generated a large literature that is philosophically complex (see Sen 2011), and a detailed discussion would put the present enquiry far off track. In other writings, I have objected to the decisionism implied by social choice, (Thompson 2010). My use of the risk-based approach should not be interpreted as an endorsement of the claim that either social choice or the risk-based approach function as a comprehensive approach to ethical analysis.

2.2 Ethics and Risk

29

Risk assessments inform the social choice framework in key areas of regulatory policymaking, but the risk-based approach can also be used to organize an ethical evaluation of technological innovations. In either case, the approach presumes that the problem situation has already been configured as one in which avoidance of harmful or unwanted outcomes is a highly significant, if not pivotal, factor for decision making. This presumption reflects a considerable narrowing of the social choice approach, but it is warranted for many questions in technological ethics. The unintended harmful consequences of many products and technical practices were recognized only after years and sometimes decades of use. Decisions to ban, recall or restrict certain drugs (Thalidomide, fen-phen, opioids) reflect a public choice based on post hoc documentation of harmful effects. Restriction of agricultural organochlorines (such as DDT) followed studies that demonstrated how their toxicity increased through bioaccumulation. Environmental and public health agencies now ban or restrict the use of asbestos as an insulation material. Most countries banned leaded gasoline (TEL) between 1990 and 2000. In each case (and there are many more), a scientific assessment established a significant correlation between physical contact with the product and human mortality and morbidity. Ex ante assessments are now required for pharmaceuticals and pesticides. These regulatory practices establish a conceptual model for technology assessment. Anticipatable impacts become data for a decision as to whether the product or process should be allowed, and if so, under what restrictions. Again, this type of decision conforms to the model of public choice, with the information assembled through risk assessment informing the evaluation of alternate courses of action. The conceptual model also applies in many domains where states have not asserted regulatory authority: personal computers, social media platforms, facial recognition technology, data mining, again, the list is endless. In these cases, risks are anticipated as part of ethical review that might be conducted by civil society organizations, academics or even by the innovators themselves. The ubiquity of the risk based approach provides a reason for using it as a starting point for the evaluation of agricultural and food biotechnologies. This is a way of thinking that has become commonplace. In fact, as Chaps. 4 through 9 will demonstrate, the vast majority of objections to GMOs and other food-oriented gene technologies can be fitted to this model. At the same time, the conceptual framework of risk assessment is open to many interpretive judgments. Choices about what data to collect and comparison cases can generate analyses that distort actual risks and conceal dangers behind a façade of uncertainty. Beyond deception, well intentioned methods in risk assessment can justify decisions that turn out to be wrongheaded in the long run. At about the same time that the first edition of this book was written, epidemiological studies that identified smoking and exposure to second hand smoke as risk factors for cardiovascular disease were leading to significant change in law and behavior, (Katz 1997). In the intervening decades, bioethicists have recognized that epidemiologically derived risk factors can contribute to health inequalities both by concealing unique characteristics in subpopulations and by obscuring more efficacious social determinants of health, (Valles 2018). These problems call for more sophisticated applications of a risk-based approach, rather than its abandonment. Indeed, working within the basic conceptual

30

2 The Presumptive Case for Food Biotechnology

elements of risk analysis is a necessary condition for spotting erroneous and abusive uses of its methods.

2.3 The Risk-Based Approach The risk assessment framework can be summarized by the following schematic (see Fig. 2.1). Hazard identification is the process of identifying what bad outcomes can occur as a result of adopting one option rather than another. In classic regulatory decision making, the option under analysis is often whether to permit the use or release of a new product, or whether to adopt rules governing some technical practice (such as generation of electricity using nuclear reactors of a given design). In these cases, hazard identification is a largely inductive process of ascertaining the possibility of an adverse outcome based on the current state of the relevant science, and estimating its severity. Simply knowing what can go wrong does not provide an estimate of risk, however. Hence risk quantification utilizes a variety of modeling, statistical, and experimental techniques to estimate the probability that a hazard will materialize, given adoption of the option under consideration. Though conceptually simple, both activities can be difficult to accomplish in practice. Some of these difficulties are philosophical and will be taken up in the balance of the book. In Chap. 3, I will discuss ways in which two philosophers, Stephen Stich and Bernard Rollin, use a risk-based approach in their discussions of gene technology. Chapters 4 through 9 undertake systematic ethical inquiry into the risks of agrifood biotechnology by considering four ontologically distinct forms of hazard. Chapter 4 examines food safety. Here the hazards are straightforward and non-controversial: Fig. 2.1 Stages of Risk Assessment

Exposure QuanƟficaƟon

Hazard IdenƟficaƟon

Risk Management

Risk communicaƟon

2.3 The Risk-Based Approach

31

illness or death from eating a GMO. Chapter 5 takes up impacts on the animals being modified through biotechnology. The 1970s and 1980s saw a rebirth and reexamination of ways in which standard concepts of benefit and harm as applied to human beings could and should also be extended to sentient animals. Chapters 6 and 7 examine environmental hazards, which are philosophically more complex and controversial. At the root is a longstanding controversy in environmental ethics as to whether standard concepts of benefit and harm, which focus on human beings, are adequate to our understanding of an environmental impact. These disputes also function as a proxy for a more comprehensive debate over the specification of key elements in the risk-based approach, itself. Natural resource economics developed approaches to the valuation of environmental outcomes that tied environmental risk assessments quite closely to utilitarian welfare economics. As such, Chaps. 6 and 7 take up basic philosophical questions about the specification of hazards, exposure and the rules for risk management. Chapters 8 and 9 consider social consequences: ways in which people are benefited and harmed economically, culturally or in terms of their social life. In this domain, economic analysis is often quite helpful, as welfare economics provides an avenue for conceptualizing a technology’s consequences in terms of risk. However, the biological training of molecular geneticists may bias them against interpreting events that harm one’s economic well-being under the heading of risk. Many regulatory regimes also reflect this bias: they provide a basis for action only when risks to health are identified. Yet economists have tended toward the opposite extreme. Graduate programs in economics offer dozens of courses dealing with risk, but it is always the economic hazards of investment, finance, insurance or management decision making that are at issue. Disciplinary biases aside, no logical reason prevents socioeconomic outcomes from being classified as hazards, though there may well be sound ethical reasons for managing these risks differently from health or environmental risks. Risk analysts will stress a subtle point at this juncture: risk should not be conflated simply with the loss or “badness” of the hazard, should it occur. Risk must also reflect the chance or probability that hazard actually will materialize, and this can only be appreciated once the various factors that contribute to a given individual’s or group’s exposure to the hazard have been taken into account in the risk quantification phase. Research in the social psychology of decision making suggests that many people tend to neglect or poorly understand the probabilistic dimension of decision-making (Sunstein 2002), hence there is an argument for discounting the public’s estimate of relative risks associated with any given activity or phenomenon. This argument has been made with respect to the environmental and food safety risks of GMOs (Hettich and Walther 2012), but my analysis will not prejudge the opponent’s position. Having an estimate of risk does not tell a regulator what to do about it. One option is always “Nothing.” Simply accept the risk. Other options include regulatory measures that mitigate the severity of hazards, that compensate losers, or that develop mechanisms to secure consent from those parties that will bear risks. The procedure for making a decision about risks is represented by the risk management element in the schema. Cost-benefit analysis is a procedure developed by economists to allow decision makers the opportunity to include not only risks (as determined by

32

2 The Presumptive Case for Food Biotechnology

hazard identification and risk quantification) but also beneficial elements associated with adopting the option. The key point to notice is that one can make a distinction between assessing (or estimating) what the risk is, and deciding what to do about it. The decision making phase is referred to as risk management. Social choice models for decision-making under risk propagated in the 1970s and 1980s tended to stop with hazard identification, risk quantification, and risk management. These models were criticized as lacking a role for public input. In response, risk communication was added to the model following an influential U.S. National Research Council study of the 1990s colloquially referred to as “the orange book” (NRC 1996). The role of risk communication within the overall framework of risk assessment and social choice continues to be a topic of confusion, contestation, and persistent debate. Some participants in those debates argue that all the interests that gave rise to controversy over GMOs should be brought into decision-making through the mechanism of risk communication (see, for example, Wickson 2007. Hartley and Millar 2014, argue a view more consistent with mine.). On this view, risk communication equips a political decision maker with a better grasp of what matters to his or her constituents. However, regulators are more narrowly constrained by the statutes that delimit their authority, which generally specify a restricted class of hazards that can be the basis for regulatory action. In the United States, for example, regulators at the Environmental Protection Agency have authority to review and restrict the human health and environmental impacts of pesticides, while authority for other food safety hazards resides at the Food and Drug Administration. The Department of Agriculture can regulate substances or activities that could pose an economic threat to agriculture, but not from agriculture. Most socio-economic effects fall entirely outside the scope of the authorizing legislation under which these agencies operate. An ethical application of the risk assessment framework need not (indeed must not) be constrained by the particular statutes or regulatory policies that are in effect at any given time. The argument developed in Chaps. 4 through 9 is thus an appeal to broaden the categories and procedures used to conceptualize risks from agricultural and food biotechnology (indeed from all agricultural technology) as much as it is an extension and application of the risk assessment framework. While the four domains of risk assessment serve to clarify and structure ethical investigation, one task for philosophical analysis is to prevent premature closure of the categories in which hazards are articulated or exposures are calculated. As argued narrowly in Chap. 3 (a case study of the debate over the animal drug recombinant bovine somatotropin) the policy framework for evaluating biotechnology has precluded even debate, much less decision making that could be responsive to legitimate social concerns.

2.4 The Logic of the Presumptive Case In its focus on unwanted outcomes, the risk-based approach adopts an implicit burden of proof that falls on the side of providing reasons to restrict, control, limit, regulate, or moderate the use of the technology. In other words, by outlining an approach

2.4 The Logic of the Presumptive Case

33

that emphasizes reasons not to allow use of a product or practice, one has indirectly implied that if no such reasons are forthcoming, the use may proceed. It is important to make the rationale for this burden of proof more explicit, for it is the source of deep ironies in the discourse on agrifood biotechnology. On one side, opponents of the technology have expended most of their energy in allegations that gene technologies pose health-oriented consequences to food consumers, agricultural workers, animals and to ecosystems. These allegations are invitations for technological innovations that mitigate or eliminate the pathways leading to unwanted outcomes. They have, in effect, told biotechnology companies how to address their objections through some form of technological fix. By accepting the exclusion of social consequences, they strengthen a political ecology that relieves innovators of the responsibility to even discuss how their products might reinforce patterns of poverty, racism and injustice. On the other side, biotechnology advocates often object to the risk-based approach, thinking that it is more important to talk about the putative benefits of their technology. While they are probably right as a principle of marketing, they fail to realize that a general ethical policy of having to prove benefits in advance would probably bring technological innovation to a crushing halt. Logic permits only three options here. I am defending an interpretation of the riskbased approach which holds that in the absence of reasons not to, the technology may be used. There is an opposite presumption, which demands a hands-down argument to justify pursuit of every innovative step. This type of moral and regulatory stance may have been in place under monarchies that required explicit and advance permission from the king before any change in a technical practice could take place (see Busch 2011). An extremely conservative approach like this stifles innovation, but perhaps there are reasons why gene technology’s opponents might prefer it, at least if it can be confined to the food system. I will return to this interpretion presently. The third option mandates case-by-case evaluation of benefits, costs and risks for every proposed use of technology. This third choice may seem appealing at first blush. Pro-biotechnology advocates seem to be calling for this when they complain about the risk-based approach’s focus on the downside of their favorite innovation. However, demonstrating benefits would actually require generating even more predictive data than the risk based approach. As such, it becomes onerous and difficult to apply in practice. In order to appreciate why, it is helpful specify the sense in which case-by-case evaluation is an alternative to the presumption of permission. I do not mean to suggest that specific tools and techniques never require case-by-case evaluation. Case-by-case evaluation challenges the presumptive approach only if it means that every time a new application or product comes up for review, evaluators must undertake a fresh appraisal to determine whether it has compelling benefits, when viewed from a social perspective. In contrast, I am claiming that the initial burden of proof is to identify a risk or harm, with the presumption that this establishes the baseline for further considerations. This does not exclude the possibility that decision-makers then evaluate benefits with an eye toward whether they are substantial enough to compensate for risks. My point is that publics in industrial democracies do not generally require innovators to demonstrate a benefit before being allowed to experiment with their approach, and that this is, in fact, an ethically justified presumption.

34

2 The Presumptive Case for Food Biotechnology

Allegedly neutral case-by-case evaluation of technology actually imposes intolerable costs on our decision-making. There is no area of life in which people weigh every possible option on a case-by-case basis. We would clearly spend all our time weighing and deliberating if we did. Instead, we rely on “filters” to determine which cases demand more careful scrutiny and deliberation. Such filters often take the form of biases or dispositions that implicitly structure the burdens of proof that we impose on others and ourselves. Although we can certainly review and rethink when faced with any given case, the idea that we will thoroughly consider every possibility is not really a viable one. The question can thus be limited to two cases: should our cognitive filters be set for or against biotechnology? Jonas’s philosophy of technology is a sophisticated argument against the view of Martin Heidegger (1889–1976), who is arguably the most influential advocate of the view that our filters should be set against technology. Heidegger developed a deep mistrust of technological responses to humanity’s problems. He viewed worry and care as existential features of the human condition that call for reflective responses typically found in religion, poetry and in the philosophical contemplation of being. Heidegger was wary that Western cultures were falling into a pattern of seeking technical responses to every obstacle, and that this was leading people to see the entire world as a resource or tool for achieving momentary satisfaction. At the extreme, people would come to regard each other and even their own being as a mere resource at the disposal of technical means, (Heidegger 1977). In some of his lectures, Heidegger singled out industrial agriculture as a particularly significant example of the tendency he feared, (Glazebrook and Story 2015). Jonas had regarded Heidegger as a mentor, but in the 1960s he made a famous and public break with him over Heidegger’s prior association with the Nazi party in Germany (Jonas 1966). While Jonas shared some of Heidegger’s concerns about the cultural tendencies of the 20th century, his philosophy of technology acknowledged that humans would need to use technical means in order to cope with growing problems such as environmental pollution, population growth and the growing power of technological capability, itself. For Jonas, this implied that philosophers would need to work more closely with scientists, engineers and other specialists in technology, rather than putting themselves at a distance from them. Jonas believed that one of the key philosophical tasks was to articulate a conception of life that was both consistent with the understanding of contemporary biologists and capable of sustaining a process of ethical valuation, (Jonas 1966). As recounted above, both utilitarian and libertarian traditions in ethics provide rationales for being favorably disposed toward technical innovation, even as one implements Jonas’s Prinzip verantwortung. However, Jonas’s philosophical aspirations face a practical problem. Scientists often do seem to be so absorbed in the technical realization of their ideas that they cannot undertake ethical reflection. For example, Jeffrey Burkhardt has questioned whether scientists’ bias toward biotechnology is adequately open to objections. He bemoans the way in which “the scientific attitude” makes those who are developing biotechnologies very insensitive to a broad range of ethical concerns. While calling for large scale cultural

2.4 The Logic of the Presumptive Case

35

change within the sciences, he expresses pessimism about the possibility that scientists will seriously entertain reasons not to go forward with applications of biotechnology any time soon (Burkhardt 1997). Weed scientist Robert Zimdahl has supplemented Burkhardt’s pessimism with a book-length study of how and why agricultural scientists fail to consider ethical arguments, as well as alternative technological approaches. Zimdahl attributes much of the problem to an unexamined positivist philosophy in the agricultural sciences. Like Burkhardt, he calls for reform, but expresses doubt that reform is at hand (Zimdahl 2006). Hugh Lacey’s detailed analysis of agricultural biotechnology as a case study in the intersection of values and objectivity. Lacey argues that extra-scientific commitments to an ethic of control have led some scientists to favor gene technologies, while scientists guided by an ethic of cooperation and sublimation to nature have chosen an alternative path, Like Burkhardt and Zimdahl, Lacey believes that scientists must be more reflective about the value judgments that steer their research agenda, (Lacey 2005). An empirical study of attitudes among molecular biologists by three University of Reading social scientists (Cook et al. 2004) provide further support for these analyses. Do the uncritical attitudes noticed by Burkhardt, Zimdahl, Lacey and others lend support to Heidegger’s pessimism? It might in the abstract, but from a practical perspective, scientists’ failure to undertake ethical reflection is counteracted by social and governmental filters (that is, institutions) that weed out a lot of bad ideas without our having to pay much attention to them. A scientist who has a “great idea” for genetically engineered rutabagas except for that unfortunate side-effect (people who eat them break out in an uncomfortable rash) will not get far in the real world of food and agriculture. The mere fact that most products will not be developed unless there is a chance of making money from them weeds out lots of bad ideas (and unfortunately, as will be discussed below, some good ones). The market is a filter. Environmental protection and food safety agencies within government provide additional filters. The threat of a liability lawsuit may be the ultimate filter for many individuals and firms that contemplate introducing new technology. An awful lot of the bad ideas in food biotechnology will be eliminated from consideration whether working scientists or ordinary citizens adopt an ethical predisposition against food biotechnology, or not. These economic, regulatory and tort-based legal filters are a part of the social imaginary that structures technological innovation. From the scientist’s perspective, we might characterize it as a standard-opearting-procedure (SOP) for new agricultural technologies. Of course it is possible that these institutions have gone awry, so noting them is not to say that they are working perfectly. That is why Jonas’s position is critical to an ethical evaluation of novel technology, and why we should, indeed, be on the lookout for risks. Nevertheless, deep pessimism about technological threats must be tempered by the recognition that any technology faces a significant set of hurdles as a matter of course. Given the range of potential beneficial applications for agrifood biotechnology, one would expect that many cases will be presented for our consideration. Given the economic and regulatory filters that are already in play, many applications will never see the light of day as practical agricultural or food technologies. It is thus reasonable to expect that food biotechnologies able to work their way through the economic and

36

2 The Presumptive Case for Food Biotechnology

legal filters described above will meet the tests of Jonas’ principle of responsibility to a considerable degree. There is thus a purely methodological reason to adopt the following view: our cognitive filters should be on the alert for bad outcomes and products, rather than the reverse. I admit that there is a certain irony in this result. Being favorably disposed toward technology in general implies that ethicists’ work consists in trying to tear it apart. The principle of responsibility tells us to look for risks and concerns, to emphasize the bad side of technology. However, if we are doing this in the spirit of Jonas’s project, rather than Heidegger’s, we are only doing what is necessary to insure that novel technologies are implemented in a responsible way. As a result, most of this book is dedicated to biotechnology’s possible problems. This reasoning may sound contrary to technology boosters and skeptics alike. If he is for biotechnology, why is he spending all this time on problems? Alternatively and contrarily, if we are concerned about problems, why do we adopt an outlook presuming that biotechnology will be good? The answer to both questions is that responsibility requires us to look for reasons not to do something, but that also implies that when all of the precautionary concerns have been satisfied, we should probably allow it.

2.5 The Social Dimension of the Presumptive Case In addition, a strong presumptive case in favor of technology still exists within industrialized and industrializing economies. It is a component of the dominant social imaginary. This means that, although things could be different, for practical purposes the presumptive case is a social fact that people hoping to navigate the world had best learn to cope with. People expect change, and even if they do not expect it to be as uniformly beneficial as they once did, the traditional, static social structures, with their rigid social hierarchies and their lack of social mobility, are a thing of the distant past. This social fact may imply that most individuals in early 21st century society are inclined to favor technological change, but even if it does not, it shows that resetting the social imaginary against any broad form of technology will be very costly. It will be the life’s work many dedicated people, and they will have to be very persuasive. Furthermore, it will compete with other large social issues that confront different aspects of the dominant imaginary, such as opposition to racism and gender bias, as well as environmentalism and world peace. As such, the case against food biotechnology needs to be compelling to justify a social movement to reverse the status quo. However, a confirmed Heideggarian might well and truly cite the deep social institutionalization of a pro-technology worldview as confirming evidence for Heidegger’s suspicions. The fact that Western cultures are so accepting of technology implies that people are disinclined to consider alternatives that require a change in our preferences or the relinquishment of power. Indeed, there are critics of agrifood biotechnology who have situated their critique in explicitly Heideggarian terms, (see Heller 2007; Schyfter 2012; Metcalf 2013). What is more, critics who

2.5 The Social Dimension of the Presumptive Case

37

bemoan the “neoliberal” slant of contemporary societies may well be objecting to just this way in which our contemporary institutions are structured, (see Pechlaner and Otero 2010; Meghani 2017). The philosophy behind this complaint owes more to first generation critical theorists such as Theodor Adorno (1903–1969) or Herbert Marcuse (1898–1979) than it does to Heidegger. A philosophically rich response to either form of criticism lies beyond the scope of this book, but it is possible to make a few observations in defense of the presumptive case for biotechnology. First, one can incorporate concerns about social institutions into a risk-based approach, and I will argue that an ethically adequate appraisal of biotechnology requires that we do so. Critics who see biotechnology aligned with neoliberalism cite the prominence of institutions such as the World Trade Organization (WTO) in structuring the pattern of technology use, or see a tendency to regard living things as exchangeable commodities as evidence of moral decline. Examples of such concerns are discussed in almost every chapter. In other words, the eventualities that are feared by these critics are viewed as legitimate hazards, albeit with highly complex social and psychological mechanisms that determine the likelihood of their realization. As such, the presumptive approach does not imply that these institutions are ethically defensible, only that we can and should address them as outcomes and potentialities that are at least partially amenable to the standpoint of social choice. Secondly, a riskbased approach does adopt a starting point that is commensurate with the institutions that structure contemporary ways of life. I am arguing that this is a mark in its favor because by working through these institutions in form of social learning, the risk-based approach offers greater hope of learning and subsequent adaptation than taking a stance of pure resistance to all forms of technical change. We start where we are and try to do better. Third, although acts of resistance can be psychically rewarding, Jonas’s point was that we are midstream with many past technologies in need of corrective action. In this situation, resetting social filters to an anti-technology stance will require a herculean effort. Diverting energies to such a task only allows the harmful, disruptive and corrupting elements in our current milieu to continue unchallenged. Finally, some overbroad interpretations of the presumptive case do, in fact, need qualification. It is to that task that the next section turns.

2.6 Making the Case for Biotechnology Badly The presumptive case outlined here should not be interpreted as a blanket endorsement of biotechnology. An approach to technological ethics that evaluates risks against the background of a presumption that innovations do not need elaborate or detailed arguments asserting their benefits can still generate persuasive reasons to reject or constrain novel technologies. Unfortunately, a number of unreflective pro-technology arguments do circulate among defenders of gene technology. This section of the chapter qualifies the presumptive case by examining pro-technology arguments that commit logical or moral fallacies.

38

2 The Presumptive Case for Food Biotechnology

During the first half of the 1980’s, scientists, venture capitalists and university fund-raisers became highly practiced at making the case for both food and medical biotechnology in economic terms. They convinced funding agencies, administrators, state governments and private investors to place large sums of money at their disposal on promises of impressive financial returns and great wealth for all (see Teitelman 1989). Some of the ethical fallout from those promises is discussed in Chap. 10, but in short, biotechnology’s boosters made their case solely in terms of economic gain. Biotechnology was good because it was going to make everyone (or everyone who got on board soon enough) very rich. Needless to say, this is not a compelling ethical argument for biotechnology. Although the importance of economic returns and benefits should not be underestimated in ethical assessments, too much of the “case for biotechnology” consisted only in economic boosterism and whining about the negativism of the critics. Biotechnology’s boosters have done even more serious damage to their own case by offering several singularly bad arguments. The balance of this chapter will take on four bad arguments that seem to have many proponents among the scientists and decision makers who will ultimately determine the fate of food biotechnology. The first of these appeals to an outdated and naive notion of technological progress. It is the Modernist Fallacy. The second fallacy assumes an inappropriate reference group for making comparisons about the relative risks of genetic engineering. It is a version of the Naturalistic Fallacy, the common moral mistake of claiming that because something is natural, it is therefore good. The third fallacy also addresses risks of genetic engineering and is an instance of the Argument from Ignorance. The final argument emphasizing world hunger is dealt with at substantially greater length. I call it the World Feeder’s Fallacy. The first three bad arguments are examples of fallacious reasoning that one hears repeatedly at scientific meetings, both from the podium and over coffee. Anyone who has been present at such meetings has heard them, and it serves no positive purpose to single out any particular individual for attribution. Casual conversation is not a propitious setting for the production of an informed and rigorous ethical argument; however it is quite likely that most of the people offering these arguments actually believe that what they are saying is establishing an important point about the ethics of food biotechnology. The following criticisms are offered in the spirit of improving the quality of debate, rather than embarrassing individuals who hold these views.

2.6.1 The Modernist Fallacy One easy way to dismiss any and all ethical concerns that might be raised about virtually anything is to reply “That’s progress.” Advocates of food biotechnology have not resisted the temptation to deploy this reasoning, (if it can actually be called reasoning by any decent standard). The universal applicability of this strategy is a good reason for giving it a harder look. Other similarly universal replies to criticism (“That’s politics,” or “That’s life.”) signal one’s reluctance to discuss the matter

2.6 Making the Case for Biotechnology Badly

39

further without also conveying one’s moral approval of the state of affairs. “That’s progress,” implies that whatever ethical concerns or consequences have just been brought forward, they are the price that must be paid for progressive social change. Now, it may be correct to conclude that some social, animal, environmental or even human costs are a price that must be paid for ethically compelling reasons. If so, it is important to state those reasons and to justify the need to accept certain costs in order to achieve them. If a new rice or potato variety really does end hunger in a region of resource poor farmers, such a result may indeed be worth some loss of local cultural institutions. If a new procedure for inspecting meat really does decrease the risk of food borne disease significantly, it may indeed justify changes in the configuration of meatpacking or inspection that costs some jobs. There may also be ways to mitigate some of these costs, so the matter does not end here. Nevertheless, there are circumstances where it is appropriate to rebut an ethical critique by pointing out the compelling reasons for accepting certain costs in exchange for progress on other fronts. The Modernist Fallacy consists in presuming that science, technology, capitalism, or maybe just history is inherently progressive, so that any change brought about by these forces is always good. Alternatively, one may believe that any resistance to science, technology, etc., is a form of traditionalism or irrationalism that must be overcome. A strong, (often justified) faith in the power of science to alleviate harms, encourage democracy and promote social justice characterized the period in philosophy and economic history that is now known as Modernism. It had a good run, beginning with the philosophical writings of Francis Bacon (1561–1626) and René Descartes (1596–1650), and becoming socially effective during the industrial revolution. During this period, the open and skeptical pattern of scientific inquiry was indeed both a force and a model for the democratization of hierarchical societies, and the technologies of the industrial revolution led to the expansion of European civilization across the expanse of the globe. People will be debating whether Modernism was a good thing for some time to come. Certainly, it was less good from the perspective of conquered peoples than it seemed to Europeans who wrote much of the history for the period, but perhaps it is too much to lay all of the blame for colonial oppression at the feet of science and technology. The point here is that surely no one can take such an attitude of unalloyed optimism toward science and technology today. If the scientific and technological achievements of the last five centuries are on balance good, they can still be made much better by attending to environmental consequences, human health consequences, and social consequences that are the unintended accompaniment of science-based technical change. While only a few intellectuals challenged the philosophical basis for modernism until recently, much of the 20th century consisted in discovering the health and environmental consequences of the old smokestack industries and of chemical technologies. Social movements and intellectual developments that undercut the supreme self-confidence of the scientific attitude accompanied these discoveries, (Harvey 1989; Beck 1992). Science and the scientific attitude are capable of thriving without the social and cultural background of European expansion and

40

2 The Presumptive Case for Food Biotechnology

colonialism. Nevertheless, some of the less savory aspects of the science’s social and intellectual milieu tar the agricultural and food technologies of the present. The modernist fallacy is particularly important because it connects some critics of biotechnology to the deeper philosophical critique of the presumptive case. Jeremy Rifkin included a popularized diatribe against Bacon and Descartes in his books Algeny (Rifkin and Perlas 1983) and Declaration of a Heretic (Rifkin 1985) as does Andrew Kimbrell in The Human Body Shop, (Kimbrell 1993). More scholarly versions of the same argument can be found in books by Maria Mies (1993), Vandana Shiva (1993) and Ruth McNally and Peter Wheale (1995). These critiques were made before GMOs were actually in the field; they attack the very idea of genetic engineering, rather than its consequences. But the argument has not gone away since the release of genetically engineered crops in the late 1990s. Mae Wan Ho echoes it in here more biologically oriented critique, which also alleges a number of specific risks, (Ho 2000). Finn Bowring (2003) produced a book-length version of it that interprets developments in medical and agricultural biotechnology as part of a grand pattern in the history of science (Bowring 2003). This theme was echoed in Hugh Lacey’s sophisticated critique. Lacey argues that philosophically modernist science is founded on an ideal of dominance and control, while adopting a positivist rhetoric intended to stifle debate over values, (Lacey 2005). Timothy Morton argues that scientific reductionism in agricultural science precipitates a death-spiral that humanity cannot escape without overcoming the worldviews of the modern era, (Morton 2012). Lere Amusan and Seyi Olalekan Olawuyi stress how the repression of value debate places agriculture and food biotechnology in service to the profit-seeking behavior of corporations, (Amusan and Olawuyi 2019). I would not endorse most of these philosophical criticisms, but that is beside the point in the present context. To reply to such criticisms with “That’s progress,” is to beg the question, to commit the logical fallacy of assuming precisely the point that needs to be proven. A philosophically defensible application of the risk-based approach will consider each of these putative hazards on their own merits. The late 20th century may have been a period of overreaction, and biotechnology may even be unfairly falling victim to an obsessive fear of science and technology. Yet even if one believes that, one should not blithely maintain the sort of faith in the progressive nature of science and technology that would permit one to simply dismiss concerns about unwanted consequences without giving them their due. The presumptive case for food biotechnology that is given above is about as far as one can go. A less critical faith in progress is indeed blind faith, and the sort of faith that has been the enemy of science in the past. How ironic that some scientists become the least scientific in their willingness to dismiss concerns and objections to biotechnology! The Modernist Fallacy is a truly bad argument. It should be expunged from even coffee table conversation.

2.6 Making the Case for Biotechnology Badly

41

2.6.2 The Naturalistic Fallacy Philosopher G.E. Moore (1873–1958) described the Naturalistic Fallacy in his 1903 book Principia Ethica. It has since entered the philosophical lexicon as the logical mistake of concluding that something is good merely from the fact that it exists, that it is part of nature, of SOP or the status quo. The fallacy is likely to be committed by certain types of conservatives as well as by those who detest change. It is given a religious backing by those who believe that the world as it is embodies God’s design, but scientists are capable of the Naturalistic Fallacy, too. The instances of the Naturalistic Fallacy that occur in debates over biotechnology are subtle and a defensible argument can be made for key claims if one cares to do it. They involve making comparisons between natural phenomena and the behavior of transgenic organisms. Such comparisons are not in themselves problematic, but if the point of the comparison is to argue that the behavior of transgenic organisms is unproblematic or in some sense “acceptable,” because the behavior of non-transgenic (or natural) organisms is similar, then the natural phenomena are being invested with normative significance. Such arguments often involve claims about risk. Here are two arguments that exemplify the problem. 1. The kind of alterations that molecular biologists are making in plants and animals are just like those that occur as a result of natural mutation. They are, therefore, an acceptable risk. 2. Modern biotechnology is just like plant or animal breeding. Since the risks of plant and animal breeding have been acceptable, the risks of biotechnology are acceptable. The first version seems to state that because risks of biotechnology are consistent with risks from natural mutation, they are ethically acceptable. The second version states that because they are consistent with historical risks of plant and animal breeding, they are acceptable. The first argument is a clear instance of the Naturalistic Fallacy. Moore’s discussion has given this logical mistake its name (though his analysis was both more subtle and more philosophically ambitious than the account given here), but both David Hume (1711–1776) and John Stuart Mill (1806–1873) had called attention to the logical mistake years before Moore. Mill’s essay Nature noted that we can derive nothing of ethical significance by comparing intentional actions performed by human beings to acts of nature. “In sober truth,” he wrote, “nearly all the things which men are hanged or imprisoned for doing to one another are nature’s everyday performances,” (1874, p. 20). David Wiggins traces Mill’s discussion to a footnote in Hume’s 1751 An Enquiry Concerning the Principles of Morals, where he states that what is natural can be opposed alternately to what is unusual, what is miraculous and what is artificial. It is the contrast to artifacts that is relevant to gene technology, a contrast that obtains within the domain of the natural as conforming to the laws of nature. Wiggins’ cites the example of feeding sheep offal to cattle—the proximate cause for Mad Cow Disease—as an example of an artifice that ought never to have

42

2 The Presumptive Case for Food Biotechnology

been done, (Wiggins 2000). The mere fact that humans must live with the risks of mutations not produced through artifice tells us nothing about whether it is ethically acceptable for someone to act in such a way as to intentionally bring about such risks. The second instance of the fallacy at least compares like and like. Plant and animal breeding are intentional actions. However, it is not clear that society at large has ever undertaken an informed debate on whether these risks are acceptable, either. Indeed, stories of mistakes in planned introductions—Chinese carp and killer bees— are a commonplace theme in literature that raises concern about the environmental risks of genetic engineering for plants and animals. It is an open question as to whether a risk assessment of animal feeds would have detected the risk of prion diseases like Mad Cow. It is clear that the potential for this type of risk falls within the scope of a risk-based approach to technological ethics. More informed critics note that plant and animal breeding are often associated with increases in fertilizer or pesticide use, creating risk through an indirect mechanism. It is likely that any well-publicized change in food and agricultural technology like biotechnology would have brought on a new debate over risk. German social theorist Ulrich Beck has argued that many social issues once debated in terms of class conflict are now debated as issues of risk (Beck 1992). Given the dramatic changes in technology and social organization that have occurred since World War II, simply assuming that historical trends on risk levels provide evidence for contemporary criteria of risk acceptability is unwarranted. It is possible that what people who offer arguments like 1 and 2 above are trying to say is that the probability of harm from food biotechnology is quite low. This is not an ethical claim. It is an attempt to infer the probability of harm from food biotechnology by analogy to a distinct but relevantly similar sample population for which experience provides good (if not statistically quantified) information about the probability of harmful environmental or food safety consequences. There is nothing fallacious in this general pattern of inference, though inference by analogy can be tricky when examined case by case. I have discussed some of the philosophical problems that have arisen in plant scientists’ attempts to use this pattern of inference elsewhere (Thompson 2003), though they have largely been omitted from the subsequent treatment of environmental issues in this book. If one is careful in stating the point, however, there can be no objection to using such analogies in estimating risks from transgenic crops. However, low probability is not in itself enough to prove that a risk is acceptable. When consequences are sufficiently high, when risks are unnecessary, or when people are needlessly prevented from participating in a decision process, even very low probability risks can be socially unacceptable.

2.6.3 The Argument from Ignorance Philosopher Kristin Shrader-Frechette is known for her studies of faulty arguments used in developing the case for nuclear power, for geological disposal of nuclear waste, and for radiation technology in general. She notes that a persistent and disturbing fallacy in that literature that “occurs when one assumes that because

2.6 Making the Case for Biotechnology Badly

43

one does not know of a way for repository failure or radionuclide migration to occur, none will occur. Such inferences are examples of the appeal to ignorance,” (Shrader-Frechette 1993, 105). Technical disparities between radiation issues and biotechnology limit the lessons that one can learn from Shrader-Frechette’s work on nuclear waste, but virtually anyone with knowledge of the arguments that boosters of biotechnology have brought forward (especially in informal settings) will find the similarities disturbing. Advocates for biotechnology report “no evidence of harm (or risk)” associated with field experiments or farmer plantings of transgenic crops when in fact there is no evidence of any kind because no one has bothered to look. Some types of harm (such as rare allergic reactions) would be very difficult to detect, so the fact that none have been reported needs to be placed in proper context. Failing to do this is apt to be misleading. The fact that the argument from ignorance can mislead links its use to the public’s lack of receptivity toward biotechnology. Here is how that link is made: Replete with assurances about the safety of chemical technology and nuclear power, boosters of those technologies forged ahead. Many of their beliefs about the probability of an accident may have been well founded, but the public has become suspicious of such assurances in the wake of accidents at Bhopal at Chernobyl. While biotechnology may differ from chemical and nuclear technology in many ways, the conduct of the science community is, from an outsider’s perspective, distressingly similar. The appeal to ignorance has failed before; perhaps it will fail again. Skepticism over the reliability of expert assurances leads to another type of argument from ignorance: since we cannot be certain about the effects of an innovation, we infer that they will be bad. This becomes a blanket argument against any new technology that draws inspiration from Jonas’ Prinzip verantwortung in the form of arguments that appeal to the precautionary principle. Specific cases in which the precautionary principle has been used in arguments against agrifood biotechnology are discussed in later chapters. Some readers of Jonas argue that his recognition that technology can be beneficial notwithstanding, he intended the Prinzip verantwortung as a general warning against the expansive growth of technology. They see this as consistent with a version of the precautionary principle that enjoins governments to take action against technological risks well before scientific certainty is achieved, (Guillamme 2019). The question is, how much before? An application of the precautionary principle based on nothing more than the speculation, “Something bad might happen,” is an appeal to ignorance that should be rejected as soundly as the advocate’s appeal to “Haven’t seen any problems so far.” However, the risk-based approach itself opens the way to precautionary arguments that are credible. If hazard identification has determined a class of harmful outcomes consistent with mechanisms of action in the relevant hazard domain, a precautionary approach dictates defensive action even before the probability that the outcome will occur has been estimated with scientific precision. There may still be disagreement about mechanisms of action and hazard domains. As this book argues, biotechnology advocates have been dismissive toward hazards in the socioeconomic domain. Yet, socioeconomic mechanisms of action may be better understood than epidemiologically and toxicologically derived risk factors, even as it is notoriously difficult to make precise

44

2 The Presumptive Case for Food Biotechnology

predictions of the timing or conditions in which these mechanisms operate. More detailed discussion of these specific precautionary arguments occurs throughout the book. As in the naturalistic fallacy, there are valid inferences that can be drawn from the fact that one cannot imagine how a harmful consequence could materialize, as well as from worries that one might materialize, even when one can’t say how. Risk assessment is a process that begins with a systematic attempt to imagine the scenarios and mechanisms that can end in harm. It is inevitable that the scenario no one thinks of will be omitted from the estimate of risk that such exercises produce. Nevertheless, when scientists work diligently to anticipate the full complement of risks, it is reasonable to conclude that the scenarios no one thought of are unlikely. It is not reasonable to think (and no judicious scientist would claim) that the unanticipated scenario does not exist, though this is what the appeal to ignorance effectively does claim. Complacency arises easily when appeals to ignorance go unchallenged, and complacency can result in the exercise of risk analysis being pursued less diligently than it should be. If a risk-based review of biotechnology is to be pursued in an ethical manner, the appeal to ignorance must be expunged from both daily practice and the public discourse on biotechnology.

2.6.4 The Argument from Hunger While modernist, naturalist and ignorance fallacies circulate over coffee whenever scientists congregate, a more complex and insidious bad argument for biotechnology has become firmly entrenched in public discourse. This is the World Feeder’s Fallacy: agricultural biotechnology is the only solution to world hunger. It is generally accompanied by the claim that those who oppose it are themselves ethically for the disease and starvation that their opposition to agrifood biotechnology is alleged to cause. I hope that no one will take issue with the three fallacies discussed above, but many clearly do think that agricultural biotechnology holds such great hope for world’s poor and dispossessed that opposing it is morally wrong. As such, it is important to devote a bit more attention to making the case for viewing the claim that the world needs biotechnology to feed itself. Tracing the history of the argument from hunger would itself be a substantial task, even if one were to confine the topic to its use as an argument for biotechnology. There has always been some hope among agricultural scientists that rDNA techniques would be useful in developing new crop varieties for the developing world. This hope started to emerge as an explicit argument for biotechnology as developed country Bt and herbicide tolerant crops began to encounter serious opposition in the 1990s. Advocates of biotechnology began to look for a “poster child”: a biotechnology that was so appealing it could be used to silence the critics. One candidate was Charles Arntzen’s plan to develop a banana capable of delivering vaccines as a means of fighting tropical disease, (Arntzen 2015). The one that eventually achieved public notoriety was Ingo Potrykus’s “Golden Rice,” the vitamin-A

2.6 Making the Case for Biotechnology Badly

45

enhanced rice variety intended as a partial response to a widespread nutritional deficiency. Potrykus appeared on the cover of Time Magazine in July 2000 and the accompanying story touted his work as an important advance in the battle against the ills of poverty (Nash 2000). The story precipitated a continuing series of exchanges between boosters and knockers debating the value of Golden Rice for meeting nutritional needs. David Castle and Michael Ruse have collected articles representing both sides of this exchange in their book Genetically Modified Foods: Debating Biotechnology (2002). The argument from hunger has been articulated in more general terms by several distinguished agricultural scientists. Per Pinstrup-Andersen, a Danish economist with long experience in international development, has directed this argument directly to a European audience. Pinstrup-Andersen holds Europe accountable for the reluctance of developing countries to adopt products of agricultural biotechnology due to either concerns about their ability to export into European or markets or more straightforward fears based on Europeans’ reluctance to accept GM crops. (Pinstrup-Anderson and Schiøler 2000). Norman Borlaug (1914–2009), who won the Nobel Peace Prize for his work on green revolution crops, has stressed biotechnology’s capacity to aid the poor and hungry people of the world in a number of fora, calling opponents of biotechnology “anti-science zealots,” (Borlaug 2000, 2001, 2002). Political scientist Robert Paarlberg has produced a book-length version of the argument. Paarlberg’s claims are specific to the African context. Like Pinstrup-Andersen, he argues that European anti-biotechnology activism has prevented adoption of GMOs that would have been used by African farmers but for the fact that Europe is the primary market for African exports, (Paarlberg 2009). Pinstrup-Anderson, Borlaug and Paarlberg each make a cluster of claims. One is that opponents of biotechnology are immoral in virtue of the harm that they are doing to needy people. This extravagant claim goes considerably beyond the suggestion that biotechnology can be used to develop crops and livestock that increase incomes and reduce food insecurity. My critique is not addressed to this more modest claim; biotechnology’s potential for hunger-reducing outcomes is not under attack. The argument from hunger makes a stronger claim: ethical objections to biotechnology (such as those reviewed in the balance of this book) are mooted by biotechnology’s capacity to address world hunger. They become first irrelevant and then immoral in themselves in virtue of the way that they retard the rapid uptake of potentially beneficial GMOs. The argument from hunger surfaced again in the summer of 2002 when several African countries refused U.S. food aid because it was not certified as “GM free.” The story received substantial play in the U.S. media, where it was generally portrayed as a case of moral insensitivity on the part of African and European leaders, allowing people to starve for fear that future export markets would be lost. There is little doubt that African rejection of even milled cornmeal (maize) went beyond reasonable precaution. However, these stories failed to note that the U.S. routinely takes pains to satisfy purely aesthetic preferences in the delivery of food aid. For example, most Africans prefer white rather than yellow maize, and the U.S. Agency for International Development (USAID) takes steps to satisfy that preference. Furthermore, since large

46

2 The Presumptive Case for Food Biotechnology

maize producing regions in the U.S. were not growing GE varieties in 2002, it would have been easy for USAID to have satisfied a preference for non-GM food aid, as well. If anyone was actually starving while all the dawdling was going on, U.S. officials were blamable as surely as African leaders. In May of 2003, the food aid episode became the centerpiece in a U.S. trade action against the European Union’s continuing reluctance to accept GM crops. The argument from hunger has been imbedded in cynical and strategic manipulations from the outset, and it is tempting to write it off entirely as a particularly odious form of deceit perpetrated to defame honest critics and dismiss legitimate concerns. Nevertheless, the argument from hunger is complex because for the first time in the history of agricultural science, the developing world is broadly positioned to make substantial use of innovative techniques. Not surprisingly, the greatest capacity for using science to develop new agricultural technology resides in Western Europe, North America and a few industrially developed countries such as Japan, Australia and New Zealand. It has been this way since the dawn of agricultural science in the 19th century laboratories of Justus von Leibig (1803–1873) and Joseph Henry Gilbert (1817–1901). The much-maligned Green Revolution was largely an attempt to adapt agricultural technologies from the sphere of European influence to growing conditions in Africa, Asia and Latin America. For a variety of reasons, scientists in these areas have a much greater capacity to use biotechnology in response to their own problems than has been the case for agricultural technologies that depend heavily on traditional chemical, mechanical and even breeding expertise. Agricultural scientists in less developed countries continue to work closely with developed country science through institutions such as the World Bank, the Rockefeller Foundation, the Gates Foundation and the Consultative Group on International Agricultural Research (CGIAR), which coordinates the activities of national and non-profit development agencies. As such, it is really true that agricultural biotechnology might well be deployed in response to some genuine problems faced by poor and hungry people in the developing world (see Rosegrant and coauthors 2001; Nuffield Council 1999, 2003). The irony is that just as the developing world has achieved this capability, other forces have conspired to frustrate its exploitation. For one thing, critics of the Green Revolution, which did in fact achieve impressive gains in agricultural yields at the occasional expense of environmental costs and the displacement of poor farmers and landless labor, have been gaining steam for three decades. There is now organized opposition to new agricultural technology in the developing world. For another thing, opposition to agricultural biotechnology in the developed world, especially Europe, has created a climate of suspicion and doubt about this technology that is slowing its adoption in developing countries. Countries that export agricultural commodities to places that have imposed a ban on GM foods have been especially susceptible to this worry, and the food aid episode of 2002 is evidence of this problem. While there is a strong case for using biotechnology in the developing world, events have transpired to create hurdles for deploying it, hurdles that did not exist thirty years ago when the capacity for indigenous scientific work was considerably less (Scoones and Glover 2009).

2.6 Making the Case for Biotechnology Badly

47

It is, however, a rather large leap in logic to move from this carefully stated claim to the claims that biotechnology holds the solution to hunger, or that opposition to biotechnology is morally irresponsible, much less the even stronger claim that opponents of biotechnology are committing acts tantamount to the murder of starving people. Yet all these immoderate claims are heard in defense of agricultural biotechnology. Biotechnology cannot be said to hold the solution to world hunger because as Amartya Sen demonstrated in the path breaking book Poverty and Famines: An Essay on Entitlement and Deprivation (1981), the misery and suffering of the poor is never due simply to a lack of food. While the techniques now in the hands of developing country scientists might increase yields and will almost certainly help developing country farmers reduce losses from disease and insect pests, solving hunger involves a reform of social institutions that deprive poor people of secure economic and political resources. Lacking these, there will still be hunger, even when there is plenty of food. In fact, some portion of the opposition to biotechnology comes from people who are arguing that social reforms must accompany technical change in developing countries. This claim is at the root of Vandana Shiva’s argument against biotechnology (Shiva 2000) and is stated repeatedly in grass roots literature coming out of India. While it is certainly true that some opposition to biotechnology has little to do with a concern for social inequality, other forms of opposition are deeply committed to addressing issues of social inequality, especially by insisting that new technologies be accompanied by needed social reforms. To tar biotechnology’s critics broadly as being unconcerned about the poor is either ignorant or cynical in the extreme. The argument from hunger is also insidious because even those who reject it often do so with an equally fallacious and irresponsible reply: the problem is not a lack of food, but a matter of distribution. Like the argument from hunger itself, this comeback has a grain of truth. Sen’s analysis supports the claim that hunger is a problem of distributive justice, but to say this is not to say that the problem would be solved by redistributing food, as if what we need are more boats and trucks. To think that hunger will be solved by exporting surplus production from industrialized countries to the developing world is just as naïve as thinking that a new potato or rice variety is the answer. Many critics of biotechnology underestimate the need to maintain and continuously improve humankind’s capacity for biologically-based responses to problems in agriculture. The productivity of industrial agriculture cannot be regarded as a permanent achievement. Not only does it involve levels of water and energy use and forms of pollution that are themselves creating problems, but diseases and pests are constantly evolving and will eventually become resistant to technologies that hold them in check. It may be unnecessary to state such obvious points in a text written for scientists and leaders in agriculture, but it is critical that the case for biotechnology be built upon this more subtle and valid foundation, rather than on a simplistic and ultimately misleading portrayal of its ability to feed the world. The argument from hunger is a bad argument not because there is no truth in claiming that rDNA techniques will be an important part of the toolkit for agricultural scientists who work to improve food production in the developing world. Nor is it false to suggest that the current climate of opposition to biotechnology is slowing the progress of work that is currently underway. Nevertheless, agricultural scientists’

48

2 The Presumptive Case for Food Biotechnology

desire to have things the way they were forty years ago is probably not a defensible position. It might not be a bad thing to have technical change go a little more slowly and more deliberately in the developing world, especially if the slowness is because people in vulnerable positions have attained a modicum of power. Once one has witnessed starvation, the imperative for change becomes paramount and impatience starts to look like a virtue. Nevertheless, the main thrust of the argument to come in the balance of the book is that meeting the concerns and criticisms of opponents is among the ethical responsibilities that agricultural scientists and decision makers must accept. Telling people to buzz off because we are busy helping the poor simply will not do. While it is certainly possible to take a different view of how far scientists, government officials and industry leaders need to go in meeting the views of critics, it is something else again to promote a simplistic view of poverty and deprivation in order to bring about better public acceptance of biotechnologies that are being used in industrial agriculture today. The argument from hunger is a bad argument because it has been deployed shamelessly and cynically in a manner that promotes continued misunderstanding of the problems of global hunger and of agricultural science’s role in addressing them.

2.7 Conclusion The presumptive case for agrifood biotechnology is not in itself an argument for biotechnology. The existence of social filters creates an expectation (at least among those who work with and develop technology) that the applications of rDNA techniques in food and agriculture that run this gauntlet are at least as likely to be beneficial as harmful. Regulation and economic competition further filter out bad ideas. This justifies a disposition or poise toward biotechnology: absent reasons not to, it should go ahead. All of which may simply be to say that at present, agricultural biotechnology is a social fact, an artifact of the dominant social imaginary. The organizations that support and govern the food system have deployed people with expertise in gene technology throughout. Such people are poised to use biotechnology and any attempt to consider the ethics of biotechnology in agriculture and the food system must begin with this fact. This is not to say that technology is always or automatically good, for one can maintain a presumptive bias in favor of technology only under the condition that scientists, government officials and the private sector make faithful attempts to evaluate technology, and to correct or mitigate its unwanted consequences. Technology must be monitored, but responding to the problems created by yesterday’s technical fix will, as often as not, require more technology, not less. Hans Jonas‘principle of responsibility articulates the basic ethical rationale for review of new technology, and the concepts of risk analysis provide a conceptual framework of completing Jonas’s project. The larger aim of this book is to work through the conditions that have been proposed to limit the presumptive case for biotechnology, discarding some,

2.7 Conclusion

49

endorsing others. This means that much of the discussion will be focused on criticisms and negatives. Yet the ethics of food and agricultural biotechnology is not simply a matter of limits and constraints, for the promise of biotechnology is real, substantial and should not be ignored.

References Amusan, L., and S.O. Olawuyi. 2019. Between food quality and quantity for all in Africa: What they refuse to tell us about GMO foods. Gender and Behaviour 17: 12284–12298. Arntzen, C. 2015. Plant-made pharmaceuticals: From ‘Edible Vaccines’ to Ebola therapeutics. Plant Biotechnology Journal 13: 1013–1016. Beck, U. 1992/1986. Risk Society: Towards a New Modernity. In: (trns) M. Ritter, London: Sage Publications. Borlaug, N. 2000. The Green Revolution Revisited and the Road Ahead. Special 30th Anniversary Lecture. Oslo: Norwegian Nobel Institute. Borlaug, N. 2001. Ending world hunger. The promise of biotechnology and the threat of antiscience zealotry. Plant Physiology 124: 487–490. Borlaug, N. 2002. Feeding a world of 10 Billion People: The miracle ahead. In Vitro Cellular and Development Biology. Plant 38: 221–228. Bowring, F. 2003. Science, Seeds and Cyborgs: Biotechnology and the Appropriation of Life. London: Verso Press. Burkhardt, J. 1997. The Inevitability of animal biotechnology? Ethics and the scientific attitude. In Animal Biotechnology and Ethics, ed. A. Holland and A. Johnson, 114–132. London: Chapman and Hall. Busch, L. 2011. Standards: Recipes for Reality. Cambridge, MA: The MIT Press. Castle, D., and M. Ruse (eds.). 2002. Genetically Modified Foods: Debating Biotechnology. Buffalo, NY: Prometheus Books. Cook, G., E. Pieri, and P.T. Robbins. 2004. ‘The scientists think and the public feels’: Expert perceptions of the discourse of GM food. Discourse & Society 15: 433–449. Gasper, D. 1997. Sen’s capability approach and Nussbaum’s capabilities ethic. Journal of International Development 9: 281–302. Glazebrook, T., and M. Story. 2015. Heidegger and international development. In Heidegger in the Twenty-First Century (eds) T. Georgakis and P.J. Ennis, pp. 121–139. Dordrecht: Springer. Guillaume, B. 2019. The roots of Hans Jonas’ ethics of the future, and precaution. Journal of Applied Ethics and Philosophy 10: 1–7. Hartley, S., and K.M. Millar. 2014. The challenges of consulting the public on science policy: Examining the development of European risk assessment policy for genetically modified animals. Review of Policy Research 31: 481–502. Harvey, D. 1989. The Condition of Postmodernity: An Enquiry into the Origins of Cultural Change. Oxford: Basil Blackwell. Heller, C. 2007. Techne versus technoscience: divergent (and ambiguous) notions of food “quality” in the French debate over GM crops. American Anthropologist 109: 603–615. Hettich, P., and S. Walther. 2012. Bleak prospects for research in GMP in Switzerland. European Journal of Risk Regulation 3: 361–371. Ho, M.-W. 2000. Genetic Engineering: Dream or Nightmare?, 2nd ed. New York: The Continuum Publishing Co. Revised and Expanded. Jonas, H. 1966. The Phenomenon of Life: Toward a Biological Philosophy. New York: Harper. Jonas, H. 1984. The Imperative of Responsibility: The Search for Ethics in a Technological Age. Chicago: U Chicago Press.

50

2 The Presumptive Case for Food Biotechnology

Katz, S. 1997. Secular morality. In Morality and Health, ed. A.M. Brandt and P. Rozin, 297–330. New York: Routledge. Kimbrell, A. 1993. The Human Body Shop. New York: HarperCollins. Lacey, H. 2005. Values and Objectivity: The Current Controversy about Transgenic Crops. Lanham, MA: Lexington Books. McNally, R., and P. Wheale. 1995. Genetic engineering, bioethics and radicalised modernity. In Contested Technology: Ethics, Risk and Public Debate, ed. R. von Schomberg, 29–50. Tilburg: International Centre for Human and Public Affairs. Meghani, Z. 2017. Genetically Engineered Animals, Drugs, and Neoliberalism: The Need for a New Biotechnology Regulatory Policy Framework. Journal of Agricultural and Environmental Ethics 30: 715–743. Metcalf, J. 2013. Meet shmeat: Food system ethics, biotechnology and re-worlding technoscience. Parallax 19: 74–87. Mies, M. 1993. New reproductive technologies: Sexist and racist implications. In Ecofeminism, eds. M. Mies and V. Shiva, 174–195. London: Zed Books. Mill, J.S. 1874. [1961] Nature. In The Philosophy of JS Mill: Ethical, Political and Religious, ed. M. Cohen. New York: Modern Library. Morton, T. 2012. The Oedipal logic of ecological awareness. Environmental Humanities 1: 7–21. Nash, J.M. 2000. Grains of hope. Time Magazine 156 5 (July 31): 38–46. NRC (United States National Research Council). 1996. Understanding Risk: Informing Decisions in a Democratic Society. Washington, D.C.: National Academies Press. Nufield Council on Bioethics. 1999. Genetically Modified Crops: The Ethical and Social Issues. London: Nufield Council on Bioethics. Nufield Council on Bioethics. 2003. The Use of Genetically Modified Crops in Developing Countries: A Follow-Up Discussion Paper to the 1999 Report. London: Nuffield Council on Bioethics. Paarlberg, R. 2009. Starved for Science: How Biotechnology Is Being Kept Out of Africa. Cambridge, MA: Harvard University Press. Pechlaner, G., and G. Otero. 2010. The neoliberal food regime: Neoregulation and the new division of labor in North America. Rural Sociology 75: 179–208. Pinstrup-Andersen, P., and E. Schiøler. 2000. Seeds of Contention: World Hunger and the Global Controversy Over GM Crops. Baltimore: Johns Hopkins University Press. Rifkin, J. 1985. Declaration of a Heretic. Boston and London: Routledge and Kegan Paul. Rifkin, J. with N. Perlas. 1983. Algeny. New York: Viking. Rosegrant, M.W., M.S. Paisner, S. Meijer, and J. Witcover. 2001. 2020 Global food outlook – trends, alternatives and choices. Washington DC: IFPRI. Schmid, A.A. 2004. Conflict and Cooperation: Institutional and Behavioral Economics. Oxford, UK and Malden, MA: Basil Blackwell. Schyfter, P. 2012. Standing reserves of function: A Heideggerian reading of biology. Philosophy & Technology 25: 199–219. Scoones, I., and D. Glover. 2009. Africa’s biotechnology battle. Nature 460: 797–798. Sen, A.K. 1981. Poverty and Famine: An Essay on Entitlement and Deprivation. Oxford, UK: Oxford U Press. Sen, A.K. 2011. The Idea of Justice. Cambridge, MA: Harvard University Press. Shiva, V. 1993. Women’s indigenous knowledge and biodiversity conservation. In Ecofeminism, ed. M. Mies and V. Shiva, 164–173. London: Zed Books. Shiva, V. 2000. Stolen Harvest: The Hijacking of the Global Food Supply. Cambridge, MA: South End Press. Shrader-Frechette, K.S. 1993. Burying Uncertainty: Risk and the Case Against Geological Disposal of Nuclear Waste. Berkeley: University of California Press. Sunstein, C. 2002. Risk and Reason: Safety, Law and the Environment. New York: Cambridge University Press.

References

51

Teitelman, R. 1989. Gene Dreams: Wall Street, Academia and the Rise of Biotechnology. New York: Basic Books. Thompson, P.B. 2003. Value judgments and risk comparisons: The case of genetically engineered crops. Plant Physiology 132: 10–16. Thompson, P.B. 2010. The Agrarian Vision: Sustainability and Environmental Ethics. Lexington: University Press of Kentucky. Valles, S.A. 2018. Philosophy of Population Health: Philosophy for a New Public Health Era. New York: Routledge. Wickson, F. 2007. From risk to uncertainty in the regulation of GMOs: social theory and Australian practice. New Genetics and Society 26: 325–339. Wiggins, D. 2000. The Presidential address: Nature, respect for nature, and the human scale of values. The Proceedings of the Aristotelean Society 100: 1–32. Zimdahl, R.L. 2006. Agriculture’s Ethical Horizon. San Diego: Academic Press.

Chapter 3

Biotechnology, Policy and the Problem of Unintended Consequences: The Case of rBST

Abstract This chapter applies the ethics framework for evaluating emerging technology developed in Chaps. 1 and 2. It illustrates the application of a risk-based approach to the ethical analysis of agrifood technology by reviewing the policy debate in the United States over the first important product of agrifood biotechnology, recombinant bovine somatotropin (rBST). The US debate anticipated European reactions to biotechnology by a decade. It formed the basis for a regulatory approach that continues to influence the governance of agrifood biotechnology in the United States to this day. At the same time that this approach established relatively strong standards for safety evaluation, it precludes governance on other ethically relevant criteria. This forces critics to mount all resistance within the narrow window of legally actionable safety standards. The rBST case thus explains why activists who were opposed to trends in mainstream industrial agriculture on environmental and socioeconomic grounds adapted their protests to include more speculative concerns about food safety. Keywords Risk · Food safety · Animal health and welfare · Social impact · Environmental impact · Ethics and governance of technology Philosopher Hans Jonas published the German edition of The Imperative of Responsibility: In Search of Ethics for the Technological Age, in 1979. As noted in Chap. 2, Jonas called for an ethic of responsibility that would neither demonize nor sanctify science and technology, but that would use science and technology as aggressively as possible in a systematic inquiry into the unintended and unwanted impacts of technological change (Jonas 1984). The book seemed unexceptional in many respects at the time of its publication. In retrospect, Jonas’s analysis was wiser than we knew. While a bald statement of the ethic of responsibility seems trivial, in calling for a view of science and technology that navigates between the rocks of overenthusiasm and the shoals of Luddism, Jonas was challenging science and society to recognize the fallacy of modernism, and to find a new way to cope with technology’s inevitable unwanted consequences. Jonas recognized that this would unavoidably involve new forms of interaction between science and government, and he devoted considerable

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_3

53

54

3 Biotechnology, Policy and the Problem of Unintended Consequences …

space in The Imperative of Responsibility to a comparison between capitalism and socialism. The early years of biotechnology saw philosophers make several attempts to flesh out Jonas’ Prinzip verantwortung, making specific applications to gene technology. Stephen Stich, the Princeton University philosopher best known for his work in cognitive science and the philosophy of mind wrote two essays on gene technology that predate his more celebrated work. Bernard Rollin, who is jointly appointed in philosophy and veterinary medicine at Colorado State University, wrote a series of articles on genetic engineering of animals, culminating in his book The Frankenstein Syndrome. In the United Kingdom, Michael Riess and Roger Straughan produced a book-length study intended to cover both biomedical and agricultural applications of biotechnology, with Straughan offering a number of subsequent contributions specific to food. As discussed below, Jeffrey Burkhardt produced a condensed argument focused on one particular technology, the recombinant bovine growth hormone. The first edition of this book could be added to that list of early efforts. Entries by Gregory Pence and Hugh Lacey appeared about a decade later, well after the first generation of bioengineered crops were in the fields. Although these treatments vary in their authors’ enthusiasm for agricultural and food biotechnology, they all share some version of the risk-based approach to technological ethics. They see ethics as contributing crucial normative information to the anticipation and management of technology’s unwanted consequences. Simultaneously, a consortium of activists came together in opposition to gene technologies. In Rachel Shurman and William Munro’s sociological study of this group, its members are motivated primarily by social concerns. Shurman and Munro describe the formation of a social movement in opposition to agrifood biotechnology emerging from interactions among individuals committed to various social reforms, only a few of which revolved around the food system. These social advocates were mobilized by what they saw as an effort to bring an entirely new domain under the control of profit-seeking corporations. Decisions permitting patents on gene processes, gene sequences and on entire organisms were central to their concern. The recombinant bovine growth hormone case was pivotal in the formation of the antiGMO social movement. It was an early public policy decision case where activists were able to mobilize significant opposition to the technology. Efforts to oppose the technology also served as a conduit for cooperation between North American and European civil society organizations. This consortium undertook much of its advocacy by framing the issues in terms of risk, but Shurman and Munro portray key individuals as objecting to the implicit intentions of innovators, rather than the potential impact of their innovations. They saw biotechnology as an attempt to commodify life itself, (Shurman and Munro 2010). The decade long debate over recombinant bovine growth hormone serves as both a central episode in establishing the terms of debate over agrifood biotechnology and as a microcosm of the terms in which the larger controversy developed. This chapter begins with a brief orientation to rBST, followed by a review of the work by Stich, Rollin and Burkhardt. Each describes a risk-based framework for evaluating the ethics of agrifood biotechnology, and Burkhardts’s is specific to the recombinant

3 Biotechnology, Policy and the Problem of Unintended Consequences …

55

growth hormone debate in the United States. The chapter then works quickly through four types of risk as applied to rBST: food safety, animal health, environmental and socioeconomic. The concluding section of the chapter is speculative. Perhaps if the case had been handled differently, the entire international debate over agricultural and food applications of gene technology might have played out differently.

3.1 What is rBST? Why Does it Matter? Bovine somatotropin (BST) is the growth hormone specific to Bos taurus, the common dairy cow. All mammals produce somatotropin or growth hormone naturally. Although the specific chemical composition of somatotropins vary, gene constructs that code for producing growth hormone were among the first to be identified owing to the enormous commercial potential for synthetically produced human growth hormones (HGH), (Mayne and coauthors 1984). A research group at the University of California at San Francisco announced that they had cloned the sequence for BST in 1980, (Miller et al. 1980). Versions of Bos taurus hormone produced by engineered microorganisms are referred to alternately as recombinant bovine somatotropin (rBST) and recombinant bovine growth hormone (rBGH). As with many issues in the agrifood biotechnology debate, choice of terminology matters. The consumer and farm advocates who opposed this technology tended to call it a growth hormone, while those who supported it called it somatotropin. I will alternate my usage throughout this chapter. Somatotropin regulates not only growth but also lactation. In postpartum females, the production of growth hormone stimulates lactation, and milk production ceases when levels of growth hormone return to normal, background levels. When administered to cows under carefully managed conditions, bovine somatotropin increases milk production and extends lactation for a period of several weeks or months. As discussed at more length below, administering bovine somatotropin to cows can cause significant increases in a dairy’s milk production. However, using natural bovine somatotropin harvested from cows for this purpose is not economical because of the high production cost of the hormone. Several animal drug companies including Elanco (a division of Eli Lilly), Monsanto (now Bayer), and Upjohn (now Pfizer) succeeded in developing a recombinant transformed microorganism to produce bovine somatotropin during the 1980s. The Monsanto version, trade-named Posilac™, was approved for use in the United States in the Fall of 1993. The story in other countries with well-developed regulatory systems (and significant levels of dairy production) is different. Canada has never approved rBGH, and the substance has been banned in Europe (Brinckman 2000; Buttel 2000). Jeffrey Burkhardt built upon Jonas’s ethic of responsibility in his important article on the ethical significance of rBST. He divided Jonas’s ethic into five components. First, there are ethical questions that must be raised with respect to any individual’s use of a tool or technique. Second, ethical principles should govern the decisions that groups (or society as a whole) make to adopt techniques. Third, there are ethical

56

3 Biotechnology, Policy and the Problem of Unintended Consequences …

issues that must be raised about how the choice to adopt or reject a technology is framed. An advocate of technology who presents technical change (or “progress”) as inevitable has not made a fair presentation of alternatives The fourth area arises with respect to decisions to research and develop specific technologies, not just to adopt or reject them. Finally, the broadest dimension concerns “the technological ethos,” or society’s disposition toward science and technology expressed as a form of culture (Burkhardt 1992, 226–231). Technical changes raise ethical questions at each of these levels, yet it seems likely that naïve readers of Jonas were thinking primarily of Burkhardt’s first or second level, at best. The unintended consequences of technical change permeate culture, and eventually include even the religious questions raised in Chap. 11. Yet, as Burkhardt noted in his 1992 analysis, it is the near term health, environmental and social impacts of biotechnology that dominate a policy debate, (Burkhardt 1988, 53). The larger cultural issues become disassociated from the specific technical application and become redefined as political questions that address how we cope with health, environment and economic risks within our governance institutions. Arguably, biotechnology has done a better job of coping than some technologies, but as Chaps. 4 through 10 demonstrate, these are complex issues. This chapter offers a synoptic treatment of the ethics of food biotechnology, confined to the case that Burkhardt discussed in 1992. While the debate over rBGH may seem like ancient history to twenty-first century readers, there are four good reasons for taking the time to look more closely at this debate. First, a tight and short focus on one case provides a roadmap for thinking ethically about other new applications of biotechnology. Although rBST differs in important respects from other biotech products that have been or will be developed, this is a defect shared by every possible case study that might be proposed. Second, although rBST has disappeared from headlines, there is an important sense in which this case is far from “over.” As the succeeding discussion shows, U.S. regulatory agencies approved rBST in 1993, but nearly thirty years later, few industrialized countries have followed suit. The issues thus remain controversial, as regulatory and other bodies—the social filters alluded to in Chap. 2—have handled this case in a very uneven fashion. The public’s reaction has also shifted over time. rBST was used extensively in the United States immediately after its approval in the early 1990s, but by 2010 consumers were willing to pay premium prices for non-rBST milk. Practice in the industry shifted accordingly, and economic filters now dominate, even in the United States, (Wolf et al. 2011). Third, the rBGH controversy has value as an object of analysis for those who study science, risk and political power. In addition to Burkhardt’s original paper and other studies discussed below, the rBGH case has been the focus of a paper by Buttel (2000) and a book length study by Guehlstorf (2004), both completed after this chapter was written in the mid 1990s. Dozens of scientific papers on rBGH continue to be published every year. rBGH also continues to surface in articles and books for the general public. For example, David Robinson Simon’s Meatonomics, an expose on practices in industrial meat production, includes a discussion of how the industry attempts to conceal the use of rBGH, (Simon 2013, pp. 59–60). As one of the first

3.1 What is rBST? Why Does it Matter?

57

products to move through the gauntlet of regulatory approval and market acceptance, it seems likely that future studies will find it useful to touch upon the rBGH debate for some time to come. Finally, it is important for readers who may not have been paying attention between 1984 and 1994 to recognize that opposition to biotechnology was not invented by Europeans in 1998. As Munro and Shurman demonstrated, Americans debated rBST intensely and vociferously leading an unprecedented Congressional action that delayed release of the product even after regulatory reviews were complete. It was in some respects a key test case for U.S. opposition to biotechnology (Shurman and Munro 2010, pp. 127–134). The genetic engineering that made rBST possible was performed on a microbe, which in turn produces rBST for use on dairy cattle. Although regulatory issues for genetically engineered food and research animals certainly differ from issues associated with genetically engineered animal drugs, the politics of the rBST case nevertheless prove a useful object lesson in thinking through the unintended consequences of food biotechnology more generally.

3.2 Biotechnology Policy and Philosophy Scholars in technology studies exhibit a depressing tendency to rediscover already well-studied topics, sometimes only a few years after the publication of seminal work. The rBGH case has largely escaped the attention of many—arguably most— social science researchers who have contributed to the scholarship on food ethics and agricultural biotechnology. American philosophers now taking up work in food ethics have been especially slow to recognize early work on the risks of gene technology. In fact, Stephen Stich published a philosophical article on the ethics of recombinant DNA controversies in 1978. Stich’s article appeared in the wake of the 1976 conference at Asilomar where leading scientists debated the risks inherent in genetic engineering. Stich reviewed “bad arguments” that surfaced in both scientific and lay debates. He defended an approach that took the ethical responsibilities of scientists seriously, but that interpreted those responsibilities largely in terms of anticipating and mitigating risks (Stich 1978). A few years later, Stich gave a more detailed formulation of his approach to ethical evaluation of gene technology. Here, Stich examines strengths and weaknesses of risk–benefit analysis as applied to gene technology in general, which would include then novel applications in medicine and drug development. Though I would argue that Stich misrepresents the way that economists operationalize the approach, his six-step characterization of the process is what matters here. Step 1 requires the enumeration of alternative policies, a potentially controversial process that characterizes public choice more generally. In Step 2, Stich discusses the characterization of hazards, which he calls “partial outcomes”. His examples include a cancer epidemic and spreading resistant bacteria. For Stich, partial outcomes include benefits, which must be combined with risks to achieve a total outcome. Step 3 for Stich is moral valuation, a process that, as applied to risks, makes a more

58

3 Biotechnology, Policy and the Problem of Unintended Consequences …

explicit statement of the hazard, the sense in which the outcome is bad or unwanted. Step 4 is determining probabilities, that is, exposure quantification. In Step 5, Stich discusses how this yields an estimate of the expected utility of the option (a procedure deferred in this book until Chap. 6). Stich’s final step 6 is selecting a policy, a phase that he cautions readers to regard as more complex than might appear at first, (Stich 1982). Step 6 corresponds to what I have called risk management. Writing more narrowly on genetic engineering of animals, Bernard Rollin echoed Stich’s message a few years later. Rollin argued that the Frankenstein metaphors applied to genetically engineered animals do not prove that it should never be done. The lesson for scientists is to avoid the fictional Dr. Frankenstein’s, “failure to foresee the dangerous consequences of his actions or even to consider the possibility of such consequences and take steps and precautions to limit them” (Rollin 1985, 1986 p. 290). Rollin’s 1995 book devoted attention to the questions that must be addressed in attempting to foresee the dangerous consequences of transforming vertebrates, with particular attention to the risks that animal biotechnology poses both to the human users of a genetically engineered animal, and to the animals themselves. Rollin is less explicit in laying out a framework that corresponds to the hazard identification, exposure quantification and risk management steps in the risk based approach, but his overall treatment of genetically transformed animals exemplifies it. Stich and Rollin devote considerable attention to the argument establishing scientists’ responsibility to consider risks or unwanted outcomes very seriously before pursuing their research. Both classify these risks and unwanted outcomes into categories that reflect a demarcation first between fact and value, and then amongst different kinds of value. Their approach organizes a vague and contentious thicket of issues by analyzing how distinct burdens of proof might be applied to different components of the controversy. Both Stich and Rollin dismiss the possibility that genetic engineering could be intrinsically wrong. Movement of genetic materials is permissible, subject to consideration of the consequences. Moreover, the types of consequence that count are familiar: human health, animal welfare, environmental quality, and distributive justice. While genetic engineering allows humanity to do things that have never been done before, Stich and Rollin define the ethical issues raised by molecular biology as familiar problems of technological risk. Significantly, however, Stich and Rollin do not fully agree on how to address the problem of unwanted consequences from an ethical perspective. In these early articles, Stich seems far more comfortable with consequentialist or optimizing solutions to the problem of technological risk. The classical characterization of the consequentialist approach follows a rubric typical for decision analysis as practiced in the twentieth century. This rubric starts by specifying a partition, or list of all possible options available to a decision maker (e.g. Stich’s Step 1, enumerating alternative policies). Decision makers base their choice on a comparison of the consequences that would ensue from each option in the partition. For utilitarians (the most familiar version of consequentialism), the comparison (the culmination of Stich’s Step 5, calculating expected utilities) should include all impacts on everyone affected. Strict utilitarians favor the option that has the optimal or “best” consequences, commonly expressed in terms of maximizing benefits: Choose the option that produces the greatest good for

3.2 Biotechnology Policy and Philosophy

59

the greatest number of affected parties. This is the phase (Step 6) that Stich describes as involving underappreciated complexity. Although consequentialism does not necessarily require a commitment to utilitarian maximizing, the approach dictates that potential outcomes from each option be predicted and then assessed or subjected to a process of valuation. This value, whether positive or negative, is then discounted by the probability that the outcome will actually occur. In combining the value of outcomes with the probability that the outcomes will occur, the consequentialist approach treats ethical decision making as an exercise of weighing the expected value of outcomes. That is, if one hopes for a benefit of 100 pleasure units, but there is only a five percent chance that the benefit will actually be realized, it is important to realize that 100 pleasure benefits dramatically overstates the value that one can expect to reap from choosing a given option. A more reasonable approach is to discount the 100 pleasure units by the 5% probability, yielding an expected value of just 5 pleasure units. When all costs and benefits of an activity are assessed in terms of expected value, these values may be summed, producing a net value. Each option in the partition can be analyzed similarly, which allows an analyst to rate options according to a sum ranking of their expected value. As noted, a utilitarian decision maker must choose the option that is expected to produce the optimal ratio of benefit to cost, or the greatest net expected value. Stich is at most committed to the spirit rather than the letter of this approach. He does not discuss how it would be implemented in a detailed fashion, but he nonetheless seems comfortable with an assessment of biotechnology that compares its costs and benefits. Rollin agrees that the ethics of animal biotechnology demand a prediction of its likely consequences. He differs from Stich in that for him there are at least some potential consequences that should not be subjected to an evaluation of cost and benefit trade-offs. Rollin notes a class of possible outcomes whose ethical significance is sufficient to determine the correct course of action irrespective other costs or benefits. His 1986 article is primarily concerned with impacts on animals. He describes the potential for creating dysfunctional animals, condemned to lives of physical pain or cognitive suffering. Rollin states that when such animals are inadvertently produced, scientists have an obligation to terminate the experiment, ending the animal’s suffering. When there is knowledge that dysfunctional animals are likely to be produced, the experiment should not be done. These are firm rules for action; scientists do not conduct a cost–benefit analysis in order to decide what to do (Rollin 1986). However, Rollin goes on to describe the potential for using genetic engineering to change an animal’s nature, so that, for example, pain does not occur, or more radically, the animal remains permanently unconscious. Such modifications are not only permissible in Rollin’s view, but also might be obligatory for scientists who wish to develop transgenic models for certain types of disease. In none of these discussions does Rollin endorse a cost–benefit type of accounting or a calculation of offsetting costs or benefits that could override the judgment that these singular outcomes determine whether or under what conditions an experiment ought to proceed. Although biotechnology raises ethical issues in virtue of its unwanted consequences, Rollin

60

3 Biotechnology, Policy and the Problem of Unintended Consequences …

shows that it is possible to bring non-consequentialist patterns of ethical reasoning to bear on the problem of unwanted consequences. The framework put forward by Stich and Rollin follows the conceptual model of risk analysis outlined in Chap. 2. The four categories of possible outcomes functions as a prima facie hazard identification. Further specification would characterize the specific hazards in each category in more detail. A qualitative description of how these hazards can materialize serves as the first step in an exposure assessment, though scientific risk quantification would go considerably further in assigning probabilities to each unwanted outcome. Finally, the difference between Stich and Rollin reflects two different approaches to risk management. Stich seems to be advocating some form of cost–benefit optimization, while Rollin is arguing for a lowest feasible risk standard. Neither Stich nor Rollin take up the problems associated with risk communication, though Rollin’s booklength study, The Frankenstein Syndrome, does discuss ways in which the public’s attitudes to genetic engineering should be taken into account at the time of decision making, Rollin (1995). The international public controversy over the approval and adoption of rBST, the hormone that increases productivity of dairy cows, provides a model for analyzing ethical issues related to other forms of agrifood biotechnology. With but few additions, the hazards mentioned in the Stich/Rollin theory of scientists’ ethical responsibility are well represented. The next section shows how a Stich/Rollin assessment might be applied to rBST. The following sections in the chapter elaborate each of the four areas of unwanted consequence. Two claims are argued in the balance of the chapter. First, although the philosophical dimensions of unwanted consequences are likely to remain controversial, the existence of a reasonably well functioning political forum for human health, animal welfare and environmental impact constitutes a political solution to these problems. Second, the lack of a political forum for debating social consequences is a serious political deficiency in biotechnology policy.

3.3 rBGH: Assessing Unwanted Consequences The social history of rBGH in the United States deserves a more extended treatment than is warranted in the present context, and readers wishing to follow it more closely should consult (Guehlstorf’s 2004) book, as well as the treatment in Shurman and Munro (2010). Readers should note that additional elements of the rBST case are discussed in other chapters. Here it must suffice to say that a complex network of interested parties opposed the technology. Most objections to rBGH that were publicly articulated fall into the categories of animal welfare, food safety and social consequences. Negative environmental impact was not a driving factor among activist objections to rBGH, but some commentators expressed worries. Thus all four types of hazard noted in the Stich/Rollin model are represented. In the rBST case, food safety was deeply contested. Concerns about the integrity of the food industry, the regulatory process, and agricultural research organizations were expressed as uncertainties about the safety of rBST milk, but scientists and regulators were adamant

3.3 rBGH: Assessing Unwanted Consequences

61

about excluding these concerns from risk assessment. This difference of opinion is crucial to the ethical analysis of agrifood biotechnology in general. Although safety became the eventual focal point, social consequences associated with restructuring in the dairy industry actually started the entire debate. In the end, social impacts were the most politically contentious elements of the rBST debate within the United States. Shurman and Munro suggest that the industry’s unwillingness to take social risks seriously was a crucial factor in uniting activists into a social movement to oppose GMOs. In this respect, controversy over rBST is a particularly apt model for considering agrifood biotechnology. Each type of hazard can be reviewed briefly.

3.3.1 Food Safety Food safety is perhaps the most obvious area of potential impact from genetic engineering in the agrifood sector. For purposes of this chapter, food borne risk is the probability that consumption of a food will produce injury or debilitating disease, or that substitution of a food for reasonable alternative foods will adversely affect a person’s health through nutritional deficiencies. Food safety policy is a paradigm example of risk governance, and it reflects ethical schisms mentioned above. Consequentialists treat risk in the manner described by Stich: measure probabilities and compute expected values, then choose the course of action that optimizes the production of good over bad. An alternative approach to risk management stresses rights. When dealing with risks to human beings, the rights approach emphasizes informed consent on the part of the person exposed to any risk, no matter how small or uncertain. The consequentialist position translates into public policy as a judgment that key decisions should be made by experts who can assemble and interpret information on risk. Informed consent requires mechanisms where individuals are exposed to food borne risk only under circumstances of their own choice.1 At present, the consensus standard is that foods produced using biotechnology must be at least as safe as conventional foods. In fact, the procedures for assessment of food products from biotechnology virtually assures that far more will be known about the risks of genetically engineered foods than from foods of more conventional origin. There is thus the logical possibility that ethics might weigh in on the side of less attention to food safety in virtue of disproportionate expenditure of resources on the assessment and mitigation of minimal risks (Johnson and Thompson 1991). Of all potential impacts from rBGH, food safety has received the greatest technical specification. It is also the one on which there is the consensus has been most firm: rBGH does not pose a measurable probability of harm to human beings who consume milk from rBGH treated cows (see Munro and Hall 1991 for a discussion). Assuming 1 This

characterization of food safety oversimplifies the contrast between a consequentialist and rights based approach. It should be regarded merely as a preliminary sketch, intended to orient the reader to issues that arise in reflecting the entire range of hazards in risk-based decision-making. More detail follows in Chap. 4.

62

3 Biotechnology, Policy and the Problem of Unintended Consequences …

that one accepts this scientific consensus, anyone inclined to take a consequentialist position on food safety risk would deem rBGH a non-issue. Despite this, food safety emerged as one of the most prominent public points of controversy in the rBST case. Samuel Epstein, a biomedical researcher at the University of Illinois expressed early concerns about potential health impacts (Epstein 1990), and a group of U.K. researchers documented increased incidence of insulinlike growth factor in the milk and mammary tissue of goats treated with growth hormone (Prosser and coauthors 1991). These concerns were rebutted in the scientific literature, and I will not take readers through the details of these debates here. A second 1990 article in Science gave special consideration to potential impact on children and reported no human health consequences associated with consumption of rBST. The article reports that rBST is biologically indistinguishable from BST that occurs naturally in cow milk (Juskevich and Guyer 1990). Manfred Kroger reiterated these findings in a review of literature on human food safety, (Kroger 1992). Even activist critics of rBGH found little technical basis for complaint with respect to consuming the product itself. In 1991, Michael Hansen of Consumers Union produced an essay on consumer concerns with rBGH that was clearly hostile to the product, yet Hansen cites only public opinion research documenting non-scientists’ concern about safety. He questioned whether rBGH is healthy for the dairy cows on which it is used, but raised no human health concerns associated with human ingestion of rBGH, Hansen (1991). A book by Sheldon Krimsky and Roger Wrubel that raised significant concerns about the environmental impact of GMOs devotes an eight-page section of their book to food safety impacts of rBGH. Their conclusion: “Unusually strong, although not universal, consensus among diverse members of the medical, veterinary, and nutritional community indicates that rBST use on cows does not pose a health risk to humans.” (Krimsky and Wrubel 1996, p. 173). Critics do raise food safety questions about rBST by linking it with collateral production practices that may indeed pose risks, a theme discussed at more length in Chap. 4. Hansen, for example, emphasized the possibility that mastitis associated with elevated levels of milk production might create human health hazards (Hansen 1991). These links were sufficient to spark a significant amount of public resistance to rBST in the United States. The Pure Food Campaign under the leadership of Jeremy Rifkin organized chefs on both coasts to protest what they termed adulteration of milk by addition of rBGH. rBGH has been repeatedly questioned with respect to its purity. However, with the exception of the few sources cited here, the vast majority of food purity concerns address factors that do not bear in any direct way on the probability of injury or other deleterious human health impacts. Purity comes to be understood as an aesthetic attribute of importance to some consumers. Chapter 4 undertakes a more detailed discussion of the ethical relationship between purity and safety. Even in the 1990s, anecdotal evidence suggested that a significant number of people do not want milk from cows treated with rBGH, and recent studies support that judgment, (Wolf et al. 2011). There are several ethical reasons why this might be the case. First, reasonable people may wish to dissociate themselves from foods produced using recombinant DNA technology on religious or aesthetic grounds.

3.3 rBGH: Assessing Unwanted Consequences

63

Nothing is more human than to adopt beliefs about the purity and authenticity of foods that would be difficult or impossible to support on scientific grounds. Is New York State Champagnean oxymoron? The French certainly think so. Avoiding impure or inauthentic foods may not be a safety issue in the narrow sense, but it can be extremely important to those who hold the relevant beliefs. Second, people routinely make consumer choices to express solidarity with other groups or political causes. This type of consideration overlaps with aesthetics to some extent, as the injunction to “Buy American” echoes the French desire for authentic champagne. In the rBGH case, however, solidarity may have more to with loyalty to small dairy producers or animal welfare concerns. In either case, it may be important for some consumers to choose so-called “non-BST” milk.2 Neither of these concerns relate to the probability of disease or injury that associated with drinking rBST milk. They function as a form of food anxiety, rather than food borne risk in the sense described above. Ironically, controversy itself creates anxiety. As questions are raised about the technology, people naturally wonder whom to believe. They may ultimately resolve this question by considering the costs of being fooled. If the critics of rBST are wrong, a consumer is losing several cents per gallon of milk purchased. Although this may add up to significant social costs, even a family purchasing a hundred or more gallons of milk every year may find the three or four dollars a year cost a reasonable price to pay for avoiding the anxiety of a new and unfamiliar form of milk. If the scientists were wrong, after all, the cost would be measured in ill health, especially to children who drink more milk than adults do. Even if one thinks it far more likely that the scientists are right, it may be rational to forego the marginal consumer price benefit in exchange for the familiarity of ordinary milk. In sum, the burdens of proof outlined in the Stich/Rollin framework imply that critics of rBGH never produced reasons to ban the product on food safety grounds. What they produced were reasons why individual milk consumers might want an alternative. This leads to a subtle but important logical point. From any individual person’s viewpoint, the driving concern may be experienced as one of food safety, but as applied to policy, this concern is not supported by existing risk assessments. You or I may observe claims about purity, or react to the fact that there is disagreement among scientific groups and regulatory bodies. Given the average person’s experience base, these observations create anxiety about this product: What should I think? Who knows? This anxiety is embedded in our own individual estimate of risk, and we are less confident in the safety of the product, as a result. This is an entirely respectable application of risk-based reasoning, but it is doubtful that we would want decision-makers charged with protecting the public to apply such an idiosyncratic approach. Public choice requires decision makers to consider the interests of everyone, including people who don’t give a whit about purity, and people who hope to benefit from use of the new technology. This, in turn, requires regulators to determine whether biologically credible mechanisms for posing harm exist 2 This

label appeared in U.S. grocery stores between 2007 and 2017. As noted, all milk contains BST so it is literally inaccurate.

64

3 Biotechnology, Policy and the Problem of Unintended Consequences …

(e.g. hazard identification) and then use data and modeling to estimate how likely these harms actually are (e.g. exposure quantification). This is a considerably more onerous decision process than most of us would (or should) be expected to follow, and it requires the decision maker to discount the relevance of putative harms that are not supported by science and data. This does not imply that food anxiety has no place in risk-based thinking. The riskbased approach does support challenges to consumer autonomy and psychological well-being, but these impacts would fall under socioeconomic consequences, rather than food safety. The policy issue, in short, is one of consent. Those who want the price savings, or who are confident in the product’s safety should have access to the product. Those who do not want it, for any reason, should have some mechanism for avoiding it. The mechanism is almost certainly a label that would allow those who want “ordinary,” milk to get it, though the actual social history of attempts to label rBGH played out in complex and unexpected ways (see Buttel 2000). A more detailed analysis of the ethics of labeling follows in Chap. 4. Here it must suffice to say that the most philosophically promising mechanism for milk and for other foods using biotechnology is a negative label, one that certifies the absence of any use of recombinant DNA technology in producing the food. Although negative labels are far from perfect in assuring consent, they represent a reasonable compromise between enabling consent for those who care about biotechnology in their food, and not stigmatizing a safe, beneficial technology for those who do not (see Thompson 2002).

3.3.2 Animal Welfare Although impact on animals may be a marginal category in some areas of science politics, it has always been prominent in discussions of animal biotechnology, and for obvious reasons. Rollin’s (1986) paper on animal biotechnology stresses the possibility that genetic engineering may produce situations that contribute to animal suffering. Certainly, this potential has been one of the most controversial topics with respect to rBST. Gary Comstock raised the issue of animal welfare impacts associated with rBST in a 1988 paper, noting stress associated with the administration and with the pharmacological effects of rBST, (Comstock 1988). Concern over the linkage of rBST to enhanced milk production (and in turn to increased incidence of mastitis) has been the subject of considerable review and worry ever since. A more detailed discussion of these indirect impacts of rBST appears in Chap. 5. Rollin introduces the concept of telos to describe the genetically encoded set of physical and psychological needs that determine “the fundamental interests central to [animals’] existences, whose thwarting or infringement matters to them” (Rollin 1995, p. 305). He suggests that any experimental or production practice which compromises an animals’ telos is morally wrong, and specifically notes that a farmer’s

3.3 rBGH: Assessing Unwanted Consequences

65

profitability (or a consumer’s price reduction) does not provide a sufficient justification for practices that violate the package of rights an animal must be accorded in virtue of its telos. For both Rollin and Comstock, these rights cash out in terms of practices that produce pain or suffering to individual animals, or that frustrate animals’ ability to behave according to their genetic endowment or “nature”. Current regulatory approaches to animal welfare vary dramatically around the globe. In the United States, ex ante assessment of animal drugs (such as rBGH) focus on efficacy (does the drug do what it claims), food safety and animal health. Impact on an animal’s nature does not figure in the risk assessment. The United States Food and Drug Administration (FDA) evaluates any change or modification in an animal’s metabolism similarly, irrespective of whether the change is induced through administration of a pharmaceutical product or a change in the animal’s genetic makeup. Unlike the procedure used in evaluating crops, regulatory review is mandatory and the agency certifies both the efficacy of the treatment and its safety for both animals and humans who consume animal products. Anti-cruelty statutes provide an opportunity for animal advocates to bring charges on behalf of abused animals, and provide a theoretical basis for ex post regulation. In practice, however, anti-cruelty statutes are rarely successful in overturning an agricultural production practice, though they have led to reforms in transport of animals. It is nevertheless easy to see how the anticipation of impacts described by Rollin and Comstock fits under the general heading of responsibilities noted by Stich, once the hazard of pain and suffering in non-human animals is recognized. Extensive physiological, behavioral and cognitive approaches to the assessment of impact on animals have progressed dramatically since the evaluation of rBGH. There can now be no ethical excuse for failing to apply them in the study of animal welfare. A full moral assessment of transgenic animals—animals whose genomes have been altered through manipulation of recombinant DNA—will be more difficult, as discussed in Chap. 5. I argue that regulatory authorities choosing to disapprove rBST differed either on their judgment of the drug’s effect on animal welfare, or have applied the kind of socioeconomic criterion specifically disallowed in the United States, (as discussed below). Studies on elevated levels of BST (it matters not whether it is the recombinant form) do indeed correlate with higher levels of mastitis. My analysis is that the FDA chose to interpret the cause of higher than normal rates of mastitis as the result of greater milk production. Dairymen have several means to increasing milk production, and many of them elevate mastitis. rBGH also increases per cow milk production, and correlatively elevates mastitis. In contrast, other regulatory agencies have concluded that that administration of rBGH is the relevant causal agent, and have concluded that the increase in disease is a sufficient reason for disapproval (see also Rollin 2011). Thus, unlike food safety, regulators do have a scientifically credible basis for rejecting rBST based on risks to animal health. It should be stressed that the different regulatory approaches turn upon differences in risk management, rather than disagreements about the likelihood that harms to animal health will follow from the use of recombinant growth hormone. It is tempting to conclude that European

66

3 Biotechnology, Policy and the Problem of Unintended Consequences …

regulators simply gave more weight to the interests of animals. Alternatively, the principle driving the U.S. decision may have been a somewhat strained interpretation of fairness to regulated parties. rBGH is a drug, and as such, FDA had clear authority to regulate. Other means of increasing milk production involve unregulated activities, such as the animal’s feed regimen. Regulators may have thought it unfair to use an outcome common in dairy practice (e.g. mastitis associated with high producing cows) to disallow a pharmaceutical agent that produces the same outcome. Most relevant for the moment, however, is the fact that the risk-based approach not only reveals this ethical issue, but that the stages of hazard, exposure and management help to identify points on which ethical issues were contested.

3.3.3 Environmental Impact Agricultural technologies are routinely assessed with respect to environmental impact, though requirements to assess such impacts have arguably been less stringently applied to agriculture than to manufacturing and energy sectors of the economy. While the technical requirements of environmental assessment are relatively well defined, the ethical significance of environmental assessment is extremely complex. There are, for example, environmental impacts that impinge on human health, but assessments also model technology’s impact on broader ecosystem processes. Impacts on these processes may be considered adverse only when they affect human life, but they may also be considered significant simply because they challenge the stability or equilibrium of an ecological zone. A growing literature in environmental ethics in agriculture provides the basis for minimizing any challenge to ecological integrity, (see Aiken 1984; Norton 1991). The rBST case is a relatively poor model for illustrating the full range of ethical issues associated with environmental impacts of biotechnology. The consensus on rBST was to regard environmental impact as one of the least serious of consequences of potential impacts associated with the technology. This was based on the assumption that rBST would reduce the number of dairy cows, and since fecal wastes were then regarded as the most serious environmental contaminant associated with dairying, the reduced number of cows was projected to produce a corresponding reduction in the total volume of waste. As such, the environmental impact of rBST was judged positive (United States Executive Branch 1994). Yet there were critics who opposed rBGH on grounds of ecological sustainability. In particular, the social consequences of restructuring the dairy industry (discussed below) were projected to have secondary environmental impact. Nutrients in animal manure cycle differently on traditional pasture-based dairies than in the intensive, concentrated dairies thought at the time to be most likely users of rBGH (Lanyon and Beegle 1989; Geisler and Lyson 1991). Nutrient cycling has subsequently become an important element in evaluating the environmental and social performance of dairies. Authors emphasizing pasture-based dairying do not typically discuss rBGH (see Dolman and coauthors 2014). Other analysts evaluate sustainability in terms

3.3 rBGH: Assessing Unwanted Consequences

67

of environmental impact measured on the per liter environmental impact of milk production, (see Von Keyserlingk and coauthors 2013). This approach tends to favor dairy systems with high producing cows because each cow consumes a base level of feed and water for maintenance. Low producing cows have disproportionately greater impact on the environment per unit of milk they produce, and authors who take this approach report favorable environmental impact from use of rBGH, (Capper and coauthors 2009). Perhaps the key document in the early debate was a collection of papers published under the title The Dairy Debate in 1993. Articles on possible health issues and consumer concerns that might arise in connection with using milk produced using rBGH were included alongside studies demonstrating that an aggressive program of rotational grazing could provide a meaningful alternative for dairymen (Feenstra 1993). The line of argument put forward in that volume held that risks to the environment (understood in terms of unwanted environmental consequences) give an insufficiently developed picture of sustainability. If rBGH contributes to intensification in the dairy industry (a socioeconomic impact) this might in turn create environmental risks. Only when a technology such as rBGH is considered in comparison with alternative approaches does a clear picture of the environmental dimension emerge, (Buttel 2000).

3.3.4 Social Consequences Social consequences are associated with all agricultural technologies. Some consequences, such as the elimination of hand labor jobs, may be intentional. Yet the most hotly debated social impacts revolve around the way that technical changes effect the economics and balance of power among farm producers. Some technologies are too costly for poor producers, but can give large or wealthy farmers significant advantages over the poor. The economic structure of agriculture in both developed and developing countries means that aggressive early adopting farmers derive short-term benefits from production enhancing technology, but that the ultimate beneficiaries are food consumers. Although animal biotechnologies may be less susceptible to a farm size bias than are mechanical and chemical technologies, it is reasonable to think that many poor producers will be unable to compete with richer competitors as a direct result of biotechnology. Economist Robert Kalter’s study of economic impacts from rBST predicted that relatively small-scale dairy producers might be disadvantaged when rBST became available. The basic idea is that the “size-distribution” of farms, (that is, the proportion small and large farms,) skews to fewer and larger farms when technologies (such as rBST) increase the productive efficiency of farming, (Kalter 1984; Kalter and coauthors 1985). This prediction is itself complex, and a substantial literature on it is summarized by Loren Tauer (1992), who followed up with empirical analyses of the social consequences that rBST actually had (Tauer and Knoblauch 1996). For the purposes of this discussion, the economic issues that arise in predicting or

68

3 Biotechnology, Policy and the Problem of Unintended Consequences …

measuring a technology’s effect on the size-distribution of farms and the make-up of rural communities are less relevant than the general question of why alleged impacts on small versus large farms might be thought ethically significant. There are at least two strategies for approaching this issue. One begins with the assumption that those adversely affected by new technology are harmed in some way analogous to impacts described above. They may be deprived of income they would have received without the technology, and may be harmed in more subtle psychological and social ways. These impacts must be weighed against benefits not only to other producers, but also to food consumers (Thompson et al. 1994, pp. 242– 245). A second strategy begins with the observation that those who make decisions about whether to develop and market a technology occupy a position of power over the small farmers affected by socio-economic change. On this more populist view, what is ethically significant is the distribution of power, not the distribution of risks and benefits. The remedies enhance affected parties’ ability to influence decisions that will have dramatic effect on their future livelihood and way of life. In this respect, it is crucial to note that in the United States no administrative agency has the authority to monitor or regulate technology based upon social consequences. Lacking an outlet for their frustrations, groups seeking rectification of income and power inequalities will politicize the regulatory process for environmental, animal welfare and human health consequences (Thompson 1992). Shurman and Munro’s study of anti-GMO activists suggests that they were primarily influenced by the second type of concern. The rBGH debate was a significant factor in forming the anti-GMO social movement because it demonstrated the longer-term potential for commodification of life, (Shurman and Munro 2010). The term commodification did not appear in common usage until the 1950s, but the idea derives from nineteenth century critiques of capitalism. Narrowly, commodification occurs when goods or services not previously obtainable for money come to be bought and sold. Broadly, commodification is a change in the worldview or mindset of a community in which the price and use value of objects obscures the way in which these values depend on, emerge out of, interpersonal social relationships. The commodification of human labor through the establishment of competitive wage rates is the paradigm case. A worldview that comes to see wage employment as a function of labor markets erases nuanced respects in which cooperative endeavor, interdependencies and the sense in which less fortunate individuals are obliged to serve elites. As the human dimension is covered over by the commodity form, elites gain in virtue of fact that oppressed members of the community have lost the cognitive resources to articulate, and in some cases even perceive, the injustice in their situation. Commodification is thus a social outcome with a subtle mechanism and potentially far reaching significance. In the case of gene technology, commodification is a further erosion of a century’s long process, in which the form of commodity exchange continues to expand into more and more dimensions of social relations and cultural forms. The predictions of economic disruption in the dairy industry provided a reason for rural activists to align with other social advocates who may have been concerned about issues such as genetic privacy or commercialization of the human genome. For them. recombinant

3.3 rBGH: Assessing Unwanted Consequences

69

bovine growth hormone is significant less for its impact on the size distribution of dairies than as the proverbial camel’s nose under the tent. It is a threat to future erosion of humane social relations that, from the activist point of view, must be nipped in the bud, (see Schurman and Munro 2010, 58–81).

3.4 Ethical Disputes, Governance and Consensus Politics This summary of the debate over rBST shows that food biotechnology was controversial in the United States long before it became a topic for discussion in Europe. Analysts who claim that gene technologies were never resisted in the U.S. are simply mistaken. Indeed, the analyses of Stich, published in 1978 and 1982, and Rollin, published between 1984 and 1995, indicate that American philosophers were thinking along the same lines as Hans Jonas, whose Das prinzip verantwortung was published in 1979 and translated into English in 1984. The categories of risk to human and animal health, to the environment and to social welfare are exemplified in the rBST debate, making it a useful case study for the risk-based approach to technological ethics. What lessons can scholars of the twenty-first century take from it? In the best of circumstances, most of the problems identified by Jonas and described in the Stich/Rollin model are amenable to governance. This is not to say that the philosophical problems are fully resolved politically. Philosophers will continue to debate whether non-human species or ecosystems themselves are morally considerable. Furthermore, I am not suggesting that solution to a political problem results in dissolution of the competing political or moral interests that may have led to a problem in the first place. What is more, different governance regimes can incorporate political and moral values in different ways, leading to conflicts at the regulatory level. The role of values in evaluating rBGH as a drug with implications for animal health are discussed in more detail in Chap. 5. However, the rBGH case is an object lesson for three more general observations. First, the standard regulatory politics of new technology often blunts the rancor of continuing philosophical and political differences even when differences are not finessed altogether. For rBST, the United States regulatory system provided forums for governance of food safety and risks to animal welfare. The animal health component of the rBST debate provides the most cogent demonstration of the way that an ethical controversy was dampened by the politics of the regulatory process. As discussed at more length in Chap. 5, animal rights activists clearly seek radical changes in society’s practice, and these changes might well eliminate many practices in animal agriculture. On this point, animal activists might take a “rights” view that challenges the consequentialist perspective of those who think that a compromise to animal health, welfare or needs fulfillment could be offset by benefits to human beings. Bernard Rollin argues that the rBST debate is evidence for the growing influence of animal rights thinking, (Rollin 2011). In the end, the FDA took a more permissive attitude to mastitis incidence than regulators in other countries. Ethicists

70

3 Biotechnology, Policy and the Problem of Unintended Consequences …

interested in pressing for greater moral consideration of animal interests might have latched on to the rBST case as a vehicle for even larger political ambitions. Yet in the United States, at least, they did not. The case of recombinant rennin (or chymosin), which was roughly contemporaneous with rBGH, provides more support for Rollin’s claim. Rennin is an enzyme essential to cheese making traditionally harvested from the inner lining of the fourth stomach of calves and other young ruminants. The calves are slaughtered to obtain rennin. An engineered bacterium was developed in the 1980s to produce a purified version of the enzyme under industrial conditions, (Emtage et al. 1983). While not strictly an “animal biotechnology,” recombinant chymosin met absolutely no resistance on either food safety or animal welfare grounds. Given its impact on animal welfare, it is easy to see why. It is a gene technology that eliminates the need to use (much less kill) animals. Recombinant rennin affects animal agriculture for the better from an animal welfare or animal rights perspective. To base ethical objections to rBGH strongly on the fact that gene technology is used would have also undercut the approval and adoption of recombinant chymosin. More generally, the existence of an institutionalized governance process shifts the moral debate away from the specific tool or technique. In one sense, the fact that rBST made it through food safety and animal health regulatory procedures provided a resolution to the debate, at least as a matter of law and policy. The effect is that even those who were not satisfied by the decision redirect their attention from the specific case toward the regulatory process itself. The food safety review for rBST illustrates this point. As already discussed, negative labels represent a compromise solution that allows those who wish to avoid rDNA products to do so. When rBST was introduced in the early 1990s, the FDA had adopted a posture of opposing food labels that fail to provide consumers with scientifically supported dietary health information. FDA viewed labels stating that a product is rBST free as making a misleading health claim. This stifled companies that wished to label a product as rBST (or rBGH) free, and precipitated numerous state-level efforts to supplement FDA’s regulatory procedure with product labels that would permit consumers to make choices based on their personal values (including socioeconomic concerns), (Kolodinsky et al. 1998). In other words, the target of debate shifts from rBST itself to the institutions governing products of gene technology. If there is a weakness in Shurman and Munro’s characterization of the rBST debate as the first salvo in the battle between industry and activists, it is that they do not examine the institutional context in government. FDA regulators may have approached the approval process with an eye toward precedent setting that could affect products like recombinant rennin, and consistency with a painstakingly policy history on food labeling that might be construed to imply health benefits or risks. Finally, the rBGH case illustrates the way that weak or non-existent governance institutions can have profound effects that extend far beyond the debate over a single product. Although the environmental risks of rBGH were downplayed in the special study conducted by the George W. Bush administration, lack of clarity about what, precisely, counts as an environmental risk has continued to plague agrifood biotechnology, as the extended discussions in Chaps. 6 and 7 attest. However, it is the

3.4 Ethical Disputes, Governance and Consensus Politics

71

missing forum for debating the socioeconomic impacts of rBGH that has had the most enduring impact on the fate of agricultural gene technologies. Kalter’s work sparked a half decade or more of outrage within the U.S. dairy industry, and a policyoriented summary published in the widely-read journal Issues in Science and Technology brought the potential social impact of massive bankruptcies to the attention of anyone with a passing interest in either agriculture or science policy, (Kalter 1985). His prediction coincided with a movement to preserve smaller, family-run farming operations, which were already experiencing an entirely separate economic crisis, (Barnett 2000). If animal activists had reasons not to single out gene technologies for rancor, small farm activists had even more powerful reasons to stoke anxieties about safety and environmental uncertainties as a way to bring an issue that lacked an effective governance mechanism to widespread public attention.

3.5 Social Consequences Redux Good policies do represent political solutions to philosophical problems when they appeal to the overlapping consensus that exists in most industrialized democracies. In an uncharacteristically pragmatic moment, John Rawls (1921–2002) argued that political philosophy must seek to identify the common principles and policies that would be endorsed even by persons having very different life philosophies, (Rawls 1987) Rawls was hopeful that the main elements of a just society could be identified by emphasizing these areas of consensus, rather than the dissent that often underlies them at the level of fundamental beliefs. Food safety, animal welfare and environmental impact represent three areas of policy where Rawls’ hopes had some chance of being fulfilled. The unwanted, unintended social consequences of animal biotechnology are less amenable to a hopeful solution. Ironically, it is in the area of distributive justice that Rawls’ appeal to the overlapping consensus seems least promising. Social consequences in agriculture must be analyzed at multiple levels (see Berlan 1991). New technologies routinely jeopardize some forms of employment, and ruin businesses that are wedded to obsolete technologies. The first level of analysis draws our attention to the individuals left without jobs and income during such transitions. Social welfare programs in most developed countries moderate the effect of these transitions, offering temporary benefits and retraining to affected parties. The second level of analysis takes up the loss of these jobs to the communities in which the relevant industries are located. Job loss in one sector translates into failed businesses, schools and hospitals across the board. From a comparative politics perspective, public policies for coping with community transition are uneven. The United States arguably does a poorer job of moderating transitions at this level than do most countries. Nevertheless, it is reasonable to claim that some policies and mechanisms exist for coping with the impoverishment and psychological harm that are associated with these aspects of technological change.

72

3 Biotechnology, Policy and the Problem of Unintended Consequences …

Another level of analysis considers the impact of technological transitions on entire regions. With respect to transitions in mining and manufacturing, as well as agriculture, entire regions of the world have been effectively depopulated. When people leave the countryside in this way, individual households and community institutions are lost. In addition, entire ways of life, and networks of kinship and mutual affection are dissolved. There is really no way to compensate the losers for these transitions, for the basic values that define what is a profit and what is a loss have been stripped away from them. To be sure, they will land somewhere else, but mere survival aside, the systems of meaning that determine value will have to be reconstructed entirely. Rawlsian liberalism holds that we should not be partisans when it comes to comparatively evaluating alternative forms of life, but placing a form of life at risk is nonetheless a social evil in the liberal system. The commodification critique goes on to say that a cascade of further consequences ensue when a domain of life is overtaken by commercial exchange relations. A mindset—a social imaginary—becomes commonplace in which the morally relevant properties of that domain are replaced by commodity exchange. Left-leaning critics of capitalism see deep problems with Rawlsian liberalism on this point. Not only does the new commodity form conceal the disproportionate growth in the social power of capitalists, the market form puts more humane ways of living at a disadvantage. Liberal neutrality is not the way to proceed in cases like this. It is thus reasonable to say that when technological transitions have such systematic effects, a loss occurs that cannot be compensated. Whether such losses should be permitted, or whether they should be resisted is a large question. Clearly, the dispute over the social consequences of rBGH centers on just such a question. The dairy farmers who have opposed this technology are acutely aware that a delicate balance of subsidy and productivity keeps them in business. If productivity increases, there will be more milk at a lower price. Small-scale producers cannot recoup losses from a reduced margin of profit by increasing volume. Furthermore, increases in volume will put political pressure on policies that keep prices at current levels—a circumstance especially crucial for dairymen in Europe and Canada. The classic dairy, with 50 to 100 or 200 cows, is at risk.3 What will go with it are the businesses, schools and hospitals of a hundred counties, but what is worse is that a form of life that is thought particularly characteristic of agriculture will disappear from the landscape. It will be replaced by industrial plants servicing 2000, 4000, even 10,000 cows in a single location, trucking the feed in and the milk and manure out (Lanyon 1994). Canada and Europe have been more willing to take social consequences seriously than has the United States, but it is at least arguable that the US exercises hegemonic influence over world policy on the matter of social consequences. Trade agreements and the sheer pressure of international competition make governments reluctant to place their own producers into an economically disadvantageous position relative 3 In writing for the third edition, it is clear that this transition has been completed in the United States,

though very small dairies continue to survive in other industrial nations, including Canada. It is less clear that rBST was the causative agent. Computerization of record keeping and automation of milking and husbandry have allow dairymen to obtain significant economic advantages by keeping large herds.

3.5 Social Consequences Redux

73

to US producers. The lack (or weakness) of regulatory procedures for social consequences represents a form of international assurance problem. If every government would regulate based on social impact, the rules of international competition would be fair. However, as long as one government does not, any government wishing to regulate technology on the basis of social impact risks losing its ability to compete in the relevant sector altogether. When the one actor is as large and dominant as the United States, the absence of political solutions to the social consequence issue is a foregone conclusion. Most significantly, the North American debate over rBST set the stage for the complex debate over the socio-economic consequences of gene technologies. This early controversy influenced later ones by fixing social scientists’ attention on production economics and market forces as the key mechanisms that cause significant social consequences. A more detailed discussion of these causal mechanisms follows in Chap. 8. As the previous analysis has shown, it also established a precedent by limiting application of the risk-based approach to biophysical hazards. In my judgement, the economists who analyzed socioeconomic effects assumed that they were talking about risks. Food and veterinary scientists assumed that the word risk applied only to health and environmental impacts, and curiously, scholars in science studies followed them, rather than the economists. They characterized socioeconomic impacts as ethical, and took this to imply that they were not risk-based and lacked scientific underpinning, (see Levidow and Carr 1997).

3.6 Learning from rBST There were, therefore, four ethical problems associated with rBST, but at best only three institutionalized forums for governance. The significance of this fact is both political and philosophical. Politically, the lack of a policy framework for even raising, much less resolving, problems associated with social consequence introduces a high degree of uncertainty into the politics of food biotechnology. Again, the rBGH case illustrates this point. Why did rBGH become an issue at the Food and Drug Administration (FDA) where questions of human and animal health were to be assessed? Why, especially when there was such unanimity among the science community that food safety risks were minimal, does it continue to raise public concern? Some of the answers have been discussed above, but the political contentiousness of rBST at FDA must have arisen partly because those interested in social consequences had nowhere else to go. The absence of a forum for debate and regulation of social consequences in either administrative or judicial branches of government leaves no alternative but the translation of these issues into trumped up and ultimately false concerns about human and animal health, or environmental consequences. In earlier editions of this book, I ended my discussion of the recombinant bovine growth hormone case with hopeful comments. In essence, I said that if the nonexistent public forum for debate over new technology had been available, much

74

3 Biotechnology, Policy and the Problem of Unintended Consequences …

of the ensuing acrimony and division over GMOs might have had a very different trajectory. A perceptive external reviewer of my revised manuscript took strong issue with these passages, going so far as to say that they made the entire manuscript seem miserably out of date. The “out of date” comment may go a bit too far. Other analysts continue to promote optimistic proposals for addressing political conflicts within a framework that has been broadened to accommodate ethical discussion, (see Hartley and coauthors 2016; Hamburger 2018; Kuhlmann et al. 2019). Yet even when participatory governance procedures have been promising in some ways, they have not succeeded in either introducing a robust forum for social consequences, or prevented unresolved disputes over the social dimension from spilling over into human health, animal welfare and environmental impact, (Brom and coauthors 2015; Bremer and coauthors 2015). At least insofar as they predicted the trajectory of debate over gene technology, the analysts who ruled socioeconomic impact out of the risk assessment court of appeals have proven correct. As such, I am compelled to retract the optimism of earlier treatments and stress more limited conceptual lessons from the rBST debate. First, I insist that the adaptation of Jonas that was undertaken by Stich, Rollin, Burkhardt and myself (even of some did not know they were adapting Jonas) does provide a framework for discussing the ethics of agrifood biotechnology. The riskbased approach suggests that enumerating hazards is an important step in proactive analysis of gene technology. Assigning probabilities to each hazard is philosophically complex in ways that are seldom recognized within the regulatory process, in part because the public choice framework implies a generalized, society-wide point of view that is necessarily insensitive to vulnerabilities and uncertainties peculiar to the standpoint of individuals and subgroups. Later chapters discuss some of these issues in more detail. Nevertheless, the risk-based approach helps not only to make these problems discernable, but also in framing hypotheses for addressing them at cognitive, regulatory and ethical levels. Second, my external reader’s pessimism reminds me that the original version of the text addressed property disputes and metaphysical concerns as addenda to the risk-based coverage of food safety, animal welfare, environmental impact and social consequences. Shurman and Munro’s analysis of how the social movement against GMOs formed suggests that concerns about the commodification of life overlap strongly with topics covered in this addendum. They are, in a sense, social consequence issues, but they fit less cleanly into an interpretation of social consequence that emphasizes economic loss, social displacement or distributive inequality. These issues may have motivated civil society advocates to join forces in opposition to the emerging forms of biotechnology during the 1980s and 1990s, and the bovine growth hormone debate would have been the first policy forum in which that opposition was tested. Nevertheless, critics articulated these concerns as general objections to gene technology, rather than as specific reasons to resist recombinant BST. As such, the case study is not a good model for examining some of the more complex arguments that are taken up in Chaps. 10, 11 and 12 of the revised edition. Finally, despite my claim that socioeconomic outcomes can be framed as hazards, isolating the social concerns—including those relating to commodification—is

3.6 Learning from rBST

75

nonetheless a philosophically effective strategy for understanding the ethics of agricultural and food biotechnology. The reason is that once one gets to quantifying risk, that is, going from a brute hazard to a concept that reflects the likelihood it will occur, causal mechanisms matter. There are ethical issues relating to food safety, animal transformation and environmental quality, and, the mechanisms that expose us to these hazards can be understood, if not always agreed upon, without resorting to sociotechnical forms of causation that are both more complex and more difficult to grasp fully. There are reasons why people with a stake in the outcome of policy debates will resist my attempt to separate risk-based questions into four discreet categories, and some of those reasons are philosophically interesting in themselves. Less interestingly, experienced political actors can exploit fear, reaction and confusion and this circumstance will always limit the philosophy’s effectiveness in a political forum. (We should never forget Socrates’ struggle with the Sophists). Nevertheless, philosophy betrays its essence and history by aligning with any interest willing to promote a political outcome at cost to sincerity and our best guess at the deeper truth. I do not fault advocates who do this—be they from industry or civil society—because that is their job. I am dedicated to different and (perhaps) more difficult path.

References Aiken, W. 1984. Ethical issues in agriculture. In Earthbound: New Introductory Essays in Environmental Ethics, ed. T. Regan, 257–288. New York: Random House. Barnett, B.J. 2000. The US farm financial crisis of the 1980s. Agricultural History 74: 366–380. Berlan, J.-P. 1991. The historical roots of our present agricultural crisis. In Towards a New Political Economy of Agriculture, ed. W.H. Friedland, L. Busch, F.H. Buttel, and A.P. Rudy, 115–136. Boulder, CO: Westview Press. Bremer, S., K. Millar, N. Wright, and M. Kaiser. 2015. Responsible techno-innovation in aquaculture: Employing ethical engagement to explore attitudes to GM salmon in Northern Europe. Aquaculture 437: 370–381. Brinckman, D. 2000. The regulation of rBST: The European case. AgBioForum, 3(2 and 3), 164–172. Accessed 11 Feb 2020 at https://www.agbioforum.org/v3n23/v3n23a15-brinckman.htm Brom, F.W., S. Chaturvedi, M. Ladikas, and W. Zhang. 2015. Institutionalizing ethical debates in science, technology and innovation policy: A comparison of Europe, India and China. In Science and Technology Governance and Ethics, eds. M. Ladikas, S. Chaturvedi, Y. Zhao and D. Stemerding, pp. 9–23. New York: Springer. Burkhardt, J. 1988. Biotechnology, ethics, and the structure of agriculture. Agriculture and Human Values 5: 53–60. Burkhardt, J. 1992. Ethics and technical change: The case of BST. Technology in Society 14: 221–243. Buttel, F. 2000. The recombinant BGH controversy in the United States: Toward a new consumption politics of food? Agriculture and Human Values 17: 5–20. Capper, J.L., R.A. Cady, and D.E. Bauman. 2009. The environmental impact of dairy production: 1944 compared with 2007. Journal of Animal Science 87: 2160–2167. Comstock, G. 1988. The case against BGH. Agriculture and Human Values 5: 36–52. Dolman, M.A., M.P.W. Sonneveld, H. Mollenhorst, and I.J.M. De Boer. 2014. Benchmarking the economic, environmental and societal performance of Dutch dairy farms aiming at internal recycling of nutrients. Journal of Cleaner Production 73: 245–252.

76

3 Biotechnology, Policy and the Problem of Unintended Consequences …

Emtage, J.S., S. Angal, M.T. Doel, T.J. Harris, B. Jenkins, G. Lilley, and P.A. Lowe. 1983. Synthesis of calf prochymosin (prorennin) in Escherichia coli. Proceedings of the National Academy of Sciences of the United States of America 80: 3671–3675. Epstein, S.S. 1990. Potential public health hazards of biosynthetic milk hormones. International Journal of Health Services 20: 73–84. Feenstra, G. 1993. Is BGH sustainable? The consumer perspective. In The Dairy Debate, ed. W.C. Liebhardt, 1–63. Sustainable Agriculture Research and Education Program: University of California, Davis. Geisler, C., and T. Lyson. 1991. The cumulative impact of dairy industry restructuring. BioScience 41: 560–567. Guelhstorf, N.P. 2004. The Political Theories of Risk Analysis. Dordrecht, NL: Springer. Hamburger, D.J.S. 2018. Normative criteria and their inclusion in a regulatory framework for new plant varieties derived from genome editing. Frontiers in Bioengineering and Biotechnology 6: Article 176. https://doi.org/10.3389/fbioe.2018.00176 Hansen, M. 1991. Consumer concerns: Give us all the data! In NABC Report 03: Agricultural Biotechnology at the Crossroads Biological Social & Institutional Concerns,, ed. J.F. Macdonald, 169–176. Ithaca NY: NABC. Accessed 11 Feb 2020 at https://ecommons.cornell.edu/handle/ 1813/49708 Hartley, S., F. Gillund, L. van Hove, and F. Wickson. 2016. Essential features of responsible governance of agricultural biotechnology. PLoS Biology 14: e1002453. Johnson, G.L., and P.B. Thompson. 1991. Ethics and values associated with agricultural biotechnology. In Agricultural Biotechnology: Issues and Choices, ed. B. Baumgardt and M. Martin, 121–137. West Lafayette, IN: Purdue Research Foundation. Jonas, H. 1984. The Imperative of Responsibility: The Search for Ethics in a Technological Age. Chicago: U Chicago Press. Juskevich, J.C., and C.G. Guyer. 1990. Bovine growth hormone: Human food safety evaluation. Science 264: 875–884. Kalter, R.J. 1984. Production cost: Commercial potential and the economic implications of administering bovine growth hormone. In Proceedings of the Cornell Nutrition Conference for Feed Manufacturers. Ithaca, NY: Cornell University. Kalter, R.J. 1985. The new biotech agriculture: Unforeseen economic consequences. Issues in Science and Technology 13: 125–133. Kalter, R.J., R. Milligan, W. Lesser, W. Magrath, L. Tauer and D. Bauman. 1985. Biotechnology and the dairy industry: production costs, commercial potential and the economic impact of the bovine growth hormone. In Agricultural Economics Research Bulletin, 85–20. Ithaca, NY: Department of Agricultural Economics, Cornell University. Kolodinsky, J.Q. Wang., and D. Conner. 1998. rBST labeling and notification: Lessons from Vermont. Choices 13: 38–40. Krimsky, S., and R. Wrubel. 1996. Agricultural Biotechnology and the Environment: Science. Policy and Social Issues: University of Illinois Press, Urbana, IL. Kroger, M. 1992. Food safety and product quality. In Bovine Somatotropin and Emerging Issues, ed. M. Hallberg, 265–270. Boulder: Westview Press. Kuhlmann, S., P. Stegmaier, and K. Konrad. 2019. The tentative governance of emerging science and technology—a conceptual introduction. Research Policy 48: 1091–1097. Lanyon, L.E. 1994. Dairy manure and plant nutrient management issues affecting water quality and the dairy industry. Journal of Dairy Science 77: 1999–2007. Lanyon, L.E., and D.B. Beegle. 1989. The role of on-farm nutrient balance assessments in an integrated approach to nutrient management. Journal of Soil and Water Conservation 44: 164– 168. Levidow, L., and S. Carr. 1997. How biotechnology regulation sets a risk/ethics boundary. Agriculture and Human Values 14: 29–43.

References

77

Mayne, N.G., H.M. Hsiung, J.D. Baxter, and R.M. BelagajeIn. 1984. Direct expression of human growth in escherichia coli with the lipoprotein promoter. In Recombinant DNA products: Insulin, Interferon, and Growth Hormone, ed. A.P. Bollon. Boca Raton, FL: CRC Press. Miller, W.L., J.A. Martial, and J.D. Baxter. 1980. Molecular cloning of DNA complementary to bovine growth hormone mRNA. Journal of Biological Chemistry 255: 7521–7524. Munro, I.C. and R.L. Hall. 1991. Food safety and quality: Assessing the impact on biotechnology, In Agricultural Biotechnology, Food Safety, and Nutritional Quality for the Consumer., ed. J.F. MacDonald, 64–73. Ithaca, NY: National Agricultural Biotechnology Council. Accessed 11 Feb 2020 at https://ecommons.cornell.edu/handle/1813/49681 Norton, B. 1991. Toward Unity among Environmentalists. New York: Oxford University Press. Prosser, C.G., C. Royle, I.R. Fleet, and T.B. Mepham. 1991. The galactopoietic effect of bovine growth hormone in goats is associated with increased concentrations of insulin-like growth factorI in milk and mammary tissue. Journal of Endocrinology 128: 457–463. Rawls, J. 1987. The idea of an overlapping consensus. Oxford Journal of Legal Studies 7: 3–24. Rollin, B.E. 1986. The Frankenstein thing. In Genetic Engineering of Animals: An Agricultural Perspective, eds. J.W. Evans and A. Hollaender, 285–298. New York: Plenum Press. Rollin, B.E. 1995. The Frankenstein Syndrome: Ethical and Social Issues in the Genetic Engineering of Animals. New York: Cambridge University Press. Rollin, B.E. 2011. Animal rights as a mainstream phenomenon. Animals 1(1): 102–115. Rollin, B.E. 1985. [republished 1990] The Frankenstein thing. In Agricultural bioethics: Implications of agricultural biotechnology., eds. S. M. Gendel A.D. Kline, D. M. Warren and F. Yates, 292–380, Ames, IA: Lowa State University Press. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists and Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press. 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. San Francisco: Conari Press. Stich, S. 1978. The recombinant DNA debate. Philosophy and Public Affairs 7: 187–205. Stich, S. 1982. Genetic engineering: How should science be controlled? In And Justice for All: New Introductory Essays in Ethics and Policy., eds. T. Regan and D. VanDeVeer, 82–119. Totowa, NJ: Rowman and Littlefield. Tauer, L.W. 1992. Impact of BST on small versus large dairy farms. In Bovine Somatotropin and Emerging Issues: An Assessment, ed. M. Hallberg, 207–217. Boulder, CO: Westview Press. Tauer, L.W., and W.A. Knoblauch. 1996. The empirical impact of bovine somatotropin on New York dairy farms. Journal of Dairy Science 80: 1092–1097. Thompson, P.B. 1992. Ethical issues and BST. In Bovine Somatotropin and Emerging Issues: An Assessment, ed. M. Hallberg, 33–50. Boulder, CO: Westview Press. Thompson, P.B. 2002. Why food biotechnology needs an opt out. In Engineering the Farm: Ethical and Social Aspects of Agricultural Biotechnology, ed. Britt Bailey and Marc Lappé, 27–44. Washington, DC: Island Press. Thompson, P.B., R.J. Matthews, and E. vanRavenswaay. 1994. Ethics, Agriculture and Public Policy. New York: Macmillan. Von Keyserlingk, M.A.G., N.P. Martin, E. Kebreab, K.F. Knowlton, R.J. Grant, M. Stephenson, C.J. Sniffen, J.P. Harner III., A.D. Wright, and S.I. Smith. 2013. Invited review: Sustainability of the US dairy industry. Journal of Dairy Science 96(9): 5405–5425. United States Executive Branch. 1994. Use of Bovine Somatotropin BST in the United States: Its Potential Effects, A Study Conducted by the Executive Branch of the Federal Government. U.S. Government Printing Office, Washington, DC, January. Accessed 11 Feb 2020 at https://www.biotech.wisc.edu/docs/default-source/outreach-documents/use-of-bovine-somatr opin.pdf?sfvrsn=999ec28_0 Wolf, C.A., G.T. Tonsor, and N.J. Olynk. 2011. Understanding US consumer demand for milk production attributes. Journal of Agricultural and Resource Economics 36: 326–342.

Chapter 4

Food Safety and the Ethics of Consent

Abstract This chapter addresses a series of philosophical questions that arise in a general consideration of food safety risks, with specific attention to products of gene transfer. The first topic is to demonstrate the sense in which modern technology has converted what were once norms of prudence and self-interest into ethical responsibilities. The next topic is a summary review of the way that food safety experts view food safety risk, followed by a discussion of how this way of thinking is applied to products of gene transfer. From this point, the chapter summarizes a different conceptual framework that shows how the history of food science has created alternative rationalities for thinking about the risks we bear in consuming food. This alternative helps to explain why communication of risks from gene transfer have been so difficult to communicate, and explains why labeling is a component of food safety policy. The chapter concludes with a discussion of how labeling could address some of the ethical tensions created by the tension between expert and lay perspectives on the risks of consuming food. Keyword Risk assessment · Food safety regulation · Mandatory and voluntary labeling · Utilitarian optimization · Rights Foods developed through biotechnology must be safe. No one disputes this, though it is certainly possible to disagree about what safety means and about whether the use of biotechnology introduces unacceptable risk. While in many areas there are extended technical and political disputes about the appropriate level of acceptable risk, the standard in food safety is de minimus, or the lowest feasible level of risk. Food safety is disputed on scientific grounds when the mechanisms of food borne illness are not understood or when there are disagreements about the probability of harm. Recent work on inductive risk stresses the role that values play in framing these disagreements, Douglas (2009). Food safety also raises philosophical issues in the management of risk because society must choose between public policies that minimize the probability of food borne illness and those that protect individual consent.. Even a quarter century since the first writing of this chapter, there is still surprisingly little work on the ethics of food safety. Several authors have examined

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_4

79

80

4 Food Safety and the Ethics of Consent

how merchantability standards might serve as a basis for legal action based on defects in food safety or unexpected nutritional quality (Buzby and Frenzen 1999; Crawford 2002), but philosophers have neglected this topic. The Dutch philosopher Michel Korthals takes up this theme in a series of papers that frame some of the issues discussed in this chapter in terms of the difference between citizen and consumer perspectives, on the one hand, and technical versus common sense attitudes, on the other (Korthals 2001, 2004). Burkhardt (2001) addresses the topic with the larger context of probing the way that the scientific attitude (discussed in Chap. 3) influences the direction of biotechnology. Jensen and Sandøe (2002) have explored attitudes toward food safety within the framework of rational attitudes toward risk. In preparing this revision of the book, I ran a Scholar Google search on the terms ethics and food safety. The vast majority of items returned were published prior to 2007. The search revealed a cluster of articles on fraud, owing especially to the melamine scandal in the People’s Republic of China (see Sharma and Paradakar 2010) or concealing evidence of violations in processing (Roman and Moore 2012). Four important articles provide what I take to be amplification and development of the approach developed in the first and second edition of this book, (see Millstone 2007; Sperling 2010; Mepham 2011; Burkhardt 2012). These articles do not analyze debates over biotechnology, though the debate looms in the background. Ben Mepham’s discussion of food additives discusses consumer sovereignty and animal welfare in addition to food safety. These two categories of hazard are discussed in other chapters of this book. For their part, food scientists have tended to treat food safety as a largely technical issue lacking any particular ethical dimension. This view may have been consistent with a perspective that sees food safety purely as a prudential norm, yet the obvious change in social circumstances noted above has clearly created fiduciary responsibilities, elements of trust and frameworks of ethical responsibility that simply did not exist in the past. However, work on inductive risk implies that even comparatively technical approaches presume value judgments that inform the way that humans have dealt with food-born health risks since before scientific approaches to pathogens and toxins became available. Among food scientists, Ralph Early argues that industry must consider ethics within its decision-making frameworks for food safety, Early (2002), and Donald Thompson, writing with Jonathan Marks, have discussed the institutional structures that govern food safety and their relationship to the food industry (Marks and Thompson 2011). Julian McClements reviews a number of epistemic and social values crucial for ascertaining food safety, (McClements 2019). None of these entries discuss the ethics of food safety for genetically engineered foods. Many biotechnology critics focus on the adequacy of regulation. In the years since the first and second editions of this book appeared, I have produced both humorous (Thompson 2007) and solemn (Thompson 2015, pp. 227–239) analyses of the ethical issues that have occurred in the history of debates over the safety of agrifood biotechnology. This chapter, in contrast, addresses the issues at a conceptual and theoretical level, with very little discussion of episodes in the debate over food safety or regulatory policy issues. My goal is to avoid debates on food law that presume knowledge of a complex regime that stretches from local health inspection to agencies such as the Food and Drug Administration (FDA) and the European

4 Food Safety and the Ethics of Consent

81

Food Safety Authority (EFSA). The food safety regime extends on to regional and international bodies such as the World Trade Organization (WTO) or the Codex Alimentarius, housed jointly at the Food and Agriculture Organization (FAO) and World Health Organization (WHO) of the United Nations. Although I hope to avoid a lengthy discussion of food safety regulation, readers will be aided by a preliminary discussion of the regulatory context. From this point, the chapter moves on to a more general philosophy of food safety. The chapter then closes with a discussion of the ethics of food labels. Although there has been recent change in U.S. labeling policy, I will defend the position that was outlined in the 1997 edition of this book.

4.1 The Ethics and Political Theory of Food Safety Regulation Advocates of biotechnology have argued persistently that the new techniques for modifying and domesticating plants and animals are not fundamentally different from traditional food technologies (see Brill 1985; Miller and Conko 2001; Kleter et al. 2019). This claim about the fundamental similarity of transgenic and traditional technologies has been revisited many times in the past quarter century and will resurface at a number of points in this book. Citizen-scientist activists have argued that there are major gaps in the regulatory systems for food safety, irrespective of whether biotechnology is involved. In some cases, the alleged gaps reflect a lack of clarity about which regulatory office will be responsible for approving a product of biotechnology, or what methods and data they will utilize in making their assessment. The United States announced its Coordinated Framework for regulating products of gene technology in 1986. This was more than a decade before the first genetically engineered crops, but after significant controversies over the authority of National Institutes of Health’s Research Advisory Committee to oversee developments outside the medical realm. Under the Coordinated Framework, U.S. Executive Branch agencies were directed to oversee new products from gene transfer under the existing suite of laws. The decision was motivated by the George H.W. Bush administration’s anti-regulatory stance, as overseen by an ad-hoc Competitiveness Council chaired by Vice President Dan Quayle. Under these laws, food safety oversight fell to FDA, which continues to operate under the Federal Food, Drug and Cosmetic Act of 1938, amended by Congress several times. The law mandates pre-market review for drugs, but foods are not required to undergo mandatory approval prior to market, (Marden 2002). From a practical standpoint, food safety was a concern for processed foods that may include unsafe ingredients or additives, but not for foods themselves. Given the focus on food processing, FDA developed a list of foods and ingredients that are Generally Recocognized as Safe (GRAS), and concentrated on testing the safety of ingredients not included on this list. Although amendments to FDA’s regulatory mandate extended the agency’s authority with respect to additives, this basic structure remains in place despite recent discussion of its inadequacy in an age of gene edited food crops and animal products, (Peck 2017). Later sections in the chapter

82

4 Food Safety and the Ethics of Consent

review the philosophical rationale for this approach, as well as reasons for revising it. Given the fact that foods are exempted from pre-market approval, GMO critics can truthfully state that the technology is unregulated in the United States. At the same time, both FDA and many other statutes provide remedies against companies that sell a product discovered to be unsafe on a post-market basis. As such, companies do have legal and economic incentives that reinforce their moral responsibility to assure the safety of foods. In addition, the Coordinated Framework functions against the background of a U.S. tradition in product liability law that exposes innovators to extreme financial risk, when litigants are successful in convincing a court that they have been injured by their use of a product. Complainants have considerably greater access for torts in the U.S. than in most industrial democracies. In order to protect itself, the biotechnology industry sought some regulatory imprimatur for the safety of GMOs, unavailable under the existing statute. The result was an approach in which regulators review novel foods through a voluntary submission process. The innovating company would notify the agency of their intent to release a product and at the same time, assert its safety. The assertion can be supported with data from toxicological studies on the gene, as well as the protein made by the gene, though proteins found in GRAS foods are also regarded as GRAS. Companies also submit data on changes to the genome and assays that demonstrate the chemical composition of the modified organism. Regulators review data to determine whether the genome is stable and whether the resulting GMO is consistent with the range of reference assays established for unmodified organisms. If satisfied, they issue a letter stating that they see no basis for rejecting the applicant’s claim of safety. To date, no known products on the U.S. market have escaped this type of review process. The Coordinated Framework invests the primary responsibility for regulating environmental impacts at the U.S. Department of Agriculture’s Animal and Plant Health Inspection Service (APHIS), a topic deferred to Chaps. 6 and 7. The FDA’s basic approach to food safety is augmented by two features. First, the U.S. Environmental Protection Agency (EPA) has the authority to regulate pesticides for both human and environmental safety under the Federal Insecticide, Fungicide and Rodenticide Act (FIFRA). Foods engineered to produce the bacillus thuringiensis (bt) toxin were reviewed at FDA to ensure that chemical components of the food (corn, potato, rice) is consistent with other varieties of the same species. The safety of the toxin itself is reviewed at EPA. Second, FDA decided to review genetically modified animals under their authority to supervise drugs intended for veterinary applications. In this instance, the food safety of a genetically modified food animal would still be handled according to the procedure summarized above, but the entire review would be conducted under the auspices of regulators who have the authority to reject an application based upon a products having unacceptable impact on animal health, (see Mandel 2006; Wozniak et al. 2012). In contrast to the United States, Canada, Australia and the European Union have each passed laws that either create new offices under existing ministries or that direct existing offices to undertake regulation of gene technologies applied to agricultural crops and food animals. This approach has several obvious advantages over the U.S.

4.1 The Ethics and Political Theory of Food Safety Regulation

83

system. First, the complexity of the U.S. system makes it extremely difficult to understand. Most Americans—including molecular biologists, agricultural scientists and most policy specialists not focused on biotechnology—would be hard pressed to recount the summary given in the preceding paragraphs. It is, in a word, nontransparent. Second, lodging authority in a single office is at least arguably a step toward closing gaps that emerge as new applications of gene technology are applied. Such gaps are a key basis for recent calls to reform the Coordinated Framework, (Kuzma 2016, Peck 2017). Finally, the rest of the world has appreciated the fact that some form of mandated regulatory review is critical to winning the public’s confidence in the technology. The U.S. approach gives the appearance of being designed to foster industry development at the expense of assuring safe applications of the technology. FDA can reject an application that has been voluntarily submitted, but they cannot require submission and, excepting animals regulated comparably to animal drugs, there is no opportunity for FDA to make a legally meaningful approval of a gene-altered food. Yet it is not clear that the U.S. is out of line with other countries when it comes to actually evaluating food safety. Although EFSA has formally abandoned the view that genes from GRAS foods are also GRAS, food safety assessments at FDA and EFSA remain remarkably similar, (Kuiper and coauthors 2002; Waigmann and coauthors 2012). The larger difference between the United States transcends food safety ethics, where, on the one hand, Europeans have long benefited from mandatory labeling, while, on the other, final approval resides in a process operating through the European Parliament that is even more obtuse and difficult to follow than the Coordinated Framework, (see Morris and Spillane 2010). Charlotta Zetterberg and Karin Edvardsson Björnberg criticize the current European approach as inflexible, arbitrary and at best ambiguously applicable to the new breeding techniques emerging in the era of gene editing. They characterize the contrast with the U.S. approach in terms of the product vs. process distinction discussed later in the chapter, (Zetterberg and Björnberg 2017). There is little doubt that debates over regulatory structure and process will continue. The focus of this chapter, however, is to articulate a basic risk-based philosophy for addressing food safety that is applicable to gene technology. While it is impossible to omit references to the regulatory debate altogether, the balance of the chapter will stress the assessment of food safety risks, and the ethical bases for taking one approach to the management of these risks, rather than another.

4.2 Safety Criteria and Biotechnology Food safety professionals emphasize three types of hazard. Acute toxicity occurs shortly after ingestion and is associated with cell death and other mechanisms that disrupt bodily function. More difficult to identify chronic effects, such as carcinogenicity, may require years to become apparent, or require continuous exposure over a period before causing harm, (see Scallan and coauthors 2011). Finally, allergens

84

4 Food Safety and the Ethics of Consent

trigger an overreaction of the body’s own defense mechanism, and affect a comparatively small proportion of the overall human population, (see Taylor 2002 and Cianferoni and Spergel 2009) for a discussion of allergenicity as related to GMOs). Nutritional effects of food consumption are implicated in diseases of obesity, but are not usually characterized as food safety hazards. In the interest of conciseness, I have omitted sections from previous editions that discuss general philosophical problems in testing for toxicity. Given the science of food safety as a starting point, the pivotal question is whether there are any facts about the safety of biotechnology in the production, processing or preservation of a food that are cause for special notice, concern or alarm. The key issue is whether risk assessments developed for previous types of agricultural and food technology are adequate for protecting the public against food safety risks associated with transgenic plants and animals. Given the expert framework discussed above, the question is whether risks from genetically modified products will be picked up by existing approaches to assessment of acute and chronic toxicity or allergenicity. When boosters of biotechnology claim that biotechnology is not fundamentally different from traditional technology, they are claiming that there is no need for markedly different risk assessments. This leads into the “product vs. process” debate noted by Zetterberg and Björnberg, as well as many others, (see, for example, Marchant and Stevens 2015). The “pro-biotechnology” position on food safety is that products of food and agricultural biotechnology can pose risks, and as such should undergo the same approval process as would be applied to other substances entering the food supply. Boosters argue that the process of transformation through genetic engineering and the other tools of biotechnology does not signal any basis for unique regulatory scrutiny. The “anti-biotechnology” position on food safety is in fact not a single position but a cluster of precautionary views. What holds them together is the claim that process matters. Some do question the presumption that hazards will be associated with the product, rather than the process, (Cranor 2003, Hilbeck and coauthors 2015). Yet as noted above, much of the criticism has been directed less at the actual safety of gene technology, but rather at the adequacy of the regulatory process. The rationale for the pro-biotechnology view is complex but goes roughly as follows: Genes either code for specific proteins or regulate the manufacture of proteins. If we assume that the proteins in existing foods are safe (and after all, humans have been eating most of them for centuries), then simply producing a safe protein in a new plant or animal should not count as introducing a new substance into the food system. It is, of course, theoretically possible to introduce a new substance through genetic engineering, but regulatory agencies have announced that such forms of transformation would be subjected to the same kinds of extensive toxicological testing and clinical studies that are currently required for novel drugs. None of the foods currently on the market are of that sort, (see McHughen 2000) for a more detailed but non-technical discussion). There are, however, some obvious weak points in this view. First of all, ordinary plant and animal breeding can elevate the levels of certain food components or cause them to be produced in new parts of the organism. Tropane alkaloids (the poison

4.2 Safety Criteria and Biotechnology

85

in deadly nightshade) are produced in the leaves of potatoes, but not the tubers of the varieties we normally eat. Plant breeding can and has caused this toxin to be produced in the tuber, so shouldn’t we worry that gene transfer could do the same thing? Secondly, what if the protein being introduced in the plant combines with something else to create a toxic effect? Finally, what if genetic engineering creates some totally unknown type of hazard that we have never experienced before? One could question the presumption that gene transfer is not inherently dangerous on any of these grounds, and questions of this sort form part of the basis for objecting to gene transfer on food safety grounds. These three questions have answers, however, at least in the minds of biotechnology’s boosters. The first two questions are answered by noticing that these are exactly the same hazards that are associated with ordinary plant and animal breeding. Plant breeding has produced hazardous plants, but they never got on the market because plant breeders know what to look for. Among the types of test that can be done, scientists can do a thorough analysis of all the chemicals that are in a given food, and compare that list (along with the amounts of each) to established reference models of what’s in a tomato, an ear of corn, or whatever. This is exactly the data that is sent to FDA to establish the claim that the genetically engineered variety is substantially equivalent to plants on the GRAS list. No one can recall a case where a product of plant breeding caused an injury to anyone. As such, the boosters claim, the long history of plant breeding equips us to answer the first two challenges in an era of biotechnology. As for the third question, the boosters counter that it is too broadly stated. Everything we do could have some unknown hazard. We may, in fact, be doing some dangerous things at any moment. Therefore, this argument does not provide a basis for treating transformation through recombinant DNA differently than anything else. I think there is some validity to these replies, but I would stress that it rests on the experience of the plant breeding community, rather than their scientific expertise. The product side of the product/process relies on a hidden ethical premise: Given these replies to reasonable questions, it would not be fair for government regulatory agencies to subject agrifood biotechnology to forms of scrutiny, while failing to require similar types of scrutiny for other technologies. Fulfilling regulations takes time and it generally costs money (sometimes quite a lot of money) to perform the tests needed to assign probabilities to the various hazards that pervade the food system. If plant breeding has the same hazards as gene transfer, requiring biotechnologists to perform these tests while exempting plant breeders puts the products of biotechnology as a significant disadvantage. From an ethical perspective, one can look at this as a matter or simple fairness (treat like as like) or through an economic/utilitarian lens: requiring tests skews the cost curve in a way that distorts the net benefit obtained from agrifood biotechnology. We get something less than the greatest good for the greatest number, as a result. Opponents of agrifood biotechnology may remain unsatisfied with these answers, but, as Zetterberg and Björnberg argue, it is important not to require an impossible burden of proof (Zetterberg and Björnberg 2017). I would stress that safety assurances rest on inductive logic. We cannot be certain that any product of food biotechnology

86

4 Food Safety and the Ethics of Consent

is safe, but we would be foolish to think that certainty obtains for any warrant of food safety. In my view, the rationale for continuing to raise doubts about the safety of food biotechnology is more political than ethical. The inductive logic behind food safety risk assessment means that skeptical arguments are always available to raise doubt, and some people will translate the lack of certainty into a perception of genuine risk (see Thompson 1987 for my initial statement and philosophical analysis of this problem). In fact, there may be good (that is, ethically defensible) reasons to do this, but they turn upon the goodness of shaking up the power structure in the industrial food system, rather than protecting ourselves against disease or injury from the consumption of food. Such political concerns will be revisited in later chapters, and especially in Chap. 12. Nevertheless, I do not see them as raising legitimate concerns about the safety of food, at least in so far as that is understood as the probability that consuming a product will cause illness or injury.

4.3 Ethical Gaps in Food Safety Governance The first edition of this book identified three additional ethical issues: the problem of “bad actors”, the collateral effects of biotechnology and the problem of social uncertainty.

4.3.1 Bad Actors Released in June of 1994, the film I Love Trouble (Touchstone Pictures 1994) featured a fictional genetically engineered hormone that an unscrupulous company was trying to bring to market despite some unfavorable research trials. During the course of the film, company thugs manufacture data and misrepresent the findings of studies designed to demonstrate the fictional product’s safety, along with displaying a willingness to engage in more pedestrian crimes and even murder, all in the name of corporate greed. This kind of behavior violates the most obvious norms of ethics, and no one needs a philosophy book to tell them why. Do such unscrupulous companies—bad actors—increase the risk of consuming foods produced using biotechnology? Most assuredly they do. Studies of China’s melamine scandal (Sharma and Paradakar 2010) and American attempts to conceal violations at a peanut butter plant (Roman and Moore 2012) have integrated this problem into the ethics literature. Are films like I Love Trouble remotely realistic? There have been instances where rogue scientists have violated both the letter and spirit of institutional policies (but not laws) in pursuing biotechnological research, and rumors of repressed rBST data circulated among critics for years. Genetic engineering continues to figure as a plot device in science fiction, but I Love Trouble has been forgotten. No film has done to agrifood biotechnology what The China Syndrome and Silkwood did to nuclear power.

4.3 Ethical Gaps in Food Safety Governance

87

More disturbing cases of bad actors in biotechnology have occurred, however. One concerned a Japanese company that introduced a new method of manufacturing the dietary supplement tryptophan using genetically engineered microorganisms. This incident figures prominently in Gary Comstock’s book Vexing Nature? Several deaths occurred in connection with impurities in the product, and opponents of genetic engineering continue to cite it as an example of the dangers of biotechnology. Comstock reports that manufacturing irregularities caused the problem, rather than genetic engineering. The tryptophan incident was a case of an industrial bad actor. In the end, Comstock concludes that that the tryptophan episode provides an object lesson in the need for regulatory oversight but does not provide a compelling ethics argument against any and all applications of biotechnology to food, (Comstock 2000). Another incident occurred in connection with field tests for a new pharmacologically active type of maize (corn) plant conducted by the Prodigene Company in 2002. The company did not take adequate steps to ensure that this grain, never intended for use as human food, was kept out of the food supply, and the U.S. Department of Agriculture (USDA) levied heavy fines that contributed to the company’s bankruptcy, (Thayer 2002). More generally, the Biotechnology Regulatory Services at the Animal and Plant Health Inspection Service (APHIS) at USDA conducted a review of compliance with its procedures, finding that 2% of authorized field tests involved potential compliance infractions. Between 1990 and 2001, after a thorough investigation of each potential compliance infraction, APHIS found that 76 percent of all potential compliance infractions were actual compliance infractions, and of those, 12 percent were referred to APHIS’ Investigative and Enforcement Services (IES) unit and were deemed violations, (APHIS 2006).

The potential for unethical conduct exists everywhere in human life, and biotechnology is no exception. Activist critic Vandana Shiva makes allegations of unethical conduct on the part of employees from biotechnology companies (as well as the United States Government) in India. These alleged activities range from theft of intellectual property to extortion and bullying (Shiva 2003, 2005). Bad actors exist, but how should one account for them in a risk-based approach? It is logically possible to examine empirical data on corruption and estimate the degree to which bad actors increase exposure to hazards. This is seldom done, however. A largely unexplored nexus of epistemological and political theory puzzles arise through what I have called “the virtue-risk feedback loop.” First, one observes blatantly unethical conduct; next one infers that the agent is not to be trusted. Then one infers that their products are not to be trusted, finally concluding that they are risky. But there is more: people who conceal the risks of their products are unethical, so the feedback loop cycles again, (Thompson 2015, 212–214, 2019). Anyone who believes that bad actors are common in corporate or academic settings is likely to believe that the risks of agrifood biotechnology are greater than the current risk assessments predict. This may or may not be true; regulatory risk assessments proceed under the assumption that regulators can judge risk apart from an agent’s intentions. Yet, these agencies also rely on data supplied almost entirely by the innovators themselves. The

88

4 Food Safety and the Ethics of Consent

failure to incorporate bad actors into risk assessment and regulatory policy is thus an ethical gap in the current system.

4.3.2 Collateral Consequences The most troublesome food safety issues associated with rBST did not involve human ingestion of rBST itself, but rBST’s potential for impact on other aspects of milk production. Sheldon Krimsky and Roger Wrubel summarize the debate over alleged links between use of rBST and mastitis, a disease of the udder that is normally treatable with antibiotics. Mastitis is clearly an animal health problem and could cause a food safety problem if antibiotic residue is allowed to contaminate milk, (Krimsky and Wrubel 1996). Whether or not one finds their criticisms of rBST convincing, they illustrate a general problem that deserves far more consideration than it typically receives. The effects of technology are systemic. Using a novel technology alters the way that many other things are done, sometimes in subtle and unpredictable ways (Tenner 1996). As discussed in Chap. 7, herbicide tolerant crops have been criticized by those who believe that they will encourage farmers to use more herbicides. Indeed, a global controversy over the safety of glyphosate, the herbicide used for genetically engineered Round-Up Ready™ crops, has broken out since the second edition of this book appeared in 2007, (Duke 2018). Controversy over glyphosate emphasizes health impacts from direct exposure by applicators, rather than a threat to food safety, though one celebrated study exposed rats to herbicide resistant potatoes (Mesnage and Antoniou 2017). It is nevertheless yet another example of how biotechnology that is benign in itself might have collateral consequences that affect food safety. It is relevant and important to consider the way that one technology will have an impact on the use of other tools and techniques (or on non-technological practices) when considering risks to health. David Collingridge’s principle contribution to technological ethics is a detailed account of the epistemological reasons why collateral consequences are difficult to anticipate. Like any scientific research activity, risk assessment requires a stopping rule, a point at which scientists judge that they have done enough, that they have a meaningful result. Stopping rules for statistical analysis require a high degree of technical sophistication and theoretical justification. There do not seem to be similar ways to develop stopping rules for interactions with other actions and practices that create a cascade of impacts. From a logical and ethical perspective, it is important to remember that this problem attends to every effort to base a decision or value judgment on some projection of future events. Human beings never have all the relevant information; that is why we introduce probabilistic thinking into our planning, in the first place, (Collingridge 1980). However, much of the impetus for technological ethics is expressed in Hans Jonas’s insistence that we must assume responsibility for collateral consequences by trying to do the best that we can.

4.3 Ethical Gaps in Food Safety Governance

89

4.3.3 Social Uncertainty From the perspective of science, (that is, from the perspective that sees the issue in terms of focused controversies, interpretation of data and methodological debates) the term “uncertainty” implies a set of standard problems well known to anyone trained in scientific procedures. It is also important to note a less technical phenomenon that might be referred to as social uncertainty. Social uncertainty is the indecision, hesitation and skepticism that often attends scientific controversy and that prevents policy makers and ordinary citizens from feeling confident about what they ought to do. Because scientific concepts, data and even experimental studies are open to multiple interpretations, it is often possible to make almost contradictory statements about what is and what is not known about a controversial phenomenon on the basis of scientific studies. Because many different scientific disciplines are involved in estimating technological risk, virtually no one is in a position to command all the information and expertise relevant to decision making. Even highly educated and interested members of the public and government officials, in particular, must thus rely on expert testimony for their interpretation of technological risks. This situation of unequal access to expertise and knowledge creates a form of uncertainty about risks that is seldom discussed in scientific studies. In effect, anyone who has not actually collected data and conducted risk analysis is in a position such that their assessment of food biotechnology’s safety reflects two factors: the content of experts’ risk estimates, and one’s confidence in the competence, neutrality and reliability of the experts themselves. Confidence derives from the social relationship between any given individual and the expert group conducts the risk analysis. In situations where confidence in the group doing the analysis is very low, statements that risks are acceptable can actually increase a person’s estimate of the risk associated with the activity in question. Social sources of uncertainty cannot be eliminated by conducting more scientific studies. One of the most significant sources of social uncertainty is grounded in the difference between local knowledge and more conventional scientific techniques. Brian Wynne (1992) conducted a series of studies on attempts to quantify and manage risk in the wake of the Chernobyl accident which revealed that scientifically trained risk managers made a number of errors in mitigating radiation contamination risks among sheep farmers because they did not understand local conditions and practices. Farmers lost confidence in the scientists when it became clear that they did not understand a number of relationships crucial to the economic viability of sheep farmers. This loss of confidence exacerbated public health risks when communications between public officials and farmers became mired in misunderstanding and distrust. If it becomes clear that scientists do not know or appreciate things that are well known by affected parties, trouble ensues, especially when the affected parties are in a position of social vulnerability to actions they know to be ill informed or narrow in perspective. There is little doubt that problems of social uncertaintyplague food and agricultural biotechnology, beginning with the debate over environmental risks associated

90

4 Food Safety and the Ethics of Consent

with genetically engineered ice-nucleating bacteria in the 1980 s (see Thompson 2017, pp. 221–228). The first edition of this book appeared in 1997 before products were on the market, and it included the speculation that social uncertainty with respect to the food safety of GMOs was being dampened by confidence in the regulatory system. Now that products of agrifood biotechnology are being profitably sold and utilized, people are actually being exposed to whatever risks might legitimately be associated with this technology. In 1997, the risks were hypothetical. Twenty years later, commentators recognize that food safety has emerged as a source of significant public concern, (Malyska et al. 2016). Social uncertainty has emerged as a significant contributor to the policy dilemmas faced by industry and food safety regulatory authorities. What is the ethical response to social uncertainty? To some extent, this question must await analysis developed in Chap. 12, for it is often the interplay amongst food safety, animal welfare, environmental and social issues that creates competing interests and places people into positions where they are unlikely to place confidence in expert views of risk. The succinct answer is to advise against the creation of situations in which great inequalities in power and access to information exist. However, it must be admitted that some such inequalities are unavoidable. As with bad actors and collateral consequences, social uncertainty arises within the bounds of reason, even if it is not recognized in expert risk assessment.

4.4 The Philosophy of Food Safety Once obvious risk-based tests are passed, why would the safety of food biotechnology raise any serious ethical questions at all? The dominant pro-biotechnology answer to this question is that there are no more ethical questions: Critics of biotechnology are hysterical or liars (or both), and the public is ill informed, easily misled and otherwise apathetic about the safety of food biotechnology. However, the dominant answer is wrong. Risk defined as probability of food borne illness or injury does not encompass the full range of issues that are traditionally associated with food safety. Socially based management of food risks has undergone three broad phases. For most of human history, the problem has been one of classification. Is it food or not? When medical scientists began to appreciate the importance of germs and other contaminants as a cause of disease, a more subtle approach arose which stressed the elimination of impurities. Only recently have scientists begun to appreciate the complexity of whole foods, recognizing that many components of foods may have toxic properties under certain circumstances, and that toxicants themselves may be responsible for benefits that offset the disease risk associated with their presence in food. This has introduced a tendency to think of food safety as an optimization problem that gives rise to the risk-based approach described above, and supports it with impressive scientific credentials. The optimization approach is correct when safety is conceived very narrowly, but the consideration of bad actors, collateral consequences and social uncertainty

4.4 The Philosophy of Food Safety

91

shows that alternative views are entirely reasonable. For food safety decisions we must make collectively (and given the complexity of our food system, that is most of them), norms of democracy demand an accommodation for all reasonable points of view. This is not to say that everyone gets whatever they want, but it does entail that collective or political decisions must be made in a manner that does not foreclose the possibility of reasonable disagreement and that when divergent viewpoints can be accommodated without imposing unreasonable costs or inconveniences on others, they should be. The case for applying these general norms of democracy to the issue of food safety and biotechnology begins by considering each of the three historical approaches to food safety in more detail, and by gaining an appreciation of the concomitant values that complicate food safety decisions. Given the diversity and complexity of these values, norms of democracy weigh heavily in favor of a policy that preserves key elements of consumer sovereignty and consent. The chapter concludes with a discussion of how labeling is one means for supplying consumers with information and protecting the autonomy of their decision-making with respect to diet.

4.4.1 Classification Human beings recognized the existence of poisonous foods in prehistory. By the time of the Greeks, such knowledge served as the basis for Socrates’ famous meditations on death and duty prior to drinking hemlock. The ability to distinguish foods with toxic constituents from microbial toxins would have confounded early efforts to manage the risk of ingesting poisons through a dietary regimen, but trial and error would have eventually ruled out acutely poisonous plants and animals. It is less sure but seems likely that humans came early to the knowledge that eating the wrong thing could have delayed effects and increase chronic health risks. Clearly, early beliefs about food safety were a mix of superstition, speculation and hard-won experience. The Pythagorean cult of which Plato was a member proscribed the eating of beans, though it is unclear whether the deadly effect of castor beans, metaphysical beliefs about the germinative power of seeds, or simply the problem of flatulence was the occasion for this food rule. Generally, such rules classify plants and animals into food and non-food groups and in some cases specify finer distinctions for preparation or serving foods. Whatever else might be said about these culturally transmitted food rules, in every known case following them results in food consumption that is safer than a diet of randomly sampled plants and animals. This is an unexceptional fact; a food culture that included acutely toxic elements would disappear rather quickly. Customary and religious dietary rules distinguish food from non-food. While adherents of a given system of food classification may associate violation of the rules with sickness, injury or death, they may also associate it with religious, spiritual and social forms of risk. Religious, ethnic and regional food rules persist today because they are constitutive of social, cultural and personal identity, because they reinforce feelings of well-being and order, and because people love them, find them pleasing,

92

4 Food Safety and the Ethics of Consent

satisfying and take gratification in following them, (Rozin 1990). Food safety is but one function of a cultural dietary regimen. This implies that upsetting a regime has the potential for impact that goes far beyond the ambit of the biological sciences.

4.4.2 Purification Anthropologist Mary Douglas (1966) has used the term ‘purity’ to describe a traditional society’s adherence to a system of social practices including food rules such as described above. What I have in mind here is a way of looking at food safety that begins with the advent of the germ theory of disease in the nineteenth century. The germ theory held that disease was the result of invisible infectious agents or germs, and that strategies of purification could control disease before it starts. The germ theory gave rise to a strategy for food safety that deployed technical means for preventing the entry of infectious agents into human food, for destroying them once they are there, and eventually for treating those afflicted by germs. Refrigeration, sterilization, irradiation and temperature monitored cooking are all weapons in the arsenal of purification. Dane Scott has linked discoveries in microbiology by Louis Pasteur and Robert Koch to immunologist Paul Ehrlich’s vision for a monocausal paradigm in biomedical research. According to Scott, this model, “…does not consider the host or the environment as diseasing-causing factors,” Scott (2018, p. 41). This assumption facilitates a scientific focus on therapies or interventions that target a specific agent that is the source of morbidity and mortality, a strategy that Ehrlich referred to as a search for “the magic bullet,” (Scott 2018, p. 40). There is little doubt that the magic bullet approach was quite effective in combating acutely toxic agents in food and drinking water, (see Latour 1988). Belief in the efficacy of purification spawned legislation that regulated the food industry and created governmental agencies for enforcing the first food safety laws. It was the beginning of the modern conception of food safety. Sinclair Lewis’s The Jungle created an uproar with its description of unsavory practices in the Chicago packing plants, among them an episode in which a hapless immigrant worker is ground into sausage as the line rolls on without missing a beat. In a ghoulish way, this episode from The Jungle is indicative of the way that food safety regulation based on a strategy of purification carries water for cultural foodways, just as Mary Douglas would predict. If we set prion diseases (unknown to the readers of The Jungle) to the side, grinding up a human being along with the sausage introduces little additional health risk for consumers. The Jungle was effective because however little regard middle class Americans may have had for immigrant workers, they were queasy about consuming parts of them in their hot dogs. The food safety regulations that came on line in most industrialized countries prior to World War II used the phrase “safe and wholesome.” They prohibited the use of dogs, cats and rodents in meat products despite the fact that all these animals are used for food in some non-European cultures. Purification introduced science, technology and government into the pursuit of food safety, but retained many cultural norms in defining when a food is pure, (Belasco 1997).

4.4 The Philosophy of Food Safety

93

Purification was still the model when a new class of food safety problems began to be discussed in the 1950 s. The U.S. Delaney Clause, for example, appears to be a model of the purification approach. The law requires the U.S. Food and Drug Administration to ban any additive found to cause cancer in humans. The law is accompanied by a list of traditional food ingredients deemed “Generally Recognized As Safe” (GRAS). It is imminently plausible to think that many of the legislators who voted for this law, if not Congressman Delaney himself, were thinking that foods included on the GRAS list are non-risky, and that hazards are associated with contaminants, (Levine 1986). They may well have thought of themselves as instructing the regulators to seek out the contaminants and ban them. As eminently plausible as this thinking may seem, it is incompatible with the new thinking on food safety that is represented by the optimization paradigm.

4.4.3 Optimization Quite recently, scientists and regulators have adopted the risk-based view of food safety. The conceptual distinction between purity and risk-based approaches to food choice hangs on one’s interpretation of “no risk,” the criterion applied to food additives under the Delaney clause. A thorough purificationist (if such ever existed) believes that whole foods consumed since time immemorial bear no risk. This cannot mean that one will never come to harm from eating them, since food pathogens and unexplained poisonings and reactions have been around for time immemorial, too. Consistent with Scott’s portrayal of the monocausal paradigm in medicine, food is simply not viewed as a disease-causing agent. Health impacts are, on this model, always caused by agents that are introduced into foods, either intentionally or nor. Toxic microbes are not intentionally introduced into food, but chemical additives are. Additives are thus risky in an ethically significant way that microbes are not. The harm they produce is the result of an action for which some individual or organization can be held morally and legally culpable. Few (if any) scientists or regulators think of risk in these terms. For them, “no risk” means zero risk, a quantity that can be reached only when the probability of harm associated with an action or choice is zero. This is an assumption that makes the Delaney Clause intellectually incoherent. At best, epidemiological studies and animal trials will reveal no statistical evidence of carcinogenicity, but “no statistical evidence” is far short of proving zero chance of harm. More detailed studies and better tests can always overturn this result, and that has indeed been the case on several occasions. In addition, the purificationist interpretation of the Delaney Clause clearly admits circumstances in which GRAS substances with risks known or strongly suspected to be higher than those of banned additives are allowed, clearly a suboptimal situation. What is worse, the risk-based approach has begun to surface evidence that many common foods fail the zero-risk test. The work of Bruce Ames (1979) is typical of this new view, though it was clearly on the horizon well before Ames became an advocate of it. Ames’ biochemistry work has shown that virtually all foods contain

94

4 Food Safety and the Ethics of Consent

mutagens—substances that increase a the statistical rate of “errors” or minor changes in the order of DNA base pairs as a cell duplicates itself. Mutagens are thought to precipitate cancerous cell growth, though their very ubiquity shows that any link between mutagenicity and carcinogenesis is complex. Cancer would be everywhere if mutagens caused cancer, tout court. Ames has promulgated the view that since naturally occurring mutagens far outnumber those associated with chemical additives and pesticide residue, it is unlikely that the rate of chronic diseases in the human population will be significantly reduced by strategies that target food additives or chemical residues (Ames 1983; Ames et al. 1987; Ames and Gold 1998). Ames’ way of thinking is now the paradigm for conceptualizing chronic toxicity in foods. A full treatment of this view would be out of place in the present context. However, it is appropriate to point out how the rationale he applies to pesticide residue has influenced thinking on the link between chronic risks (such as cancer) and many possible contaminants in food. One can test for mutagenicity or carcinogenicity by isolating a substance and exposing laboratory animals to high doses, but what does that tell you? If Ames’ thinking is right, many foods contain mutagens, but we should not conclude that they should be banned. The same would apply to genetically engineered foods, which can be expected to contain both mutagens and anti-mutagens. Since a genetically engineered food includes all the components (proteins and other chemicals) that were in the food before adding or deleting genes, it can, like any food, be expected to contain mutagens and anti-mutagens. There is no point in breaking a GMO down to all of its biochemical constituents and testing each of them for mutagenicity, since doing so will provide no basis for distinguishing the GMO from unmodified foods from the same species. Under the Delaney Clause, discovering that a food contains mutagens (which are candidates for cancer-causing agents) could lead to a ban on the food with no benefit in total risk reduction. Under this scenario, the appropriate strategy for addressing chronic risks in foods has been something akin to “Don’t ask; don’t tell.”

4.5 Classification and Purification Versus Risk-Based Optimization Since mutagens are everywhere in human foods, the new view drastically revises the importance of the food/non-food distinction. Foods (tomatoes, beets, beef) are as capable of introducing the fatal mutagen into the body as are additives or non-foods. People do not get cancer all the time because somehow these mutagens are kept in check, either by the body’s defense mechanisms, or perhaps even by each other. The risks of chronic diseases such as cancer cannot be controlled by purifying foods, for the foods themselves are not benign. Instead, we must seek a balance point, not yet well understood, in which the mutagenicity of what we eat is balanced or blocked (as much as possible) by all the other factors (good nutrition, exercise, cognitive stimulation) that create health.

4.5 Classification and Purification Versus Risk-Based Optimization

95

This way of conceptualizing food health risks fits hand in glove with the risk/benefit trade-off thinking that had already begun to emerge in the wake of attempts to regulate agricultural insecticides, herbicides and other chemically based pest control technologies. Indeed, some such technologies, notably rodenticides and insecticides used to control pest infestations of harvested grain, had been introduced in order to prevent contamination. This is a clear application of the purification philosophy. It was evident that when adequate substitutes were unavailable, a ban on these technologies would be followed by a resurgence of pest infestations and attendant health problems. In such cases trade-offs are inevitable, and the trade-off involves food safety, not merely economic losses. Sometimes the interests of human health will be better served by accepting the risks associated with the technology, sometimes it will be better to accept the pests. Both Ames’ work and the trade-off logic of chemical technology support a reconceptualization of food safety along the lines of risk optimization. Risk comes to be seen as pervasive and the presumptions of purification (that risk is due to impurities only) come to be seen as naïve. We must accept some chance of harm, and the norm that we should apply to the problem of food safety is to define an optimum, a balance point, and to regulate food production and processing so as to approximate the optimum as closely as possible. There is no logical reason why this way of thinking should not be applied to acute toxins, as well as mutagens and other chronic agents. Optimization is clearly a tricky business; it is not simply a matter of minimizing risk. It requires one to answer questions such as How should we compare a statistically high level of risk to a few agricultural field laborers, to a very low probability of harm that may fall on food consumers? Optimization strategies for food safety demand a high level of philosophical and ethical sophistication in their treatment of risk, but the point here is simply to note how dramatically the optimization approach differs from that of either classification or purification. Seen as strategies bound by the limited knowledge of their respective moments in history, classification and purification can be interpreted as worthy attempts at risk optimization that have become obsolete. This amounts to the claim that ethnic or religious foodways and the purification technologies that followed were conceived and adopted as a way to strike the proper balance between risk and benefit for food safety. I do not think that this claim is plausible, though I know of no historical or anthropological research that could disprove it. What seems more likely is that the normative philosophy of optimization evolved at the same time as the relatively recent scientific theories and political problems to which it is so admirably suited. If that is right, then the values and experience of previous generations and of many in the present day do not provide adequate support for a strategy of optimization, and if this is true, then reasonable, intelligent people are likely to conceptualize food safety more along the lines of classification and purification than as an optimization problem.

96

4 Food Safety and the Ethics of Consent

4.6 Food Safety and Ethics The most obvious way to move from risk-based optimization to an ethical philosophy for food safety is to interpret it as an extremely sophisticated extension of utilitarianism. The main problem with risk-based optimization is also a traditional problem of utilitarianism: too little attention to consent. The consent-based approach harks back to a system of caveat emptor. Individuals accept risks to which they have given informed consent, but coercion or concealment of relevant information is morally unacceptable. Critics of consumers’ right to know argue that this ethic is unrealistically demanding for a modern food system, but the justifiable demands of such rights are less onerous than is sometimes thought. The existing food system in most parts of the world is surprisingly close to what the rights view would hold as ideal. The ethics of food safety is framed by the dialectical opposition between riskbased optimization, on the one hand, and an alternative ethical system that emphasizes consumer choice, citizen autonomy and consent, on the other. I regard this as a genuine dilemma in which there are compelling points to be argued on each side. But the dialectic is complicated. First, few specialists in food safety appear to recognize that there is a dilemma at all. Indeed, they write and speak as if risk-based optimization is pure science, rather than being shot through with (often defensible) philosophical assumptions. As such, it may be necessary to overstate the case for autonomy and consent. Second, there is slippery intellectual turf separating the two poles of this dialectic. Given certain plausible views on uncertainty, the authority of science and the individual’s right to believe what they want, it is possible to slide from optimization over to consent without noticing it. However, to launch this dialectical exercise, it is useful to articulate the links between risk-based optimization and utilitarian ethics more explicitly. One strength of the utilitarian approach is that in emphasizing the likely consequences of a given action or policy, it represents the most obvious way to bring predictive science to bear on ethical decision making. Indeed, many scientists gravitate so easily to the view that public policy should apply the best science to predict the consequences of several policy options, then choose the option with the best outcomes, they fail to notice that they are applying a philosophical framework at all. The modern food system has evolved under circumstances in which the public was generally happy to have many decisions about the ingredients or composition of food left to the experts (Knorr and Clancy 1984). A utilitarian might interpret this state of affairs as ethically justified because the cost of staying informed far outweighs any benefits the information would bring to the mass of people. Nevertheless, the problems with a too narrow emphasis on risk management and trade-offs are also generally the problems of utilitarian approaches. Utilitarianism has always been criticized because it appears to make short work of rights in too many instances, depriving people of the opportunity to make their own choices in the name of doing what is good for them. In the case of food safety policy, the claim that policy should be risk-based, (or simply science-based) neglects the fact that science provides little insight into many of the dimensions that influence individual food

4.6 Food Safety and Ethics

97

choices. These dimensions offer no reason to ban or regulate food biotechnology as a matter of public policy, but any of them might well provide an individual with defensible reasons for preferring not to eat these new foods. Is “food safety,” a purely scientific concept, to be defined and controlled exclusively by food scientists, or is “food safety,” a term of ordinary language? If it is the latter, then it becomes relevant to see what non-scientists mean when they use the concept of safe food. Food journalist Robin Mather puts in this way in writing for a popular audience: Because food is so important to us on so many levels, we must generally trust blindly in government’s promises of a safe food supply and in the safe practices of those who produce the foods we buy. We wade doggedly through complicated, confusing and often contradictory information about the foods we eat. The reason is that most of us realize on a fundamental level that food choice is one of the last arenas in which we have some measure of control. Mather (1995, 5).

Here Mather characterizes “control” as a primary dimension of food safety. Most of her book discusses the social organization of agricultural production (see Chap. 8), but she links the social impact of biotechnology to this paragraph on food safety by noting how difficult it is for food consumers to “track where our food comes from” (p. 5). Building on Mather’s comment, the problem with the risk-based approach is not that it is wrong as far as it goes, but that it does not go far enough, ignoring the “many levels” on which food is important to us. Mather’s complaint with government regulation of food safety is that it is being used to undermine “one of the last arenas in which we have some measure of control.” Of course, the Department of Transportation may be “undermining our control” when they mandate air bags or speed limits. In one sense, every public policy affects individuals’ control over their lives. This is not in itself a powerful argument. What must be shown is that food consumers have rationally defensible ends in view in wanting alternatives to biotechnology-derived foods. There are at least three examples of non-science-based but rationally defensible ends. 1. Religious and Ethnic Beliefs. As already noted, the cultural history of food beliefs has produced a rich array of religious and ethnically based beliefs about what is and is not food. These beliefs are imperfectly correlated with scientific probabilities concerning illness or injury, at best. Yet no one challenges the right of religious and ethnic minorities to stipulate food rules that are far more restrictive than those stipulated in legal codes. Jewish and Muslim practices, among the most widespread, demand both dietary restrictions and special procedures for slaughter and preparation. These rules remain under the constant supervision of each groups’ ecclesiastical authorities. Any intentional or de facto attempt to decide this question on their behalf would violate minority rights that are well established throughout the industrialized world and covered by the International Declaration on Human Rights, (see Brunk and Coward 2009). 2. Latent Purificationism. Few lay persons and even many scientifically trained individuals can make ready sense of the risk-based optimization approach that is

98

4 Food Safety and the Ethics of Consent

de rigour among toxicologists, epidemiologists, biochemists and food regulators. Since the first edition of this book, research in cognitive science suggests that most people rely on heuristics, rather than risk optimization, to make most everyday decisions, Kahneman (2011). The purification model that has been around for a century is consistent with this. The idea that crossing genes is a violation of purity rules comes readily to many people, and they are at least psychologically justified in feeling somewhat queasy about genetically engineered food. 3. Distrust of Science. A subset of the queasy are angry, too. Many of them purchase organic foods, often at great inconvenience to themselves, and they write angry letters to activist magazines, complaining about “Franken-foods.” Clearly it is impossible to isolate oneself completely from science and its impact on our world. Claiming a right to do that would be preposterous. But people have not previously been forced to choose between total reliance on science or subsistence agriculture. Food choices and alternatives have been the norm, if not a right, and people are justifiably resentful (and suspicious!) of the forces that threaten this valued status quo. Philosophical positions that promote distrust of science and technology are discussed in Chap. 14. The virtue-risk feedback loop amplifies the importance of this feature.

4.7 Food Labels and Consent Many of the things that people want to know about food have nothing to do with science and are only marginally related to safety (as conceived as the probability of illness or injury). But people want to feel good about their food choices and if this means knowing that their Champagne comes from France rather than California, or that their hot sauce is made in Texas, NOT New York City, then having the ability to discriminate on the basis of such information contributes to their feelings of wellbeing and satisfaction. To the extent that “safety” connotes a feeling of security and well-being, such information contributes to food safety. This stretches the word ‘safety’ in a way that food safety specialists would reject, but what is clear is that people will feel suspicious of and at-risk from individuals or groups who try to deprive them of information they deem valuable. And then, the virtue-risk feedback loop is starting to cycle. The point of food safety policy is not merely to make foods safe, but to provide the public with reasonable assurances of food safety. Ironically, it may not be possible to accomplish this latter objective without providing information that has little bearing on the probability of harm from consuming the food in question. Europe’s approach to agrifood biotechnology stipulated mandatory labels for GMOs almost from the outset. In the United States debates over labeling raged for over two decades, culminating in Congress mandating that the U.S. Department of Agriculture develop some form of label in 2016. The U.S. approach has not been fully implemented as of this writing, but the law is receiving criticisms from many different directions, (see Strauss 2018; Westerman 2018). In other writings, I have argued that the food industry

4.7 Food Labels and Consent

99

has a responsibility to provide consumers who want to avoid consumption of GMOs or other products of gene technology with the means to do so. Thompson (1997, 2002). However, many of the arguments offered for mandatory labeling genetically engineered foods do not stand up to close examination. The most straightforward argument for requiring labels proclaiming the use of biotechnology would be if there are demonstrated health risks associated with consumption of products derived from biotechnology, as there are for alcohol and tobacco products. Such labels would warn consumers of such risk. Clearly, from what has gone before in this chapter, no evidence exists for health risks, so this argument relies on a false premise. One might also claim that since biotechnology creates products that are not readily recognizable as processed, GMOs may be confused with natural foods, meats or grains. Consumers who want a product free from the use of gene transfer or editing have no way to recognize such products. As such, they are indeed deprived of any right they have to satisfy that preference by unlabeled produce, meats and grains. I believe that there is cogency to this argument, but one must careful. Way back in 1996, Michael J. Reiss and Roger Straughan argued that the natural/artificial distinctions is the source of many ethical blunders, and we ought not to encourage more of them, (Reiss and Straughan 1996). Bioethicists have seen the concept of naturalness deployed to suppress women’s rights, racial minorities and especially against lesbian, bisexual, gay and transgender individuals. The ethical stress in the pro-labeling argument must be laid on individual preference, rather than the alleged unnaturalness of a genetically engineered food. However, the preference argument itself needs qualification. Data on public opinion demonstrate an abiding interest in labeling of genetically engineered foods in virtually every population surveyed. (Hoban and Kendall 1993, Frewer et al. 1997, Pew Initiative 2004). These surveys alone do not establish ethical reasons for requiring that the producers of biotechnological foods label their products because given the question, “Would you like more information (or labels) on X?” many people are likely to respond affirmatively, without regard to what X is, or whether they have any legitimate interest in having the information. The argument needs supplementation with an account of why consumers might have reasonable preferences, which they would exercise if the information were available. Many preferences are illegitimate, and should not be supported by mandatory product labels. However, considerations discussed earlier do supplement preference surveys by showing how sensible and widely held views on food could lead a reasonable person to experience anxiety about genetically engineered food. Consumer preferences against GMOs are ethically legitimate, even when they are falsely conceived. The most compelling argument against mandatory labels is that they compel speech. Freedom of speech implies the liberty to express utterances, but also protects persons’ right to hold their tongue, to withhold speech in situations of their choosing. John Stuart Mill’s argument in On Liberty examines exceptions to this principle, concluding that it is overridden when the acts of one person harm (or have the potential to harm) another person. Agents can be compelled to disclose information material to risks, but the U.S. Supreme Court has held that even commercial speech (e.g. speech pertinent to the promotion or sale of commercial goods) is protected by

100

4 Food Safety and the Ethics of Consent

the First Amendment. Food companies can indeed be compelled to disclose risks to health, but they cannot be compelled to disclose much else, (Pomeranz and coauthors 2019). It remains to be seen whether the new U.S. labeling law is consistent with the First Amendment, but I remain content with the view that mandatory GMO labels are not supported by ethics. The 1997 edition of Food Biotechnology in Ethical Perspective was interpreted by some readers as implying that science trumps the consumer’s right to choose (see Jackson 2000; Streiffer and Rubel 2004). However, alternatives to mandatory labels exist, and they are capable of resolving the ethical problems with approaches that preclude consumer exit. Beyond the compulsion of speech issue, the most serious ethical objection to mandatory labels is that they stigmatize products of biotechnology unjustly. Although there are reasonable concerns that may lead some to avoid genetically engineered foods, it is at least as reasonable to accept them as beneficial. This is, of course, especially the case for foods engineered to boost nutritional value, but even crops primarily of value to farmers can be beneficial additions to human diets. Stigmatization of agrifood biotechnology groundlessly reduces the commercial viability of genetically engineered foods, and this can plausibly be interpreted as interference in the rights of the food industry, its investors, and non-profit biotechnology researchers. Robert Streiffer and Alan Rubel criticize this last point in two papers by noting that no one has a right to any level of guaranteed economic return on a transaction or investment, Streiffer and Rubel (2004), Rubel and Streiffer (2005). It is certainly true that economic returns are not protected by rights. However, those who offer products for sale to the public do have the right to fair conditions of competition. Suppose a country (call it Govingia) passes a law requiring that all imported beer bear a large label with a skull and crossbones, then indicates somewhere on the Govingian Ministry of Trade’s website that the skull and crossbones is not intended imply any defect or hazard but is simply the symbol adopted to indicate an import product. We would not be inclined to call this fair. It is not at all implausible to diagnose this unfairness as a violation of the importing company’s right to fair conditions of competition. Streiffer and Rubel also do not address the possibility that there might be unfairness to non-profit biotechnology researchers, people who have invested a life’s work in a set of technologies and who may have no personal financial stake in their success. Ingo Potyrkus has complained bitterly about what he takes to be the lack of fairness with which his work on Golden Rice has been received (Potrykus 2001). So although Strieffer and Rubel are certainly correct to note that no one is entitled to success, economic or personal, I reiterate that the stigmatization mandatory labels might create for agrifood biotechnology could in fact be ethically problematic in being unfair to the people who develop this technology. The 1997 edition also suggested that mandatory labels for genetically engineered foods would be very difficult to enforce because at that time no test existed that could reliably detect whether genetic engineering had been used. That is a situation that changed fairly soon after the book was published, and technical monitoring issues have presented little barrier to segregation or sourcing of GM from non-GM products. Things are changing again in the era of gene editing, however, so the issue of technical enforceability is with us once again. This is yet another consideration in favor of a

4.7 Food Labels and Consent

101

voluntary, industry-led approach. A voluntary non-GM label would almost certainly sell at a higher price than commodity grade “may contain” GM foods. Part of these price premiums would certainly go toward the costs of segregating and labeling the non-GM product. The cost burdens for any kind of mandatory label will be more broadly distributed across the food system, probably raising the costs of all foods slightly. These cost increases may be very slight indeed when widely distributed, but it is nonetheless the case that increased costs for food fall most heavily on the poor, for the same reason that lower costs benefit them more. Economic arguments such as these are made by Per Pinstrup-Andersen and Ebbe Schiøler, (Pinstrup-Andersen and Schiøler 2000), and by Gary Marchant and Guy Cardineau (Marchant and Cardineau 2013). I argue that the ability to avoid genetically engineered foods is what matters from an ethics perspective. In consumer decision making, the principle of consent is protected by the availability of alternatives. These alternative foods give food consumers the right of exit from a system of food transactions that they find objectionable. If there are identifiable alternatives to the products of biotechnology, then consumer sovereignty and the principles of consent are protected, (Thompson 2002). There are several ways in which the principle of exit can be protected, and the most obvious of them all involve labels that identify a product as “biotech free.” One approach is to specify that “green” or “organic” products may not utilize biotechnology, so that concerned consumers may opt out of the biotechnologicallydominated mainline food system by shifting over to an alternative segment that exists independently of the GMO question. This is the solution that appeared most likely in 1997 and that eventually became United States policy when the USDA organic standard disallowed the use of crop or livestock varieties modified through recombinant tools. Yet this solution is far from ideal. Pam Ronald and Raoul Adamchak argue that green and organic buyers should be the most enthusiastic buyers of the environmentally friendly products that could be developed using biotechnology (Ronald and Adamchak 2017). The move to a “Green equals biotech free,” policy both stigmatizes these products and undercuts the potential market for applications of gene technology that are intended to support environmental benefits. A voluntary label placed on products where no biotechnology is used also supports exit. The label would be negative in that it would proclaim the absence of biotechnology, rather than its presence. Since all foods produced prior to the first genetically engineered crops could qualify, it would be odd to suggest that affixing a negative label to them should be anything other than voluntary. Yet, if a sufficient number of products begin to use negative or “no biotech” labels, then those who wish to avoid biotechnology can do so. Such negative product claims began to appear in the United States just as the second edition of this book went to press. This chapter has gone on too long to warrant a philosophically adequate treatment of them, and a brief discussion must suffice. In 2009, Robin Jane Roff gave an analysis of the Non-GMO Project that (ironically) supports the approach I took in earlier editions of this book. Roff draws a distinction between labels that function to incentivize a particular farming system and those that are designed to present alternatives to a mainstream or dominant practice.

102

4 Food Safety and the Ethics of Consent

She characterizes the latter approach as one oriented to servicing “consumer choice,” though I have argued that exit is a more ethically precise term. Roff argues that the Non-GMO Project was originally conceived as a way to enroll farmers and consumers in a cooperative project to fight against the growing use of gene-engineered crops. Learning from the experience with the organic label, organizers resisted a certification process that allowed farmers or food companies use both GMO and non-GMO to participate. Gradually, however, the need to cultivate markets for the label led to greater influence by food industry firms. To her dismay, Roff concludes that the label is serving the interests of the food industry, rather than a social movement dedicated toward halting the growth of agrifood biotechnology, (Roff 2009), see also Bain and Silfa (2017). Roff despairs at this turn of events in part because she sees it as weakening the effectiveness of a social movement against biotechnology. As stated above, I regard explicitly social concerns as ontologically distinct from food safety hazards. They become relevant to the extent that they provoke distrust or moral outrage, thus initiating a virtue-risk feedback that emerges in a consumer’s concern about the safety of a product. It is not irrational to be concerned about the food safety of GMOs, but the question as to how these individually rational perspectives should be reflected in public choice is also ontologically distinct from the hazards addressed by food safety governance. As such, I defer discussion of these problems to later chapters. Roff also sees choice-facilitating food labels as implicated in the regrettable growth of neoliberalism. This, too, is ontologically distinct from food safety hazards, and is taken up especially in Chap. 14. I agree with Roff’s analysis of why the mainstream food industry benefits from a voluntary “biotech free” label, but unlike her, I see such labels as ethically justified in light of the need to protect consumer consent. The question of whether the critique of neoliberalism abrogates the legitimacy of consumer consent also awaits developments in later Chapters.

4.8 Food Safety Risks in Ethical Perspective When viewed from the risk-based perspective on food safety, biotechnology scores as well as alternative methods for developing plant and animal varieties with the traits that farmers, food processors and consumers want. Chapter 2 offers an argument for thinking that innovations in food production should not be require an ethical justification in terms of future benefits. We should instead focus on risks, harms and unwanted consequences. Together with the review provided in this chapter and an addendum to the food safety debate in Chap. 13, all this implies that food safety risks do not provide compelling arguments against the use of rDNA techniques for modifying plant and animal genomes. Lest this generalization be misinterpreted, it does not imply that every product of gene technology is safe. Rather, it states that the process itself is not appreciably more risky than other forms of plant and animal breeding (discussed in Chap. 1). In this specific sense, I am committed to a product not process perspective.

4.8 Food Safety Risks in Ethical Perspective

103

However, consumers do retain the right to give or withhold consent to the residual risks, however small, associated with the use of gene technology. Scientists and regulators must not abandon the view that their primary responsibility is to ensure that biotechnology does not endanger public health, but there are ways to do this without coercive manipulation of the food system (Korthals 2014). When it comes to protecting the consumer’s right of exit, the use of gene technology does matter. It matters in part because, from the debate over bovine growth hormone onward, gene technologies are associated with additional risks beyond human health. Virtuous actors would not stifle debate over these social risks; they would take part in reasonable efforts to govern them. That industry and, to a large extent, even university scientists have failed to engage these non-health related issues tarnishes their character. This launches the virtue-risk feedback loop, leading to the reasonable judgment that they might be hiding something. It is not unreasonable to worry that the thing they might be hiding is that the products of gene transfer are not safe to eat. In this specific sense, I am committed to a process, not product perspective. My position looks overly subtle to colleagues on both sides of the agrifood biotechnology debate. I adhere to it in part because, as noted throughout this chapter, as a scholar aspiring to the role of honest broker, (see Pielke 2007). I cannot sublimate my understanding of the truth about safety to strategic objectives in the social realm. I also believe that the failure to appreciate these ironies continues to block achievable progress in the agrifood biotechnology debate. People who come to be genuinely mistrustful of the individuals and organizations pursuing biotechnology view every consideration adduced to reassure with suspicion. The virtue-risk feedback loop prevents them from thinking more deeply about food safety risks. Scientists wedded to a product-not-process perspective are unable to see that, biological hazards notwithstanding, the process is embedded in social institutions that are themselves potential sources of risk. As I see it, this was the position that I was arguing even in 1997. I hope that I have learned something about how to communicate it, even if I have not learned anything that has led me to change it.

References Ames, B.N. 1979. Identifying environmental chemicals causing mutations and cancer. Science 204: 587–593. Ames, B.N. 1983 Dietary carcinogens and anticarcinogens. Science 1256–1263 Ames, B.N. and L.S. Gold. 1998. Misconceptions about Environmental Pollution, Pesticides and the Causes of Cancer. NCPA Policy Report No 214. National Center for Policy Analysis, Dallas, TX Ames, B.N., R. Magaw, and L.S. Gold. 1987. Ranking possible carcinogenic hazards. Science 236: 271–280. APHIS (Animal and Plant Health Inspection Service, U.S. Department of Agriculture). 2006. What is the compliance history with APHIS’ biotechnology regulations? https://www.aphis.usda.gov/ brs/compliance9.html Accessed 6 April 2006. Bain, C., and T. Selfa. 2017. Non-GMO vs organic labels: Purity or process guarantees in a GMO contaminated landscape. Agriculture and Human Values 34: 805–818.

104

4 Food Safety and the Ethics of Consent

Belasco, W. 1997. Food, morality and social reform. In Morality and Health, ed. A.M. Brandt and P. Rozin, 185–200. New York: Routledge. Brill, W.J. 1985. Safety concerns and genetic engineering in agriculture. Science 227: 381–384. Brunk, C., and H. Coward (eds.). 2009. Acceptable Genes?: Religious Traditions and Genetically Modified Foods. Albany, NY: SUNY Press. Burkhardt, J. 2001. Agricultural biotechnology and the future benefits argument. Journal of Agricultural and Environmental Ethics 14: 135–145. Burkhardt, J. 2012. The ethics of food safety in the twenty-first century: Who keeps the public good? In The Philosophy of Food, ed. D. Kaplan, 140–160. Berkeley: University of California Press. Buzby, J.C., and P.D. Frenzen. 1999. Food safety and product liability. Food Policy 24: 637–651. Cianferoni, A., and J.M. Spergel. 2009. Food allergy: Review, classification and diagnosis. Allergology International 58: 457–466. Collingridge, D. 1980. The Social Control of Technology. New York: St. Martin’s Press. Comstock, G.L. 2000. Vexing nature?: On the ethical case against agricultural biotechnology. Dordrecht, NL: Springer. Cranor, C.F. 2003. How should society approach the real and potential risks posed by new technologies? Plant Physiology 133: 3–9. Crawford, F.E. 2002. Fit for its ordinary purpose: Tobacco, fast food, and the implied warranty of merchantability. Ohio State Law Journal 63: 1165–1224. Douglas, H. 2009. Is Science Value Free? Science. Policy and the Value Free Ideal: University of Pittsburg Press, Pittsburg. Douglas, M. 1966. Purity and Danger: An Analysis of the Concepts of Pollution and Taboo. New York: Praeger Press. Duke, S.O. 2018. The history and current status of glyphosate. Pest Management Science 74: 1027–1034. Early, R. 2002. Food ethics: A decision making tool for the food industry? International Journal of Food Science & Technology 37: 339–349. Frewer, L.J., C. Howard, and R. Shepherd. 1997. Public concerns in the united kingdom about general and specific applications of genetic engineering: Risk, benefit and ethics. Science, Technology and Human Values 22: 98–124. Hilbeck, A., R. Binimelis, N. Defarge, R. Steinbrecher, A. Székács, F. Wickson, M. Antoniou, P.A. Bereano, E.A. Clark, M. Hansen, E. Novotny, J. Heinemann, H. Meyer, V. Shiva and B. Wynne. 2015. No scientific consensus on GMO safety. Environmental Sciences Europe 27: Article 4. Accessed 12 Feb 2020 at https://enveurope.springeropen.com/articles/, https://doi.org/10.1186/ s12302-014-0034-1 Hoban, T.J., and P. Kendall. 1993. Consumer Attitudes about Food Biotechnology. Raleigh, NC: North Carolina Cooperative Extension Service. Jackson, Debra. 2000. Labeling products of biotechnology: Towards communication and consent. Journal of Agricultural and Environmental Ethics 12: 319–330. Jensen, K.K., and P. Sandøe. 2002. Food safety and ethics: The interplay between science and values. Journal of Agricultural and Environmental Ethics 15: 245–253. Kahneman, D. 2011. Thinking—Fast and Slow. Macmillan, New York: Farrar, Straus and Giroux. Kleter, G.A., H.A. Kuiper, and E.J. Kok. 2019. Gene-edited crops: Towards a harmonized safety assessment. Trends in Biotechnology 37: 443–447. Knorr, D. and K. Clancy. 1984. Safety aspects of processed foods. In Food Security in the United States, eds. L. Busch and W. Lacy. Boulder, CO: Westview Press. Korthals, M. 2001. Taking consumers seriously: Two concepts of consumer sovereignty. Journal of Agricultural and Environmental Ethics 14: 201–215. Korthals, M. 2004. Ethics of differences in risk perceptions and views on food safety. Food Protection Trends 24(7): 30–35.

References

105

Korthals, M. 2014. The food we eat: The right to be informed and the duty to inform. In The Right to Know and the Right Not to Know: Genetic Privacy and Responsibility, eds. R. Chadwick, M. Lefitt and D. Shickle, 196–208, Cambridge, UK: Cambridge University Press. Krimsky, S., and R. Wrubel. 1996. Agricultural Biotechnology and the Environment: Science. Policy and Social Issues: University of Illinois Press, Urbana, IL. Kuiper, H.A., H.P. Noteborn, E.J. Kok, and G.A. Kleter. 2002. Safety aspects of novel foods. Food Research International 35: 267–271. Kuzma, J. 2016. A missed opportunity for US biotechnology regulation. Science 353: 1211–1213. Latour, B. 1988. The Pasteurization of France. A. Sheridan and J. Law, tr. Harvard University Press, Cambridge, MA. Levine, J. 1986. Hearts and minds: The politics of diet and heart disease. In Consuming Fears: The Politics of Product Risks, ed. H.M. Sapolsky, 40–79. New York: Basic Books. Malyska, A., R. Bolla, and T. Twardowski. 2016. The role of public opinion in shaping trajectories of agricultural biotechnology. Trends in Biotechnology 34: 530–534. Mandel, G.N. 2006. Toward rational regulation of genetically modified food. Santa Clara Journal of International Law 4. Accessed 4 July 2020 at https://digitalcommons.law.scu.edu/scujil/vol4/ iss1/2/ Marchant, G.E., and G.A. Cardineau. 2013. The labeling debate in the United States. GM Crops & Food 4: 126–134. Marchant, G.E., and Y.A. Stevens. 2015. A new window of opportunity to reject process-based biotechnology regulation. GM Crops & Food 6: 233–242. Marden, E. 2002. Risk and regulation: US regulatory policy on genetically modified food and agriculture. Boston College Law Review 44: 733–789. Marks, J.H., and D.B. Thompson. 2011. Shifting the focus: Conflict of interest and the food industry. American Journal of Bioethics. 11: 44–46. Mather, R. 1995. A Garden of Unearthly Delights: Bioengineering and the Future of Food. New York: Dutton. McClements, D.J. 2019. Future Foods: How Modern Science is Transforming the Way We Eat. Cham, CH: Springer Nature Switzerland AG. McHughen, A. 2000. Pandora’s Picnic Basket: The Potential and Hazards of Genetically Modified Foods. New York: Oxford University Press. Mepham, B. 2011. Food additives: An ethical evaluation. British Medical Bulletin 99: 7–23. Mesnage, R. and M.N. Antoniou. 2017. Facts and fallacies in the debate on glyphosate toxicity. Frontiers in Public Health 5, Article 316. Accessed 12 Feb 2020 at https://www.frontiersin.org/ articles/, https://doi.org/10.3389/fpubh.2017.00316/full Miller, H.I., and G. Conko. 2001. Precaution without principle. Nature Biotechnology 19: 302–303. Millstone, E. 2007. Can food safety policy-making be both scientifically and democratically legitimated? If so, how? Journal of Agricultural and Environmental Ethics 20: 483–508. Morris, S.H., and C. Spillane. 2010. EU GM crop regulation: A road to resolution or a regulatory roundabout? European Journal of Risk Regulation 4: 359–369. Peck, A. 2017. Re-framing biotechnology regulation. Food and Drug Law Journal 72: 314–340. Pew Initiative on Food and Biotechnology. 2004. Americans’ Opinions About Genetically Modified Foods Remain Divided, But Majority Want A Strong Regulatory System(2). Accessed 13 Feb 2020 at: https://www.pewtrusts.org/en/about/news-room/press-releases-and-statem ents/2004/12/08/americans-opinions-about-genetically-modified-foods-remain-divided-but-maj ority-want-a-strong-regulatory-system2 Pielke, R.A., Jr. 2007. The Honest Broker: Making Sense of Science in Policy and Politics. New York: Cambridge University Press. Pinstrup-Andersen, P., and E. Schiøler. 2000. Seeds of Contention: World Hunger and the Global Controversy over GM Crops. Baltimore: Johns Hopkins University Press. Pomeranz, J.L., P. Wilde, D. Mozaffarian, and R. Micha. 2019. Mandating front-of-package food labels in the US–What are the First Amendment obstacles? Food Policy 86: 101722. Potrykus, I. 2001. Golden rice and beyond. Plant Physiology 125: 1157–1161.

106

4 Food Safety and the Ethics of Consent

Reiss, M.J., and R. Straughan. 1996. Improving Nature? The Science and Ethics of Genetic Engineering: Cambridge U Press, Cambridge, UK. Roff, R.J. 2009. No alternative? The politics and history of non-GMO certification. Agriculture and Human Values 26: 351–363. Roman, K.D., and M.A. Moore. 2012. A case study of crisis communication, image restoration and utilitarian ethics: A recall of contaminated peanut butter examined. Journal of Business Case Studies 8: 311–324. Ronald, P.C., and R.W. Adamchak. 2017. Tomorrow’s Table: Organic Farming, Genetics, and the Future of Food, 2nd ed. New York: Oxford University Press. Rozin, P. 1990. Social and moral aspects of food and eating. In The Legacy of Solomon Asch: Essays in Cognition and Social Psychology, ed. I. Rock, 97–110. New York: Laurence Erlbaum. Rubel, A., and R. Streiffer. 2005. Respecting the autonomy of European and American consumers: Defending positive labels on GM foods. Journal of Agricultural and Environmental Ethics 18: 75–84. Scallan, E., R.M. Hoekstra, F.J. Angulo, R.V. Tauxe, M.-A. Widdowson, S.L. Roy, J.L. Jones, and P.M. Griffin. 2011. Foodborne illness acquired in the United States—Major pathogens. Emerging Infectious Diseases 17: 7–15. Scott, N.D. 2018. Food. Springer, New York: Genetic Engineering and Philosophy of Technology. Sharma, K., and M. Paradakar. 2010. The melamine adulteration scandal. Food Security 2: 97–107. Shiva, V. 2003. Biotech Wars: Food Freedom versus Food Slavery. ZCommentaries June 25, 2003. Accessed 12 Feb 2020 at https://zcomm.org/zcommentary/biotech-wars-food-freedom-vs-foodslavery-by-vandana2-shiva-1/ Shiva, V. 2005. Terrorism, Agriculture and US India Cooperation. ZNET Daily Commentaries, August 10, 2005. Accessed 12 Feb 2020 at https://www.rightlivelihoodaward.org/fileadmin/ Files/PDF/Literature_Recipients/Shiva/Shiva_-_Terrorism__Agriculture_and_U.S_India_Coop eration.pdf Sperling, D. 2010. Food law, ethics, and food safety regulation: Roles, justifications, and expected limits. Journal of Agricultural and Environmental Ethics 23: 267–278. Strauss, D.M. 2018. Genetically modified organisms in food: Ethical tensions and the labeling initiative. In Ethical Tensions from New Technology: The Case of Agricultural Biotechnology, ed. H.S. James, 83–97. Boston: CABI International. Streiffer, R., and A. Rubel. 2004. Democratic principles and mandatory labeling of genetically engineered food. Public Affairs Quarterly 18: 223–248. Taylor, S.L. 2002. Protein allergenicity assessment of foods produced through agricultural biotechnology. Annual Review of Pharmacology and Toxicology 42: 99–112. Tenner, E. 1996. Why Things Bite Back: Technology and the Revenge of Unintended Consequences. New York: Alfred A Knopf. Thayer, A. 2002. USDA fines ProdiGene; Industry reacts to farm interests. Chemical and Engineering News 80 (50): 12. Thompson, P.B. 1987. Agricultural biotechnology and the rhetoric of risk: Some conceptual issues. the Environmental Professional 9: 316–326. Thompson, P.B. 1997. Science policy and moral purity: the case of animal biotechnology. Agriculture and Human Values 14: 11–27. Thompson, P.B. 2002. Why food biotechnology needs an opt out. In Engineering the Farm: Ethical and Social Aspects of Agricultural Biotechnology, ed. B. Bailey and M. Lappé, 27–44. Washington, DC: Island Press. Thompson, P.B. 2007. Shall we dine? Confronting the strange and horrifying story of GMOs in our food. In Food and Philosophy: Eat, ed. F. Allhoff and D. Mason, 208–220. Oxford, UK: Think and Be Merry, Basil Blackwell. Thompson, P.B. 2015. From Field to Fork: Food Ethics for Everyone. New York: Oxford University Press. Thompson, P.B. 2017. The Spirit of the Soil: Agriculture and Environmental Ethics, 2nd ed. New York: Routledge.

References

107

Thompson, P.B. 2019. Talking about Gene Edited Crops, in “Forum”. Issues in Science and Technology 35(2) https://issues.org/forum-41/ Touchstone Pictures. 1994. I Love Trouble. Burbank, CA: Touchstone Home Video. Waigmann, E., C. Paoletti, H. Davies, J. Perry, S. Kärenlampi, and H. Kuiper. 2012. Risk assessment of Genetically Modified Organisms (GMOs). EFSA Journal 10: s1008. https://doi.org/10.2903/ j.efsa.2012.s1008. Westerman, L.A. 2018. Consumer choice or confusion: That GMO label doesn’t mean what you think it means. Drake Journal of Agricultural Law 23: 199–228. Wozniak, C.A., A.F. Waggoner and C. Reilly. 2012. An introduction to agricultural biotechnology regulation in the United States, In Regulation of Agricultural Biotechnology: The United States and Canada, eds. C.A. Wozniak and A. McHughen, 1–14. New York: Springer. Wynne, B. 1992. Sheep farming after Chernobyl: a case study in communicating scientific information. In When Science Meets the Public Washington, ed. B.V. Lewenstein, 43–68. American Association for the Advancement of Science: DC. Zetterberg, C., and K.E. Björnberg. 2017. Time for a new EU regulatory framework for GM crops? Journal of Agricultural and Environmental Ethics 30: 325–347.

Chapter 5

Animal Health and Welfare

Abstract This chapter examines the ethical significance of gene technology on the health and well-being of livestock, poultry and any other animal species kept for agricultural purposes. Agricultural biotechnologies include drugs and feeds developed for use on livestock, as well as genetic transformations and cloning. Key applications are reviewed and examples are given. Bernard Rollin’s early work on this topic is summarized and used as a basis for further analysis. Philosophical alternatives to Rollin’s approach to understanding the basis of human obligations to other animals are discussed, including the welfarist approach of Peter Singer and the rights approach of Tom Regan. Though not a welfarist in general, Rollin argues that impact on the welfare of the transformed animal is the sole criterion for evaluating the ethics of genetically engineered animals. Additional literature on the ethics of using genetic engineering tools on animals is reviewed, with emphasis on views laying stress on the inherent wrongness of transforming an animal’s nature, irrespective of the impact on pain, suffering or disease. Although many arguments against any and all applications of animal biotechnology are philosophically flawed, they cannot simply be dismissed. Only a more extensive philosophical debate can clarify when a genetic change in an agricultural animal’s nature is inappropriate. Keywords Animal ethics · Animal welfare · Telos · Animal integrity · Transgenic animals · Animal natures · Intrinsic objections to genetic engineering Like any animal production technology, drugs or feeds derived from biotechnology can have adverse effects on animal health. Whether achieved through breeding or through transgenic methods, genetic modification of animals can result in dysfunctions severe enough to constitute cruelty (Broom 1995). Researchers at the U.S. Agricultural Research Service’s (ARS) Beltsville, Maryland station inserted the gene for human growth hormone into pig embryos in one of the early experiments to apply biotechnology to food animals. The animals experienced a painful arthritic condition that ultimately led researchers to terminate the experiment and to euthanize the pigs. Critics of food biotechnology were quick to seize upon these experiments

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_5

109

110

5 Animal Health and Welfare

as evidence for the unacceptability of genetic engineering in animals. (Fox 1992a; Kimbrell 1993). In fact, a considerable literature on the genetic transformation of non-human vertebrates had already accumulated well before the first genetically engineered field crops were grown in the late 1990s. Frequent critiques continued to appear in writings by philosophers, social scientists, journalists, animal activists and creative artists well into the 21st century. The position I endorse in this chapter has three main elements: • Genetic change certainly can cause suffering and harm animal health. Humans have prima facie duties to reject such uses of biotechnology. • The risk-based approach provides a framework for analyzing these prima facie duties. • Philosophers should be more cautious in linking claims about the intrinsic or metaphysical wrongness of genetic engineering to claims about animal health or welfare. Much of the chapter is taken up with this debate. I will argue that philosophers have searched without success for the appropriate terminology with which to express their moral umbrage over genetic engineering. While I sympathize with the sense of moral outrage expressed in some articles on genetic engineering of food animals, I believe these arguments are unhelpful to animals themselves. The United States Food and Drug Association (FDA) approved the first genetically engineered food animal, a variety of fast-growing salmon, in 2019. Thus, no actual exemplars of a commercialized transgenic food animal existed throughout a lengthy period of ethical critique that began in the 1980s. In the meantime, a series of scholars have entered the fray, usually writing with neglect (and probably ignorance) of the debate that has gone before. I have not attempted a thorough update of the more recent literature for the current edition, though I believe it is fair to say that no truly novel themes have emerged since the first edition was published in 1997. What is more, a review of this earlier literature testifies to the fact that issues were identified quite early. Scientists and research administrators (including those at funding agencies) had many opportunities to anticipate and more thoroughly investigate the ethics of genetically engineered food animals, yet arguably did not do so. At the same time, more recent scholars’ tendency to write as if they were the first people to think about these questions reinforces the research community’s tendency to regard this domain of the applied ethics literature as the carping of poorly informed dilettantes. I hope to counteract that tendency by helping younger scholars navigate the 20th century literature.

5.1 Animal Biotechnology and Food

111

5.1 Animal Biotechnology and Food Animal biotechnology is the application of recombinant DNA techniques to animals. Genetic engineering (including gene editing) and cloning are the two leading forms of animal biotechnology. As with plants, genetic engineering involves the introduction of genetic constructs into the DNA of an animal, or the use of genetic tools to alter the regulation of already existing genes. In the case of gene editing, this may involve removing a gene. As with plants, the potential for such introductions does not appear to be limited in principle by the source (e.g. the species) of the transgene. For cloning, there is no direct analog with plants because asexual reproduction of plant tissue is a routine process, practiced by home gardeners who work with plant cuttings. During the early 1990s animal cloning was discussed frequently alongside genetic engineering as a promising application for agriculture, but the technique being discussed was the physical separation or splitting of at the blastocyst stage. Both halves of a split embryo (indeed more splits can be made up to a limit of about six) have the potential to develop into genetically identical animals, or clones. Animal cloning took on a different meaning and attained great significance in the public mind after the 1997 announcement that researchers at the Roslyn Institute had produced the sheep “Dolly”, (see Thompson 1999). In theory, both genetic engineering and cloning can be applied to animals of virtually any species. Most of the past interest in animal biotechnology has focused on vertebrate species, though gene drives—genetic constructs that, once released, move through a population quickly—are currently being developed to transform arthropods. Among vertebrate species, much of, (if not most of), the genetic engineering and cloning research involves rodents. This research aims either to achieve basic advances in genetics and genetic manipulation, or to further applications of relevance to human medicine and public health, and can be classified to as medical biotechnology. Bernard Rollin’s discussion of animal biotechnology in The Frankenstein Syndrome was focused equally on medical and agricultural applications. This chapter limits the scope of discussion to genetic engineering of animals intended for food and fiber production, bypassing the issue that Rollin identified as the most difficult ethical dilemma in agricultural biotechnology: Medical researchers might transform animals so that they exhibit the symptoms of painful genetic disease in order to search for possible cures, (Rollin 1995). There are a few cases of genetically engineered animals that might be thought to trouble the distinction between agriculture and medical research. For example, researchers propose to genetically engineer animals as bioreactors that produce some commercially important substance in their blood, milk or bodily tissues, (see Houdebine 2000). Researchers might also use genetic engineering to transform animals so that their organs could be transplanted into humans without triggering immunological rejection. The idea of animals as organ donors has long been a dream for biomedical research, and new hope for it lives in the era of gene editing, (Belmonte 2016). These biomedical applications of gene transfer could involve livestock species. For example, gene transfer and cloning were used to develop a small herd of cattle

112

5 Animal Health and Welfare

producing human immunoglobulin G, which is crucial for safe and effective blood transfusions, (Robl 2007). This was one of the first commercially viable bioreactors. Pigs are the primary target for developing transplantable organs owing to their physiological similarity to human beings, (Denner 2017). Medical bioethicists continue to pursue the risks and other ethical issues associated with these questions, and they will not be taken up here, (see Bourret et al. 2016). The overwhelming majority of people see no ethical problems with using animals for food, but as Rollin notes, people in industrialized countries are becoming increasingly less tolerant of animal production practices that subject animals to pain, suffering, fear and stress. Early public opinion studies that attempted to rank the degree of ethical concern associated with various forms of biotechnology reported a result that surprised many: a greater percentage of respondents reported ethical concerns in connection with animal genetic engineering than with respect to genetic engineering of human beings (Hoban and Kendall 1993). Although subsequent polling has not replicated the methods of this early study in a manner that would permit direct comparison, high levels of public concern with both genetic engineering and animal cloning have continued to be supported by more recent polls (Pew Initiative 2004; Van Eenennaam and Young 2018). It is reasonable to surmise that part of the reason people found animal biotechnology morally problematic in the early poll was associated with current and future biomedical uses of animals. However, more recent studies provide ample evidence for the suggestion that people wonder whether the genetic transformation of food animals has the potential to create new or exacerbate current food animal production practices that are ethically questionable, if not clearly unacceptable. The plight of the food animal is at least one of the main bases for this concern.

5.2 Harm to Animals and the Risk-Based Approach The risk assessment paradigm stipulates that unintended harms and other consequences are specified through an inductive process of hazard identification. There is nothing in the paradigm that limits the type of harm to human beings. Bernard Rollin, the leading philosophical analyst of animal biotechnology, writes, “opponents of genetic engineering of animals are right to fear that such engineering will proliferate animal suffering, though they are wrong in thinking that it must do so,” (Rollin 1995, 181). In this instance, the root issues concern the basis of our moral concern for animals, and the nature of our obligations to them. These deep and important philosophical issues are intertwined with some of the most complex, pervasive and enduring philosophical matters: the nature of morality itself; the nature of consciousness; and the relationship between human spirituality and the material world. It is important to inquire into these matters, but it is unreasonable to think they can be settled to the general satisfaction of all interested inquirers. What is more, for many of us, such questions will trouble our own conscience.

5.2 Harm to Animals and the Risk-Based Approach

113

Rollin’s describes his view of the root issues as “the consensus social ethic for animals,” and doubtless there is wide agreement that “the plight of the animal,” to use Rollin’s phrase, must be part of any ethical evaluation of the genetic engineering of food animals. The effects of genetic engineering on animal health were extremely controversial during early phases of public debate on biotechnology, in part because rBST raised a constellation of problems documented in Chap. 3. These included social consequences, labeling and food safety, as well as animal health. When products from crop biotechnology appeared on grocery shelves in the late 1990’s, the debate over animal biotechnology had shifted to cloning. Dolly, the first successfully cloned mammal, was announced in 1997. By the time of Dolly’s publicized death in 2004, philosophers writing on animal biotechnology regarded it largely as a medical technology. Despite many efforts, neither gene transfer nor cloning of food and fiber animals have emerged as important technologies in the food system, though with the approval of fast growing salmon, this may change in the future. As with other issues discussed in the book, the ethical issues are tangled up with regulatory policy. As the Beltsville pigs faded from memory, Rollin’s concern for “the plight of the creature,” did not become a major issue in the regulation of food biotechnology. Yet the approval of the Aqua Bounty™ salmon and the emergence of gene editing (discussed at more length in Chap. 13) are bringing the old ethics debate back into relevance. Gene edits, in particular, may prove to be highly significant for livestock, as this more precise form of altering genes can be used to develop animals with improved resistance to disease. Other modifications, such as altering the nutritional components of meat or other animal products, may follow, (Tan et al. 2016). Some practitioners of animal biotechnology have criticized the current premarket regulatory approval process, which requires innovators to submit data on the genetic modification’s impact on animal health. Generating such data is expensive and builds delay into the process of regulatory review. They argue that gene edited livestock should be treated more like genetically engineered crops, (Van Eenennaam et al. 2019) The problem with this argument is that, so far as we know, plants do not feel pain or experience suffering, (see Shriver 2020). Biotechnology can affect how animals fare (the plight of the creature) in two distinctive ways. First, animal drugs (such as rBST) and possible feeds and feed additives can be produced using biotechnology. Like drugs or feeds produced through conventional means, these products have the potential to affect animal health and nutrition. Second, a direct change in animal genomes can be effected by genetic engineering, including gene editing. By altering genomes, biotechnology is capable of making significant modifications of animal phenotypes, hence it is not surprising that such changes can have attendant effects upon animal welfare. In the U.S. regulatory system, oversight for veterinary medices resides at the Food and Drug Administration (FDA). FDA applies its authority to oversee technologies that affect an animal’s metabolic processes (e.g. drugs) to changes in an animal’s genome. The agency requires that innovators supply data on efficacy (does the product do what is claimed), human health (would consumption of meat, milk or other animal products cause harm to humans, and animal health. This does not imply that animal health is an

114

5 Animal Health and Welfare

overriding concern, but it certainly creates a regulatory requirement that innovators find inconvenient.

5.3 Drugs and Animal Feeds from Biotechnology Pharmaceuticals represent one of the largest and most lucrative uses for biotechnology to date. Human drug development is covered by standard research ethics (Resnik 2015), though even here I find few examples that single out the use of recombinant tools over the last twenty years. Animal drug development has received virtually no attention from either scholars or activists over the last decade. As such, the case of rBST continues to be the best illustrative example. rBST generated a great deal of reaction from advocates of animal welfare. Their criticism took a sophisticated philosophical form quite early in the debate and continued long past 1992 when rBST was approved in the United States. Gary Comstock published a detailed description of animal welfare impacts both from the topical administration of rBST and from increased susceptibility to stress-related bovine diseases such as mastitis (Comstock 1988). Sheldon Krimsky and Roger Wrubel summarized the controversy over rBST and animal health in their 1996 book, giving prominence to the opinions of David Kronfeld, who cites a litany of pathologic changes in cows associated with the use of rBST (Krimsky and Wrubel 1996, pp. 176–179; see also Kronfeld 1993). If such allegations are true, how is it that rBST was approved for use in the United States? The answer to this question lies partly in the technical literature on rBST and animal health and partly in the way that U.S. regulators applied values to their legal mandate at the time that rBST was reviewed. Dale Moore and Lawrence Hutchinson summarize a large technical literature on rBST and animal health with the conclusion, “When animal-health effects have been documented in BST studies, they have generally been shown to be secondary to increased milk production, indicating the importance of excellent nutrition and management if BST is used to enhance production,” (Moore and Hutchinson 1992, 122). The boosters who defended rBST argued that rBST increases milk production. They did not deny that increasing milk production is linked to detrimental impact on animal health. That is, high producing animals tend to have health problems, though these problems can be minimized with “excellent nutrition and management.” Administering a dose of rBST puts an animal that might not otherwise be a high milk producer into the high producing group. Once in that group, they tend to exhibit the health problems of high producing animals. This point, which is of critical significance in the analyses of Comstock and Kronfeld, is not disputed by the defenders of rBST. But does rBST cause problems for animal health? Here we must parse the causal claims carefully. It seems clear that rBST causes an increase in milk production. Furthermore, it seems clear that something in the physiology of high producing dairy cattle causes a susceptibility to the so-called production diseases (such as mastitis) of concern to Comstock and Kronfeld. Here we have a case where X causes Y and Y causes Z. Z is production disease, a class of outcomes

5.3 Drugs and Animal Feeds from Biotechnology

115

of clear significance with respect to animal welfare. Y is increased milk production, not necessarily of moral significance and X, of course, is rBST. The critics of rBST seem to be saying that the causal chain is transitive: since Y causes Z and X causes Y, then X is a contributing cause of Z. This is the reasoning that appears to have persuaded a number of governments to ban rBST. However, the argument in defense of rBST offers a different interpretation of the causal chain. Although this argument is seldom made explicitly, it might go something like this: • Since there are other ways of increasing milk production (such as feed regimens or conventional breeding) that are legal, it would be prejudicial to ban rBST. • Furthermore, there are ways to control the incidence of disease through careful management, and to treat resulting diseases using standard veterinary approaches. • Therefore, no animal health affects (of regulatory significance) are attributable to rBST. This argument does not dispute the causal claims (e.g. X causes Y and Y causes Z), but adds some normatively significant information, namely that there are other possible causes of Y (conventional breeding and feed regimens). Furthermore, these causes (call them XCB and XFR ) are legally permitted. The argument now takes a different form. Either XCB or XFR or our original X (e.g. rBST) causes Y. Y causes Z, and Z (the animal health effect) is the basis for normative action. Although rBST, XCB and XFR are sufficient conditions for Y, it is clear than none is a necessary condition. FDA appears to have interpreted causality as implying both necessary and sufficient conditions for the actionable outcome Z. However, it is far from clear that the argument holds up if ‘ethical’ is substituted for ‘regulatory’. Animal producers may have the legal right to try and increase milk production through manipulating feed regimens or genetics, but it does not follow that they are morally justified in either activity if doing so places their animals at substantially increased risk from production diseases. The proper moral conclusion may well be to rethink the entire complex of productivity enhancing technologies. However, neither XCB nor XFR are animal drugs. FDA has no regulatory authority over XCB and XFR . Regulators reviewing rBST did not have the option of rethinking the entire complex of technical options being used in the dairy industry. Regulators must make value judgments that are consistent with their legal mandate, but it is still not clear that regulators at FDA were forced to approve rBST. A regulatory policy guided by the norm of protecting animal health would not feel compelled to approve a practice that increases the burden of disease simply because there are other factors that will continue to cause the disease. However, it appears that FDA regulators were influenced by a different norm, a version of “equality before the law.” If the agency were reviewing three animal drugs, the norm of treating applicants equally would certainly require that if two approved agents contribute indirectly to an decrease in animal health, it would be unfair to ban a third applicant on an identical ground. The best face that can be put on FDA’s action is that they believed XCB and XFR had established a baseline for mastitis risk. As such, a regulatory ban on rBST would have been at best arbitrary and possibly discriminatory.

116

5 Animal Health and Welfare

It may be naively idealistic to think that scientists, animal drug or feed companies or producers themselves will alter their behavior in accord with such ethical considerations. Dairy and meat industries alike are extremely competitive, so one does not deny oneself advantages that less scrupulous competitors are free to exploit. Nevertheless, the example of rBST illustrates the kind of ethical issues that can arise in conjunction with animal drugs, and they are quite unlike anything discussed in human medicine. Many dairy cooperatives in the United States have agreed to ban their members from using rBST. They believe that being able to claim that their milk is “BST free” makes it more attractive to consumers, and with an oversupply of milk, there are few incentives to increase production, (Wolf, Tonsor and Olynk 2011).

5.4 Engineered Animals A genetically engineered animal is any animal whose genome has been modified using rDNA methods, including microinjection of gene sequences or gene editing. Animals with genomes that include sequences derived from other species are transgenic. Animals whose genome is edited to delete constructs tied to congenital disease are engineered, but not transgenic. Part of the rationale for weakening the standards for regulatory approval is that non-transgenic gene edited animals have no novel proteins that could affect metabolic function, (Van Eenennaam 2018). However, putatively non-transgenic hornless cows were found to be carrying sequences unknown to the group that edited their genes, (Regalado 2019). Furthermore, as discussed in Chap. 14, the process of gene editing causes off-target alterations in plant and animal genomes. There does not appear to be a good reason to neglect hazards to animal health and welfare in a risk-based evaluation of genetically transformed livestock. The transgenic/non-transgenic distinction is of little significance at this juncture. At the same time, many experimental modifications of animal genomes appear to have had little impact on animal health or cognitive stress, leading Ian Wilmut to note that, with the exception of the Beltsville pigs, “the effects of genetic change on animal welfare are usually trivial,” (Wilmut 1995, 241). Rollin begins his less optimistic discussion of “the plight of the creature” with a history of attitudes toward the moral status of animals up to the present day consensus ethic that is the basis for his philosophical position. Rollin’s eighty page chapter takes up three specific ethical issues associated with genetic engineering of animals: the welfare of agricultural animals, the engineering of animal models for human disease, and ethical issues in the patenting of animals (Rollin 1995, 137–218). The issues of patenting and intellectual property are discussed in Chap. 10. As already noted, ethical issues involving the use of animals in medical research fall beyond the purview of a book on agrifood biotechnology. This leaves Rollin with a relatively short (seven pages) discussion of the actual effects of biotechnology on food animals. The section begins with description of the Beltsville pig experiments described above, along with sheep experiments also done by ARS, and a third experiment on cattle (pp. 188–189). In each case, dysfunctional

5.4 Engineered Animals

117

results led to disease and indisputable suffering on the part of the animals. More recent reviews of recombinantly transformed animals report that while relatively little data on the welfare of engineered animals is available, observable welfare deficits (such as those mentioned by Rollin) are not common, (Ormandy et al. 2011). However, it should be noted that this generalization applies to biotechnologically transformed breeds that have been stabilized, rather than individuals produced in the research process. The results cited by Rollin do demonstrate the potential for harmful impacts on welfare, and more recent studies do not controvert this possibility. Based on such results Rollin stipulates norms for research and commercial production of genetically engineered animals. Rollin argues that institutional review boards (IRB’s) “should demand that, in pilot research on agricultural animals, a small number of animals be used and that early end points for euthanasia of animals be established in advance and implemented at the first sign of suffering or problems that lead to suffering, unless such suffering or disease can be medically managed,” (Rollin 1995, p. 189). For commercial production, Rollin notes that genetic engineering may make animal suffering far more profitable than it currently is. One reason is that dysfunctional animals might prove useful as producers of valuable products, such as drugs or biologics. Another is that genetic engineering may be used to make animals more tolerant of cold or dehydration, with unknown effects on animal suffering. (Rollin 1995; p. 192; see also Fox 1992b, p. 217). In response to these possibilities, Rollin calls for applying a principle of “conservation of welfare” to commercial food animal production: no genetic engineering will be permitted that makes the animal worse off than a non-genetically engineered animal in comparable circumstances, (Rollin 1995, p. 179). Rollin’s reliance on the conservation of welfare principle leads to some surprising ethical conclusions. For example, “if we could genetically engineer essentially decerebrate food animals, animals that have merely a vegetative life but no experiences, I believe it would be better to do this than to put conscious beings into environments in which they are miserable, though again this seems aesthetically abhorrent to us,” (Rollin 1995, 193). By decerebrate, Rollin means an animal that is genetically incapable of feeling pain. Such an animal would be more like an oyster or an insect than normal sentient vertebrate such as a cow, a pig or a chicken (though we cannot be totally sure that oysters and insects are incapable of feeling pain). Rollin’s claim is that animals incapable of experiencing pain or frustration would be better off than sentient members of livestock species living in conditions that are commonplace in concentrated animal feeding operations (CAFOs), today. The potential for improving animal welfare by reducing an animal’s ability to experience suffering has generated a significant literature by philosophers. Bernice Bovenkirk, Frans Brom and Babs van den Bergh’s seminal paper links the morality of altering an animal’s genetic endowment to the dystopian vision of Brave New World by Aldous Huxley (1894–1963). They offer ‘animal integrity’ (discussed below) as a grounding concept for this moral judgment. Peter Sandøe showed that the philosophical problems do not depend on biotechnology. Congenitally blind hens produced through conventional breeding might be less troubled by the crowded conditions that exist in industrial egg CAFOs, yet we would not find this to be a morally appropriate solution to those conditions, (Sandøe et al. 1999; Gamborg and Sandøe 2002). I

118

5 Animal Health and Welfare

connected the arguments of Rollin and Sandøe in a 2008 paper in which I describe the genetic strategies for improving welfare as “disenhancement”, arguing that intuitions about the ethical justifiability of doing this to animals were relevant to the ongoing discussions on using genetics to enhance valued traits in humans, (Thompson 2008). Disenhanced animals also figure prominently in the dystopian fiction of Margaret Atwood, (Atwood 2003). There is now an explosive literature on the philosophical puzzles associated with animal disenhancement. A full discussion of this literature would require much more than a simple revision of the 2007 version of this chapter, and readers have other avenues for satisfying their curiosity on the debate. Candace Croney (writing with four co-authors) published an overview discussion that I will not attempt to replicate here, (Croney et al. 2018). Two summary points are relevant. As Clare Palmer argues, the philosophical dilemma lies in the fact that the moral ontology of disenhancement differs from the case where someone harms an individual cow, pig or chicken by disabling a capacity. In blinding an animal, and arguably with non-genetic mutilations like tail-docking or beak-trimming, disablement is a harm above and beyond any pain the animal might experience as a result of the procedure. In the case of genetic disenhancement, not only is there no pain, but the resulting individual never had the capability to begin with. It is difficult to see why anyone would say that this individual is harmed, because an animal having the capacity would be a different individual. Palmer sees this a special case of the non-identity problem, (Palmer 2011). Derek Parfit (1942– 2017) described this class of problems in ethical theory. In Parfit’s work on future generations, the problem arises when the person affected by an action would not have been born if the action had not been taken. Palmer is saying that the disenchanted animal would not have been born but for the genetic change. Nevertheless, other philosophers have no trouble saying that the animal is harmed. Ariana Ferrari (2012) and Adam Henschke (2012) each argue that since harm is pervasive in the larger context of industrial animal production, even suggesting that individual animals are not harmed is a form dissembling. Their claim that Palmer and I fail to acknowledge the reality of existing practice notwithstanding, both Ferrari and Henschke analogize acts that clearly do affect actually existing individual animals and those that cause very different individuals to be born. Korinn Murphy and William Kabasenche extend the analogy further, suggesting that the entire project of industrial animal production is one of domination, (Murphy and Kabasenche 2018). As such, the more comprehensive harm of industrial animal production covers or comprehends acts that reduce capacity through genetic change, as well as by directly affecting a living animal. Puzzles of the non-identity problem to the side, I am not sure I understand how these claims rebut my own analysis. Both Rollin and I had claimed that disenhancement would be rejected on relational grounds. People will not stand for it because it disrupts the way that people think of themselves and their relationship to the animals they eat. However, our point, (along with Palmer) was that these arguments also direct attention away from the pain and suffering of living animals, whether or not they have been genetically modified. What is more, these analyses are so broadly framed that they do not, in fact, speak to the transformation of animals

5.4 Engineered Animals

119

through gene technology, at all. The mere fact that Ferrari and Henschke use analogies to non-genetic technologies in animal production testifies to this fact. The shock value of calling out routine practices in industrial animal production engages moral emotions, and appropriately so. Ferrari and Henschke are mobilizing this emotional engagement in opposition to genetic modification. However, genetic disenhancement produces a different kind of shock, and while Rollin, Palmer and I are simply calling attention to the difference, other authors have written that gene technology should, in fact, be used. As discussed below, Adam Shriver has become the leading advocate for disenhancement, adapting Peter Singer’s utilitarian view to the issue, (Shriver 2009; Shriver and McConnachie 2018). Daniel Wawrzyniak argues that while cognitive disenhancement is unacceptable for the reasons that Rollin and I state, there my nonetheless opportunities to apply gene technology in making positive improvements in animal lives, (Wawrzyniak 2020). The shock value of these technologies takes us away from thinking closely about what we owe to farmed animals, or what would benefit and harm them. The shock of genetic disenhancement raises ethically legitimate issues, but they concern conceptually distinct metaphysical questions about the nature of morality and the human condition, (see also Weisberg 2015). As discussed later in the chapter, the early debate on animal transformation revolved around the alteration of an animal’s telos, a concept introduced by Rollin, but reinterpreted by those who wanted to mount a categorical argument against animal biotechnology. Rollin’s 1995 discussion of the principle of conservation of welfare endorsed disenhancement of research animals, and appears to support the same for farmed animals. He has qualified his support for the practice in subsequent publications, emphasizing the moral character of those who would contemplate such a practice, rather than its impact on the animal, (Rollin 1998, 2003). In summary, genetic engineering of food animals can lead to at least four kinds of impact on the lives of animals: 1. Inadvertent and unwanted dysfunctional states. Such impacts would presumably occur in laboratory settings, and would call for euthanasia of the experimental animals. 2. Unwanted but anticipated dysfunctions that are accepted by commercial producers because of the commercial value of the affected animals. Here the use of genetic engineering would be comparable to other production technologies (such as CAFOs) that are detrimental to animals, but accepted because they are profitable for producers. 3. Intentional transformations that cause uncertain and controversial changes in the quality of animal experience. 4. No welfare impact at all, from the animal’s point of view. The risk-based approach applies criteria to each of these categories that define hazards in terms of harm to the animals’ veterinary health (morbidity, mortality, metabolic function) as well as to pain and suffering. Rollin has given us an application of what he takes to be the consensus social ethic with respect to each of these, but people taking a different ethical stance might disagree. In fact, the disagreements proceed in opposite directions. While the critics discussed above think that this

120

5 Animal Health and Welfare

characterization of hazards is to narrow, those who would deny any moral status to non-human animals believe that it is too broad.

5.5 The Moral Status of Animals A half century ago, debate over the moral status of non-human animals and the ethical significance of human use of these animals was infrequent. Ruth Harrison (1920– 2000) published Animal Machines in 1964 precipitating five decades of intense debate, and a proliferation of philosophical treatments of the moral status of animals, (Harrison 1964). The philosophical writings of Peter Singer and Tom Regan (1938– 2017) shaped the main strands in the animal rights movement. Singer and Regan are still the most frequently cited philosophers on animal ethics, but a plethora of modifications and alternative approaches have appeared on the scene since 1997 when this book originally appeared. Perhaps too succinctly stated, Singer provides the rationale for attending to animal welfare, while Regan argues for animal rights. Animal welfare (sometimes rendered animal well-being) includes the veterinary health of an animal (e.g. disease, physiological dysfunction, morbidity, mortality) as well as conscious or affective states that an animal experiences as satisfying or troubling (e.g. pain, anxiety, fear, comfort, satiation). Prior to the 1970s, the view that animals simply do not have conscious experiences (e.g. that they are not sentient) was not uncommon among biomedical researchers, including veterinarians. The origins of this view are often traced to the early modern philosopher René Descartes, though it is doubtful that this was Descartes’ actual view, (see Harrison 1992). Singer attacked this view in several high-profile venues (Singer 1973, 1975), while Rollin provided a detailed history and scientific rebuttal of it in his book The Unheeded Cry, (Rollin 1989). Singer’s emphasis was on the ethical implications that follow from recognizing that many animals (certainly all vertebrates) are sentient. One cannot, on pain of inconsistency, fail to consider an animal’s suffering when evaluating the ethical justifiability of any practice that affects animals. Both Singer and Rollin accept the possibility that animal suffering can be offset by compelling benefits in some cases, though the philosophical argumentation they use to reach this conclusion differs significantly, (Singer 1979; Rollin 1981). This is the point on which Regan holds a different view. Regan argues that the trade-off logic sanctioned by Singer is philosophically untenable because it would sanction many actions (including invasive medical research on children incapable of giving consent) that we rightly take to be morally indefensible, (Regan 1986, 2003). Regan argues that deciding the fate of another by weighing the goods and evils that would accrue if a particular course of action is pursued is to treat them merely as the means to an end. Using an argument that mimics the logic (but not the deeper views on animals) of Immanuel Kant (1724–1804), Regan claims that this violates the most basic principle of morality. While Singer, Rollin and Regan agree that animals from all livestock species are sentient, Regan makes the further claim that they are subjects-of -a-life,

5.5 The Moral Status of Animals

121

(Regan 1983). That is, they possess the kind of consciousness that makes it wrong to put their suffering up for grabs in a trade-off optimizing calculation. Earlier editions of this book included sections that summarize many details in the debate between Singer and Regan, as well as critiques of their views offered by R.G. Frey (1941–2012) and Carl Cohen (see Frey 1980; Cohen and Regan 2001). At the dawn of third decade in the 21st century, the philosophical debates on the moral standing of animals have expanded to include dozens more relevant positions, but the Singer/Regan debate illustrates a point that is key for analyzing genetic modification: There are different approaches to deciding on the ethical justifiability of applying recombinant gene transfer or gene editing to sentient animals. For Singer, the potential of unwanted outcomes in experimental settings provides a rationale for dramatically curtailing much of the animal research that is currently done, and this would presumably include genetic engineering. However, it might still be possible to rationalize some gene transfer work if a risk assessment included the potential for animal suffering in its calculation of hazards. In recent publications, Adam Shriver has argued that genetic modification that does reduce animal suffering would be justified on the grounds that Singer puts forward, (Shriver 2009). Singer believes that his views support vegetarianism for most people, though he does allow exceptions for people who would endure serious hardships from following a vegetarian diet. It is highly doubtful that he would support farming practices that add to animal suffering and virtually certain that he would not find farmer profitability to be an outcome that could offset a detriment to animal welfare. Singer does not appear to have published on the disenhancement problem, though the most straightforward way of applying his utilitarian reasoning would suggest that reduction of suffering in farmed animals would be a good thing. Whether it would be justified will depend on the total package of benefits and harms (including any residual suffering). A totally decerebrate organism producing the meat, milk and eggs humans like to eat would almost certainly find support from Singer’s approach, (see, for example, Schaefer and Savulescu 2014). The “no effect” outcome would have no impact on whether Singer would approve of a practice on moral grounds. If the modification has no impact on animal welfare, the moral evaluation of the transformation would depend entirely on whether its benefits to humans (or other sentient creatures) were significant enough to offset animal welfare deficits that derive from other sources (such as the animal’s husbandry or housing). In Regan’s view, an animal rights position requires moral vegetarianism. It is right to take an animal’s life for human food only in life threatening circumstances. That applies to few of us, and even then, rarely. This makes much of food animal agriculture thoroughly beside the point, and there is, in fact, virtually no discussion of how animals fare in differing production settings in Regan’s 1983 book. It might still be meaningful to ask whether genetic engineering could be performed on dairy cows or laying hens without violating their rights, but the spirit (if not the letter) of animal rights views would appear to preclude any genetic manipulation that was not (as with human genetic manipulation) intended solely for the benefit of the animal itself. Unfortunately, Regan was less than forthcoming with respect to this question even in a publication nominally devoted to animal biotechnology (see Regan 1995).

122

5 Animal Health and Welfare

However, it would certainly appear that performing gene transfer or knockouts on any animal treats the animal as a means to some further end. Steve Sapontzis develops an animal rights analysis of animal biotechnology that makes the point clear: Overcoming our species prejudice and creating a world in which we treat those who are powerless against us sympathetically and fairly is what “animal rights” is about. … So, if animals do not yet have rights, it is not due to an inadequacy on their part but to a failure on ours—our failure to be fully moral agents. Overcoming our instinctive human chauvinism to adopt an animal-respecting moral perspective is needed to erase that failure. Part of that transition would be acknowledging that the generic [sic] identity of animals is not a resource to manipulate for human taste, profit, curiosity or health without respect for the well-being of the animals themselves (Sapontzis 1991, 184–5).

Sapontizis’ emphasis on the genetic identity of animals anticipates the discussion of animal telos, below. Critics of Singer and Regan, including both Frey and Cohen, defend versions of the view that human interests exist on an entirely different moral plane from the sentient experiences of brute animals. Very few philosophers have discussed whether this includes the interest that animal producers have in continuing to farm. Charles Blatz has defended a related view, however, arguing in more classically Kantian language that the difference between humans and livestock species derives from the fact that language and rational thought give human beings a capacity for autonomous choice. Blatz applied this view in an argument intended to show the permissibility of genetically engineering food animals for the kind of production traits that would be economically valuable (Blatz 1991). Non-philosophers have also attempted arguments to stake out a similar position. In an early public meeting called to debate the ethics of animal biotechnology, David Meeker, then of the U.S. National Pork Producers’ Council, argued that while animal welfare counts, even relatively trivial human interests (his example was testing his teenage daughter’s make-up on laboratory animals) override animal welfare concerns. (Meeker 1992). Unlike Meeker, Frey, Cohen and Blatz agree that many current human uses of animals are unacceptable. Genetic engineering is acceptable to Blatz because the “immiseration” (his word) of pigs does them no harm in itself. However, when harm is done for no overriding human purpose (in Blatz’s view the creation of suffering animals like the Beltsville pigs is an example) genetic research cannot be considered part of an ethically defensible project. Blatz writes, “When the best we can say about an endeavor is that accidentally it might pay off in and ethically compelling way, while at the same time that endeavor is expected … to involve costs which we should avoid (other things being equal), then we should not engage in that endeavor,” (Blatz 1991, 173). Presumably, Blatz would find the desire to wear make-up a less than ethically compelling endeavor, as well.

5.6 Rollin’s Consensus Morality

123

5.6 Rollin’s Consensus Morality This survey of approaches to the root problem in animal biotechnology can be summarized as follows: Singer’s utilitarianism accords with Rollin’s principle of conservation of welfare, and not surprisingly, given the welfare orientation of sentience utilitarianism. An engineered animal that is no worse off from a welfare standpoint is no more or less acceptable (on ethical grounds) than the non-engineered comparator. Shriver has also supported Rollin’s principle of welfare conservation in a recent article, arguing that it is relevant for gene-edited livestock, (Shriver 2020). A strong rights view proscribes the use of animals for food, so food animal biotechnology becomes moot, at least for Regan’s version of animal rights. Blatz’s neo-Kantian criteria for judging an ethical project to be of “overriding importance,” would likely find experiments or commercial release of genetically engineered animals acceptable so long as they, too, were consistent with Rollin’s principle. Furthermore, since Blatz does not find immiseration of animals problematic, he has left little room for objecting to animal disenhancement. Blatz’s view leaves the window open for commercial applications that are detrimental to animal welfare, so long as they are part of an ethical project of overriding importance. Relieving hunger might constitute such a project, but the trade-off analysis that would support such an inference will be complex and hotly contested. This means that Blatz’s view, like Singer’s, collapses into a prescription that is wholly consistent with Rollin’s principle of the conservation of welfare, despite the dramatic philosophical differences between Singer’s utilitarian and Blatz’s neo-Kantian views. It is possible that this is exactly what Rollin means when he talks about a “consensus social ethic,” though in many contexts he seems to be referring to something more like “received public opinion.” Excepting the most extreme philosophical positions, authors beginning from different starting points converge on norms for the use of animals that would probably be shared by the majority of people, even if they have given little thought to the problem. On the face of it, Rollin seems to be describing a kind of moral conventionalism here: what is ethical is what we agree on. Yet Rollin’s earlier work on animal rights (Rollin 1981) belies that interpretation. There Rollin develops a position that, like Frey’s and Blatz’s, relies heavily on an analysis of interests and agency. His analysis results in the view that animals have a moral right to life—not an absolute right, but one that demands careful analysis and justification whenever it is abridged (p 49). The 1981 book was also where Rollin first used the notion of telos to flesh out the content of our obligations to animals. Borrowed from Aristotle (385 BCE–322 BCE), an animal’s telos is “a nature, a function, a set of activities intrinsic to it, evolutionarily determined and genetically imprinted,” (Rollin 1981, p 39). As developed in Animal Rights and Human Morality, Rollins view was that we should think in terms of animal rights because: • rights direct us toward proper respect for the interests of individual animals, as distinct from utilitarian approaches that aggregate welfare;

124

5 Animal Health and Welfare

• the term “rights,” conveys the seriousness with which we should deliberate in choosing actions that are contrary to animals’ interests; and • the laws needed to protect animals’ interests would establish rights that could be claimed by advocates on animals’ behalf. None of this offers much in the way of advice as to what sort of consideration is owed to any specific animal. Unlike the view developed by Regan, Rollins’ version of animals’ rights are quite unlike human rights, and this is so because animals’ telos are quite different from the purposes and ends that we associate with human nature. Telos specifies the content of animal rights and helps us come to terms with our respective duties toward them. This philosophical position remains consistent with the view that Rollin espouses in The Frankenstein Syndrome. Although Rollin has continued to refer both to rights and to telos, it is questionable whether he would emphasize these terms had he the opportunity to begin anew. They have led to widespread misunderstanding. Rollin’s commitment to rights as the fundamental moral notion is far weaker than Regan’s, for example, and though Rollin is clearly influenced by Aristotle (385 BCE-323 BCE), he recommends the notion of telos more as a heuristic for considering animal needs and interests than as a naturalistic foundation for morality. Given this orientation, Rollin can find nothing wrong with using genetic engineering to change an animal’s telos. This was, at least, the position that Rollin defended in his first writings on “the Frankenstein thing,” (Rollin 1986). He expanded on this position in the 1995 book, and reiterated in a 1998 paper. By 2003, Rollin’s position had started to soften a bit. He has apparently been persuaded by an argument that goes something like this: The loss of a capability that would have contributed positively to a creature’s well-being constitutes a form of harm, even if that loss occurs before the creature has had any opportunity to exercise the capability. Thus, decerebrate animals (or even blind hens) are worse off than normal animals of their species. This result almost certainly brings Rollin much more closely in line with the “consensus social ethics”. Few people presented with the blind hen argument embrace the thought of blind hens, and Sandøe himself has concluded that they are definitively not compatible with his utilitarian view of duties to animals, (Sandøe et al. 2014). However, Rollin’s 2003 view only complicates rather than settling the issue with respect to animal transformation through genetic engineering. The loss of a capability is indeed a form of harm, but it is not a decisive harm. It may tilt the balance of our moral deliberations against genetic engineering, but it also may not (Rollin 2003). Rollin has also clarified his use of the Aristotelean notion of telos in a 1998 paper. For Aristotle himself, there is an important difference between human telos and that of non-human entities such as rocks, plants and animals. For the non-human world, telos is a principle of explanation that accounts for certain dynamic processes in nature. The telos of an acorn is the oak tree it will become. Although this usage is at least superficially quite similar to the “genetic potential” sense that Rollin seems to have in mind, for Aristotle only human beings have a telos that gives rise to moral significance. Only the human telos—fulfilling our potential for rational life—involves moral dedication in its very being. Rollin acknowledges this point

5.6 Rollin’s Consensus Morality

125

and also puts some distance between himself and some of the others who have embraced the idea of animal telos (discussed below). He traces his use of telos to lectures given by the Columbia University pragmatist John Herman Randall (1899– 1980). Like John Dewey (1859–1952), Randall was particularly impressed with Aristotle’s practice of drawing philosophical principles from particular situations. Both Dewey and Randall tended to use Aristotelean concepts as if Aristotle had coined them as purely contingent responses to practical problems (Rollin 1998). Whether or not this adaptation of Aristotle can be defended, Rollin’s 1998 essay clarifies the philosophical pragmatism inherent in his general approach, giving even more support to the reading given of his “social consensus ethic” discussed above. Others have borrowed Rollin’s terminology to articulate the strongest objections to genetic engineering.

5.7 Animal Telos, Animal Integrity and Objections to Genetically Engineered Animals The notion that animals have a “nature,” with which humans should not tamper has broad appeal. Nevertheless, it is devilishly difficult to specify what this nature is, and why it would have moral implications for biotechnology. Neither rights nor utility arguments provide an easy account of why humans should respect animal natures. To the extent that one can claim rights for individual animals, the argument provides a philosophical foundation for proscribing actions that harm an existing animal. Yet do unborn animals have rights to a particular constitution or telos? As Regan has argued, utilitarians make organisms into “vessels of sentient welfare,” (Regan 1986). What matters is how these vessels are filled with experiences of satisfaction or suffering, not the shape or nature of the vessel itself. Therefore, although there is something less than unanimity on terminology, many critics of transgenesis have searched for something like telos to characterize what is at issue. Others have laid emphasis on the notion of animal integrity. Although Bernard Rollin takes credit for introducing the notion of animal telos, it became influential in early debates over genetically engineered animals through criticisms leveled by Michael W. Fox, the veterinarian and animal protectionist. Rachel Schurman and William Munro identify Fox as one of the important figures in the “handful” of civil society advocates who launched the social movement against GMOs, (Schurman and Munro 2010, p. 680). Jeremy Rifkin, the most vocal of biotechnology’s early critics and another player in Schurman and Munro’s cabal of early advocates (see p. 77), also used the concept of telos in testimony before the Recombinant DNA Advisory Committee of the U.S. National Institutes of Health in 1985: “The crossing of species borders…represents a fundamental assault on the principle of species integrity … such an intrusion violates the telos of each species and is to be condemned as morally reprehensible,” (quoted in Mauron 1989, 252). Fox proposes ethical limits on genetic engineering, rejecting the idea that “we may

126

5 Animal Health and Welfare

alter the telos of an animal provided that there is no suffering.” Fox defines telos as the ‘beingness’ of an animal, “its intrinsic nature coupled with the environment in which it is able to develop and experience life.” He lists ways of harming telos and goes on to say, “To contend that we can enhance the telos of an animal—and thus by extension believe that we can improve upon nature—is hubris,” (Fox 1990, 32) In the same article he writes, The organism and its environment are one, and we recognize that unity and harmony as health and the full expression of the animal’s telos. The telos is in part preconditioned (if not predestined) for, and dependent upon, a particular environmental niche and optimal conditions for its normal development and expression, which in turn means health and fulfillment for the animal. To deny such health and fulfillment by keeping the animal under impoverished and even stressful environmental conditions (as on a factory farm) is to cause harm. (Fox 1990, 34).

Fox’s 1992 book against agrifood biotechnology, Superpigs and Wondercorn, does not mention telos in the chapter on ethics. Elsewhere in the book he defines the term simply as “a Greek word meaning ‘end’ or ‘aim’,” (Fox 1992b, 22). He notes that scientists have ridiculed the notion of telos, and quotes M.J. Osborn (1927–2019) to the effect that the idea is “contrary to any evidence provided by biology and belongs rather in the realm of mysticism,” (quoted in Fox 1992b, p. 23). Fox then concludes: “But the debate about telos is a matter of semantics. The real issue is whether living things have inherent natural qualities that we tamper with at our peril. I believe that they do. If this is mysticism, so be it.” (Fox 1992b, p. 24). Rifkin’s testimony also cites “species integrity,” but that is another unpromising candidate. There are at least three distinct moral claims that one could attempt in appealing to species integrity. One is that naturally evolved species are either valuable in themselves, or contribute to ecological stability in subtle ways. A second is that human transgression of species boundaries is itself wrong. The third is that species integrity captures what is important about telos. Biologist Robert Colwell offered a version of the first argument quite early in the biotech debate, though his usage called attention to environmental risks of food and agricultural biotechnology, rather than impact on animals, (Colwell 1989). As Alan Holland has noted, Colwell’s argument applies exclusively to species that evolve under natural conditions, and not to domesticated species. As such, it cannot have implications for the topic of transgenic farm animals (Holland 1995, 299–300). Quoting Vernon Pursel of the Beltsville pig experiments, theologian Andrew Linzey proposes that genetic engineers reject the notion of species integrity, finding all genetic material the same, “from worms to humans.” (Linzey 1990, p. 184). From this Linzey claims to deduce that if it is acceptable to create transgenic animals, it must also be acceptable to create transgenic humans. Since this is, on Linzey’s view, an absurd proposal, he concludes that Pursell’s rejection of species integrity is mistaken. Since rejecting species integrity leads to transgenic humans, it is unacceptable to reject species integrity. Thus, Linzey concludes that the concept of species integrity is valid. Noting some of the problems with other formulations, Balzer et al. (2000) suggest we redescribe the problem as an affront to the animal’s dignity. The 1997 edition of Food Biotechnology in Ethical Perspective concluded the discussion of species integrity by finding that many of these arguments actually turn upon

5.7 Animal Telos, Animal Integrity and Objections …

127

claims to the effect that the environment is harmed (see Chap. 7), or that metaphysical boundaries are violated (see Chap. 10). As such, integrity, dignity or telos arguments are thus not really about animals at all. Bovenkerk, Brom and van den Bergh argue that however flawed the term ‘species integrity’ is with respect to identifying something biologically meaningful or of direct relevance to the morally significant traits of animals, it nevertheless serves as a way to move public debate over the appropriateness of animal biotechnology forward (Bovenkirk and Bergh 2001). The philosophical literature on biological, genetic or species integrity has continued to grow, suggesting that that they were right. Michael Hauskeller produced a booklength study of the concept, (Hauskeller 2007). Animal integrity is also discussed (not altogether favorably) as applied to human-animal chimeras by Alison Harvey and Brian Salter (2012). Other researchers have expanded the scope of its application beyond biotechnology to other forms of animal use, including conventional lab animals, (Röcklinsberg et al. 2014). In a widely reprinted article, Traci Warkentin (2006) relied upon the term integrity in recounting a strong sense of disapproval over what she describes as the “dis/integration” of food animals through the applications of agrifood biotechnology. All these approaches succeed in conveying the authors’ repugnance, and they serve to indicate that something beyond welfare or Tom Regan-style animal rights must be at stake in the genetic transformation of an animal. Many become vague in indicating exactly how the target of disapproval lies has an effect on any actual animal or its quality of life. As with Warkintin’s article, something like integrity or dignity becomes most compelling as an articulation of what is troubling about animal disenhancement, another topic that continues to accumulate commentary.

5.8 Against Changing Telos or Species Integrity of Food Animals The presumptions of some of my critics to the contrary, I am not myself entirely comfortable with radical genetic modification of animals. To begin my argument, it is useful to consider Henk Verhoog’s attempt to link the integrity/telos concern to a form of harm that is actually experienced by a living creature. Verhoog interprets telos as implicit in the background assumptions for accounts of abnormality and suffering. Anticipating an argument that Christopher Preston made against synthetic biology (Preston 2008), Verhoog claims that modifications rob animals of their being as the product of evolutionary history. Unlike Preston, Verhoog calls the epistemic authority of biology as the arbiter of questions regarding life into question, stating that those who use a scientifically based definition of species have simply begged the key moral question (Verhoog 1996). Verhoog notes how plant and animal species serve as natural kinds that ground ordinary language. His position is that animals coevolved with humans into distinct species through a conceptual as well as a biological process (Verhoog 1992). The implication is that part of what it means to be human is to live among well-defined animal species. He laments the loss of a personal

128

5 Animal Health and Welfare

relation between the biologist and his object of study: “There seems to be a reverse relationship between the degree of reductive objectivation and the degree of moral relevance of the entities studied,” (Verhoog 1993, p. 94). Rollin’s 2003 paper (in which his position on changing telos has most significantly been modified) was responding to Jason Robert and Françoise Baylis (2003), who had argued that human biotechnology is controversial because the boundary between humans and other animals is integral to our moral vocabulary. Technologies that challenge this boundary destabilize our ability to make moral judgments at all. There is an obvious similarity between this argument and the one put forward by Verhoog. Oddly, however, Robert and Baylis also write that modifications of nonhuman animals are non-controversial, though perhaps they are thinking of the mouse biotechnology that has become commonplace in medical research. Rollin extends their argument to any modification of telos, and on this basis admits that contrary to views he expressed earlier, changes to telos could, on the grounds Robert and Baylis note, involve serious ethical issues. One might question whether this modification of Rollin’s view raises ethical issues that really have anything to do with the animals themselves. Rather, the problem lies in the way that the appearance of these disturbing animals has challenged our ability to think and communicate with one another. The harm here is to ourselves, or at least to other human beings and the human moral community. It is only when one takes the further step that Verhoog takes that this has anything to do with the plight of the creature. In saying that ordinary language and ordinary conceptions of species boundaries have moral priority over the theories of biologists, Verhoog is arguing that human-animal relations are properly constituted, conceptualized and regulated in conformity with these ordinary language conceptions. When our capacity to conceptualize human-animal relationships is challenged by new technology, this does violence to the relationship, itself. Thus, even if animals do not suffer in the sense of enduring pain or disease as a result of this change, their moral standing is challenged, and their capability of appearing to us as moral subjects is potentially threatened. Verhoog is arguing that in robbing animals of their ability to be seen by us as whole beings, representative of a natural kind, biotechnology is having an ontological impact on animals, at least in so far as they are capable of entering into moral relationships with human beings. As such, it is Verhoog’s earlier arguments that pose a greater philosophical challenge to Rollin’s view that changing telos does no harm to the animals themselves. Alan Holland also offered an argument against changing animal telos in a single paragraph at the end of a long and mostly critical article evaluating Fox, Rollin, Verhoog and others who have taken up the question of telos. There he states that even Rollin’s claim that changing telos to relieve animal suffering, turns out to fall foul of something akin to Kant’s proscription against treating rational natures, which are ends in themselves, as means—even as means which could be regarded as beneficial to the animal in question. It was on grounds just such as these that Kant condemned suicide. …changing an animal’s nature for the sake of rendering it less susceptible to disease is less than respectful of that animal’s nature, since it would involve subordinating the whole nature to the cause of relief from disease. Essentially, it puts respect for the states of a subject above respect for the subject (Holland 1995, 304).

5.8 Against Changing Telos or Species Integrity of Food Animals

129

This passage is put forward within a carefully crafted context that provides all the proper disclaimers dissociating the argument from what Kant might have actually thought, given his view that we have no direct duties to animals at all. Holland’s argument presumes that respecting an animal as an individual subject is co-terminal with respecting its nature, something it derives in virtue of being an individual of a certain species. Speaking now for myself, I find an adaptation of Marx’s notion of “species being,” to suggest a more promising line of argument against changes in an animal’s genetic make-up. Karl Marx (1818–1893) introduced the idea of “species being,” in his manuscripts on alienated or estranged labor. He lists four ways in which the capitalist institution of wage labor harms the worker through estrangement. First, workers are estranged from what they make, which belongs not to them but to the person for whom they work. Second, workers are estranged from that portion of their life spent at work, as they come to see only the weekend, holidays and retirement as the times when they can realize their autonomously chosen life goals. Third, they are estranged from one another, since they must regard one another as competitors for jobs. Finally, they are estranged from their species being. Human species being is to be the organism, the being that realizes itself through productive work. Marx argues that wage labor separates workers from what it means to be most fundamentally human (Marx 1988, pp. 74–78). There is some risk that this reference to Marx will cause even more mischief than Rollin’s references to Aristotle. Marx clearly thought that species being is possible only for creatures capable of having an intellectual awareness of themselves as members of a species. Humans do this but it is doubtful that cows, pigs or chickens do as well. Nevertheless, Marx is useful in the present context because each form of estrangement that he discusses is both psychological and material. Estrangement is psychological in that it is experienced as anxiety and anomie; it creates a kind of existential angst. Arguably, many workers never experience the angst of estrangement that Marx describes. Yet it is the material fact of separation or estrangement that is most significant for Marx, as it has been for other social critics. Aldous Huxley’s Brave New World offers a nightmare vision in which the psychological peril of existential angst is relieved through drugs and (ironically) genetic technology, but the moral lesson of Brave New World is that such relief only makes the moral problem worse. It is wrong to educate, acclimatize or behaviorally condition humans so that they are estranged from their humanity. It would be equally wrong to attempt (or inadvertently affect) this feat with genetic engineering. A great deal of moral philosophy that has been done on behalf of animals extends concepts that have traditionally been thought to apply only to humans beyond the human community. A similar move is being made here. Does it matter that in the present context we are speaking not of human nature or human species being, but of non-human animals? Does one do unacceptable violence to Kant, Marx and Huxley in making an analogous argument against genetic engineering of animals? Given that we routinely speak of animal natures, given that those who tend to sheep, cows, pigs or horses develop a fine appreciation of “the sheepness of the sheep, the pigness of the pig,” it would seem that the burden of proof for falls on the side of denying

130

5 Animal Health and Welfare

animal telos, understood in this restricted sense. If telos is meaningful in this sense, why should we not also conclude that it would be wrong to genetically engineer animals into a creature that is incapable of participating in the telos characteristic of their species? Rollin may have reached the conclusion that such engineering is not wrong by adopting a radically individualistic notion of telos. Each organism has functional needs, and having a given set of functional needs may be typical of animals in given species. Knowing the needs typical of a given species then becomes a way of knowing the functional needs of any individual. Here, one’s knowledge of moral obligation derives from the species, but the obligations themselves are bound up entirely in meeting the needs of individual organisms. Nothing follows about whether it would be permissible to deliberately bring into being a creature with an entirely novel set of functional needs. If this is Rollin’s position, it has two possible unsettling implications. First, it may follow that if it is permissible to estrange an animal from a given set of functional needs with genetic manipulation, it is also permissible to estrange an animal from functional needs through behavioral conditioning or even surgery. It is not clear how strongly telos, or the functional needs that define it, is genetically determined. If genetic determination is very strong, such non-genetic forms of estrangement are unlikely to be fully successful. Attempting them would constitute ordinary cruelty. If, however, functional needs are, as Fox and Verhoog argue, fixed by the interaction between genes and environment, Rollin’s original view permits much broader manipulation of animals than it might have seemed. Sandøe’s blind hens bring home the practical implications of a permissive view on genetic modification, and the objection to it still holds, even if blind hens can’t be justified given Sandøe’s.utilitarian approach (Sandøe et al. 2014). Whatever we finally decide about a proposal for transforming animals, there is a problem with any philosophical principle that makes the case for transformation this easy, this one sided. Indeed, in the debate over utilitarianism between Bernard Williams (1929–2003) and J.C.C. Smart (1920–2012), it was the ease with which utilitarians reach certain conclusions Williams’ cited as one of his principle arguments against it, (Smart and Williams 1973). As Holland notes, we seem to be in a mindset in which the animal’s suffering is the only thing that prevents us from regarding it as a moral nullity, entirely at our disposal for the satisfaction of any need or desire. Dispose of suffering and we may indeed feel like Gods. Second, if telos is radically individualistic, what would be wrong with the genetic modifications Huxley described in Brave New World? Such humans would not strictly be humans at all. They would lack ordinary human functional needs, and it would not be wrong to create such sub-human creatures. Holland’s Kantian argument brought forward, along with the Marxist twist added here, provides an account of why the Brave New World modifications are wrong. But an interpretation of telos, human nature or species being that is radically individualistic robs this argument of its moral force. Yet it seems morally arbitrary to attribute significance only to human telos, human nature or human species being. It ignores the received practice among animal caregivers of recognizing highly analogous traits characteristic of other species.

5.8 Against Changing Telos or Species Integrity of Food Animals

131

Clearly, the telos that is characteristic of any species (including humans) is instantiated only in the individuals of the species. If we recognize immorality in acts that would modify a human genome to the point that the resulting individual would no longer be characteristic of the human species, why is it not also immoral to modify the genome of other animals so that the resulting individuals are uncharacteristic of their species? Until someone can offer a non-arbitrary reason for making this distinction, radical forms of transgenesis for animals should be regarded as morally problematic. My only hesitancy in reiterating this conclusion, drawn initially in 1997, is the obvious point that there are many things we do to animals that we would regard as deeply problematic if they were done to humans. Eating them, for example. However, it is important to emphasize that here we are considering acts that are wrong not because of their effect on individual humans, but because they so vitiate our ability to make sense of humanity. On this point, my argument links with Verhoog’s. The problem may lie in practices that disturb our ability to conceptualize our relationships with animals in moral terms. This is not a harm done to any individual animal, to be sure. It relates, nonetheless, to the plight of the creature, as surely as it relates to who we are, to human beings’ conception of themselves as moral agents.

5.9 Animal Biotechnology and Moral Obligation The above analysis supports the view that alienating an individual food animal from the functional needs, the telos characteristic of its species is morally questionable and likely immoral. The fact that such technology is used to relieve the potential for suffering becomes irrelevant. However, this principle may proscribe less than is initially thought. In the first place, few extant examples of transgenic animals appear to have such a dramatic effect on the individual animals. There may be little reason in the foreseeable future to apply such a radical form of genetic engineering to food animals, in the first place. The more difficult cases that Rollin considers arise in the arena of biomedical research. As noted at the outset, these cases involved compelling human needs. They are negotiated under an aura of emergency conditions and exceptional circumstances that simply do not apply to the discussion of food animals. Thus while the principle stated above would prohibit the use of genetic engineering to create a breed of decerebrate poultry, intended for intensive factory farming, it would not rule out every case in which genetic engineering might be used to relieve the suffering of animals being used in the exceptional circumstances of biomedical research. Such practices should not, it would seem, become too routine, but they would have to be subjected to a very different philosophical analysis, in any case. The majority of moral duties relevant to food animal biotechnology are captured by Rollin’s principle of conservation of welfare applied within the context of a riskbased approach. We should not initiate production practices (using transgenics or other forms of biotechnology) that make food animals worse off than they are now. If that status quo can be maintained (at least), products to increase food quality or

132

5 Animal Health and Welfare

productivity are acceptable. If animal well-being can be improved through biotechnology, all to the good. Evaluating animal welfare is not easy (see Broom 1995), but the products of biotechnology do not present unique challenges. With Rollin (as with most mainstream animal protectionists and the general public), what counts is sentient experience of pain, fear, suffering and stress, along with traditional measures of animal health. While it would be inappropriate to sacrifice important human needs to the improvement of animal well-being, a strict logic of comparing costs and benefits to humans and animals should not be employed to rationalize actions that make food animals worse off than they currently are. The consensus morality is that food producers can do better, and they have a moral duty to try. Philosophers who have taken up these issues since the second edition appeared in 2007 do not cite Rollin, much less Verhoog or early critics like Fox or Rifkin. I believe that the first round of debate has philosophical resources that are being neglected, and I hope that this chapter will bring them to the new generation’s attention. Verhoog, especially, is neglected by those who suspect a hidden theology in his position. Yet, the reading I make of it is secular, even if laden with metaphysical commitments. My own views on key philosophical points have evolved in ways that this chapter could not have been revised to reflect. I see connections between this chapter’s discussion of telos and species being, and my more recent work on animal natures (see Thompson 2015, pp. 137–146). I do find it troubling when opportunities to improve the lives of farmed animals are foregone because of highly abstruse philosophy. Yet, like Bernard Williams (discussed above) I am less troubled by acts that promote better welfare than by the ease with which philosophers arguing for these technologies elide the consideration of animal natures, including human nature. But that is a story that has little to do with animals themselves.

References Atwood, M. 2003. Oryx and Crake. New York: Doubleday. Balzer, P., K.P. Rippe, and P. Schaber. 2000. Two concepts of dignity for humans and non-human organisms in the context of genetic engineering. Journal of Agricultural and Environmental Ethics 13: 7–27. Belmonte, J.C.I. 2016. Human organs from animal bodies. Scientific American 315 (5): 32–37. Blatz, C. 1991. It is morally permissible to manipulate the genome of domestic hogs. The Journal of Agricultural and Environmental Ethics 4: 166–176. Bourret, R., E. Martinez, F. Vialla, C. Giquel, A. Thonnat-Marin, and J. De Vos. 2016. Human– animal chimeras: ethical issues about farming chimeric animals bearing human organs. Stem Cell Research & Therapy 7: 87. Bovenkirk, B. F.W.A. Brom and B.J. van den Bergh. (2001) Brave new birds: The use of integrity in animal ethics. Hastings Center Report 32(1): 16–22. Broom, D.M. 1995. Measuring the effects of management methods, systems, high production efficiency and biotechnology on farm animal welfare. ed. by T.B. Mepham, G.A. Tucker and J. Wiseman. Issues in Agricultural Bioethics, 312–334. Nottingham, UK: University of Notingham Press. Cohen, C., and T. Regan. 2001. The Animal Rights Debate. Lanham, MA: Rowman and Littlefield.

References

133

Colwell, R.K. 1989. Natural and unnatural history: Biological diversity and genetic engineering. In Scientists and their Responsibilities, ed. W.R. Shea and B. Sitter, 1–40. Canton, OH: Watson Publishing International. Comstock, G. 1988. The case against BGH. Agriculture and Human Values 5: 36–52. Croney, C., W. Muir, J.Q. Ni, N.O. Widmar, and G. Varner. 2018. An overview of engineering approaches to improving agricultural animal welfare. Journal of Agricultural and Environmental Ethics 31: 143–159. Denner, J. 2017. Advances in organ transplant from pigs. Science 357: 1238–1239. Ferrari, A. 2012. Animal disenhancement for animal welfare: The apparent philosophical conundrums and the real exploitation of animals. A response to Thompson and Palmer. NanoEthics 6: 65–76. Fox, M.W. 1990. Transgenic animals: ethical and animal welfare concerns. In The Bio-Revolution: Cornucopia or Pandora’s Box, ed. P. Wheale and R. McNally, 31–54. London: Pluto Press. Fox, M.W. 1992a. The new creation: An update on animal gene engineering. In Animal Biotechnology: Opportunities and Challenges, ed. J.F. MacDonald, 49–56. Ithaca, NY: National Agricultural Biotechnology Council. Fox, M.W. 1992b. Superpigs and Wondercorn. London: Lyons and Burford. Frey, R. 1980. Interests and Rights. Oxford, UK: Claredon Press. Gamborg, C., and P. Sandøe. 2002. Breeding and biotechnology in farm animals. ed. R. Levinson and M.J. Reiss. Key Issues in Bioethics: A Guide for Teachers, 133–142. London: Routledge Falmer. Harrison, P. 1992. Descartes on animals. The Philosophical Quarterly (1950-), 42(167): 219–227. Harrison, R. 1964. Animal Machines: The New Factory Farming Industry. London: V Stuart. Hauskeller, M. 2007. Biotechnology and the Integrity of Life. Aldershot, UK: Ashgate. Harvey, A., and B. Salter. 2012. Anticipatory governance: Bioethical expertise for human/animal chimeras. Science as Culture 21: 291–313. Henschke, A. 2012. Making sense of animal disenhancement. NanoEthics 6: 55–64. Hoban, T.J., and P. Kendall. 1993. Consumer Attitudes about Food Biotechnology. Raleigh, NC: North Carolina Cooperative Extension Service. Holland, A. 1995. Artificial lives: Philosophical dimensions of farm animal biotechnology. ed. by T.B. Mepham, G.A. Tucker and J. Wiseman. Issues in Agricultural Bioethics, 293–306. Nottingham, UK: University of Nottingham Press. Houdebine, L.M. 2000. Transgenic animal bioreactors. Transgenic Research 9: 305–320. Kimbrell, A. 1993. The Human Body Shop. New York: HarperCollins. Krimsky, S., and R. Wrubel. 1996. Agricultural Biotechnology and the Environment: Science. Policy and Social Issues: University of Illinois Press, Urbana, IL. Kronfeld, D.S. 1993. Recombinant Bovine Growth Hormone: Cow responses delay drug approval and impact public health. The Dairy Debate, ed. W. Leibhardt, 65–112. Davis, CA: University of California Sustainable Agriculture Research and Education Program. Linzey, A. 1990. Human and animal slavery: A theological critique of genetic engineering. In The Bio-Revolution: Cornucopia or Pandora’s Box, ed. P. Wheale and R. McNally, 175–188. London: Pluto Press. Marx, K. 1988. Economic and Philosophic Manuscripts of 1844, M. Miligan (trans) Buffalo, NY: Prometheus Books. Mauron, A. 1989. Ethics and the ordinary molecular biologist. In Scientists and Their Responsibilities, ed. by W.R. Shea abd B. Sitter, 249–265. Canton, OH: Watson Publishing International. Meeker, D. 1992. Animal well-being and biotechnology. Animal Biotechnology: Opportunities and Challenges, ed. J.F. MacDonald, 77–84. Ithaca, NY: National Agricultural Biotechnology Council. Moore, D.A., and L.J. Hutchinson. 1992. BST and animal health. Bovine Somatotropin and Emerging Issues, ed. M. Hallberg, 99–132. Boulder: Westview Press. Murphy, K.N., and W.P. Kabasenche. 2018. Animal disenhancement in moral context. NanoEthics 12: 225–236.

134

5 Animal Health and Welfare

Ormandy, E.H., J. Dale, and G. Griffin. 2011. Genetic engineering of animals: ethical issues, including welfare concerns. The Canadian Veterinary Journal 52: 544–550. Palmer, C. 2011. Animal disenhancement and the non-identity problem: a response to Thompson. NanoEthics 5: 43–48. Pew Initiative on Food and Biotechnology. 2004. Americans’ Opinions About Genetically Modified Foods Remain Divided, But Majority Want A Strong Regulatory System(2). Accessed Feb. 13, 2020 at: https://www.pewtrusts.org/en/about/news-room/press-releases-and-statem ents/2004/12/08/americans-opinions-about-genetically-modified-foods-remain-divided-but-maj ority-want-a-strong-regulatory-system2. Preston, C.J. 2008. Synthetic biology: drawing a line in Darwin’s sand. Environmental Values 17: 23–39. Regalado, A. 2019. Gene-edited cattle have a major screwup in their DNA. MIT Technology Review. Aug. 29. Accessed July 7, 2020 at https://www.technologyreview.com/2019/08/29/ 65364/recombinetics-gene-edited-hornless-cattle-major-dna-screwup/. Regan, T. 1983. The Case for Animal Rights. Berkeley: The University of California Press. Regan, T. 1986. The case for animal rights. In In Defense of Animals, ed. P. Singer, 13–26. New York: Harper. Regan, T. 1995. Obligations to animals are based on rights. Journal of Agricultural and Environmental Ethics 8: 171–180. Regan, T. 2003. Animal Rights, Human Wrongs: An Introduction to Moral Philosophy. Totowa, NJ: Rowman and Littlefield. Resnik, D.B. 2015. What is ethics in research & why is it important. Your Environment, Your Health. National Institute of Environmental Health Sciences. Accessed July 7, 2020 at http:// nook.cs.ucdavis.edu/~koehl/Teaching/ECS188/Reprints/Ethics_Research_NIH.pdf. Robert, J.S., and F. Baylis. 2003. Crossing species boundaries. American Journal of Bioethics 3: 1–13. Robl, J.M. 2007. Application of cloning technology for production of human polyclonal antibodies in cattle. Cloning and Stem Cells 9: 12–16. Röcklinsberg, H., C. Gamborg, and M. Gjerris. 2014. A case for integrity: gains from including more than animal welfare in animal ethics committee deliberations. Laboratory Animals 48: 61–71. Rollin, B.E. 1981. Animal Rights and Human Morality. New York: Prometheus Press. Rollin, B.E. 1986. The Frankenstein thing. Genetic Engineering of Animals: An Agricultural Perspective, ed. J.W. Evans and A. Hollaender, 285–298. New York: Plenum Press. Rollin, B.E. 1989. The Unheeded Cry: Animal Consciousness. Animal Pain and Science: Oxford University Press, New York. Rollin, B.E. 1995. The Frankenstein Syndrome: Ethical and Social Issues in the Genetic Engineering of Animals. New York: Cambridge University Press. Rollin, B.E. 1998. On telos and genetic engineering. Animal Biotechnology and Ethics, ed. A. Holland and A. Johnson, 156–187. London: Chapman and Hall. Rollin, B.E. 2003. Ethics and species integrity. American Journal of Bioethics 3 (3): 15–17. Sandøe, P., B.L. Nielsen, L.G. Christensen, and P. Sørensen. 1999. Staying good while playing god: The ethics of breeding farm animals. Animal Welfare 8: 313–328. Sandøe, P., P.M. Hocking, B. Förkman, K. Haldane, H.H. Kristensen, and C. Palmer. 2014. The blind hens’ challenge: does it undermine the view that only welfare matters in our dealings with animals? Environmental Values 23: 727–742. Sapontzis, S.F. 1991. We should not manipulate the genome of domestic hogs. The Journal of Agricultural & Environmental Ethics 4: 177–185. Schaefer, G.O., and J. Savulescu. 2014. The ethics of producing in vitro meat. Journal of Applied Philosophy 31: 188–202. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists versus Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press.

References

135

Shriver, A. 2009. Knocking out pain in livestock: Can technology succeed where morality has stalled? Neuroethics 2: 115–124. Shriver, A. 2020. Prioritizing the protection of welfare in gene-edited livestock. Animal Frontiers 10: 39–44. Shriver, A., and E. McConnachie. 2018. Genetically modifying livestock for improved welfare: a path forward. Journal of Agricultural and Environmental Ethics 31: 161–180. Singer, P. 1973. Animal liberation: Review of Godlovitch, Godlovitch and Harris, Animals, Men and Morals. New York Review of Books, April 5, 1973, Accessed Feb. 16, 2020 at https://www.nybooks.com/articles/1973/04/05/animal-liberation/. Singer, P. 1975. Animal Liberation: A New Ethic for Our Treatment of Animals. New York: Avon Books. Singer, P. 1979. Practical Ethics. Cambridge, UK: Cambridge University Press. Smart, J.J.C. and B. Williams. 1973. Utilitarianism: For and Against. Cambridge, UK: Cambridge University Press. Tan, W., C. Proudfoot, S.G. Lillico, and C.B.A. Whitelaw. 2016. Gene targeting, genome editing: from Dolly to editors. Transgenic Research 25: 273–287. Thompson, P.B. 1999. Ethical issues in livestock cloning. Journal of Agricultural and Environmental Ethics 11: 197–217. Thompson, P.B. 2008. The opposite of human enhancement: Nanotechnology and the blind chicken problem. NanoEthics 2: 305–316. Thompson, P.B. 2015. From Field to Fork: Food Ethics for Everyone. New York: Oxford University Press. Van Eenennaam, A.L. 2018. The importance of a novel product risk-based trigger for geneediting regulation in food animal species. The CRISPR Journal 1: 101–106. Van Eenennaam, A.L., and A.E. Young. 2018. Public perception of animal biotechnology. Animal Biotechnology, Vol. 2. ed. H. Niemann and C. Wrenzycki. Cham, CH: Springer. https://doi.org/10.1007/978-3-319-92348-2_13. Van Eenennaam, A.L., K.D. Wells and J.D. Murray. 2019. Proposed US regulation of geneedited food animals is not fit for purpose. npj Science of Food 3: 1–7. Verhoog, H. 1992. The concept of intrinsic value and transgenic animals. Journal of Agricultural and Environmental Ethics 5: 147–160. Verhoog, H. 1993. Biotechnology and ethics. Controversial Science ed. T. Brante, T.S. Fuller and W. Lynch, 83–106. Albany, New York: SUNY Press. Verhoog, H. 1996. Genetic modification of animals: Should science and ethics be integrated? The Monist 79: 247–263. Warkentin, T. 2006. Dis/integrating animals: ethical dimensions of the genetic engineering of animals for human consumption. AI & SOCIETY 20: 82–102. Wawrzyniak, D. (2020). Why fitting animals is ethically dubious, LANDBAUFORSCH · The Journal of Sustainable and Organic Agricultural Systems 70(1): 1–4. Accessed July 7, 2020 at: https://www.landbauforschung.net/fileadmin/landbauforschung/Pdf_Papers_Reviews/ LBF-70-01-1_PP_Wawrzyniak_200619.pdf. Weisberg, Z. 2015. Biotechnology as end game: Ontological and ethical collapse in the “biotech century”. NanoEthics 9: 39–54. Wilmut, I. (1995). Modification of farm animals by genetic engineering and immunomodulation. Issues in Agricultural Bioethics, ed. by T. B. Mepham, G. A. Tucker and J. Wiseman, 229–246. Nottingham, UK: Nottingham University Press. Wolf, C.A., G.T. Tonsor, and N.J. Olynk. 2011. Understanding US consumer demand for milk production attributes. Journal of Agricultural and Resource Economics 36: 326–342.

Chapter 6

Ethics and Environmental Risk Assessment

Abstract In its most common form, environmental risk assessment is an adaptation of consequentialist ethical theory. Hazards are identified as significant through careful articulation of the values (axiology) that determine why outcomes are considered to be bad, harmful or adverse, and exposure quantification is used to characterize risk as an expected value. Several leading examples of how this framework is operationalized in the characterization of environmental risks from agrifood biotechnology are discussed, including risks to biodiversity, weediness and acquired resistance to the effectiveness of pesticides. Epistemological uncertainties plague the quantification of expected values, however, and feed both public doubts and deeper controversy. As risk communication strives to assuage those doubts, an ironic cycle of mistrust emerges, even among those who apply rigorous standards of risk analysis: Assuming that others are less careful, they regard their contrary findings as evidence of error, rather than prompting a further check on the initial assessment. For anyone inclined to see the technology as risky, the alleged carelessness of analysts amplifies the evidence for risk. Keywords Agriculture · Biotechnology · Uncertainty · Expected-value · Environmental values · Utilitarianism Consumer safety may be the most fundamental ethical responsibility for food biotechnologists, and the ethics of transforming or cloning agricultural animals is certainly the most philosophically controversial. Yet ethical responsibility for the environmental impacts of food and agricultural biotechnology is arguably the most widely discussed, and this has been true even since food and agricultural biotechnology was only a speculative possibility. This chapter and the next probe the philosophical dimensions of the debate over environmental risk, with this chapter emphasizing the ethical and interpretative questions that arise in operationalizing the risk based approach and the next addressing the issues in placing value on environmental hazards. This chapter introduces the topic of environmental risk with a brief review of public debates over the environmental hazards of gene technology. The chapter

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_6

137

138

6 Ethics and Environmental Risk Assessment

presupposes the risk-based approach as outlined in Chap. 2, and provides a more detailed discussion of the way that this approach is most typically interpreted in environmental risk analysis. Here, risk is interpreted as an expected value. Measurement of expected value is implicit within utilitarian accounts of decision-making. Critics of environmental risk assessment (and critics of agrifood biotechnology among them) have presumed that a risk-based approach is committed to the optimization of expected value implied by the utilitarian maxim: “greatest good for the greatest number.” One key message of the chapter as that this presumption is not only unwarranted, but also undercuts effect use of the risk-based approach. As a corrective, the chapter summarizes and criticizes various attempts to characterize both hazard and exposure for environmental risks from gene technologies in the food system.

6.1 The Environmental Debate Rachel Schurman and William Munro describe scientists and executives from industry and anti-biotechnology activists living in two distinct worlds. For their part, industry possessed an overwhelming sense of moral confidence. They were sure that agricultural science was dedicated to the good. The goal of increasing food production on a global basis was the driving imperative, and other things—farm bankruptcies and the environmental impacts of industrial production systems—were side issues. What is more, scientists believed that biotechnology would actually provide a technical fix for the acknowledged problems associated with soil exhaustion and agricultural chemicals. The activists were less unified in their concerns. Concern that gene technology would eventually be used on humans motivated some, and they were desperate to bring any progress in the tools of genetic manipulation to a halt. These activists aligned with rural advocacy groups who were concerned about farm bankruptcies and environmental impacts within the agricultural sector, (Schurman and Monroe 2010). The industry group never grasped the arguments of those who saw agricultural gene technologies as a slippery slope leading to the race purification goals of the eugenics movement. In addition, they viewed farm bankruptcy as a consequence of well-functioning market forces. As such, it was not an appropriate target for ethical concern. They did understand the environmental concerns, though they differed with activists in how gene technology would affect the environment. Environmental impact, thus, became a centerpiece in the debate because it was the only topic that both sides understood as debatable. For example, Jeremy Rifkin made his entry into publishing with a 1977 book he co-authored with liberal social entrepreneur Ted Howard. Who Should Play God? The Artificial Creation of Life and What It Means for the Human Race was one of the first books to follow the Asilomar Conference on Recombinant DNA. As mentioned in Chap. 3, molecular biolgists convened at Asilomar Conference Center in Pacific Grove, CA to discuss the risks of genetic engineering. This 1975 conference provoked further considerations by journalists, moralists, religious authorities and social activists. In Who Should Play God? Howard and Rifkin emphasize impact on human health, and caution against the temptation to use gene modification to intervene in human heredity, (Howard and Rifkin 1977). By the 1980s, Rifkin’s

6.1 The Environmental Debate

139

books Algeny (1983) and Declaration of a Heretic (1985) interlaced speculation about environmental catastrophe with a broad philosophical critique of genetic engineering. Environmental themes were revisited in Rifkin’s 1998 book The Biotech Century, though in this effort the philosophical dimension was significantly reduced. Yet even in Rifkin’s more overtly philosophical writings the basis for his environmental concerns remain obscure. The following passages are typical: “when it comes to advancing our power and control over the forces of nature, our species has shown little willingness, of late, to temper its technological prowess by debating whether or not to proceed.” and “With genetic technology we assume control over the hereditary blueprints of life itself. Can any reasonable person believe for a moment that such unprecedented power is without risk?” (Rifkin 1985, 44). Rifkin was regarded as the bête noir of biotechnology during the years when his organization, The Foundation on Economic Trends, was filing (and winning) lawsuits against the National Institutes of Health and enlisting a broad coalition of religious leaders in protest against animal patenting (Schurman and Munro 2010). As products of agricultural biotechnology entered more active stages of development and commercial use, environmental risk also became the focus of extended discussion and debate within the scientific community. No less than seven study committees from the United States National Research Council addressed various dimensions of the environmental risks from agricultural biotechnology between 1987 and 2014. These studies highlight the need for attention to the environmental impact of gene technology, but they also argue that the environmental risks of GMOs are not different in kind from other types of modern agricultural technology. They largely adopt an implicit and unspoken stance toward the root issues of environmental risk: what is an environmental hazard, and what are our responsibilities (both singling and collectively) to avoid, mitigate or in other ways respond to environmental risk? Many critics of agricultural biotechnology discuss environmental risks and link them to ethics, but in almost every case, the connection between environmental risk and ethical responsibility remains obscure. Proponents of agricultural biotechnology are unlikely to link environmental risks to ethics, and in fact tend to discuss environmental risk as if it were a purely technical issue. Despite this marked difference between the language of boosters and knockers, much of the dispute over environmental risks from transgenic plants or animals concerns matters of fact. Do farmers use more or less herbicide with herbicide tolerant crops? Do Bt crops increase the probability that insects will develop resistance to Bt, and if so, what are the ecological affects? Even though current science has not resolved these issues to everyone’s satisfaction, the dispute continues to center on factual content, not ethical values. However, empirical disputes are often interwoven with theoretical or intellectual differences of opinion and practice that are, at bottom, philosophical. Some of the key philosophical disputes concern the way that we understand the process of acquiring knowledge through a combination of logic and evidence. For example, what kinds of evidence can be used to estimate the probability of an unwanted environmental impact? What is the appropriate baseline for assessing the probability of an unwanted environmental impact? These questions only become meaningful within the analytic framework of expected-value risk assessment.

140

6 Ethics and Environmental Risk Assessment

6.2 Environmental Risk: An Expected Value Approach When risk is defined as a function of the probability and value of unwanted events (such as insect resistance to the Bt toxin, the movement of herbicide tolerance into weeds, or loss of endangered species as a result of agricultural encroachment) it becomes possible to interpret environmental impact in terms of expected value. When philosophers and mathematicians developed the idea of expected value hundreds of years ago, they were trying to develop a rational theory that would tell them whether to take any given bet offered in a gambling game. As the term suggests, an expected value is the value one expects to realize at some time in the future. For gamblers the relevant future begins immediately after the game is played; for modern environmental risk analysts it is the real and indefinite future that extends before us into infinity. In both cases, the bets must be placed before these outcomes are known. The two key factors in conceptualizing an expected value are the value V (e.g. benefit or harm) that would be experienced if the event were to occur, and one’s expectation that the event actually will occur. This expectation can be represented as the likelihood or probability P of the event. Expected value is a function of V and P. For gambling games the values (V) at issue involve the reward of associated with a win, and all the probabilities can be computed mathematically. In classical theories of expected value, the function for combining V and P was derived by imagining repeated play of the same game. Eventually the probabilities would play out, and simply multiplying V times P produces the present value of a bet, which is to say, the value one can expect to earn from playing the game. Of course, on limited plays one can win or lose (which is why they call it gambling, after all). Readers will recognize that V and P correspond to the hazard identification and exposure quantification phases of risk assessment. One can certainly challenge the suggestion that gambling games model environmental risks in a philosophically appropriate way (see Hannson 2009). Some objections to the gaming model for environmental impact assessment are discussed in Chap. 7, and readers should be aware that my commitments to it are highly qualified. Nevertheless, a great deal of contemporary social science presumes that a generalization of the gambling analysis captures the essential elements of rational decision-making. It is therefore important to explore the strengths of this approach before discussing its weaknesses. Rational behavior is, on the gaming model, understood to be implicit in virtually every aspect of life consistent with the goal of making choices that “pay off.” A rational choice is one that has the best chance of realizing the ends sought by the decision maker. An expected-value approach to risk emerges naturally in this general orientation to decision making. At the same time, the function that should be used to combine V and P is not obvious outside the well-defined structure of gamble. Assigning a cardinal value (or a probability distribution) for either V or P may be far less important than having a general sense of what one stands to lose or gain (e.g. V) and the chances that the losing or winning outcomes will actually occur (P). Yet thinking of environmental risk as an expected value lends itself nicely to a scientific approach, because science has many tools for measuring probabilities.

6.2 Environmental Risk: An Expected Value Approach

141

Comparing expected values as a means of ethical decision-making has one overwhelming strength: it combines the predictive power of science with the commonsense ethical maxim, “act so as to bring about the best consequences, all things considered.” If one has a complete list of expected values for each of the principle options under consideration, one could apply a decision rule such as “Choose the option that maximizes benefits, relative to costs,” to make the choice. Approaches to food safety and to animal welfare that were described as ‘utilitarian,’ in previous chapters apply this general strategy, and it has many applications to environmental risk, as well. Utilitarian approaches also have some problems, however, some of which have already become evident in previous chapters. Chapter 4 on food safety discusses the tension between a utilitarian’s emphasis on welfare improvement and a rights theorist’s emphasis on consent. There, I argued that a risk-based approach can accommodate a deontological ethic by stipulating a risk management procedure based on informed consent. While consent critiera suggest an alternative to optimizing expected values, this chapter takes up problems internal to the valueoptimizing orientation of utilitarianism. Some problems in environmental applications will be discussed in due time, but the simple common sense inherent in weighing the consequences of one’s actions, and in using science to estimate their likelihood demands that this approach be given strong consideration.

6.3 Expected Value and the Consequentialist Framework Utilitarianism is the most typical and certainly the best-developed example of a general approach to making ethical decisions that involves the comparison of expected values. The name for the general approach is consequentialism. All consequentialist ethical theories evaluate expected values in light of a decision rule that determines which option from those available should be chosen. The decision rule corresponds to the risk management phase in risk assessment. However, not all consequentialist theories evaluate expected values in the same way, and not all of them use the same decision rule. Utilitarianism uses an optimizing decision rule called the utilitarian maxim. It is popularly expressed as “Do the greatest good for the greatest number.” Utilitarianism requires the decision maker to sum benefits against risks (and other costs) for each option. The utilitarian maxim can then be applied to these sums. It instructs the decicion maker to choose the “best” course of action, defined as the one with the greatest net value, once the risks and benefits for each option have been summed. Other forms of consequentialism may use the same procedure for summing risks and benefits, but apply a different decision rule. For example, in food safety, decision makers do not take benefits into account; they choose the option with the lowest feasible risk, irrespective of benefits. A similar approach might be pursued for environmental risks. Instead of seeking an optimal ratio of benefit to risk, decision-making for environmental impact might adopt a decision rule that considers only risks. This is not consistent with the dictates of utilitarian ethics, but it would still be consequentialist in its method and overall logic, (see Sagoff 2004).

142

6 Ethics and Environmental Risk Assessment

In fact, many different methods for assessing probabilities, evaluating outcomes, and integrating expected values under decision rules are used in environmental decision-making. There is thus the potential for methodological and philosophical disagreement about matters such as whether benefits should count, even among those committed to consequentialist ethics. Too often critics have presumed that one version of the approach characterizes the entire field. In a widely cited paper, Steven Kelman takes cost-benefit analysis to task for requiring a common unit of measure for risks and benefits, usually money (Kelman 1981). Yet while some economists do use common units when estimating expected values, doing so is not a logical or philosophical requirement of consequentialist ethics, nor is commitment to the idea that one must maximize the ratio of benefit to cost. A decision rule that eliminates all options having low-probability/high-consequence risks from consideration, for example, is entirely consistent with the expected value approach, even though doing so might exclude the option that has the greatest expected value. Procedures for risk-benefit and cost-benefit analysis can be specified in cookbook detail. Doing so commits anyone who uses the procedure to an entrenched philosophical view on environmental risk without the benefit of opportunities to examine and debate the assumptions and decisions incorporated therein, (MacLean 1986). Alternatively, one may defend expected value style risk analysis as a way of organizing information that is relatively objective, and that allows everyone to see where key philosophical decisions have been made, (Leonard and Zeckhauser 1986; Railton 1990). Misunderstanding may also arise in connection with the affinity between the expected value approach and economics. Economic cost-benefit analysis often uses the principle of potential Pareto improvement (the proposed course of action must produce social benefits that are sufficient to compensate costs to losers, or simply, benefits must outweigh costs) as its decision rule, and interprets the result as a form of social efficiency. Yet there are many instances in which we make choices that sublimate efficiency to other values, and a vast philosophical literature criticizing decision rules that emphasize efficiency, (see, for example, Gibbard 1986; MacLean 1990). The important point to emphasize in connection to the authors cited above is that each either endorses or criticizes a specific interpretation of the expected value approach, but the expected value approach is itself capable of comprehending all of these differences. The expected value approach to environmental risk was proposed very early on as a way to understand risks associated with agricultural biotechnology. A controversy over genetically engineered ice-nucleating bacteria in the 1980s set the backdrop for this debate. In an article written for policy makers, Martin Alexander argued that six factors determined the potential for ecological harm as a result of research on genetically engineered organisms. They are: (1) the probability of release, (2) the probability of survival, (3) the probability of multiplication, (4) the probability of dispersion, the (5) probability of gene flow, and (6) the probability that any of these events are harmful, (Alexander 1985). Though Alexander may have intended this argument as implicit assurance that the environmental risks of genetic engineering are very low, it was soon interpreted as the basis for estimating exposure in risk assessments of releasing genetically engineered plants into the environment. The

6.3 Expected Value and the Consequentialist Framework

143

book Risk Assessment in Genetic Engineering (Levin and Strauss 1991) includes a series of sophisticated overview papers by scientists, all of whom presume that risk assessment means identification of potentially harmful events, and quantification of the probability of those events. While making this assumption falls short of explicitly endorsing the expected value framework, discussions of the environmental risk of food biotechnology repeatedly appear committed to this framework, generally without argument, (see Davis 1987; Sharples 1987; Huttner 1993; Hino 1994).

6.4 Understanding Hazards and Harm As noted in Chap. 1, risk analysts make a distinction between hazard and risk. A hazard is a situation with the potential for harm. Hazard does not reflect any characterization of the likelihood that this harm will actually occur. To speak of risk requires an analysis of exposure as well as hazard, where exposure is a characterization (usually quantitative) of the course of events that must occur for the harm to materialize. Thus, Alexander’s rough argument offered in the mid 80s was in fact a broad characterization of the exposure pathways for environmental risk from agricultural biotechnology. Hazard identification and exposure quantification thus represent the two key technical phases in environmental risk analysis, (see Chap. 2). Hazard identification is an activity where risk analysts compile a catalog of situations that have the potential for harm. Hazard identification is often characterized as a purely technical element of environmental risk analysis, one in which value judgments do not enter. Nevertheless, to characterize a hazard, one must have some idea of possible harm in mind. Harm is a normative concept. It implies the value judgment that the potential events being anticipated are unwanted, are adverse or represent damage or loss. The ethical component of hazard identification specifies why and in what sense the events in question are considered to be harmful or unwanted. In many cases, the “badness” of the events in question is not particularly controversial. No controversy surrounds the judgement that negative effects on human health, especially death, are harmful. Yet even in such obvious cases, the badness or evil that is associated with death or injury reflects an ethical rather than a scientific judgment. Some environmental hazards (notably broad-spectrum toxins in chemical pesticides) also revert to forms of harm that involve human mortality and morbidity. Contamination of food supplies or even environmental exposure to pharmaceuticals or industrial products being produced through engineered crops could have direct effects on human health, but for the most part, environmental risks associated with agricultural biotechnology have involved more subtle and indirect forms of harm. When hazards do not involve harm to human health, it is particularly important to articulate and specify the reason potential events are viewed as possible forms of harm. One can apply different values to a number of the environmental hazards of agricultural biotechnology. Some values are simply prudential: it would be foolish for anyone to neglect potential outcomes that would frustrate the very aims and interests

144

6 Ethics and Environmental Risk Assessment

that support development or use of a biotech crop. Other values are more clearly ethical either because the aims and interests are of a general or more fundamental nature or because the harm is visited on someone other than the party that undertakes the action in question. Hazards such as weediness or acquired pest resistance may be associated with both prudential and ethical types of harm. No farmer wants to create new weeds; that is just more trouble for farmers. Overlooking these potential outcomes is imprudent from the farmer’s perspective. However, weediness and resistant pests also frustrate some efforts to control invasive species and protect biodiversity. These goals reflect the kind of ethical commitments that environmental philosophy articulates and defends. Although it may be obvious to a farmer or an agricultural scientist why weediness or acquired resistance is a bad thing, it may not be at all obvious to someone whose interest in the biotechnology debate derives from concern about the polluting effects of agricultural chemicals or impact on wild nature. There is thus a need for more explicit attention to the ethical underpinnings of hazard identification. Here, briefly, are some of the things that might be said about the ethics of some frequently mentioned environmental hazards.

6.4.1 Acquired Pest Resistance One of the main hazards associated with Bt crops is the possibility that constant exposure to the Bt toxin will cause resistance. Although the empirical evidence pertaining to the likelihood that this hazard will be realized has become substantially more mixed over time, it is nevertheless worth devoting some effort to a clear characterization of the harm underlying the classification of acquired pest resistance as a hazard. The value question is this: Is there any ethical significance to the possibility that insect pests might become resistant to Bt as a result of genetic engineering? Clearly companies that produce Bt crops do not desire this event, since it would negate much of the value of their product. On this account, avoiding acquired pest resistance looks more like an issue of prudence than of morality. Surely, it would be difficult to argue that the insects themselves are harmed by an evolutionary development that increases their survival rate. Though there is little evidence of concern about harm to the insects themselves in the literature on acquired pest resistance, by any standard it cannot be said that insects that acquire resistance are harmed more than the insects that are killed by Bt. Two possibilities present themselves for understanding this eventuality as an ethical problem. One stresses ecosystem impact: it is not the insects, but the ecosystem that is harmed. ‘Harm to ecosystems’ is an important class of environmental hazards that has been associated with genetic diversity as well as acquired pest resistance, and the underpinnings of seeing ecosystem impact as a form of ethically significant form of harm will be taken up below. A second possibility notes that many firms are producing seed for Bt crops and many farmers will buy it. Each firm has an incentive for utilizing strategies (such as mixing Bt seed with non-Bt seed) that minimize the risk of acquired pest resistance, as does each farmer. However, if

6.4 Understanding Hazards and Harm

145

other firms and farmers are following this strategy, then one can acquire an economic advantage by not following it, and most of the environmental benefits will still be intact. Of course since everyone knows that competing firms and farmers face this opportunity, no one wants to accept the increased risk brought on by defectors from the strategy without also getting some of the benefits. The strategy thus collapses in an instance of the assurance problem. (Thompson 2017). Since firms and farmers cannot rely on naked self-interest and prudential values to protect themselves from each other, they must resort to an enforceable norm. They must recognize that they have a duty to resist narrow self-interest in this case, and to engage in the cooperative strategy. The duty may need to be enforced by a trade association or by the government in order to be effective, but just as laws against theft are enforced by the state, the fact of enforcement does not undercut the moral basis for the norm. Note here that the norm is justified because it is necessary to help individuals act collectively to bring about the best consequences. This is a classical instance of the consequentialist pattern for endorsing norms as constraints on self-interest. It is a persuasive reason for regarding acquired pest resistance as an ethical, rather than simply as a prudential, harm. In this case, a clear statement of the ethical rationale underscores the importance of cooperative efforts both in mitigating the likelihood that the hazard will materialize, and in creating an understanding of which interests will be the basis for regulatory or compensatory action.

6.4.2 Weediness One possible hazard is that genetically engineered crops will become weeds, or that genes for herbicide tolerance will flow to wild relatives and then these herbicide resistant relatives will become weeds. To some extent, the problem of weediness exhibits the same pattern as acquired pest resistance. No farmer or biotechnology company wants weeds to become resistant to herbicides, but people must cooperate in order to assure that it does not happen. But other than that, what’s ethically bad about weeds? A weed is just a plant in the wrong place, but whether a plant is in the wrong or right place depends on a value judgment. For the most part, the weediness of herbicide tolerant plants is a straightforward form of harm to other people, rather than to the environment at large. People who are dependent on chemical herbicides for the food crops they grow (or eat) will be harmed if herbicide tolerant weeds plague those crops. Food prices will increase, and if farmers substitute more toxic herbicides, those living near fields will be harmed from that, too (Lebaron 1989). If firms and farmers make decisions about which technologies to develop and deploy based on economic returns, they are not likely to include impacts that affect others. These externalities—costs that are borne by others—must be included in a complete analysis of consequences, and doing so requires an enforceable norm. This norm stipulates that costs to all affected parties must be included in the evaluation of risk. Some have argued that the problem with herbicide tolerant crops is continued reliance on herbicides at the expense of more sustainable strategies (Comstock 1989).

146

6 Ethics and Environmental Risk Assessment

This is a subtle but crucial ethical point here. Risk assessment is comparative, not absolute. The expected value approach demands a judgment about which options to assess, and typically, the assessment is made in comparison to doing nothing, accepting the status quo. Economic cost-benefit analysis was developed to assess large public works projects. If a dam or a public policy produces more benefit than cost, it is considered justifiable because cost and benefit are measured relative to the status quo.. Yet when environmental risks of biotechnology are assessed, the relevant comparison to the status quo may be more complex than it is for a project such as a dam, or a highway. Specifically, the status quo may also be fraught with risk from chemical agriculture or from famine and pestilence and a third alternative may be far more attractive. An expected value argument is clearly vulnerable when it omits the consideration of options that are both reasonable alternatives and have substantially different expected value, (Railton 1990, 57). Although tracing out these harms from weediness as we have done above is in some respects is an unexceptional piece of deduction, making the chain of reasoning explicit makes some important (and more general) points more obvious. First, there are two kinds of completeness problems: hazard identification should specify all potentially affected parties, and a comparative risk analysis should consider all relevant alternatives. Second, the rationale for emphasizing completeness is genuinely consequential; completeness is important because it is necessary to understanding whether a technology will tend to bring about the best consequences. Finally, the harms noted so far are harms to people. Environment is the route of exposure, but the harms described are morally very familiar: economic loss, damage to health, and hunger. They do not pose the philosophical challenges of harm to non-human animals, for example. The chapter has not thus far articulated why biotechnology should have attracted the interest of people primarily concerned with protecting the environment for its own sake.

6.4.3 Genetic Diversity Adverse impacts on the diversity of organisms in a given ecosystem or on the diversity of alleles existing within an interbreeding population of organisms represents a much more broad and ill-defined class of environmental hazards, but such impacts may also be of much greater ethical significance than weediness or acquired pest resistance. Here it may be useful to discuss examples of both types of diversity. Vaccines for use on livestock that will protect against trypanosomiasis, or sleeping sickness have been sought for decades, and recombinant methods have been attempted since the 1980s (Cazorla et al. 2009). If successful, this product of biotechnology will allow the expansion of pastoral livestock production in Africa. Once livestock species can be vaccinated against trypanosomiasis it will be possible to introduce herds into areas that are not currently used for animal production. This expansion is predicted to be a substantial benefit to human beings in Eastern Africa, but will also convert wild landscapes into human-managed ecosystems. As the area of unmanaged wilderness

6.4 Understanding Hazards and Harm

147

shrinks, the habitat of threatened and endangered species shrinks with it. Large carnivores and herbaceous species can need relatively large and contiguous habits to meet their food needs. If habitat is given over to farming these species will disappear from the environment, and in some cases may become extinct. This is the classic form of biodiversity loss that is viewed as a significant environmental threat on a global basis, (Mace et al. 2014). The other form of biodiversity concerns diversity of genetic constructs within a breeding population. Impact on diversity of alleles within a species can be illustrated by the speculation that Bt maize has the potential to reduce the genetic diversity of Mexican landraces of Maize. The Mexican landraces are open-pollinated varieties that have been grown continuously by Mexican campesino farmers (e.g. small-scale farmers producing for both subsistence and limited commercial use) for centuries. There are important cultural and economic issues associated with the potential for impact on landraces of Maize, but here the environmental hazards will be the exclusive focus. These landraces maintain a much more diverse population of alleles than are found even in the foundation stocks of Maize maintained by commercial seed companies. While there is dispute about the likelihood that introgression of the Bt gene into these landraces will actually have an adverse effect on the diversity alleles, the focus here is not on exposure, but simply on hazard, that is, why a loss in the diversity of alleles might be a bad thing. The harm associated with either type of impact on diversity might be evaluated in many different ways. First, Mexican maize varieties and the larger wild environment may contain species and genes whose application for medicine or agriculture is currently unknown. Such potential applications represent speculative use value. Second, in the case of recombinant vaccines, people who enjoy hunting or observing wildlife would be deprived of this pleasure. The wildlife has a recreational value. Third, people who simply enjoy knowing that the animals or the traditional varieties are there would be harmed if they are not. The wildlife and traditional crops have existence value. Clearly, measuring these values is challenging, but speculative use value, recreational value and existence value are all recognized as relevant to human welfare, and are standard types of value studied in environmental economics. There is an additional possibility, however. The ecosystems may have intrinsic value totally apart from any human use or aesthetic appreciation of them. The suggestion that this is the case has been at the heart of philosophical debates in environmental ethics for two decades. Holmes Rolston is one of the leading proponents of this view. Rolston describes four ways in which traditional ethical thinking should be extended beyond human beings. The first of these, extension to higher animals, is consistent with philosophical views on animals described in Chap. 5, but Rolston also argues that all living organisms are “evaluative systems” that conserve interests of their own and are therefore deserving of our respect. Rolston extends moral concern beyond organisms to species, arguing, “the life the individual has is something passing through the individual as much as something it intrinsically possesses. The individual is subordinate to the species, not the other way around,” (Rolston 1991, p. 84) Rolston introduces this claim not to protest against alteration of animal telos (see Chap. 5), but as part of a chain of reasoning that ends by attributing moral

148

6 Ethics and Environmental Risk Assessment

significance to ecosystems themselves. He resolves the tension between human and ecosystem values as follows: “Humans count enough to have the right to flourish in ecosystems, but not so much that they have the right to degrade or shut down ecosystems, not at least without a burden of proof that there is an overriding cultural gain,” (Rolston 1991, p 92), and “Intrinsic value is a part in a whole and is not to be fragmented by valuing it in isolation,” (Rolston 1991, p. 95). If one follows Rolston, the expected value analysis is not complete until the ecosystem is treated as possessing intrinsic value itself. The key philosophical issue is whether to assign value to consequences that do not refer back, however elliptically, to a consequence (psychological, physical or economic) on human beings. The philosophical literature on this question is huge. Bernard Rollin rejects Rolston’s ecocentric approach soundly in his book on genetic engineering of animals, (Rollin 1995). Bryan Norton has written several detailed studies of this philosophical problem and he takes Rolston’s view more seriously than Rollin. Norton concludes that socalled anthropocentric or (human-centered) approaches may differ from eco-centric or intrinsic value approaches in terms of their cognitive content, but that when anthropocentric ethical systems are properly inclusive and far-sighted, they tend to produce the same ethical prescriptions as ecocentric ones, (Norton 1987, 1991). If Norton is right, hazards to non-human animals, to plants, to species (rather than individual organisms) or to ecosystems are isomorphic to hazards defined in terms of the way that human beings value these entities. Anyone performing an expected value risk analysis will capture these hazards one way or the other. There may yet be philosophical reasons to criticize anthropocentrism, especially as the central orientation of one’s worldview. Anthropocentrism may reinforce a person’s morally problematic understanding of their relationship to the more than human world, for example, (McShane 2007). There is also a pattern in which critics of anthropocentrism presume that it favors biotechnology (Chiarelli 2007; Marchesini 2019), while critics of biotechnology presume that someone in favor of it is anthropocentric, (Sharon 2013; Balsmeier 2019). One sociological study of Americans attempted to determine whether support and opposition to GMOs correlates with anthropocentric vs. ecocentric values, and, if so, in what direction. The results were mixed, with no clear correlation (Knight 2007). Chapter 7 will examine some hidden metaphysical commitments in the anthropocentrism debate, while Chaps. 11, 12 and 14 take up philosophical themes that might be interpreted to support the association between anthropocentrism and a favorable view of gene technologies. As for the moment, note that people who see deep metaphysical associations between anthropocentrism and agrifood biotechnology are unlikely to have much interest in incorporating their concerns in an expected value risk assessment.

6.5 Probability, Precaution and the Quantification of Exposure

149

6.5 Probability, Precaution and the Quantification of Exposure Obtaining an expected-value estimate of risk requires us to take the likelihood that hazards will actually occur into consideration, or some form of exposure quantification. While risk quantification is not typically thought of as involving ethical or philosophical dimensions, there is, in fact, a long history of philosophers and philosophically minded scientists or mathematicians noting ethical problems that can arise in the process of characterizing and quantifying exposure. Going all the way back to Stephen Stich’s article, the ethical literature on expected-value approaches to genetic engineering focused on problems in understanding probability. Stich himself raises the possibility that recombinant DNA research may lead to negative consequences that are entirely unanticipated, and notes that “it is doubtful whether there is any clear empirical sense to be made of objective probability assignments to contingencies like those we are considering,” (Stich 1978, reprinted 1989, p. 235 italics in original). Though he characterizes this problem as serious, he produces a list of factors that anticipate many of those noted by Alexander in 1985 (see above), and notes that since the eventual probability of harm is the product of the probability of all these factors, it is possible to place an upper limit on the probability of an unanticipated and catastrophic event. Stich’s 1982 article argues skeptical worries notwithstanding; experts’ subjective judgment of probabilities provides a reasonable basis for quantifying risk (Stich 1982, pp. 104–106). First, there are philosophical interpretations of uncertainty that are well known within science and risk assessment. One fairly narrow interpretation of the word “uncertainty,” relates to the statistically measurable margin of error that exists when an inference based on a particular sample is generalized to the entire population of similar cases. This narrow sense of uncertainty is often broadened to include the possibility of error that might exist due to modeling errors. Such types of uncertainty are not measurable, but the meaning is similar to that of statistical uncertainty. In either case, an estimate of risk may be off because the procedure for quantifying exposure fails to describe reality accurately. Both statistical and modeling uncertainties arise all the time in expected-value approaches to risk, and an entire repertoire of responses can be made in response to them. Choosing which response to make is an ethical issue, but, like the issues involved in relating hazard and harm, it is an ethical issue that is often addressed within the process of specifying and applying the expected-value approach to risk, (Douglas 2009; Hansson 2013). These issues are taken up in this chapter as dimensions of exposure quantification. Authors with an expressed interest in the ethical dimensions of uncertainty have contributed a great deal to risk studies, but prior to the late 1990s there were few philosophical discussions that pertained specifically to agricultural biotechnology. Since that time, a large literature has emerged in connection with the precautionary principle (or the precautionary approach), a philosophy of technology derived from Hans Jonas’ Prinzip verantwortung, (see Chap. 2). While this literature makes frequent use of the word “uncertainty” in rationalizing a preferred approach to the environmental

150

6 Ethics and Environmental Risk Assessment

(and also food safety) risks of biotechnology, a number of different arguments are made. There is first of all the distinct possibility that “attending to the uncertainties of environmental risk,” might mean exactly the same thing as attempting to address hazards and their likelihood in systematic fashion, which, of course, is just what the expected value approach attempts to do. If so, then the risk assessment and risk management methods in routine use for addressing environmental risk simply are an example of “the precautionary approach.” This is clearly what is at work in Indur Goklany’s white paper, “Applying the Precautionary Principle to Genetically Modified Crops,” (2000) though Goklany would not be regarded as a critic of biotechnology. However, most of the critics who have called for application of the precautionary principle or a precautionary approach to agrifood gene technology understand themselves to be calling for something that is not a standard component of environmental risk assessment. Carl Cranor has appealed to the precautionary principle as a specific burden of proof in tort actions for environmental damage, where plaintiffs have faced a very difficult challenge in proving that harm was the result of a specific environmental insult. This sense has been broadened slightly to advocate for the view that proof of harm should not be required before government regulatory agencies take action to ban or control a possible hazard, (Cranor 1993). This is, in fact, the language that was used in a European Community directive advocating the precautionary approach. Yet this does not constitute a deviation from standard regulatory practice using environmental risk assessment. Regulatory risk assessments do not, as a matter of principle, require “full certainty,” whatever that is. The European directive launched a full-scale international debate over the precautionary principle, or, alternatively, precautionary approach. Gary Comstock gave philosophical voice to the putatively American side of this debate, arguing, in effect that since everything involves uncertainty, the precautionary approach is vacuous, (Comstock 2000, see also Pence 2002). Richard Soule offered a more nuanced interpretation narrowly focused on environmental impact. He drew a distinction between a weak version of precaution, in which decision makers are permitted to avoid risk as opposed to balancing risk against benefits, and a strong version in which avoidance of environmental risk trumps everything,. Soule focused on trade implications related to gene-modified plants, worrying that even a weak version of precaution could give nations a rationale for favoring domestic interests that might suffer economic, rather than environmental, harm, (Soule 2000; see also Charlier and Rainelli 2002). Soule’s emphasis on using the precautionary approach as a pretext, rather than a decision principle, reflects the opinion of the United States biotechnology industry in the early years of the debate. Europeans who were arguing for a precautionary approach took a considerably more nuanced stand, arguing that precaution marks the appropriate stance between permissiveness and prevention. Precautionary acts are justified when there are reasons to think a seriously undesirable outcome might occur, and when the act would make that outcome less likely, (Sandin 2004). It is plausible to suppose that some European advocates of a precautionary approach would be entirely satisfied with a wellconducted expected-value analysis of the environmental risks from gene technology. They might argue for a decision-rule that is more prejudicial to downside outcomes

6.5 Probability, Precaution and the Quantification of Exposure

151

than the utilitarian maxim or the Pareto Principle, (e.g. Soule’s weak version of the precautionary principle) but they would have no objections to an expected-value characterization of environmental risk. Yet others interpret the precautionary principle as calling for social and cultural impacts to be reflected in regulatory decision making, along with food safety and environmental impact. They see this (correctly) as in direct opposition to U.S. policy (Levidow 2001; Tait 2001; Carr 2002). If this view of precaution calls for expanding the range of hazards examined in a risk-based approach (see Devos et al. 2008), it is consistent with the argument put forward in every edition of this book. However, if it is viewed as a rejection of risk-based ethical analysis, (see Levidow and Carr 1997), I am opposed. Although the precautionary debate is, with some fairness, characterized as dispute between America and Europe, more contextual analysis of the U.S. perspective demonstrates that U.S. regulatory agencies are more receptive to precaution than generally presumed. Cranor’s initial study focused on U.S. tort cases, where juries were instructed to judge whether the preponderance of evidence supports the claim that a defendant’s action caused harm. He argued that this was being interpreted as a burden of proof that no standard toxicological or carcinogenic risk analysis could meet, given the problem of scientific uncertainty. Courts needed to revise their understanding of actionable certainty. The context for regulation is more complex. The U.S. Environmental Protection Agency’s (EPA) authority to regulate chemicals under the U.S. Toxic Substances Control Act (ToSCA) of 1976 is rather weak. Cranor argues that in too many cases, EPA needs evidence of post-market injuries before taking action under ToSCA, (Cranor 1993). However, the pre-market evaluation of agricultural and food biotechnologies is regulated not under ToSCA, but under a suite of laws referred to as the Coordinated Framework (see Chap. 4). The U.S. Department of Agriculture’s Animal and Plant Health Inspection Service (APHIS) is the U.S. agency charged with the broadest responsibility for protecting the environment from the risks of transgenic plants. Their regulatory approach for biotechnology has been derived from almost a century of conducting risk analyses and making regulatory decisions regarding the importation of exotic plants and animals into the U.S. and attempting to control the inadvertent import of plant and animal disease. APHIS has never applied a standard remotely approaching scientific certainty of harm in making these regulatory decisions. It is, of course, possible to argue that APHIS should have applied even more precaution than they historically have, or at least that they should be more cautious as they undertake decisions involving transgenic plants. Notice, however, that this critique concerns the relative level of precaution, rather than an entirely different approach to risk. Other regulatory agencies apply very different standards in their decision rules, and it is quite plausible to regard some of the precautionary principle rhetoric as an inchoate critique of these principles. Under the Coordinated Framework, the EPA regulates the pesticide aspects of crops such as Bt maize or cotton (discussed in Chap. 1) under a statute that requires them to consider the benefits of the pesticide to U.S. farmers. Although they are given some leeway in evaluating the trade-off between farmer benefit and environmental risk, risk assessments

152

6 Ethics and Environmental Risk Assessment

conducted under the auspices of the Federal Insecticide, Fungicide and Rodenticide Act (FIFRA) of 1972 do reflect uncertainties into the computation of expected value prior to making such comparisons. If what one objects to is the possibility that economic benefits to farmers could offset risks to the environment, the target of one’s objection is the decision-rule, not the practice of using expected values to characterize environmental risks. Food safety risks (discussed at length in Chap. 4) could also be the focus of a precautionary principle critique. The Organization for Economic Cooperation and Development and the U.S. Food and Drug Administration (FDA), for example, have each developed a principle of “substantial equivalence” for determining which new foods and food products (including those derived from biotechnology) need to undergo strenuous regulatory review for food safety. These are, in effect, pre-risk assessment risk assessments, whereby some broad principles of guidance are applied to determine whether more detailed processes of hazard identification and exposure quantification are warranted. There is considerable debate as to whether this policy adequately protects public health (see Millstone, Brunner and Mayer 1999 and Miller 1999 for two perspectives on this debate). It is reasonable for critics of substantial equivalence to describe themselves as taking a more precautionary approach. They are arguing for collection and evaluation of data that is not currently required in the U.S. (again, see Chap. 4). Yet it is still difficult to interpret such a claim as calling for an alternative to expected-value assessments, since someone who rejects the principle of substantial equivalence is calling for more scrutiny using expected-value risk assessment. What is more, FDA’s practice of applying the de minimus decision rule certainly qualifies as a precautionary approach in Soule’s weak sense. FDA does not weigh benefit against risk but instead regulates food safety on the basis of lowest feasible risk. In this respect, at least, U.S. food safety policy conform to what Per Sandin has characterized as a precautionary approach, (Sandin 2006). In sum, Americans who were arguing for the precautionary principle were criticizing regulatory practices guided by ToSCA and FIFRA, and may not have appreciated how practices at APHIS and FDA were already applying a version of the precautionary approach. Europeans observing this debate might not have understood this either. U.S. rhetoric blasting the precautionary principle over what some Americans saw as a protectionist trade action did not help. As such, the increasingly verbal dispute obscured a set of persistent philosophical problems that arise in operationalizing a risk-based approach. Some years after this controversy had passed its zenith, Sven Ove Hanssen argued that a consistent application of the precautionary approach now calls for stance that is more favorable to genetically engineered crops, (Hanssen 2016). None of this should be taken to imply that there are no problems with an expected-value interpretation of risk. The balance of this chapter will discuss conceptual weaknesses that must be guarded against, while Chap. 7 will take up ethical issues that challenge the approach more deeply. At the same time, I am at a loss to explain why anyone whose objections to expected value track closely with mine would abandon a risk-based approach altogether. I now have a better understanding of why they describe these objections in terms of the precautionary principle than I did in 1997 or 2007. I also take some comfort in the work of philosophers

6.5 Probability, Precaution and the Quantification of Exposure

153

who agree that advocating precaution as if none were being taken already is not the most propitious way to address the values that are necessary for deploying an expected-value risk analysis, (see van den Belt 2003; Scott 2018).

6.6 Exposure Quantification: Further Issues There are also some interesting and ethically significant aspects of exposure quantification that escaped the attention of the precautionary crowd. Robert Wachbroit made a detailed and undeservedly overlooked analysis in 1991, but some of the key points were anticipated by Stich in 1978. Wachbroit makes three basic criticisms of the way that expected-value risk assessments are applied to gene technology. First, he notes that potential harms may be “thin,” (confined to human mortality and morbidity) or “thick,” (including social harms that accrue as psychological costs, even when none of the harmful events are realized). Wachbroit then points out cognitive problems in deciding how evidence bears on probability assignments. Finally, Wachbroit concludes by noting that even when risks are assessed adequately, the communication of expected value results may be politically and ethically problematic, (Wachbroit 1991, 367–377). Using the expected value or risk-benefit framework requires key philosophical value commitments in each of the problem areas Wachbroit notes. His comment about “thick” hazard identification is that social and cultural hazards should be included along with regulatory actionable hazards like threats to health. This is a major theme of this book, reflected in the way that I have organized the discussion around four ontologically distinct categories of hazard. I agree with Wachbroit on this point. Furthermore, it has been a continuing basis of complaint, surfacing again in discussions about the incorporation of “non-safety” concerns into Norwegian biotechnology assessments, (Myskja and Myhr 2020). Concern over the thickness of GMO hazard identification (to use Wachbroit’s language) is valid, but I will not discuss it further in this chapter. Wachbroit’s third point emphasizes communication issues, discussed in the following section, as well as Chap. 12. It is his second point of criticism that concerns us here. Wachbroit notes two conceptual issues associated with assigning probabilities to hazards that are not typical of precautionary principle arguments that stress uncertainty. First, he cites cognitive science research on bias in eliciting subjective probabilities. Even experts appear to be vulnerable to these biases. Wachbroit argues that the potential for expert bias introduces a different kind of uncertainty into the expected value approach, one with ethical implications for the use risk-benefit comparisons, (Wachbroit 1991). Second, he notes that the approach to exposure quantification outlined by Alexander proposes a series of independent events that must take place before harm occurs. He echoes a point made by Stich. People who think that many different things have to happen before a failure occurs tend to think that the chance of all of them happening makes failure very unlikely, but that is not necessarily the case, (Stich 1978, reprinted 1989).

154

6 Ethics and Environmental Risk Assessment

Heuristics and Bias in Probability Assignments. Although it is possible to conduct empirical studies and collect statistics, many of the probabilities for environmental risk of food biotechnology are derived from expert judgments. These judgments are informed both by experimental results from studies of the mechanisms believed to underlie environmental risk, but in fact they rely quite heavily on experience with plant, animal and microbial systems. In 1991, Wachbroit argued that objective or statistical notions of probability are simply not meaningful for evaluating the environmental risk of genetic engineering because statisticians understand probability as the long-run frequency for events of a given type. One can assess frequency for gene expression or gene flow, and these frequencies bear on the risk that a specific harmful event will occur, but there is no frequency for a single event. As such, experts are using this information to form opinions about the risk of an event. If experts disagree, as they clearly do, then “we are left in the dark about the probability of a single case. And the probability of a single case may matter,” (Wachbroit 1991, 374). What is more, well known psychological studies have documented that human beings incorporate heuristic tools in their subjective risk judgments, and these tools are capable of introducing systematic bias into the assessment of risk. Although experts may perform better than lay persons in forming risk estimates when they are dealing with matters closely related to their area of expertise, they fall victim to these biases when they attempt to synthesize their knowledge beyond their expertise (Tversky and Kahneman 1982; Hollander 1991). The upshot of all these problems is that probabilities are subject to error, and that the expected value paradigm does not give us clear guidelines about how to cope with this problem. This problem led Kristin Shradrer-Frechette to argue that decision makers should be obligated to consider multiple risk assessments developed by groups with substantially different interests. She writes that risk decisions can be made more democratic if something like a science court is used. Scientifically, statistically competent judges would weigh competing assessments of risk, and would be required to write opinions stating why they have chosen to favor one assessment over the other, (Shrader-Frechette 1991). European research groups have experimented with these approaches, and some of their results will be discussed in later chapters, (see Hennen 2012). Shrader-Frechette’s proposal is important in the present context because it stresses the need for articulated reasons behind the judgments about subjectively derived probabilities. However, reason giving can readily be incorporated within the expected-value framework, so this is not an objection to the risk-based approach. The issue of cognitive bias is implicit in more widely known criticisms made by Brian Wynne. Wynne’s justly celebrated study of errors in exposure to radiological hazards after the Chernobyl accident showed not only that expert judgment can be highly fallible, but also that the overconfidence of experts can actually be a source of risk for the wider public, (Wynne 1992). He extended this argument to make inferences about the predictions of risk from genetically engineered crops, (see Wynne 2001; Hoffmann-Riem and Wynne 2002). Wynne’s writings from the early 2000s were read as endorsing a precautionary approach, but the point being argued by Wachbroit qualifies the applicability of statistical exposure assessments specifically with respect to single case or one-off events. Here, experts must rely on an inductive

6.6 Exposure Quantification: Further Issues

155

inference that can be easily influenced by cognitive biases. Indeed, the philosophical problem has recently been characterized in terms of inductive risk, (Douglas 2009). Europeans had followed Sven Ove Hansson’s development of similar points. A more detailed follow-up on this problem would need to examine the grounds for relying upon experience with traditional crop breeding or mutation breeding (see Chap. 1) for inductions on the environmental risks of transgenic or gene-edited crops. The Small Probabilities Argument. Wachbroit describes how sources of uncertainty should lead us to be cautions in the confidence with which we regard expected values. However, there are also uncertainty/probability arguments that lead people to argue for less caution with respect to agricultural biotechnology. Stephen Stich reviewed the possibility that a human health epidemic might be caused an enfeebled strain of E. coli. This organism was produced in the wake of the 1976 Asilomar conference and was one of the first genetically engineered organisms ever created. Stich concluded that since the probability that this organism would survive and replicate outside the lab is clearly very low, even the highest estimate for the probability of a “worst case” scenario must be very low. If the benefits outweigh risks in the most likely scenarios for a worst-case result, then “lower estimates of the same probabilities will, of course, yield the same conclusion,” (Stich 1978, reprinted 1989, 236). It is clear that many scientists who have reviewed the risks of agricultural genetic engineering (including Alexander himself (1985)), have employed a similar pattern of reasoning, (see Davis 1987; Curtiss 1988; Adelberg 1988). What is ethically important about this form of argument is that if one adopts an expected value approach to ethical issues, and if the probability of any harmful environmental consequences is exceedingly low, then there is little point in debating many of the philosophical questions described above. When the probability of harm is a function of many independent (or logically redundant) probabilities, it is sure be very low. This is the mathematical result of multiplying any series of numbers between zero and 1. Even a relative large probability (.9 or 90%) falls rapidly when a sequence of six independent events are combined (e.g. 0.9 × 0.9 × 0.9 × 0.9 × 0.9 × 0.9 = 0.53, or close to only 50%). If independent probabilities are in the range of one in a hundred to start with, the combined probability discounts their expected value to zero, or near zero, even when the projected harms themselves are significant. When probabilities shrink sufficiently, the event may be ignored in risk-benefit decisionmaking. Taking this view of the risks leads scientists to conclude that debate is a waste of time, (see Trewavas 1999; Conko and Miller 2011; Conko et al. 2016). Of course, it matters a great deal whether the probabilities really are low, and this fact may explain why so much of the technical literature has focused on that question. Nassim Taleb has popularized the tern “black swan” to describe very low probability events which, he argues, tend to be neglected to our peril, (Taleb 2010). Several authors have applied the idea in discussing the potential for low-probability but high impact events from agrifood genetic engineering, (Raybould 2010; Murphy and Connor 2012). However, it is also possible that small probabilities are an artifact of complexity and redundancy in the way that one conceptualizes or models exposure. If one thinks

156

6 Ethics and Environmental Risk Assessment

that exposure is the result of a long chain of independent events, each with a probability less than 1, then the length of chain itself begins to create the impression that exposure is unlikely. The final occurrence of the harm depends upon each event happening in succession, and each event’s happening is conditional upon every previous event already having happened. This means that quantification process is a long multiplication problem with each unit in the problem having a value between 0 and 1. Even if each event has a probability of 50% in Alexander’s six-link chain, the final probability of harm will be 0.015625 or less than 2%. As chains become even longer because they are seen as composing more and more independent events, exposure becomes less and less simply in virtue of modeling complexity. Should this bother us? As far as I can discern, this remains a relatively undiscussed problem in the literature on risk from biotechnology, but doubts of this sort may be behind the inchoate worries expressed by ecologists who decry reductionist thinking.

6.7 Philosophy of Science and Risk Communication Given expected-value quantification’s systematic vulnerability to error, anyone hoping to deploy a risk-based approach must take measures to ensure that probability estimates are as accurate as possible. Although statistical methods are a critical element of these measures, this is not a purely mathematical or value free exercise. The philosophical ideas needed to implement these measures derive from epistemology and philosophy of science as much or more than they do from ethics. These fields of philosophical inquiry changed significantly since the first edition of the book. One problem is that scientists are trained to minimize the chance of accepting a false claim. This norm protects the integrity of the theory building process, but is it appropriate when assessing risk? Two American philosophers of science, Carl Cranor (1993) and Kristin Shrader-Frechette (1991), argued that the appropriate norm in risk assessment is to minimize the chance of failing to accept a true claim, or the chance of failing to anticipate and prepare for a risk that might be present, even when the data is insufficient to prove that it is. This problem was formulated in epistemological terms by philosopher Nicholas Rescher in the 1980s, though Heather Douglas argues that philosophers of science such as Richard Rudner and C. West Churchman had been discussing it in the 1960s (Douglas 2009). Rescher argued that uncertainties can plague not only the statistical significance of data, but our knowledge of what might happen, or how harmful any given event might be in the long run. Uncertainty is not merely a problem in quantifying the probability that a hazard will occur, it is a problem in the inductive identification of hazards, (Rescher 1983, 94–95). Although this situation need not lead one to reject the expected value approach, ethical decision making under this form of uncertainty “becomes a matter of comparing not expected values as such, but ranges of expectation,” (Rescher 1983, 102) Rescher presented three alternatives for coping with this problem: make an expected value decision based on the median of the range, on the worst case scenario, or on the best case

6.7 Philosophy of Science and Risk Communication

157

projection. It is not obvious which of these alternatives is most consistent with the consequentialist desire to pursue the best outcome. Individuals undoubtedly make their own choices based on whether their personalities tend toward risk-taking or risk aversion, but it is not immediately clear how to translate such personality factors into the sphere of public choice. Rescher’s work shows analytically why many probiotechnology scientists would be ready to move ahead with gene technology, while others would not, but it provides no ethical basis for deciding which of them is right. In the years since the last edition of Food Biotechnology in Ethical Perspective was updated, this general class of problems has become much more widely recognized among American philosophers of science. A heavily cited paper by Douglas in the journal Philosophy of Science described how normal science relies on value judgments to resolve basic epistemological uncertainties, (Douglas 2000). Douglas’s article has invigorated the study of agricultural and food biotechnology from a philosophy of science standpoint, and many authors are focusing on risks, (see Coutellec and Doussan 2012; Hicks 2015; Rocca and Anderson 2017; Biddle 2018). This new round of research is especially useful for the way in which conceptual issues discussed in this book are fleshed out with empirical details drawn from risk assessments of specific applications of gene transfer. As American philosophers of science have become more cognizant of the role that these value judgments play in the philosophy of science, the earlier work of Churchman and Rudner is becoming more fully appreciated. An even more historically sensitive treatment would incorporate the earlier work by Stich, Rollin and Wachbroit into their analyses. What is more, parallel developments were taking place in other philosophical discourse communities. Brian Wynne was producing work that touched upon many of these same issues from the perspective of critical science studies, (Wynne 1983, 1988) My own work on biotechnology was preceded by a set of papers on conceptual and philosophical issues in assessing technological risks published in the 1980s, (see Thompson 1983–84, 1986, 1987; Thompson and Parkinson 1984). At the same time, Sven Ove Hansson was launching a research program on the interconnection between epistemic and ethico-political values based on his work in decision theory, (see Hansson 1987, 1989). There was very little cross citation between these research streams in philosophy of science, science studies, philosophy of technology and decision theory, creating a situation emblematic of deep irony in the epistemology of risk. Scholars in one discipline are unaware of related work in another discipline. They presume that they themselves are on the forefront, as researchers from other quarters are unknown to them (and often lack credentials such as affiliations in top programs or personal connections to well-known individuals in the relevant field). Since scholars in other areas are unaware of their work, they infer that they are incompetent, unable to apply and perhaps even understand the issues. A culture of mistrust among scholars mimics the virtue-risk feedback loop discussed in Chap. 4. Speaking for myself, although I was familiar with Rescher, Shrader-Frechette and Cranor when I wrote the first edition in 1997, I was at best dimly aware of work by Wynne. I did not become familiar with Hansson’s work until most of the revisions for the second edition were complete.

158

6 Ethics and Environmental Risk Assessment

If I can use the phrase “culture of mistrust” to describe a phenomenon that generates yet another form of uncertainty in risk studies, these overlapping but isolated communities of risk scholars reflect the deeply reflexive nature of risk. Many conceptual and philosophical advantages of the expected value framework evaporate under circumstances where a large number of people must endorse decisions. In the case at hand, I do not think that philosophers’ skepticism of each other’s work had a great deal of influence on the biotechnology debate, but the lack of a unified or even mutually cognizant voice coming from philosophers cannot have helped. Biological scientists would have seen a motley group lacking in the type of rigor indicated by a well-structure pattern of interaction and citation. Why would they infer that this disorganized group of humanities specialists would say anything relevant to the risk of gene technologies? What is more, cultures of mistrust within the knowledge-production establishment erode public confidence, fueling a virtue-risk feedback loop: If these idiots can’t even talk to each other, they can’t be trusted. If they cannot be trusted, their products cannot be trusted, and so on. Simply put, ill-informed people can disrupt the orderly development and deployment of any technology, including technologies that have, after thorough risk analysis, been shown to have very attractive risk-benefit ratios when compared to available options. Scientists and policy makers working with chemical technology and nuclear power learned this lesson through cruel experience (Covello, Sandman and Slovic 1991). Most participants in research on or commercial development of food biotechnology have already learned that public opposition can cause delays and can even sabotage promising products. Schurman and Munro report that between 1990 and 2010—the most acrimonious years of debate over gene technology—scientists and industry insiders believed that their opponents were ignorant and ill-disposed toward even thinking about the complexities in risk-based thinking, (Schurman and Munro 2010, pp. 28–32). However, Schurman and Munro themselves omit all discussion of what I take to be the most important U.S. effort to engage in a communicative discourse on the risks of agricultural biotechnology. From the early 1990s through the mid-2010s, a consortium of universities conducting research in agrifood biotechnology held annual meetings under the auspices of the National Agricultural Biotechnology Council, (NABC). In the early years, these meetings convened discussions with both advocates and critics of agricultural biotechnology. Virtually of these meetings concluded by stressing the need for more public education. Both boosters and knockers supported increasing public understanding of the risks from agrifood biotechnology; it was, in fact, one of the few things on which they could agree. I attended most of the annual NABC meetings between 1989 and 2004, as well as other consensus meetings organized by the Keystone Foundation and, later the Pew Charitable Trusts, (see Ten Eyck et al. 2001 and Thompson 2008 for a more detailed discussion). The experience contributed to the more optimistic view I expressed in the first edition of that book. I believe that cultures of mistrust may have undercut the potential of these efforts, as they were completely ignored and disrespected by scholars in science studies. More pertinent in the present context, the expected value approach to ethical responsibility

6.7 Philosophy of Science and Risk Communication

159

produces a very different rationale for achieving consensus than do other philosophical approaches. For the expected-value theorist, risk communication is needed because ill-informed people will ruin the best laid plans. It is not that people are entitled to this information, nor are they entitled to participate in the information seeking elements of an expected value analysis. From a pure consequentialist perspective, people must be informed only because failing to inform them introduces unnecessary costs. This lack of entitlement to information under the expected value approach is philosophically crucial for understanding debates over risks associated with agrifood biotechnology. The expected value approach provides broad guidelines for comparing the chance of harmful outcomes from food biotechnology to the chance of beneficial outcomes. It is philosophically committed to the idea that the ethical course of action is the one that brings about the best outcome. It would seem that to execute this approach, a decision maker must simply assemble the best information available, and then do the right thing. There is no obvious place where the expected value approach requires a decision maker to share information with others, even if they are affected parties. Indeed, the question of whether information should be shared (or communication undertaken) must be subjected to the utilitarian maxim: Does doing so increase the amount of good for the greatest number? Someone who discharges the ethical responsibilities demanded by the expected value approach will have taken their interests into account already. Affected parties might have been surveyed to determine their preferences, for example. A decision maker will want to do that because this kind of information helps determine whether an outcome is beneficial or not. Alternatively, one might assume that market acceptance of the end product is an adequate indicator of people’s values. In any case, communicating with the public may or may not involve informing them about risks. Wachbroit makes a further point in this regard, noting that the expected value approach relegates communication to a role of “handling” the public,” rather than informing them, (Wachbroit 1991, 374). He calls this formulation of the ethical responsibilities associated with genetic engineering “tendentious.” Clearly he is right to note that people do not like to be handled. They react with justifiable suspicion when they think that representatives of the science community are patronizing their concerns. If this tendentiousness is experienced as arrogance on the part of the food biotechnology community, preventing the erosion of public confidence will be made all the more difficult (see Thompson 1995, for my take on this problem). Ironically, following out the general prescription of the expected value approach may result in conduct which produces anything but the best consequences! Consequential or expected value decision making adopts a scientific notion of objectivity in the sense that it strives for an impartial account of the best outcome, but it ends in paradox when it fails to recognize the strategic dimensions of acting in pursuit of the best outcome. In this context, a choice is “strategic” whenever the outcomes depend not only on what the decision maker selects, but on how other people act in response. Applying the expected value approach to environmental risk questions seems at first to be an instance of non-strategic choice. One examines the chance that any given product of food biotechnology will result in harm, either to other

160

6 Ethics and Environmental Risk Assessment

humans to the environment itself. If the assessment is that the probability of harm is very, very low, ethics seems to weigh in favor of pursing the technology, provided that expected benefits outweigh the standard costs of research and development. However, acting on this analysis, objective and unbiased though it may be, appears to provoke a ruinous response from the public. This leads next to the educational strategy: Let the public be informed! Ironically, this course of action has the strategic consequence of infuriating them further. The public does not like to be educated solely for the strategic purpose of “handling” their dissent. Their resentment erodes their trust in the handlers, which is to say that it leads them to mistrust science. What was at first only ruinous with respect to a specific product threatens to become ruinous for the entire food biotechnology industry. What can one do now? This is an extraordinary question that has bedeviled many analysts of science and risk (see, for example, van Dommelen 1995) Arguably, what one must do (now for consequential reasons) is what one would have done if one would have never been tempted by the expected value framework, in the first place. In the years since the first edition of this book, scholars in science communication have taken up these themes, emphasizing deficiencies in “the deficit model” (see Trench 2008 for an overview). As discussed at some length in Chap. 12, I view my work in earlier versions of this book as contributing to critiques of the deficit model, even if I did not use that terminology.

6.8 Conclusion The argument in this chapter builds upon a risk-based approach to technological ethics. Ethicists should consider what hazards (e.g. bad outcomes) might follow the introduction of a GMO or gene-edited plant or animal. This is not philosophically trivial, and one weakness in scientific discourse has been the simplistic way in which the badness or adversity of an outcome has been analyzed, especially in the environmental domain. But ethicists should also consider how likely it is that these outcomes will actually materialize. Such considerations usually lead risk analysts to an expected-value interpretation of risk. Only then is one in a position to have a further debate about what should be done (e.g. risk management). I have argued that this approach is considerably more flexible and hence more capable than many critics have alleged. However, I have also followed Robert Wachbroit in noting key points at which the approach can be abused, if analysts and other participants in an ethical conversation are not mindful. Although I have expanded this argument to include developments since the publication of the second edition in 2007, this general framing of the chapter remains true even to the first edition. This chapter has surveyed both strengths and weaknesses of expected-value risk assessment as a tool for anticipating and governing the environmental impact of transgenic crops and livestock, as well as gene-edited organisms that may appear in the future without identifiable transgenes. The focus of this survey has been to link epistemological problems with normative questions about the environmental impact

6.8 Conclusion

161

of gene technologies in agriculture and food production. The survey does not reflect worries associated with the expansion of agriculture into uncultivated areas (though a livestock-related example of this problem is discussed in Chap. 5). In addition the revision does not update its review of environmental concerns to include more radical forms of biotechnology such as gene drives or synthetic biology, (examples are discussed in Chap. 13). In fact, many of the philosophical points made throughout the chapter might be applied beyond environmental impacts to include aspects of food safety, animal health and even socioeconomic transitions. Although expected-value risk assessment shares important features with utilitarian moral theory, there are important differences between the consequentialism implicit in risk assessments and the classic form of welfare maximizing utilitarianism. Much advocacy for greater precaution or the precautionary principle is fully compatible with the logic and epistemic norms of expected-value risk assessment. Critiques that are mounted against gene technology citing environmental risks would be more focused (and arguably more persuasive) if the extent to which precautionary measures are incorporated into existing governance mechanisms were better understood. At the same time, pro-biotechnology advocates seldom recognize the weaknesses in the expected value approach, and often argue as if it were simply an expression of rationality tout court. Recognizing some weaknesses in the approach would, for them, lead to more convincing arguments that better address the points that are actually being contested. Recent philosophical work on inductive risk is bringing new clarity to this dimension of the GMO debate, and this update of Food Biotechnology in Ethical Perspective has not attempted to incorporate all of the important observations that studies published after 1997 have made. As with some other chapters in the 3rd edition, much of the value that the preceding survey will have for future scholars in the ethics of science and technology consists in reminding the new generation of scholars of important work that was neglected by mainstream philosophy of science in the 1980s, 1990s and even through the most contentious period of debate over GMOs prior to 2010.

References Adelberg, E.A. 1988. Closing address. In The Release of Genetically-engineered Micro-organisms, ed. M. Sussman, C.H. Collins, F.A. Skinner, and D.E. Stewart-Tull, 293–297. London: Academic Press. Alexander, M. 1985. Ecological consequences: Reducing the uncertainties. Issues in Science and Technology 13: 57–68. Balsmeier, P. 2019. Towards a posthumanist conceptualization of society: Biotechnology in Margaret Atwood’s MaddAddam trilogy and Ruth Ozeki’s All Over Creation. In: Representations of Science in Twenty-First-Century Fiction, ed. by N. Englehardt and J. Hoydis, 93–111, New York: Palgrave Macmillan. Biddle, J.B. 2018. “Antiscience zealotry”? Values, epistemic risk, and the GMO debate. Philosophy of Science 85: 360–379.

162

6 Ethics and Environmental Risk Assessment

Carr, S. 2002. Ethical and value-based aspects of the European Commission’s precautionary principle. Journal of Agricultural and Environmental Ethics 15: 31–38. Charlier, C., and M. Rainelli. 2002. Hormones, risk management, precaution and protectionism: An analysis of the dispute on hormone-treated beef between the European Union and the United States. European Journal of Law and Economics 14: 83–97. Cazorla, S.I., F.M. Frank, and E.L. Malchiodi. 2009. Vaccination approaches against Trypanosoma cruzi infection.”. Expert Review of Vaccines 8: 921–935. Chiarelli, B. 2007. Ethical anthropocentrism: Humanistic ethics and the need for a new global bioethics. Mankind Quarterly 47: 105–115. Comstock, G. 1989. Is Genetically engineered herbicide-resistance GEHR compatible with lowinput sustainable agriculture LISA? In Biotechnology and Sustainable Agriculture: Policy Alternatives, ed. J.F. MacDonald, 111–123. Ithaca, NY: National Agricultural Biotechnology Council. Comstock, G. 2000. Vexing Nature?. On the Ethical Case against Agricultural Biotechnology: Kluwer Academic Publishers, Boston. Conko, G. and H.I. Miller. (2011) The rush to condemn genetically modified crops. Policy Review 165:. 69+. Gale Academic OneFile, https://link.gale.com/apps/doc/A249312478/AONE?u=msu_ main&sid=AONE&xid=65a71529. Accessed 18 Feb. 2020. Conko, G., D.L. Kershen, H.I. Miller, and W.A. Parrott. 2016. A risk-based approach to the regulation of genetically engineered organisms. Nature Biotechnology 34: 493–503. Coutellec, L. and I. Doussan. 2012. Legal and ethical apprehensions regarding relational object. The case of genetically modified fish. Journal of Agricultural and Environmental Ethics 25: 861–875. Cranor, C.F. 1993. Regulating Toxic Substances: A Philosophy of Science and the Law. New York: Oxford University Press. Curtiss III, R. 1988. Engineering organisms for safety: What is necessary? In The Release of Genetically-engineered Micro-organisms, ed. M. Sussman, C.H. Collins, F.A. Skinner, and D.E. Stewart-Tull, 7–20. London: Academic Press. Covello, V., P. Sandman, and P. Slovic. 1991. Guidelines for communicating information about chemical risks effectively and responsibly. In Acceptable Evidence: Science and Values in Risk Management, ed. D.G. Mayo and R.D. Hollander, 66–90. New York: Oxford University Press. Davis, B. 1987. Bacterial domestication: Underlying assumptions. Science 235: 1329–1335. Devos, Y., P. Maeseele, D. Rehuel, L. Van Speybroeck, and D. de Waele. 2008. Ethics in the societal debate on genetically modified organisms: A (re)quest for Sense and Sensibility. Journal of Agricultural and Environmental Ethics 21: 29–61. Douglas, H. 2000. Inductive risk and values in science. Philosophy of Science 67: 559–579. Douglas, H. 2009. Science, Policy and the Value-Free Ideal. Pittsburg, PA: University of Pittsburg Press. Gibbard, A. 1986. Risk and value. In Values at Risk, ed. D. MacLean, 94–112. Totowa, NJ: Rowman and Allanheld. Goklany, I. 2000. reprinted 2002) Applying the precautionary principle to genetically modified crops. In Genetically Modified Foods: Debating Biotechnology, ed. M. Ruse and D. Castle, 265–291. Amherst, NY: Prometheus Books. Hansson, S.O. 1987. Risk decisions and nuclear waste (No. SKN–19). National Board for Spent Nuclear Fuel, Stockholm. Hansson, S.O. 1989. Dimensions of risk. Risk Analysis 9: 107–112. Hansson, S.O. 2009. From the casino to the jungle. Synthese 168: 423–432. Hansson, S.O. 2013. The Ethics of Risk: Ethical Analysis in an Uncertain World. New York: Palgrave Macmillan. Hansson, S.O. 2016. How to be cautious but open to learning: Time to update biotechnology and GMO legislation. Risk Analysis 36: 1513–1517. Hennen, L. 2012. Parliamentary technology assessment in Europe and the role of public participation, In: Public Engagement and Emerging Technologies, ed. by K. O’Doherty and E. Einseidel, 27–44, Vancouver: The UBC Press.

References

163

Hicks, D.J. 2015. Epistemological depth in a GM crops controversy. Studies in History and Philosophy of Science Part C: Studies in History and Philosophy of Biological and Biomedical Sciences 50: 1–12. Hino, A. 1994. Safety assessment system of field tests in Japan. In: Proceedings of the 3rd International Symposium on Biosafety Results of Field Tests of Genetically Modified Plants and Microorganisms ed. by D.D. Jones, 177, 182. University of California, Division of Agriculture and Natural Resources, Oakland, CA. Hoffmann-Riem, H. and B. Wynne. 2002. In risk assessment, one has to admit ignorance. Nature 416: 123–123. Hollander, R.D. 1991. Expert claims and social decisions: Science politics and responsibility. In Acceptable Evidence: Science and Values in Risk Management, ed. D.G. Mayo and R.D. Hollander, 160–173. New York: Oxford University Press. Howard, T. and J. Rifkin. 1977. Who Should Play God? The Artificial Creation of Life and What It Means for the Human Race. New York: Delacorte Press. Huttner, S. 1993. Risk and reason: An assessment of APHIS. In US Agricultural Research: Strategic Opportunities and Options, ed. R.D. Weaver, 155–168. Bethesda, MD: Agricultural Research Institute. Kelman, S. 1981. Cost Benefit Analysis: An ethical critique. Regulation, 74–82. Knight, A. 2007. Do worldviews matter? Post-materialist, environmental, and scientific/technological worldviews and support for agricultural biotechnology applications. Journal of Risk Research 10: 1047–1063. LeBaron, H. 1989. Herbicide resistance in plants. In Biotechnology and Sustainable Agriculture: Policy Alternatives, ed. J.F. MacDonald, 92–102. Ithaca, NY: National Agricultural Biotechnology Council. Leonard, H.B., and R.J. Zeckhauser. 1986. Cost-benefit analysis applied to risks: Its philosophy and legitimacy. In Values at Risk, ed. D. MacLean, 31–48. Totowa, NJ: Rowman & Allanheld. Levidow, L. 2001. Precautionary uncertainty: Regulating GM crops in Europe. Social Studies of Science 31: 842–874. Levidow, L., and S. Carr. 1997. How biotechnology regulation sets a risk/ethics boundary. Agriculture and Human Values 14: 29–43. Levin, M.A., and H. Strauss (eds.). 1991. Risk Assessment in Genetic Engineering. New York: McGraw-Hill. Mace, G.M., B. Reyers, R. Alkemade, R. Biggs, F.S. Chapin III, S.E. Cornell, S. Díaz, S. Jennings, P. Leadley, P.J. Mumby, A. Purvis, R.J. Scholes, A.W.R. Seddon, M. Solan, W. Steffen, and G. Woodward. 2014. Approaches to defining a planetary boundary for biodiversity. Global Environmental Change 28: 289–297. MacLean, D. 1986. Risk and consent: Philosophical issues for centralized decisions. In Values at Risk, ed. D. MacLean, 1–16. Totowa, NJ: Rowman and Allanheld. MacLean, D. 1990. Comparing values in environmental policies: Moral issues and moral arguments. In Valuing Health Risks, Costs, and Benefits for Environmental Decision Making, ed. P.B. Hammond and R. Coppock, 83–106. Washington, DC: National Academy Press. Marchesini, R. 2019. Beyond Anthropocentrism: Thoughts for a Post-Human Philosophy. Milan: Mimesis International. McShane, K. 2007. Anthropocentrism vs. nonanthropocentrism: Why should we care? Environmental Values 16: 169–185. Miller, H.I. 1999. Substantial equivalence: its uses and abuses. Nature Biotechnology 17: 1042– 1043. Millstone, E., E. Brunner, and S. Mayer. 1999. Beyond ‘substantial equivalence’. Nature 401: 525–526. Murphy, J.F., and J. Conner. 2012. Beware of the black swan: The limitations of risk analysis for predicting the extreme impact of rare process safety incidents. Process Safety Progress 31: 330–333.

164

6 Ethics and Environmental Risk Assessment

Myskja, B.K. and A.I. Myhr. 2020. Non-safety assessments of genome-edited organisms: Should they be included in regulation?. Science and Engineering Ethics Published Online first, May 20: https://doi.org/10.1007/s11948-020-00222-4. Norton, B. 1987. Why Preserve Natural Variety?. Princeton, NJ: Princeton University Press. Norton, B. 1991. Toward Unity among Environmentalists. New York: Oxford University Press. Pence, G.E. 2002. Designer Food: Mutant Harvest or Breadbasket of the World?. Lanham, MD: Rowman and Littlefield. Railton, P. 1990. Benefit-Cost Analysis as a source of information about welfare. In Valuing Health Risks, Costs, and Benefits for Environmental Decision Making, ed. P.B. Hammond and R. Coppock, 55–82. Washington, DC: National Acadamy Press. Raybould, A. 2010. Reducing uncertainty in regulatory decision-making for transgenic crops: More ecological research or clearer environmental risk assessment? GM Crops and Food 1: 25–31. Rescher, N. 1983. Risk: A Philosophical Introduction to the Theory of Risk Evaluation and Management. Lanham, MD: University Press of America. Rifkin, J. 1985. Declaration of a Heretic. Boston and London: Routledge and Kegan Paul. Rifkin, J. 1998. The Biotech Century: Harnessing the Gene and Remaking the World. Tarcher/Putnam, New York: Jeremy P. Rifkin, J., and P. Nicanor. 1983. Algeny. New York: Viking. Rocca, E., and F. Andersen. 2017. How biological background assumptions influence scientific risk evaluation of stacked genetically modified plants: An analysis of research hypotheses and argumentations. Life Science and Social Policy 13: 11. https://doi.org/10.1186/s40504-0170057-7. Rollin, B.E. 1995. The Frankenstein Syndrome: Ethical and Social Issues in the Genetic Engineering of Animals. New York: Cambridge University Press. Rolston III, H. 1991. Environmental ethics: Values in and duties to the natural world. In Ecology, Economics, Ethics: The Broken Circle, ed. F.H. Bormann and S.R. Kellert, 73–97. New Haven, CT: Yale University Press. Sagoff, M. 2004. Price, Principle and the Environment. Cambridge, UK: Cambridge University Press. Sandin, P. 2004. The precautionary principle and the concept of precaution. Environmental Values 13: 461–475. Sandin, P. 2006. The precautionary principle and food safety. Journal für Verbraucherschutz und Lebensmittelsicherheit 1: 2–4. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists Versus Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press. Scott, N.D. 2018. Food, Genetic Engineering and Philosophy of Technology. New York: Springer. Sharon, T. 2013. Human Nature in an Age of Biotechnology: The Case for Mediated Posthumanism. Dordrecht, NL: Springer. Sharples, F. 1987. Regulation of products from biotechnology. Science 235: 1329–1335. Shrader-Frechette, K.S. 1991. Risk and Rationality. Berkeley: University of California Press. Soule, E. 2000. Assessing the precautionary principle. Public Affairs Quarterly 14: 309–329. Stich, S. 1978. [1989] The recombinant DNA debate. In Philosophy of Biology, ed. M. Ruse, 229–243. New York: Macmillan. Stich, S. 1982. Genetic engineering: How should science be controlled? In: And Justice for All: New Introductory Essays in Ethics and Policy, ed. by T. Regan and D. VanDeVeer, 82–119. Totowa, NJ: Rowman and Littlefield. Taleb, N.N. 2010. The Black Swan: The Impact of the Highly Improbable, 2nd ed. New York: Random House. Tait, J. 2001. More Faust than Frankenstein: The European debate about the precautionary principle and risk regulation for genetically modified crops. Journal of Risk Research 4: 175–189. Ten Eyck, T.A., P.B. Thompson, and S.H. Priest. 2001. Biotechnology in the United States: Mad or Moral Science?” In: Biotechnology 1996–2000: The Years of Controversy, ed. by G. Gaskell and M. W. Bauer, 307–318. London: The Science Museum.

References

165

Thompson, P.B. 1983–84. Risk, ethics and agriculture. The Journal of Environmental Systems, 13:137–155. Thompson, P.B. 1986. Uncertainty arguments in environmental issues. Environmental Ethics 8: 59–75. Thompson, P.B. 1987. Collective action and the analysis of risk. Public Affairs Quarterly 1: 23–42. Thompson, P.B. 1995. Risk and responsibilities in modern agriculture. In: Issues in Agricultural Bioethics, ed. by T.B. Mepham G.A. Tucker and J. Wiseman, 31–45. Nottingham: Nottingham University Press. Thompson, P.B. 2008. Nano and bio: How are they alike how are they different? In: What Can Nanotechnology Learn from Biotechnology: Social and Ethical Lessons from the Controversy over Agrifood Biotechnology and GMOs, ed. by K. David and P.B. Thompson, 125–155. Burlington, MA: Academic Press. Thompson, P.B. 2017. The Spirit of the Soil: Agriculture and Environmental Ethics, 2nd ed. New York: Routledge. Thompson, P.B., and W.J. Parkinson. 1984. Situation specific indicators for distinguishing between high-consequence/ low-probability risk and low-consequence/high-probability risk”. In LowProbability/High-Consequence Risk Analysis, ed. R. Waller and V. Covello, 551–567. New York: Plenum Press. Trench B. 2008. Towards an Analytical Framework of Science Communication Models. In: Communicating Science in Social Contexts, ed. by D. Cheng, M. Claessens, T. Gascoigne., J. Metcalfe, B. Schiele, and S. Shi, 119–135. Dordrecht, NL: Springer. Trewavas, A. 1999. Much food, many problems. Nature 402: 231–232. Tversky, A., and D. Kahneman. 1982. Judgement under certainty: Heuristics and biases. In Judgement Under Uncertainty: Heuristics and Biases, ed. D. Kahneman, P. Slovic, and A. Tversky. Cambridge, UK: Cambridge University Press. van den Belt, H. 2003. Debating the Precautionary Principle: Guilty until proven innocent or innocent until proven guilty? Plant Physiology 132: 1122–1126. van Dommelen, A. 1995. Quality of risk assessment: Artificial and fundamental controversies. In Contested Technology: Ethics, ed. R. von Schomberg, 193–208. NL: Risk and Public Debate, International Centre for Human and Public Affairs, Tilburg. Wachbroit, R. 1991. Describing risk. In Risk Assessment in Genetic Engineering, ed. M.A. Levin and H.S. Strauss, 368–377. New York: McGraw-Hill. Wynne, B. 1983. Redefining the issues of risk and public acceptance: The social viability of technology. Futures 15: 13–32. Wynne, B. 1988. Unruly technology: Practical rules, impractical discourses and public understanding. Social Studies of Science 18: 147–167. Wynne, B. 1992. Misunderstood misunderstanding: social identities and public uptake of science. Public Understanding of Science 1: 281–304. Wynne, B. 2001. Creating public alienation: expert cultures of risk and ethics on GMOs. Science as Culture 10: 445–481.

Chapter 7

Environmental Impact and Environmental Values

Abstract This chapter completes coverage of environmental risks begun in Chap. 6, which emphasized both the philosophical rationale for expected-value risk analysis, along with weaknesses in the way that approach has been applied to agrifood gene technology. This chapter discusses ethical objections to expected value analysis and takes up classical questions in environmental ethics. These include the basis for associating moral value with non-sentient entities such as plants, collectivities such as species or ecosystems and also for nature or the environment itself. The chapter proposes a novel approach to these problems based on the standpoint or attitude of the valuing subject. Classic approaches that stress intrinsic or instrumental valuation presume that valuation proceeds from the perspective of a spectator standing aloof from nature. Although classic approaches have not presumed that this spectator is a human being, the good of any entity derives from the spectator’s gaze. This is consistent with the notion of value as a consumption activity. In contrast, a more engaged, involved or truly environed approach can be elicited by taking the perspective of a producer. This is an especially fortuitous approach for developing an environmental ethics for agriculture and food. Keywords Expected-value · Environmental risk · Intrinsic value · Anthropocentrism · Environmental virtues · Agrarianism · Eco-feminism The expected value approach (discussed in Chap. 6), remains powerfully influential among scientists and public policy analysts. To recap, it takes risk to be a function of both hazard, (the evil, harm, cost or bad outcome the might occur) and also exposure (the likelihood that the hazard will materialize, given relevant contextual considerations). When combined, risk can be represented as the value one should expect to realize, if an option is acted upon. Economists and utilitarian philosophers have emphasized decision rules (e.g. risk management strategies) that seek optimal strategies for realizing benefits, once expected costs have been taken into consideration. Nevertheless, the risk-based approach is consistent with a wide range of rules for managing risks, including, for example, de minimus and informed consent. The de minimus rule minimizes risk, irrespective of benefits foregone by doing so. Informed

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_7

167

168

7 Environmental Impact and Environmental Values

consent holds that parties exposed to the potential for harm are informed of risks, and have the opportunity to opt out. A risk is socially acceptable when this procedure is in place, unacceptable when it is not. While I will defend a risk-based approach against those who wish to discard it altogether, I also hold that taking it too seriously has a corrupting effect on our (e.g. human beings) grasp of environmental ethics. These are views that I have articulated more directly in other writings, but the first edition of this book contained a statement of them that should temper the enthusiasm of readers who wish to pigeon-hole me as a utilitarian or as a typical natural resource economist. The chapter makes a sharp turn at midpoint, leaving aside the argument that an expected-value analysis of risk is compatible with a much broader array of philosophical positions than sometimes thought. At this point, I criticize the decisionism of the expected value approach, which takes the act of choosing to be the ontological pivot point for ethical theorizing. In presenting conduct as the result of choice, expected-value utilitarians share a problematic meta-theoretical assumption with environmental philosophers who derive duties from the intrinsic value of animals, endangered species or ecosystems. In this chapter, I offer a brief sketch of an alternative agrarian approach that is developed at more length in my book The Agrarian Vision, (Thompson 2010). In the sections that follow, I will use monarch butterflies as an exemplary species to illustrate the meaning and significance of environmental impacts. Two years after Bt crops were commercialized (discussed below and in Chaps. 1 and 6), Cornell University scientists published a short note in Nature demonstrating that pollen from these crops was toxic to monarch butterflies (Losey et al. 1999). Very little about what happened next is uncontroversial. Even the publication of this note has generated controversy, but the potential for damage to monarch butterfly populations became a vehicle for many in the public to appreciate the environmental risks of agricultural and food biotechnology (Marks and coauthors 2007). A detailed study of the controversy over monarch butterflies would take the chapter far off track. Nonetheless, I will use this species to illustrate the significance of environmental values and the philosophical importance of environmental risks.

7.1 Environmental Impacts from Agrifood Biotechnology Policy-oriented studies of the environmental risks associated with agrifood biotechnology adopt a dual-track ontology for categorizing environmental hazards that do not clearly involve harm to human beings. Hazards discussed in Chap. 6 (acquired resistance, weediness) are harmful to agriculture, and hence, to humans, but this chapter will take up impact on non-target species and ecosystem effects (see NRC 2000, 2002). These two categories of hazard are distinguished from gene flow, which is a mechanism of exposure. Before discussing these concepts, it is useful to recall that agricultural technology is intended to affect key organisms in a farm environment. Specifically, gene modifications alter the crops that farmers grow and the animals they raise. In addition, like insecticide, herbicide, antibiotics and other agricultural

7.1 Environmental Impacts from Agrifood Biotechnology

169

chemicals, gene technology affects the pest organisms that frustrate farmers’ efforts to extract usable yields from their crops and livestock. The two most widely grown GMOs, Bt crops and herbicide tolerant crops (see Chap. 1) fit this model. Bt crops have been genetically altered to produce bacillus thuringiensis an insecticide that is toxic to Lepidoptera (e.g. caterpillars) that damage crops. Bacillus thuringiensis is a natural product of microorganisms and approved for use in organic farming when derived from these natural sources. Herbicide tolerance protects the crop when the farmer sprays weed-killing chemicals that remove unwanted plants from the field. These are the intended effects of biotechnology. They are viewed as beneficial within the context of agriculture, but the death of caterpillars and weedy plant species can be thought of as an environmental impact. If one were to deploy insecticides or herbicides in a protected natural area, they might well be viewed as destructive. Nevertheless, these intended effects are not characterized as morally significant in themselves in the context of agricultural practice. It would only be additional, unwanted impacts (on the health of food consumers and agricultural workers, for example) that would be seen as morally significant. Agriculture itself is exceedingly disruptive to whatever native plants and animals occupy a farmer’s fields and pastures prior to cultivation. Although one can certainly question whether agriculture should be practiced in a given region, the impact of agricultural technology is evaluated against the background assumption that intended effects of the technology are not be classified as harmful environmental impacts. Given this orientation, the category of non-target effects simply describes any impact on the entire class of organisms that were not affected intentionally, or, to put is slightly differently, non-target effects include all impacts beyond those that were the purpose for which the technology is implemented. As such, the category also includes unintended side effects on targeted organisms (e.g. insect pests, fungi or weeds). In agriculture, the concept of non-target impact derives from a well-studied side effect of chemical pesticides. Insecticides are designed to kill pests, but they often kill other insects (as well as birds and mammals) that are predators of pests. The loss of beneficials, as they are called, can rebound to the detriment of the farmer, because the species on which they prey can themselves become pests once predators are removed. In other cases, loss of a non-target species may have little impact on farming, but will be regarded as a significant environmental impact. The monarch butterfly, for example, is an iconic species for naturalists and environmentalists. The monarch exemplifies the problem of non-target impact from biotechnology due to concern that it is threatened either by exposure to Bt or by habitat loss from herbicides (Gustafsson and coauthors 2015). The terminology of targets and off-target effects has morphed in the discussion of gene editing. In this context, the target is the gene construct that a technology developer intends to modify. However, many techniques of cellular manipulation are associated with small changes in the gene sequence, apparently at random locations in the genome. These are referred to as off-target effects. In both cases the word target indicates intended effects. In the environmental context, the intended effect is reducing the population of pest organisms. In gene editing, the intended effect is modification at a specific location on the genome. In the environmental context, a

170

7 Environmental Impact and Environmental Values

non-target hazard is a deleterious effect on any organism that is not the target. In gene editing, an off target effect is any change in the sequence of base pairs not specifically intended by the technology developer. In both cases, it is important to say more about why non-target or off-target effects would be a cause for concern. This chapter will focus exclusively on non-target hazards as defined in the environmental context. Off target effects from gene editing are discussed in Chap. 13. In addition to non-target effects, impact on ecosystem integrity is an equally broad class of hazards that involve disruption or alteration of ecosystem processes. For example, soil metabolism is the process that builds both soil nutrients (nitrogen, phosphorus and potassium) and structure (or tilth) that can retain water and support plant growth. This process can be disrupted by certain types of chemicals or by physical compaction and over-cultivation of soils. It is not, in any obvious sense, an impact on some untargeted organism (though microorganisms are almost certainly affected by disruptions to soil), so the category of ecosystem integrity is intended to capture the ways in which an agricultural technology can alter an ecosystem in unintended, unwanted and unexpected ways that extend beyond the destruction of nonpest organisms. Impacts on ecological integrity would include disruption of nutrient or hydrological cycles, for example. Ongoing changes to global climate systems both affect and are affected by agriculture. What is more, impacts on ecological integrity can be knock-on effects of non-target impact. The loss of a keystone species would be expected to have significant impact on the biodiversity within an ecosystem, for example. These ecosystem effects may not be captured by a focus on non-target impacts because a) they may not be obvious and b) impacts on target organisms can also have these impacts (Siddig and coauthors 2016). While farming practices that include products of biotechnology can have impacts on the integrity of ecosystems, it is less clear how a GMO would disrupt ecological integrity beyond its role in intensifying or perpetuating those practices. Some authors write as any appearance of products from gene technology would disrupt ecological integrity (see Westra 2011) but authors who identify specific ecological processes emphasize the entire farming system within a social context. The disruptive nature of gene technology may derive from either social or ecological integration (Norton 2016). Gene flow was an early concern in agrifood biotechnology risk assessment. Gene flow, or horizontal gene transfer, occurs when genes move from one species or subpopulation to another. The concern was that transgenes introduced through biotechnology might move to wild-types of plants that were related to agricultural crops. While observers of biotechnology might have thought that this was a rare event, research showed that movement of genes was relatively common in plants (see Ellstrand and Rieseberg 2016). From the standpoint of ethics, the key question is, why should this matter? Here is an early answer: If one believes that gene flow is rare in nature, but that the appearance of biotechnology, it will become common, it is possible to construe gene movement as an impact on the environment. Doctrines in environmental ethics that interpret pristine nature as intrinsic value might then have suggested that the simple fact of any human impact on wild plant genetics was ethically problematic (Levidow and Carr 1997). This idea has become less plausible given the way that genomics has improved understanding of agricultural crops’ effect

7.1 Environmental Impacts from Agrifood Biotechnology

171

on the genetics of wild types, without regard to whether genetic engineering has been used. Nevertheless, movement of genes could be a mechanism for impacts on nontargets and ecological integrity, especially in connection with biodiversity. Such an impact was suspected when gene constructs for Bt were putatively discovered among Mexican land races of maize. Although no instance of actual gene flow has ever been documented in this case, the concern was that maize plants having the transgene would outcompete plants that did not have it, leading to a significant loss in diversity of alleles among the total population (Kuzma and Besley 2008). Though theoretical, the concern is nonetheless ethically meaningful. Nevertheless, it is important to see that the mere fact of gene flow was not a basis for concern. Rather, in this case, movement of the transgene could have further impacts on ecological integrity as it depends on biodiversity.

7.2 Environmental Hazards and Environmental Values The larger task for ethics is to explain why impacts on non-target species or indicators of ecosystem integrity are hazards, as distinct from value-less or neutral happenings. The answers to this question derive from the debates that launched environmental ethics in the 1970s and 1980s. On the one hand, humans both enjoy and depend upon the existence of species other than the agricultural crops and animals they produce. The iconic status of the monarch butterfly is a case in point. In addition, humans assuredly rely on ecosystem functions for a plethora of services that are probably even now not fully understood. Dane Scott’s book on agrifood biotechnology calls upon the planetary boundaries work of Johan Rockström to illustrate the complexity and thoroughness of this dependence (Scott 2018). What is more, future generations of humans will also want to enjoy nature, and we can expect that they will also be dependent on ecosystem services. All of these rationales for valuing impact on nontarget species or ecosystem integrity advert back to human beings. They are, in that sense, anthropocentric. On the other hand, perhaps monarch butterflies have value completely independent from the pleasure that humans take in seeing them. Perhaps this value derives from the pleasure that the butterflies themselves take in life, or perhaps it derives simply from the fact that they are alive. Ecosystems would then have value not simply because current and future human beings depend upon them, but because all sentient creatures depend upon them. Perhaps it is possible to go further. As noted in Chap. 6, Holmes Rolston has argued that systemic organization of species within an ecosystem has value, and that the organization of genetic information within a breeding population has value (Rolston 1991, 1999). When environmental philosophers characterize these as intrinsic values, they mean that the worth of a deleterious impact does not depend on human beings’ recognition of their value, much less any pleasure or use that humans make of them. There is quite a bit of disagreement among those arguing for ethical positions that extend environmental values beyond the human species. In this

172

7 Environmental Impact and Environmental Values

chapter, I will simplify a philosophical debate among bio-centrists, animists, ecocentrists, eco-feminists and others to describe all these perspectives as ecocentric. However, it is important to note an important qualification. As argued at some length in Chap. 5, the impact of genetic modification on farmed animals is widely, and in my view appropriately viewed as morally significant in terms of its impact on the animal, not simply how people feel about animals. Much of the literature criticizing anthropocentric approaches to evaluating the impact of biotechnology focuses on animals (see, for example, Kaebnick 2007; Frédéric 2018). My use of the term ecocentric is intended to stress perspectives that also ascribe non-instrumental value to non-sentient entities like plants, species or ecosystems. Anthropocentric and ecocentric strands of environmental ethics have distinct rationales for assigning value to impacts on non-target species or to ecosystem integrity. An ecocentrist, for example, might be less impressed by arguments that stress ecosystem services that are tied to supporting the food supply for human beings. It is not obvious, however, that these differences have great bearing on the ethics of agrifood biotechnology, as distinct from the differences that they imply for all forms of agriculture. Ned Hettinger has drawn upon ecocentric environmental ethics to analyze ownership regimes for genetic constructs (Hettinger 1995), as does Christopher Preston’s critique of synthetic biology (Preston 2008). Nevertheless, with the exception, of concerns about animal health and welfare, explicit appeals to non-anthropcentric environmental ethics are rare in the literature on agrifood biotechnology. Even when the terminology of ecocentism or non-anthropocentric value appears, it may not imply that non-sentient organisms or abstract entities like species and ecosystems should be valued for their own sake. Fern Wickson reviewed objections to risk assessment methodologies used to analyze agrifood biotechnology in 2014. Though her review discusses intrinsic value, all of the omissions and flaws she associates with GMO risk assessment identify hazards that could be justified in anthropocentric terms (see Wickson 2014). At the same time, ecocentric environmental ethics and appeals to intrinsic value may be embedded in criticisms that take issue with applications of risk assessment that hold too closely to the expected value model discussed in Chap. 6. Wickson’s article suggests that the intrinsic value claims can be interpreted as support for certain cultural services traditionally associated with foods and agricultural ecosystems. While threats to these services can be supported by anthropocentric arguments, they are seldom recognized as hazards in environmental risk assessments. She notes (as did the first edition of this book) that quantification of risks from products of gene technology is typically confined to impacts that would be consistent with utilitarian assumptions of welfare economics, and attendant applications of the expected value approach (Wickson 2014). Previous chapters have argued that basic logic and methods of expected value risk assessment could be reinterpreted to accommodate a much broader class of hazards than is typical. Nevertheless, ecocentrists might prefer to articulate their concerns about agrifood biotechnology through a series of objections to utilitarian, cost–benefit interpretations of expected value. Such an understanding of the ecocentrist perspective draws support from the history of environmental philosophy.

7.3 Environmental Philosophy For and Against Expected Value

173

7.3 Environmental Philosophy For and Against Expected Value Does ecocentrism—the view that non-sentient organisms and possibly ecosystems or species have a value independent of the value that humans attach to them—imply a rejection of the expected-value approach? Talk of intrinsic or ecocentric values need not be conceived as inconsistent with a broadened form of consequentialist, expected value thinking, nor need it be thought of as utterly inconsistent with goals of human use (or even economic development). At the same time, there are many theorists (and many non-theorists, too) who do in fact mount powerful objections to the consequentialist/expected-value approach. Grasping the difference between these two points of view presupposes a distinction between moral theory and ordinary moral discourse. Moral theories introduce concepts and terminology in much the same way as theories in other domains of the biophysical or social sciences. R. M. Hare (1919–2002) was one of the most influential defenders of consequentialism in the twentieth century. Hare acknowledges the difference between moral heuristics and moral theory. The consequentialist, expected-value tradition is a moral theory: it is intended to give a philosophically defensible (e.g. true) account of right action, just as physical theory is intended to give a true account of the world’s structure. We should no more expect ordinary people making ordinary moral judgments to utilize moral theory than we should expect carpenters and tradesmen to utilize theory developed by physicists, despite the obvious fact that they are expert in some key physical properties of the materials that they use in their trade. Hare argued that most ethical issues can be discussed and even decided with heuristics that are logically imprecise, using many vague and overlapping terms. However, philosophically detailed moral theories (and for Hare, the correct moral theory is a consequentialist one) have practical application when heuristics are inapplicable, or when they lead to conflicting recommendations (Hare 1981). Only the most intractable or complex issues require that one actually delve into the tasks of placing value on possible consequences and quantifying the probability that these values will actually be realized. Understanding the difference between heuristics and theory as two modes or levels of ethical analysis is called “two-level” utilitarianism (Singer 1979; Hooker 2000). Given this distinction, a consequentialist philosopher might advocate nonconsequentialist reasoning at the level of heuristics and ordinary conversation. Hare’s approach has led many contemporary consequentialist philosophers to formulate arguments that appeal to rights, duties and moral character whenever they are writing for a popular audience. This is the language in which moral heuristics are communicated outside the halls of academia. If consequentialists want to influence human conduct, they had better learn to speak that language. In fact, two-level utilitarianism explains why doing so is justified by consequentialist considerations. In a nutshell, detailed consequentialist argumentation has a cost of its own. It becomes tiring and difficult to follow. If you actually want to produce the greatest good for the greatest number, it is better to use less precise terminology that nonetheless gets people to the right answer most of the time. Hare’s thinking has received further support from

174

7 Environmental Impact and Environmental Values

studies in cognitive science that support the idea people choose to use simple heuristics in place of hard, careful risk-based thinking (Kahneman 2011). Gary Varner has provided a philosophically thorough exposition of the implications of this view in environmental ethics (Varner 2012). Most philosophers working in ethics would agree with Hare’s claim that philosophical ethics requires a degree of precision and care in its development of terminology and key ideas than does ordinary discourse or conversation among nonspecialists. However, they might not agree with his claim that some form of utilitarianism is what we need at the level of moral theory. Chap. 4, on food safety and Chap. 8, on social consequences recount arguments against the expected value framework as it is typically applied in health and economic settings. In these instances, failure to involve and consult with affected parties is thought to violate fundamental moral rights to give and withhold consent, or principles of democratic participation. A concern for participation also arises in connection with environmental policymaking. For example, Kloppenburg (1989), Margaret Mellon (1992), and Les Levidow (1995) were among the early environmentally-oriented critics of agrifood biotechnology who laid stress on the value of citizen participation in policymaking. Such calls have only mounted in the intervening years. Responsible Research and Innovation is, for example, a structured effort to involve a wider scope of social perspectives in biotechnology (Bogner and Torgersen 2018). These participatory arguments rely on ethical claims about the role of science in democratic government that are philosophically similar to those that establish the case for an ethics of consent with respect to food safety, or to an ethics of participation with respect to social consequences. It is plausible to interpret their arguments as rejection of the expected-value approach to environmental decision-making. On this reading of their argument, they would be claiming that a right to participate overrides or supersedes the process of tallying the cost and benefit of alternative outcomes that is the heart of expected-value ethics. A strong interpretation of participatory rights insists that they are not simply a heuristic, but rather an element of moral theory that captures what is ethically significant at its deepest level. These philosophical differences do have policy implications. In contrast to arguments that rely on a philosophically deep commitment to rights, consent and participation, Wickson’s criticisms of GMO risk assessments stress ways in which ecosystems and ecological processes are valued. She does not denounce the process of risk assessment or assigning expected-values for its lack of inclusiveness or respect, nor does she claim that people exposed to these risks have been wronged because they were able to withhold consent. As such, her characterization of ecocentric critiques enunciates a set of values that could be included along with more standard types of health, welfare and preference satisfaction values that are currently reflected in expected-value risk assessment. Critiques not focused specifically on biotechnology have noted that mainstream economics fails to acknowledge ecological limits to human use and exploitation of the natural environment. To the extent that expected value approaches are committed to economic methods for assigning value, they may fail to assess some of the most serious environmental risks (Daly and Cobb 1989). This type of critique was extended to research on food biotechnology by

7.3 Environmental Philosophy For and Against Expected Value

175

Gary Comstock well before any actual crops appeared in the field. (Comstock 1989, 1990). One could address Wickson’s complaints by broadening the scope of hazards identified through a risk assessment to include threats to the services that people derive from a well-functioning environment. (This general strategy is discussed at more length in Chap. 6.) Critiques that are more fundamental claim that the very process of conducting an expected value risk assessment is itself morally objectionable. When critics argue that affected parties have a right to greater inclusion in decision-making, they are making an ethical claim that cannot be incorporated into an expected-value analysis by simply representing their exclusion as a hazard, and then estimating how likely they are to complain. In this case, only full recognition of participatory rights would satisfy their claim. Yet it is difficult to see how this kind of political right might be advanced in the context of environmental impact. Philosopher Annette Baier (1921–2012) argues that expected value approaches to risk “poison the wells,” by introducing a means-end language into ethical discourse that obscures the importance of virtue and respect for nature (Baier 1986). One might say that expected-value thinking reduces the significance of non-humans and the broader environment. Their moral importance is constrained to “bad things that might occur” (e.g. hazards) even when we allow that the evils may be borne by non-humans, and even ecosystems themselves. Baier is echoing the thought of naturalist Aldo Leopold (1887–1948), who deplored the tendency to find economic rationales for actions taken on behalf of the environment (Leopold 1949, 210). Leopold believed that those who work on environmental issues divide into two groups. Group (A), “regards the land as soil, and its function as commodity-production; another group (B) regards the land as biota, and its function as something broader,” (Leopold 1949, 221). The ‘something broader’ is ecosystem integrity or health. In pointing beyond the kind of value that economic analysis is so capable of capturing, Leopold indicates the domain that is the primary topic in the balance of this chapter. The interpretation of Leopold’s claim here is quite sensitive to the fine distinctions noted above. If, on the one hand, Leopold is simply complaining about economists’ tendency to emphasize outcomes with clear monetary values, that is a critique that might be redressed by reforming an expected-value approach, rather than rejecting it. Leopold might also be criticizing the anthropocentrism of economic thinking, its tendency to limit value to the interests of human beings. That, too, is a critique that might be incorporated within a broadened expected-value risk assessment. On the other hand, Leopold’s comment also synthesizes the force of Baier’s concern. Adopting the expected value paradigm (e.g. economic thinking) reveals a character flaw: a lack of disquiet with or appreciation of humanity’s limited ability to foresee consequences. The scientists in Leopold’s A group are a corrupting influence; they “poison the wells,” to use Baier’s phrase, by systematically ignoring reasons that reflect a deeper commitment to environmental values. Here, environmental philosophy is also laying a foundation for critiques that call the scientific community’s moral commitment to anticipating hazards into question (see Hoffmann-Riem and Wynne 2002). Environmental philosophers frequently interpret Leopold as rejecting a utilitarian or expected-value approach to ethics in these remarks (see Hargrove

176

7 Environmental Impact and Environmental Values

1989, 102–4; also 1994). They claim that in directing our attention to a concept of ecosystem health, Leopold means to reject some of the more narrow forms of consequentialist ethics, and that he means to attribute intrinsic value to ecosystems, as discussed above. A Sand County Almanac has spawned an explosion of philosophical reflection that coincides with, but rarely intersects, the rise of agrifood biotechnology and its attendant environmental risks. Rachel Carson’s Silent Spring, published in 1962, sparked environmentalism as a U.S. political movement. Carson deftly combined at least two distinct ethical arguments in her critique of DDT, a pesticide widely used in agriculture during the 1950s. On the one hand, Carson laid out the case for bioaccumulation of toxins, undercutting faulty exposure assessments that had been made in risk assessment. As explained in Chap. 4, the acute toxicity of a chemical such as DDT is a function of dose, and risk assessments for DDT had suggested that the dose capable of killing insect pests would be safe for other species (including humans). Silent Spring explained how the dose would accumulate in birds who ate these insects, and that this accumulation would continue up the food chain, leading to quite concentrated doses animals near the top (which includes humans). Although it was not clear that humans were part of the bioaccumulation argument for DDT, Carson’s argument was sufficient to persuade many people that the risk assessments for agricultural chemicals had not been adequate. They might be more dangerous than we thought. Once again, we see a critique that calls for reform and expansion of expected value, rather than rejection of it. At the same time, Silent Spring made an eloquent appeal on behalf of the birds. The very title of the book conveyed a coming world deprived of birdsong. In making this argument the centerpiece of her book, Carson adopted a philosophical strategy that Leopold had discussed some two decades earlier. In decrying reliance on economic analysis, Leopold writes, When one of these non-economic categories is threatened, and if we happen to love it, we invent subterfuges to give it economic importance. At the beginning of the century, songbirds were supposed to be disappearing. Ornithologists jumped to the rescue with some distinctly shaky evidence to the effect that insects would eat us up if the birds failed to control them. (Leopold 1949, p. 210).

Silent Spring called attention to the socio-economic and health consequences of threats to songbirds, but Carson did not hesitate to suggest that we should save the birds simply because we love them (Carson 1962). Love for nature is anthropocentric, someone might say, but expressions of love always stress the value of the beloved, and never the payoff for the lover. In that sense, denunciations of agrifood biotechnology that reject economic appeals and expected-value risk analysis are in the spirit of ecocentrism, even if they do not adhere to the letter of ecocentric argumentation. Below, I will argue that environmental philosophers’ thinking remains tainted by a moral ontology derived from economics, even as they formulate what they take to be alternatives to the economists’ expected-value framework. Theorists (mostly consequentialists) in welfare economics developed this moral ontology over the course of the 19th and early twentieth century. It frames valuation as an exercise in which

7.3 Environmental Philosophy For and Against Expected Value

177

decision-makers choose from a menu of options. While most environmental philosophers reject the welfare economist’s view that maximizing pleasure or satisfaction should be the basis for their selection of a given option on the menu, they remain committed to a moral ontology that isolates choice as a process in which humans assign values to each option before making a decision. Like ordering food from a menu, this stance is passive insofar as the decision maker steps back from precipice of action to survey the world from a spectator’s point of view. This construes choice as a form of consumption (or so I will argue). Agrarian traditions of ethics construe humans as actively engaged with their environments, and as such, unable to distance themselves in the manner that the menu analogy suggests. The argument for this analysis of environmental values holds that many mainstream approaches in environmental ethics have a consumptive orientation. This makes them poorly suited to address questions in agricultural ethics, including those that arise in connection with agrifood biotechnology.

7.4 The Conceptual Landscape of Environmental Ethics Academic environmental philosophy considers themes that are as old as philosophy itself, but its current practice reflects a much narrower foundation. In the 1970s, philosophers trained in logical analysis began to question whether the dominant ethical theories—utilitarianism and some form of neo-Kantian deontology—could support the idea of duties to non-humans. Chapter 5 reviews how Peter Singer and Tom Regan extended these theoretical approaches to sentient animals, including farmed animals. Yet it was less clear that these theories could provide support for the preservation of ecosystems or the protection of endangered species. Introductory and historical treatments of environmental ethics are widely available, so there is no need to provide that here. Still, readers who lack prior familiarity with the last fifty years of thinking on the philosophy of the environment may find a brief survey of key terms and concepts helpful. The notion of anthropocentrism has already been introduced in Chap. 5 and summarized above. Philosophies that reduce all ethical claims to something matters for human beings are anthropocentric. Anthropocentric environmental philosophy provides a rationale for protecting ecosystems or species by extending its key terms (welfare, rights) to future generations. Theorists like Singer and Regan extend them to all sentient beings, that is, creatures capable of experiencing satisfaction, dissatisfaction, pleasure or pain. These are both extensionist in retaining a traditional basis for valuing the welfare of a creature or arguing that we have duties with respect to them. An approach that takes all living creatures to be morally significant is biocentric, and one that goes even further to argue that non-living entities (such as an ecosystem or a planet) or abstract entities (such as a species) have value in themselves are ecocentric. A related set of terms distinguish between things that are valuable because they function as tools or means for achieving some further end (e.g. instrumental value) from things that are valuable in themselves (e.g. intrinsic value).

178

7 Environmental Impact and Environmental Values

For example, an anthropocentrist might argue that endangered species have instrumental value because future generations may want or need them, while a biocentrist or ecocentrist would say that these organisms have intrinsic value. They are valuable irrespective of what any human being now or in the future thinks about them. John Passmore (1914–2004) was an Australian philosopher whose early book Man’s Responsibility for Nature approached these ideas by drawing a distinction between conservationism and preservationism. The conservationist wants to save natural resources because they have or might have instrumental value at some time in the future. The preservationist wants to preserve them for their own sake, and hence will be much more attentive to maintaining all the systemic elements on which flora and fauna of any kind depend, (Passmore 1974). Philosophers took quite some time sorting out the overlaps and distinctiveness of each term in this vocabulary, and the task is not complete to everyone’s satisfaction, even in the second decade of the twenty-first century. More recent work has taken a turn to virtue ethics (see Sandler 2009). Here, the emphasis shifts away from valuing things in terms of either our impact on them or our duties to them. Moral language and environmental responsibilities are recontextualized in terms of the way that conduct reflects or cultivates moral excellence with respect to environmental affairs. The original 1997 treatment in this book anticipated the turn to virtue ethics. The balance of the chapter reprints and only slightly amplifies that analysis.

7.5 Environmental Ethics in Consumption Conservationists typically adopt a view compatible with the anthropocentric interpretation of the expected value approach, and argue that ethical responsibilities regarding nature require making wise use of nature, especially when the exposure of future generations is included. Against them are preservationists who reject the implicit assumption that nature is there for human consumption, if not now then later. However, there is a more fundamental sense in which both conservationists and preservationists posed the key philosophical issues from the standpoint of consumption. The question of whether to build a dam or preserve a wild area is posed as one of options that mimic the array of choices facing any shopper in the candy store or supermarket. Environmental philosophers are encouraging us to recognize that some of these options have value beyond any pleasure or satisfaction we might derive from using, them. To the extent that one equates consumption with depletion, they are anticonsumptive, but there is an ontologically deeper way to understand consumption. Consumption is built into the shopper’s standpoint. On this view, life is a matter of selecting those goods that one values, and the role of philosophy is to help the shopper develop a more sophisticated approach to the process of valuing his or her options. One adopts the shopper’s standpoint whenever one abstracts or “stands back” from the moment of action and considers the value associated with each of the several options one faces. This standpoint is readily evident in utilitarian ethics, where the

7.5 Environmental Ethics in Consumption

179

decision maker is not only instructed to calculate the expected value of each option, but also to economize in making the final choice. However, rule-following deontologists also adopt the standpoint, differing only in applying decision rules that foreswear the comparative quantification of values. A Kantian, for example, would derive the value of his or her respective options by considering whether choosing that option would involve treating anyone solely as a means, but the Kantian still views ethics as a matter of deliberating over which of the various options can meet the ethical criteria for morally justified selection. The crucial difference between consequentialism and deontology consists in their respective decision rules. The metaphor of shopping, of standing before an array of products and evaluating the value associated with each of them prior to purchase, aptly captures the standpoint of both consequentialist and deontological ethics. This is especially true in environmental ethics, where environmental impact can be interpreted as the result of selecting one option, rather than another. For example, following the writings of philosopher Arne Naess, the preservationist tradition in environmental philosophy has matured into a movement called “deep ecology,” (see Devall and Sessions 1985; Sessions 1995). Deep ecology commits to preservation of ecosystems or natural areas that are unaffected by humanity (if any such areas still exist), and for continual reduction of human impacts, wherever they occur. As characterized by its most extreme exponents, deep ecology becomes misanthropic, recommending preservation of global ecosystems through eventual human extinction. Though mainstream proponents of deep ecology deny the charge of misanthropy (Sessions 1995, xiii), the persistent image is that wild areas are “lost” when humans use them. For deep ecologists, this is a bad outcome; it has negative value. At the extreme, extinction of humans would have positive value because it would restore the dominance of natural (e.g. non-human) processes. A more moderate position might hold that human use also has value, but these values are deficient when compared to the value associated with a natural process. Use values are, for example, transitory and exhausted once the short-term satisfaction of use has dissipated. This would be as true of recreational uses as it is of uses that damage or deplete the environment. Depletion fails to respect the value of these wild places to non-humans, too, so we can choose not to use them because their value to non-human others is judged to be greater than the value of human use. The comparison between value to humans (e.g. anthropocentric) versus values that derive either from use by non-humans or from some intrinsic value that a wild ecosystem exhibits simply by existing (e.g. ecocentric values) is nonetheless made from a perspective of choice. The chooser is outside this ecosystem. The activity of valuation is unengaged with its processes. “To use or not to use,” is, in this sense, to address the question from the standpoint of consumption. The consumer standpoint portrays all problems in ethics as if they arise at the moment of decision. Under this interpretation, performing an ethical evaluation necessarily involves reflecting on one’s possible course of action, and assigning value to each option based on the particular criteria of a given ethical theory. In classical utilitarian ethics, these criteria involve changes in the health, wealth or well-being of sentient subjects (possibly including non-humans, possibly not), and the decision

180

7 Environmental Impact and Environmental Values

rule instructs the decision maker to optimize or “get the best value” among choices available. In neo-Kantian approaches, one’s options are evaluated in terms of whether they would show proper respect to affected parties, or would respect their rights or fulfill one’s duties to them. Critics have argued that both perspectives owe a debt to the liberal tradition in characterizing the decision maker as an autonomous judge who practices ethics by delivering verdicts on every possible course of action (Korthals 2001, 2004). The widespread (if implicit) influence of the consumer standpoint is evident in social science research where opinions and attitudes are unreflectively attributed to research subjects identified as consumers. One response to this phenomenon has been to theorize a distinction between “consumer” and “citizen”. Mark Sagoff argued that most people can make a distinction between their preferences, which would dictate their behavior under conditions of comparative choice (e.g. shopping) and their values, which indicate they aspirations that they have toward leading a good life. Values, not preferences, indicate how we should live in a moral sense. They are pursued through political action, rather than consumer choice. Sagoff argues that preferences often conflict with values, and that it is an error to infer anyone’s authentic moral commitments from their consumption behavior (Sagoff 1986, 1988, 2007). Michel Korthals argues for a similar distinction, claiming that when one follows what Sagoff would have called one’s preferences, one is acting as a consumer. When one follows one’s values (again, in Sagoff’s sense) one is acting as a citizen (Korthals 2001). The distinction has been employed by others working in food studies to call attention to the unreflective (and hence morally nondefinitive) nature of consumption choices (Verbeke and coauthors 2010; Alphonce and coauthors 2014). While Sagoff was stressing the difference between acting as a consumer and acting as a citizen, it is been far more common to argue that one can and should express citizenship values in making a consumption choice. Here, people do reflect on their everyday food choices, whether shopping at a grocery market or ordering at a restaurant. Rather than allowing conventional consumer tastes to determine their selection (e.g. taste and price), they evaluate their food options by considering a much broader range of values. Were farmers and employees of food companies treated fairly? What are the impacts on climate change? Did the food product cause pollution? Were animals treated humanely? (Wilkins 2005; Gilson 2014). In incorporating all of these considerations into a simple food purchase decision, the food consumer moves far beyond the type of preferences that Sagoff described in his 1986 article. Such consumers are evaluating their consumption decisions from the standpoint of citizen values, expressing a deeper sense of the world in which they would like to live. The shift from consumer to citizen reflects the aim of public choice, where the decision making agent is expected to reflect the interests of society as a whole. Yet it is not clear that this is also a shift away from a decisionist moral ontology. In many formulations of the citizen-consumer, the ethical injunction appears to be straightforwardly consequentialist. One should act to bring about the best consequences, and the analysis should include the socio-economic impacts of spending one’s money (Anonymous 2000; Ankeny 2019). Korthals, however, stresses the emergence of a new type of moral subject, the citizen-consumer. Korthals argues

7.5 Environmental Ethics in Consumption

181

that such behavior has revolutionized the traditional moral ontology of the liberal subject, re-injecting deeper philosophical considerations into consumer behavior (Korthals 2001, 2015). Although food studies researchers have questioned the efficacy of the citizen-consumer as a change agent, the citizen-consumer model has influenced many analysts of the contemporary food system (de Bakker and Dagevos 2012; Lehner 2013). One cannot dispute that many contributors to contemporary food ethics adopt the shopper’s standpoint as the proper way to characterize moral agency in ontological terms. It is also clear that the notion of ethical consumption is, for many people, something new. They have hitherto been content to economize on quality and price, never thinking of the way their market behavior implicates them in complex systems of practice. My claim, in distinction from others who are contributing to this literature, is that the expansion of the consumption frame is only an intensification of the ontology that has characterized the liberal subject for at least 200 years. This book affords only a cursory opportunity to develop this claim, and the ensuing treatment emphasizes agrifood biotechnology to the exclusion of all other considerations. To that end, it is worth noting once again how environmental philosophers have (mostly) been strangely silent with respect to that part of the world’s land mass dedicated to agricultural production. It is as if once land is given over to production, deep ecology advocates lose interest in its role in ecosystem health. They believe that human use so thoroughly ruins and pollutes nature that there is little point in even specifying norms for productive use of land. I see nothing in intrinsic value theories that entails this conclusion. It may simply be inattention to agricultural issues, rather than antagonism, which has led to this state of affairs. An agricultural-environmental ethic, however, would not be an ethic of consumption. A truly environmental ethic for food production would change move away from “A” side of the cleavage described by Leopold, where soil is understood solely in light of its contribution to commodity production, but it would be an ethic of production, nonetheless (Thompson 2010, 2017).

7.6 Environmental Ethics in Production Farming and food production inherently make productive use of nature. Any act of production transforms and in that sense consumes its inputs, but an ethic framed only in terms of constraints on consumption will never get to the heart of the production process. In many regions of the world, the farmer is understood to occupy a unique position of productive moral agency, though that ethic is complex and has often been implicit in religious beliefs, folklore and farming practice. The ethic presumes that the farmer (meaning the entire farming community) exists in symbiotic relationship with nature (usually articulated simply as “the land”). As farmers bring forth the commodities needed to sustain themselves, they must respect a complex system of natural constraints. They must preserve soil fertility, they must conserve water, they must limit erosion, they may not overgraze their pastures, etc. These constraints

182

7 Environmental Impact and Environmental Values

define a duty of stewardship for farmers that is consistent with long run self-interest but which may diverge from former interests viewed over the short term. The potential for tension between short and long-term self-interest rests at the heart of traditional moral wisdom for agriculture and the food system. The children’s fable of the ant and the grasshopper contrasts the hard working ant, who stores food for winter, with the lazy grasshopper who lives for today. Such stories place industriousness and self-reliance at the heart of morality. The good person calibrates the burdens they impose on others according to their own capabilities and the socioecological affordances of their environment. Of course, the industrious are more than willing to help others, for everyone appreciates a helping hand now and then. Such simple ethical principles tend to be lost in the shuffle when philosophers frame ethical theory in terms of rights or utility. The argument I present here does not use virtue as an important term of art, though I have made this connection more explicit in other contexts (see Thompson 2018). It is therefore worth stressing how the agrarian view of environmental ethics emphasizes hard work and self-reliance within a framework of food production and community obligations. For traditional farmers, stewardship duties coexist with other duties that emerge out of symbiotic relationships. Each member of the family depends on every other for survival, so carrying out one’s chores reinforces both personal loyalties and a virtue of industriousness. The network of loyalty extends to the community level, as neighbors help one another in time of need. Virtues of stewardship, industriousness and charity interact and mutually reinforce one another. They are all supported by self-interest, since failure to perform the duties of stewardship, industry and charity bring on ruin. The intricate network of these virtues also constrains self-interest in an ecological fashion. While traditional farmers have the same drive to produce more and to make good trades as anyone, an unrelenting emphasis on this drive creates negative feedback. One’s standing in the community falls. Soil fertility may decline. Hard work and productivity are virtues when they are held in balance with other virtues; they translate into the vice of greed when self-interest is allowed to drive this one dimension of farm life, unchecked by others (Thompson 2017). Arguably, the environmental problem with scientific agriculture, of which biotechnology is a part, is that it has undercut or obscured the feedback loops that bind stewardship and the other virtues of traditional agriculture together. In most instances, technology has lengthened the feedback loops that constrained traditional agriculture, rather than eliminating them. Nitrogen fertilizers have lengthened the time lag between abusive land use and eventual soil depletion, for example, as modern irrigation systems (especially those that pump groundwater) lengthen the time lag between overproduction and water shortage. Other effects on feedback are more complex. Traditional farm communities would constrain their food choices according to seasonal cycles. Although they might want foods out of season, they were content with a seasonally determined diet for they had some sense of the costs involved in producing or procuring the foods they ate. Modern grocery stores have made the feedback loops between effective consumer demand and the environmental costs of food production all but invisible. Price is the only signal that food shoppers get,

7.6 Environmental Ethics in Production

183

and they have no way of knowing whether low prices conceal short and irreversible environmental exploitation, or not (Clancy 1997). Agronomist Les Lanyon has argued that transportation, fertilizers and other production technologies have lengthened both the spatial and temporal dimensions of feedback loops in the nitrogen cycle. The cycling of nitrogen through soil, into crops, from crops to animals (including humans) and back into soil is, perhaps, the most basic ecological principle of food production. When this activity takes place in a relatively constrained geographical area, feedback on nitrogen cycling will occur first with depletion of soil fertility, and declining crop production. However, when crops that are fed to animals are hauled thousands of miles from the point of production (as they are in the United States), and nitrogen in animal manure is disposed of as waste, the feedback is more likely to appear as nitrogen pollution in the watersheds where animals are concentrated (Lanyon and Beegle 1989; Nord and Lanyon 2003). One must admit that the moral significance of this change in feedback is ambiguous. Lengthening feedback loops is not inherently bad. It can introduce flexibility into the food production system, and increase the number and type of responses that humans may undertake in discharging their duties of stewardship. An expected value approach would assess the costs of pollution or the risks environmental damage and would attempt to weigh them against the benefits. The environmental ethic of production sees the moral significance not simply in the costs accruing from fragile feedback loops, but in the deterioration of decision making capability that occurs when feedback becomes invisible, and when actions appear to have no consequences. When this kind of decline becomes so pervasive that it becomes typical of farmers, policymakers, scientists and other key actors, it is appropriate to use the moral language of virtue and character to describe what has gone wrong, to say that the ecology of the virtues has given way to the narrow pursuit of self-interest. The problem is not just that there may be environmental damages, but that this transformation of food production is creating a society of people who are incapable of moderating their activity, even when the consequences are pointed out to them. This latter point is crucial to the evaluation of biotechnology, for plant, animal and microbial biotechnology’s contribution to the probability of environmental insults may be quite small, especially when compared to chemical and mechanical farm technologies. Yet if biotechnology continues to lengthen and obscure feedback loops in our food system, and if preoccupation with biotechnology blinds scientists and public administrators to the environmental dimension of agriculture, its effect on the moral character of farmers, food consumers and public administrators will be regrettable. Something like this sentiment may lie at the heart of agro-ecologist Wes Jackson’s initial animosity toward biotechnology. After reviewing controversy over ice-minus and the Beltsville pigs in passing, Jackson writes, “Some gene splicers will explain that what that hog needs is some more fine tuning to make it right— they clamor for more research. Quite frankly, I am concerned less about this hog monster than about the human monster, created by our culture, the monster who sees nothing wrong with creating such a hog,” (Jackson 1991, 207–8). Jackson goes on to argue that scientific reductionism has led us astray, and that we should, in food and agriculture, concentrate instead on building an ethic that makes us “native

184

7 Environmental Impact and Environmental Values

to our place” (210). This phrase ties Jackson’s dim view of biotechnology to his other writings in the agrarian tradition (see Thompson 2017, 123–6). The task of agricultural science must be to illuminate, rather than obscure, the system feedback loops that bind person to community, and community to land (Jackson 1994). Contemporary philosophy is typically hostile to agrarian argumentation when it does not regard it with dumbfounded non-comprehension. It is more common for the moral claims that have historically been situated within an agrarian worldview to be articulated by authors who describe their view as anti-reductionist, feminist or both. For example, Ruth Hubbard (Hubbard 1990; Hubbard and Wald 1993), argued that reductionism and genetic determinism make molecular biologists insensitive to evolutionary and ecological dimensions of biology. Regine Kolleck presses the issue more deeply and in direct connection to ecological risk. She argues that scientists have fused a Cartesian, reductionist image of the world with blindness to the influence of commercial interests in order to rationalize the release of genetically engineered organisms (Kolleck 1993, 1995). Kolleck’s critique is advanced as a component of ecofeminist philosophy that is itself complex and multidimensional (see Davion 1994). Among ecofeminists who have specifically addressed biotechnology and genetic engineering, Maria Mies cites historical links between science and military or imperialistic projects, and writes, “Without selection and elimination, this technology would be quite different, hence, it cannot claim to be neutral; nor is it free from the sexist racist and ultimately fascist biases in our societies. These biases are built into the technology itself, they are not merely a matter of its application,” (Mies 1993). Vandana Shiva portrays food biotechnology as an extension of the green revolution that, in her analysis, obliterated and systematically destroyed indigenous women’s more ecologically sensitive knowledge and control of farming techniques (Shiva 1993). While Kolleck, Mies and Shiva argue from a theoretical perspective beyond or other than that of a practicing biological scientist, similar critiques are made from within biology itself. Richard Lewontin echoes the concern that molecular biology is inherently reductionist and insensitive to ecological context, (Lewontin 1992). Evelyn Fox Keller argues that molecular biology is built upon three intellectual shifts: (1) Biologists shifted their understanding of the basis of life from complex organismenvironment relations to the physical–chemical activity of the gene; (2) they redefined life as the information encoded in genes; and (3) they recast the goals of biology from observation to experiment. Keller links these shifts to a preoccupation with mastery and the penetration of nature that was characteristic of male dominated science (Keller 1990). These analyses provide a bridge between the explicitly environmental concerns of Mies and Shiva and the more abstract reasoning of Kolleck. Molecular biologist Martha Crouch worked in plant transformation labs early in her career, but ultimately decided that this work was ill-conceived and could not be reconciled with her ethical commitments. She argued that the structure of scientific research militates against pursuit of environmental goals. Although much of her argument stresses the interpenetration of commercial forces into scientific disciplines (a topic taken up in Chap. 7), she also appeals to feminist principles. For example, she compares the network of connections that are bound together and

7.6 Environmental Ethics in Production

185

embodied in her home grown tomato with the network of a genetically engineered tomato. The latter network includes many experts and organizations that have no intrinsic interest in Crouch or her tomato. Unlike the friends and neighbors who are bound together in her garden tomato, these experts and organizations cease to have any concern with Crouch after she has purchased their product (Crouch 1991, 1995). However, after listing impacts on women and children Crouch concludes with a statement that sounds more consequentialist than feminist: “None of these effects are desirable. Therefore biotechnology should be discouraged,” (Crouch 1995, 107). Arguably, it is her emphasis on the whole network rather than comparing costs and benefits, that places her in the ecofeminist camp. There are three points to stress in reviewing these early critiques of food and agricultural biotechnology. The first is simply the way that they have become invisible to scholars working in the twenty-first century, even those who take a nominally feminist position. For example, none of them are even mentioned in Rachel Ankeny and Heather Bray’s survey article on genetically modified food for The Oxford Handbook of Food Ethics, (Ankeny and Bray 2018), while only Shiva is discussed by any of the contributors to Erinn Gilson and Sarah Kenehan’s Food, Environment and Climate Change: Justice at the Intersections (2019). Deep scholarship in food ethics needs to recover these important sources. Second, these early critics stress environmental issues and characterize their work as a form of ecofeminism. I am following their expressed aims in representing the arguments as feminist, but feminism is a complex philosophical movement, and there are doubtlessly others who would take different perspectives. In addition, these critics do not draw sharp boundaries between genetically engineered foods and applications of gene technology in medicine. Whether this should be seen as a strength or a weakness awaits a more careful and nuanced analysis of this literature than has been provided here. It is worth noting that Rachel Schurman and William Munro’s sociological analysis of opposition to agricultural biotechnology finds that some of the activists who energized this social movement were motivated by concerns about genetic modification of human beings. They just found opposing GMOs to be a more inviting political opportunity (Schurman and Munro 2010). Finally, what these feminist critiques share with the agrarian and virtue-theoretic analysis is a concern with moral character. What they lack is a clear statement of how practice relates to moral character, to the formation of virtue and vice. Such an account is available in other strands of feminism. Annette Baier was a significant influence on my own approach. She built upon the work of psychologist Carol Gilligan in claiming that a feminist ethic emphasizes relationships, in contrast to utilitarian and rightsbased approaches to ethics that emphasize individuals apart from their social network (Baier 1994, 20–5). Agrarianism is also a relational ethic, deriving moral content from the manner in which individuals are imbedded in families, families in farm communities, and communities in the natural world. The relationships that emerge in farm production shape the virtues of stewardship, industry and charity in a manner that cannot be captured by theories of utility or rights (Thompson et al. 1994, 242– 57). Jim Cheney tied this relational, virtue-oriented theme in feminism to the need to respect diversity and broad themes in environmental ethics (Cheney 1994). Baier and

186

7 Environmental Impact and Environmental Values

Cheney both make more sweeping claims than the agrarian critique, however. Their conclusions about virtue extend to many areas of modern life, not just environmental stewardship. In linking her argument to peasant agricultural systems in India, it is Shiva who replicates many of the points made about lengthening and concealing feedback loops in the agrarian analysis, above, and who brings ecofeminism closest to the agrarian critique.

7.7 Ethics and Environmental Responsibility for Food Biotechnology Whether agrarian, virtue theory or ecofeminist analyses are used to assess food biotechnology, critics must admit that corruption of moral character is not a necessary consequence of genetic engineering in agriculture. Scientists are people. Human beings share the characteristic of being environed organisms with plants and other animals. To quote John Dewey (1859–1952), organisms are in their environments not in the sense that a marble is in a box but, “as events are in history, in a moving, growing never finished process,” (Dewey 1925, 224). Moral character reflects the sense in which all organisms—not just humans—are situated within their environments as poised. They are poised for response to a change in their environment by their genes, but also by phenotypic adaptations or developments (such as learning) that supervene on genetic capacities. At the biophysical level, poise is the organism’s ability to capture the energy needed to perform metabolic functions. Some organisms incorporate defensive preparedness into their poise, as well. The environed orientation of many creatures reflects a complex system of interaction—an ecology—that comprises both the inorganic elements and the other creatures residing at the place they live. For humans, at least, the acquired elements of their poise includes not only their personal experience, but also an accretion of cognitive and institutional capabilities that are the product of history. I am not prepared to undertake reflection on what makes character—the poise of a person or group of people at a given time—good or bad, moral or immoral. It should be evident that I do not regard character as something that is completely under any individual’s control. In the present context, my claim is that the virtuous use of gene technology cannot be determined by looking narrowly at tools for genetic manipulation, or even at the purposes for which it is used. A truly environmental ethic for agricultural and food biotechnology must take into account the way in which innovators are poised toward challenges, opportunities, incentives, penalties and rewards that the institutional milieu in which they work has habituated them to expect (And the same goes for biotechnology’s critics, as well.). It should be possible and even fruitful to utilize the techniques of rDNA within a sociotechnical environment characterized by agrarian or ecofeminist affordances, if it is possible to use science at all. Schurman and Munro describe biotechnology’s opponents as holding the view that the capitalist milieu has structured the habits

7.7 Ethics and Environmental Responsibility for Food Biotechnology

187

of scientists badly, a theme that takes the argument into chapters that make up the balance of the book. I admit that there is little in the science of food biotechnology that will direct its practitioners to an ecology of virtue (excepting possibly the study of evolutionary processes at the molecular level). It will take a conscious and dedicated effort to integrate deliberative consideration of environmental values and stewardship into the scientific institutions (universities, companies, professional societies and government agencies) in order to recreate the understanding of humanity’s place in nature that came naturally to traditional farmers. That understanding was implicit and it has eroded quickly where farming has embraced unrestricted technological expansion. To the extent that food biotechnology is simply part of that expansionist attitude, it contributes to humanity’s malaise. If you are not part of the solution, a wise environmentalist slogan goes, you are part of the problem. That sentiment captures the central environmental moral responsibility for food biotechnology. Scientists and decision makers trained in economics or politics may gravitate to an expected value analysis of the environmental risks associated with genetic engineering in the food system. Such gravitational pull is understandable—prediction is the long suit of science, after all—but it should be resisted. It is, in the first place, self-defeating on its own terms. The thoroughgoing expected-value risk analyst is forced to abandon the tools and concepts of expected value when it comes time to communicate with the public. What is more the expected-value approach moves the locus of ethical deliberation away from the ecology of virtue, away from our attempt to understand how our food production practices are embedded in a web of social and ecological relations. When efforts to anticipate consequences become detached from the ecology of virtue altogether, it is arguable that they become corrupting, a theme that will be taken up again in Chap. 12. None of this is to suggest that we can do without predictions, or without attempts to understand how food and agricultural biotechnology will affect the environment. It is important to have information on the risks of food biotechnology, and it is equally important to have information about its potential benefits. Characterization of the benefits, like characterization of risks, is an empirical and contested matter. Susanne Huttner and two co-authors think, “the potential benefits of biotechnology applied to agriculture are broad—encompassing virtually the entire food-production system,” (Huttner et al. 1995, 38). In contrast, Sheldon Krimsky and Roger Wrubel conclude their study by saying: Our research indicates that there is little basis for the claim that biotechnology has been burdened with overregulation and that such regulation has thwarted innovation. Some evidence suggests that regulatory inaction or confusion has kept firms from investing in transgenic organisms. Furthermore, there is little doubt that biotechnology is having a significant impact on agricultural research, that it is responsible for inducing structural change in sectors involved with plant germ plasm, but that there are no signs of significant change in the refashioning of agriculture toward environmental goals (Krimsky and Wrubel 1996, 252).

Empirical disputes will not be settled by ethical analysis, though there is little doubt that people with different ethical values also lean toward accepting the version of contested empirical claims that most supports their philosophical inclinations.

188

7 Environmental Impact and Environmental Values

Many people involved in scientific agriculture and in commercial development of agricultural technology see nothing amiss in the environmental implications of the path that has been taken on both fronts since World War II. Few, if any, of these people have failed to support the development of food biotechnology. Others, including the author (Thompson 2017), see these trends as disturbing. Many of those who would like to redirect agriculture have come out in opposition to food biotechnology. Sometimes this opposition is based on their projection of the true environmental impacts of biotechnology. Sometimes opposition is based on the belief that better investments of research funds could be made in low-input or sustainable agriculture. Sometimes the opposition seems to come from force of habit: “if biotechnology is supported by my enemies, I’m against it,” or so the reasoning seems to go. (The poise of opponents can stand some critique in this one respect.) The tools of recombinant DNA are certainly not a sufficient basis for the redirection of agriculture, and it is always difficult to determine which will be the most reliable means of doing so. All of these qualifications notwithstanding, but it is difficult for me to justify organized agricultural and food research directed toward any cause, including environmental ones, that denies itself the tools of biotechnology. Other philosophers, such as Hugh Lacey (2005), disagree. It is possible, then, to draw the following conclusion. Research and regulation should assiduously pursue the goal of making agriculture and food production more sustainable, and of making the environmental impacts of the food system easier for everyone to understand. There is no reason why techniques of recombinant DNA should be singled out, however. This is an imperative that applies to all food technology. Where there is conceptual evidence that transfer of genetic materials might result in ecological impacts that differ from those of traditionally modified plants and animals, research should be performed to empirically test these hypotheses. To attack government programs that support this research is ethically unconscionable. Nevertheless, it is probable that excessive opposition to biotechnology has provoked otherwise reasonable people to make such attacks. It is time for advocates of sustainable agriculture to refocus their efforts toward support of food biotechnology that advances an environmental agenda, and to abandon the reactive strategy of unilateral opposition. Risk assessments will be most useful when they are integrated into ecosystem models. There we will be able to see how feedback loops are affected at the ecosystem level. But there are other feedback loops that matter just as much. These are the loops that integrate our conceptions of private and public interest into an integrated conception of moral virtue, and that make good environmental practice seem like nothing more than enlightened self-interest. Even virtuous farmers are generally unaware of how their practice reinforces their moral character (and this is why linking farming with virtue is often naive and misleading). It is not clear that anything happens “automatically” in the complex and highly articulated system of feedback loops that comprise modern life. The virtues that came naturally to farmers of the past may have to be taken up and promoted explicitly. The science and business community has been reluctant to do this, and though that reluctance is not surprising, it is nonetheless disturbing.

References

189

References Alphonce, R., F. Alfnes, and A. Sharma. 2014. Consumer vs. citizen willingness to pay for restaurant food safety. Food Policy 49: 160–166. Ankeny, R.A. 2019. From food consumers to food citizens: Reconceptualizing environmentally conscious food decision-making. In Food, environment and climate change: Justice at the intersections, ed. E. Gilson and S. Kenehen, 267–279. London: Rowman and Littlefield. Ankeny, R.A., and H. Bray. 2018. Genetically modified food. In The Oxford handbook of food ethics, ed. A. Barnhill, M. Budolfson, and T. Doggett, 95–111. New York: Oxford University Press. Anonymous. 2000. Defending ‘consumerist’ ethics (interview with Peter Singer). The Philosopher’s Magazine 9: 60–61. Baier, A. 1986. Poisoning the wells. In Values at risk, ed. D. MacLean, 49–74. Totowa, NJ: Rowman and Allanheld. Baier, A. 1994. Moral prejudices: Essays on ethics. Cambridge, MA: Harvard University Press. Bogner, A., and H. Torgersen. 2018. Precaution, responsible innovation and beyond–in search of a sustainable agricultural biotechnology policy. Frontiers in Plant Science 9: 1884. https://doi.org/ 10.3389/fpls.2018.01884. Carson, R. 1962. Silent spring. Boston: Houghton Mifflin. Cheney, J. 1994. Nature/theory/difference: ecofeminism and the reconstruction of environmental ethics. In Ecological feminism, ed. K.J. Warren, 158–178. London and New York: Routledge Publishing Co. Clancy, K. 1997. 1996 Presidential address to the agriculture. Food and human values society. Agriculture and Human Values 14: 111–114. Comstock, G. 1989. Genetically engineered herbicide resistance, part one. Journal of Agricultural and Environmental Ethics 2: 263–306. Comstock, G. 1990. Genetically engineered herbicide resistance, part two. Journal of Agricultural and Environmental Ethics 3: 114–146. Crouch, M. 1991. The very structure of scientific research mitigates against developing products to help the environment, the poor, and the hungry. The Journal of Agricultural and Environmental Ethics 4: 151–158. Crouch, M. 1995. Biotechnology is not compatible with sustainable agriculture. The Journal of Agricultural and Environmental Ethics 8: 98–111. Daly, H., and J.J. Cobb. 1989. For the common good. Boston: Beacon Press. Davion, V. 1994. Is ecofeminism feminist? In Ecological feminism, ed. K.J. Warren, 8–27. London and New York: Routledge Publishing Co. de Bakker, E., and H. Dagevos. 2012. Reducing meat consumption in today’s consumer society: Questioning the citizen-consumer gap. Journal of Agricultural and Environmental Ethics 25: 877–894. Dewey, J. 1925[1981]. John Dewey the later works, 1925–1953, V. 1: Experience and Nature. Carbondale: Southern Illinois University Press. Devall, B., and G. Sessions. 1985. Deep ecology: Living as if nature mattered. Salt Lake City, UT: Peregrine Smith. Ellstrand, N.C., and L.H. Rieseberg. 2016. When gene flow really matters: Gene flow in applied evolutionary biology. Evolutionary Applications 9: 833–836. Frédéric, T. 2018. Ethics of biotechnological research. In The paths of ethics in research in Laos and the Mekong Countries: Health, environment, societies, eds. A.M. Moulin, B. Oupathana, M. Souphanthong, and B. Taverne Bernard, 103–110. Dakar: L’Harmattan-Sénégal. Gilson, E. 2014. Vote with your fork? Responsibility for food justice. Social Philosophy Today. 30: 113–130. Gilson, E., and S. Kenehan (eds.). 2019. Food, environment and climate change: Justice at the intersections. London: Rowman and Littlefield.

190

7 Environmental Impact and Environmental Values

Gustafsson, K.M., A.A. Agrawal, B.V. Lewenstein, and S.A. Wolf. 2015. The monarch butterfly through time and space: The social construction of an icon. BioScience 65: 612–622. Hare, R.M. 1981. Moral thinking: Its method, levels and point. Oxford, UK: Oxford University Press. Hargrove, E.C. 1989. Foundations of environmental ethics. Englewood-Cliffs, NJ: Prentice-Hall. Hargrove, E.C. 1994. Science, ethics and the care of ecosystems. In Natural protected areas and wilderness, ed. J. Peepre and B. Jickling, 44–60. Whitehorse, Yukon: Canadian Parks and Wilderness Society. Hettinger, N. 1995. Patenting life: Biotechnology, intellectual property, and environmental ethics. Boston College Environmental Affairs Law Review 22: 267–275. Hoffmann-Riem, H., and B. Wynne. 2002. In risk assessment, one has to admit ignorance. Nature 416: 123–123. Hooker, B. 2000. Ideal code, real world: A rule-consequentialist theory of morality. Oxford, UK: Oxford University Press. Hubbard, R. 1990. The politics of womens’ biology. New Brunswick, NJ: Rutgers University Press. Hubbard, R., and E. Wald. 1993. Exploding the gene myth: How genetic information is produced and manipulated by scientists, physicians, employers, insurance companies, educators, and law enforcers. Boston: Beacon Press. Huttner, S., H.I. Miller, and P.G. Lemaux. 1995. U.S. agricultural biotechnology: Status and prospects. Technological Forecasting and Social Change 50: 24–39. Jackson, W. 1991. Our vision for the agricultural sciences need not include agricultural biotechnology. Journal of Agricultural and Environmental Ethics 4: 207–215. Jackson, W. 1994. Becoming native to this place. Lexington: University of Kentucky Press. Kaebnick, G.E. 2007. Putting concerns about nature in context: The case of agricultural biotechnology. Perspectives in Biology and Medicine 50: 572–584. Kahneman, D. 2011. Thinking, fast and slow. New York: Macmillan. Keller, E.F. 1990. Physics and the emergence of molecular biology. Journal of the History of Biology 23: 390–409. Kloppenburg, J., Jr. 1989. Biopesticides and economic democracy. In Biotechnology and sustainable agriculture: Policy alternatives, ed. J.F. MacDonald, 75–81. Ithaca, NY: National Agricultural Biotechnology Council. Kolleck, R. 1993. Controversies about risks and their relation to different paradigms in biological research. In Science, politics and morality: Scientific uncertainty and decision making, ed. R. von Schomberg, 27–42. Dordrecht: Kluwer Academic. Kolleck. 1995. The Limits of experimental knowledge: A feminist perspective on the ecological risks of genetic engineering. In Biopolitics: A feminist and ecological reader on biotechnology, ed. V. Shiva and I. Moser, 95–111. London: Zed Books. Korthals, M. 2001. Taking consumers seriously: Two concepts of consumer sovereignty. Journal of Agricultural and Environmental Ethics 14: 201–215. Korthals, M. 2004. Ethics of differences in risk perceptions and views on food safety. Food Protection Trends 24 (7): 30–35. Korthals, M. 2015. Ethics of food production and consumption. In The Oxford handbook of food, politics, and society, ed. R. Herring, 231–252. New York: Oxford University Press. Krimsky, S., and R.P. Wrubel. 1996. Agricultural biotechnology and the environment: Science, policy, and social issues. Urbana, IL: University of Illinois Press. Kuzma, J., and J.C. Besley. 2008. Ethics of risk analysis and regulatory review: From bio-to nanotechnology. Nanoethics 2: 149–162. Lacey, H. 2005. Values and objectivity in science: The current controversy about transgenic crops. Lanham, MD: Lexington Books. Lanyon, L.E., and D.B. Beegle. 1989. The role of on-farm nutrient balance assessments in an integrated approach to nutrient management. Journal of Soil and Water Conservation 44: 164– 168.

References

191

Lehner, M. 2013. Alternative food systems and the citizen-consumer. Journal of Agriculture, Food Systems, and Community Development 3 (4): 49–53. Leopold, A. 1949. A sand county almanac and sketches here and there. New York: Oxford University Press. Levidow, L. 1995. Whose ethics for agricultural biotechnology. In Biopolitics: A feminist and ecological reader on biotechnology, ed. V. Shiva and I. Moser, 175–190. London: Zed Books. Levidow, L., and S. Carr. 1997. How biotechnology regulation sets a risk/ethics boundary. Agriculture and Human Values 14: 29–43. Lewontin, R.C. 1992. Biology as ideology. New York: Harper Perennial. Losey, J.E., L.S. Rayor, and M.E. Carter. 1999. Transgenic pollen harms monarch larvae. Nature 399: 214–214. Marks, L.A., N. Kalaitzandonakes, L. Wilkins, and L. Zakharova. 2007. Mass media framing of biotechnology news. Public Understanding of Science 16: 183–203. https://doi.org/10.1177/096 3662506065054. Mellon, M. 1992. The regulation of genetically engineered animals: Going from bad to worse. In Animal biotechnology: Opportunities and challenges, ed. J.F. MacDonald, 165–169. Ithaca, NY: National Agricultural Biotechnology Council. Mies, M. 1993. New reproductive technologies: Sexist and racist implications. In Ecofeminism, ed. M. Mies and V. Shiva, 174–195. London: Zed Books. Nord, E.A., and L.E. Lanyon. 2003. Managing material transfer and nutrient flow in an agricultural watershed. Journal of Environmental Quality 32: 562–570. Norton, L.R. 2016. Is it time for a socio-ecological revolution in agriculture? Agriculture, Ecosystems & Environment 235: 13–16. NRC (U.S. National Research Council). 2000. Genetically modified pest protected plants: Science and regulation. Washington DC: National Academy Press. NRC (U.S. National Research Council). 2002. Environmental effects of transgenic plants: The scope and adequacy of regulation. Washington DC: National Academy Press. Passmore, J. 1974. Man’s responsibility for nature. London: Duckworth. Preston, C.J. 2008. Synthetic biology: Drawing a line in Darwin’s sand. Environmental Values 17: 23–39. Rolston, H., III. 1991. Environmental ethics: Values in and duties to the natural world. In Ecology, economics, ethics: The broken circle, ed. F.H. Bormann and S.R. Kellert, 73–97. New Haven, CT: Yale University Press. Rolston, III, H. 1999. Genes, Genesis, and God: Values and Their Origins in Natural and Human History. New York: Cambridge University Press. Sagoff, M. 1986. Values and preferences. Ethics 96: 301–316. Sagoff, M. 1988. Biotechnology and the environment: What is at risk? Agriculture and Human Values 5 (3): 26. Sagoff, M. 2007. The economy of the earth: Philosophy, law, and the environment, 2nd ed. New York: Cambridge University Press. Sandler, R.L. 2009. Character and environment: A virtue-oriented approach to environmental ethics. New York: Columbia University Press. Schurman, R., and W.A. Munro. 2010. Fighting for the future of food: Activists versus agribusiness in the struggle over biotechnology. Minneapolis: University of Minnesota Press. Scott, N.D. 2018. Food, genetic engineering and philosophy of technology. New York: Springer. Sessions, G. (ed.). 1995. Deep ecology for the 21st century: Readings on the philosophy and practice of the new environmentalism. Boston: Shambabla Publications Inc. Shiva, V. 1993. Monocultures of the mind: Perspectives on biodiversity and biotechnology. London: Zed Books. Siddig, A.A., A.M. Ellison, A. Ochs, C. Villar-Leeman, and M.K. Lau. 2016. How do ecologists select and use indicator species to monitor ecological change? Insights from 14 years of publication in Ecological Indicators. Ecological Indicators 60: 223–230. Singer, P. 1979. Practical ethics. Cambridge, UK: Cambridge University Press.

192

7 Environmental Impact and Environmental Values

Thompson, P.B. 2010. The agrarian vision: Sustainability and environmental ethics. Lexington: University Press of Kentucky. Thompson, P.B. 2017. The spirit of the soil: Agriculture and environmental ethics, 2nd ed. New York: Routledge. Thompson, P.B. 2018. Farming, the virtues and agrarian philosophy. In The Oxford handbook of food ethics, eds. A. Barnhill, M. Budolfson, and T. Doggett, 53–66. New York: Oxford University Press. Thompson, P.B., R. Matthews, and E. van Ravenswaay. 1994. Ethics, Agriculture and public policy. New York: Macmillan. Varner, G.E. 2012. Personhood, ethics, and animal cognition: Situating animals in Hare’s two-level utilitarianism. New York: Oxford University Press. Verbeke, W., F.J. Pérez-Cueto, M.D. de Barcellos, A. Krystallis, and K.G. Grunert. 2010. European citizen and consumer attitudes and preferences regarding beef and pork. Meat Science, 84: 284– 292. Westra, L. 2011. Ecological integrity and biological integrity: The right to food. In Sustaining soil productivity in response to global climate change: Science, policy, and ethics, ed. T.J. Sauer, J.M. Norman, and M.V.K. Sivakumar, 103–115. Chichester, West Suffix UK: Wiley-Blackwell. Wickson, F. 2014. Environmental protection goals, policy & publics in the European regulation of GMOs. Ecological Economics 108: 269–273. Wilkins, J.L. 2005. Eating right here: Moving from consumer to food citizen. Agriculture and Human Values 22: 269–273.

Chapter 8

Social Impact and the Technology Treadmill

Abstract The chapter introduces the analysis of socioeconomic impact from agrifood biotechnology through a review of the technology treadmill. In this model, farmers experience ethically significant harmful impacts, while secondary beneficial effects accrue to consumers (as well as technology developers). The chapter then reviews how these social impacts would be evaluated from a number of philosophical perspectives and theories of social justice. Rights theories, utilitarianism and virtue theories are sketched, and the chapter discusses how each approach might provide an analysis of the treadmill phenomenon. Analyses that link structural injustice to feminist or Marxist social theory anticipate the further discussion of social consequences in Chap. 9. Keywords Social justice · Agricultural innovation · Distributive justice · Fairness · Utilitarianism · Marxism · Feminism · Structural injustice · Food justice Previous chapters have considered food and agricultural biotechnology’s potential for unwanted impact on the health and safety of individual human beings, of nonhuman animals and on the broader environment. This chapter introduces the last category of unwanted impacts: Socioeconomic hazards include those that adversely affect individuals’ economic welfare and daily practice, as well as impacts on human relationships, including households, communities, organizations and other human institutions. For any technology, social impact can be markedly dispersed in both space and time, and accrues through a tremendous variety of mechanisms. Agricultural technologies can have particularly significant impacts. For example, Karl Wittfogel (1896–1988) studied innovations in irrigation technology, demonstrating how dramatically they changed human relationships. Construction of reservoirs or aqueducts shifts the balance of power between upstream and downstream users. Large-scale water management can require extensive coordination of individual activities, which in turn create capacities for coordination of human action that penetrate throughout the fabric of a society. Wittfogel stressed connections between the centralized water management of ancient China and its rigidly hierarchical political organization (Wittfogel 1958, 1962).

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_8

193

194

8 Social Impact and the Technology Treadmill

Wittfogel’s study of agricultural infrastructure directs attention to the questions philosophers should be asking when they study innovation. They are not exclusively ethical questions. Particular innovations in agricultural production occurred in geographical areas with specific soil types, climates and availability of water, but similar technologies are now spread throughout the world. The diffusion of agricultural technology also raises ethical questions because production systems that work well in place can cause ecological and social disruptions when they are transferred to a new location, (see Crosby 1986; Hugill 1993). In fact, changes in technical infrastructure can induce deep alteration in the way that human beings experience their world at the deepest level. The development of timekeeping technology revolutionized social organization, making spatially discontinuous coordination of bureaucratic activities possible and paving the way for the creation of modern states, (Mumford 1934; Landes 1983). Recent historical and sociological studies of technology have linked such disparate events as the rise of psychoanalysis to the development of the steam engine (Edge 1973), and the sexual revolution to the automobile (Jeansonne 1974). My general approach in philosophy of technology draws on this background literature and this chapter takes pains to articulate key social mechanisms for understanding innovation and social change.

8.1 Technology, Politics and the Prediction of Social Change As discussed in Chap. 2, Hans Jonas developed the philosophical rationale for predicting and managing the risks of technological innovation in The Principle of Responsibility. Though foresighted in calling attention to health and environmental impacts, Jonas was primarily interested in the way that technology functions as an infrastructure for interpersonal relations. He stressed the way that communication and transportation technologies were transforming the idea of responsibility itself. Throughout much of human history, ethical evaluations of human conduct presumed that face-to-face interaction would predominate. Adam Smith had noticed initial transformations of the day-to-day giving rise to anonymous transactions and contracting. Jonas argued that by the 1970s, the very nature of the ethical domain had already been transformed by what we would today call globalization, (Jonas 1984). Jonas argued that attempts to manage technology’s social consequences is controversial in part because the mechanisms that link technological innovation to its eventual impact are generally opaque to non-specialists (including many of the scientists, engineers and administrators who bring about the innovation), and often obscure even to scholars of technology. This basic conceptual problem would limit our ability to discuss and debate the social consequences of technology in the best of circumstances, but agrifood biotechnologies were not introduced under the best of circumstances. Cold War politics complicated debate over technology during the last half of the twentieth century. The less industrialized agricultural economies of Africa, Asia and Latin America became a battleground. Competing strategies for promoting economic development in nations that were gaining independence

8.1 Technology, Politics and the Prediction of Social Change

195

from colonial powers reflected the debate between capitalism and socialism. The small group of social activists at the core of the anti-GMO social movement were critics of capitalism, while scientists working with corporate support were committed implicitly to market mechanisms for diffusion of their innovations, (Schurman and Munro 2010). The debate over GMOs was prefigured as a debate over social philosophy, with boosters taking a dismissive attitude toward any claim that smelled of Marxist influence. Yet, from a philosophical perspective, omitting Marx’s thought from any discussion of technology’s social consequences is either naïve or intellectually dishonest. The following analysis is indebted to his work. Nevertheless, many regard Marx’s ideas as suspect, and the use of his name inevitably colors the manner in which assessments of technology’s social impact is received, especially in the United States. The upshot is that simply predicting the social consequences of food biotechnology can spark controversy, irrespective of the norms or values that are applied in evaluating the ethical significance of those consequences. Schurman and Munro’s analysis of the social movement to oppose gene technology describes antibiotechnology advocates as resisting the growing power of multinational corporations. They wanted to shield small farmers from corporations’ ability to dominate markets for farm commodities, (Schurman and Munro 2010). However, Schurman and Munro do not discuss the socioeconomic mechanisms that link agricultural production technology to harmful impact on small farmers. Some sort of account that portrays introduction of the technology as a cause, and social disruption as the effect is critical for a risk-based approach. If no credible causal mechanism can be produced, the unwanted social outcome will not survive the hazard identification process in risk assessment. Writing at a more general level, Jonas also treats the political argument as a contest between socialist and liberal theories of society. In Jonas’s interpretation, rights theories and utilitarianism illustrate the liberal tradition of social justice. He asks whether capitalist or socialist political systems are more likely to defend these liberal virtues, ending by claiming that circa 1975, when his book was written, neither approach is doing a good job of implementing responsible oversight of technology. Jonas discusses biophysical mechanisms (nuclear war, eugenics) that link technology to health and environmental outcomes, but he is much less forthcoming in explaining how technical change causes social change. Jonas seems to accept the view that when capital is managed by profit-seekers, the interests of third parties will be underserved. Rapacious factory owners do create jobs, but they also pollute and exploit their workers. Socialism has failed to redress the problem, in Jonas’s view, both because centralized power structures are easily corrupted and because social impacts are themselves very complex and difficult to manage. Jonas advocates philosophically informed risk assessment as a way to constrain the failures of both systems, (Jonas 1984). As noted in previous chapters, this book builds upon Jonas’s approach. I will add virtue-based agrarian approaches that provide a challenge to liberal assumptions not unlike those of feminist and postcolonial orientations that have emerged with greater clarity since the first edition of this book was written. When approaches that

196

8 Social Impact and the Technology Treadmill

stress structural causes of injustice are included, a matrix for examining the ethical significance of social consequences begins to emerge. On one axis are the respective approaches to justice, on the other are the main types of social consequence that have been associated with food biotechnology: impact on small or family farms, impact on agriculture in developing countries, and impact on the organization and structure of science itself. In addition to taking a broadened normative framework, my adaptation of Jonas is also more attentive to specific socioeconomic relationships that cause unwanted outcomes. The chapter includes a review of key mechanisms that link technological innovation to social change in the context of agriculture and food systems. The next section explores the five theoretical positions on technology and social justice listed above in more detail. The balance of the chapter fills in the matrix of utilitarian, rights-based and virtue/agrarian/feminist ethics by speculating on how each of these five theoretical positions might be applied to the three main types of social change. It must be repeated that technology’s capacity for unanticipated social impact makes any effort to anticipate social consequences subject to a high level of uncertainty and incompleteness. The effort reflected in this chapter is no exception.

8.2 The Social Consequences of Food Biotechnology The economics of food and agricultural production drive the technological changes and attendant social consequences that are the focus of this chapter. Farmers are always looking for ways to do things a little better. As societies become organized on the industrial model, it becomes possible to make a living (sometimes a very good living) by making things that help farmers do a little bit better and selling them. Agricultural economist Willard Cochrane (1914–2012) developed these unexceptional observations into an analysis of the technological treadmill in agriculture: When a new production technology allows farmers to reduce the cost of production, early adopters of the technology reap substantial profits. They can produce more than their neighbors can with a comparable investment of time, labor and capital. As long as commodity prices are stable, this extra production is translated into extra profit. However, as more and more people adopt the new technology, total food production begins to rise, and commodity prices begin to fall. This usually happens because the world can only use so much food. When prices fall, those who continue to use the old technology find themselves operating at a loss, and many go out of business. Those who adopted the new technology find that higher profits disappear; they are running harder (e.g. producing more volume of food) to stay in the same place (e.g. retain an income level comparable to what they had before the new technology came along), (see Cochrane 1979, 389–90; Browne and co-authors 1992, 56). Cochrane popularized the technological treadmill in the United States during the 1950s and 1960s, although the idea that farmers were on a treadmill of some sort was commonplace even in the 1930s (see Griswold 1948). He made a concept central to Marx’s analysis of technical change acceptable to conventional economists and

8.2 The Social Consequences of Food Biotechnology

197

to the conservative American farming community by toning down the rhetoric and by applying it to an industry (e.g. farming) where the tension between ownership and the wage rate for labor was more psychological than social. Marx himself had characterized the phenomenon that later became known as the treadmill this way: During the transition period when the use of machinery is a sort of monopoly, the profits are exceptional, and the capitalist endeavors to exploit thoroughly “the sunny time of this his first love,” by prolonging the working day as much as possible. The magnitude of the profit whets his appetite for more profit. As the use of machinery becomes more general in a particular industry, the social value of the product sinks down to its individual value, and the law that surplus-value does not arise from the labour-power that has been replaced by the machinery, but from the labour-power actually employed in working with the machinery, asserts itself, (Marx 1867, 405).

These passages from Das Kapital may state Marx’s “law” in very general terms, but they indicate that Marx was aware of the technological treadmill one hundred years before Cochrane. However, agriculture has yet to become so concentrated that speaking of monopolies is appropriate. In agricultural economies with competitive land markets, the treadmill produces a slightly different effect. Early adopters invest their windfall profits into land, buying up land holdings from the failing smaller farms. The net effect of productivity enhancing technology is summarized by the phrase, “fewer and larger farms.” Mainstream agricultural economists (who hardly think of themselves as Marxist) have now accepted this economic analysis. This dictum of “fewer and larger farms” was applied to biotechnology in Cornell University economist Robert Kalter’s theoretically unexceptional, but politically ground shaking, study of recombinant bovine somatotropin (rBST). Kalter fed the productivity increase predicted for rBST into economic models of the dairy sector, and to no social scientist’s surprise, out came the “fewer and larger farms,” result, (Kalter 1984, 1985; Kalter and co-authors 1985). Publication of the result precipitated uproar, however. The rBST case may have been the first time that producers realized the likely impact of production enhancing technology, and organized to fight it, (Buttel 1986, 2000; Browne 1987). Kalter’s early studies also sparked a debate over the mechanisms of technical change among economists. The socio-economic mechanisms linking technical change and social transformation among farmers are more complex than a simply statement of the treadmill analysis might suggest. One point of dispute arose because Kalter’s study came on the heels of controversy over adoption of mechanical tomato harvesters in California. In that case, only relatively large farms could afford to adopt the new technology; a tomato harvester is an expensive piece of equipment that is uneconomical to operate on a small plot of land. However, the reduction in market price for fresh and canning tomatoes did indeed have the effect of putting many small growers out of business, in this case leading not simply to “fewer and larger” but in fact to “no small, and even larger large” (Schmitz and Seckler 1970; Berardi and Geisler 1984; Ruttan 1991). Small farm activists knew the tomato harvester case well, and they might have organized to fight any significant technical change that came along in about 1985. American farmers were facing a number of financial difficulties in the 1980s, leading to a surge of bankruptcies. Economists now blame the 1980s farm crisis on bad

198

8 Social Impact and the Technology Treadmill

credit policies in the savings and loan sector of the U.S. economy. Easy credit induced farmers into technological investments that increased the total productivity of the farm sector, and this, in turn, precipitated a downturn in farm prices, just as Cochrane predicted. Farmers carrying large debts were highly vulnerable. Both bankruptcy and farmer suicide were the result. Some estimates suggest that more than one third of the farmers operating in 1980 had been forced to leave farming by 1990 (Iowa Public Broadcasting 2013). Farmers were supported by political activists who called attention to their plight and argued for reforms in farm policy (including agricultural research) that would target benefits more effectively toward smaller-scale, economically vulnerable farmers. Biotechnology came to be thought of as very significant technical change in this intellectual environment largely because that was the way that the scientists developing agrifood biotechnology described it. In fact, there is little evidence that most kinds of agricultural biotechnology now in the field or contemplated for development possess the “size bias,” exemplified by the tomato harvester. This point was made by economists who disputed Kalter’s prediction (Yonkers and co-authors 1986), implicitly defending the biotechnology industry. In this way, the seeds were sown for a debate over the social consequences of biotechnology that involved seemingly arcane disputes among economists. Cochrane’s treadmill concept makes no appeal to size-related efficiencies, however, though size efficiencies could clearly exacerbate the trend he predicted. Nevertheless, the “fewer and larger,” consequence is not a necessary consequence of economic theory, even when size-bias is absent. Farmers might also capture savings by reducing inputs and continuing to produce the same volume of output. Such behavior would have little effect on prices, but farmers would share a small savings from reduced production cost. Economist Loren Tauer summarized the complex strands in this economic debate as it applied to rBST, noting that even if biotechnology does reduce the profitability of dairying, many small dairies will simply accept a lower return and remain in business. Many other technological forces were affecting the economics of milk production, not the least of which are automated milk and animal health monitoring systems that are far from scale neutral. Tauer concluded that it is impossible to measure the effects of biotechnology on small vs. large farms, but “to argue that BST will have no differential impact by farm size is tenuous at best. The issue is the extent of the impact,” (Tauer 1992). A further complication of the technological treadmill argument concerns the rate at which farmers adopt new technology. Part of the treadmill logic is that early adopters reap windfall profits that they then reinvest in more land as the late adopters go bankrupt. If everyone were to adopt the technology all at once, price adjustments would be immediate, there would be no windfall profits and no one would go broke. Everyone would be making less money, but that would only be a problem in a world where agricultural subsidies do not make up the difference, anyway. Thus, the next round of debate over agricultural biotechnology concerned adoption rates, and a number of economists undertook studies of this problem. As one might expect, it turns out to be complex. Farmers in some parts of the U.S.could benefit from Bt maize, but not others. Would there be regional adjustments in profitability? And everything depends on how much the companies charge for the new biotechnologies. If their

8.2 The Social Consequences of Food Biotechnology

199

economists are very sharp, these firms will be able to set the seed prices low enough so that farmers need to buy transgenic seeds, but high enough that most of the windfall goes to the biotechnology company. If the profits go to the technology provider, this has the ironic result of reducing the treadmill effect. Studies on the economic impact of biotechnology continue to be released every year, generally identifying reasons why the impact of the technological treadmill has been (or will be) more muted than might have originally been expected. Though “fewer and larger farms,” exhausts the meaning of the treadmill for many economists who study technical change, sociologists have always been interested in the same changes that interested Marx: the structure and character of ownership and labor relations. If there are fewer farms, where do the farmers go? The Marxist assumption is that they go into labor markets as wage laborers. The treadmill is thus an account of how societies that consist of many independent, owner-operated farms become societies that consist of a few land and capital owning investors, and legions of workers who must accept the going rate for wage labor. The social transition described by the treadmill is, thus, a change in social structure. A society of owneroperators, each with individual control over their work activity and relatively equal economic opportunity to succeed, gradually becomes a society of capital owning bosses who control the work life of laborers, and who determine the future direction of society through their investment decisions, (Kloppenburg 2004). Clearly, genetic engineering is not the only or even the most important technology implicated in this transition, and the transition itself was arguably complete in industrialized countries long before 1980. However, sociologists who have studied biotechnology conclude that biotechnology is implicated in the technological changes that bring this transformation of social structure to peasant agricultural economies in the developing world (Buttel and Barker 1985; Kenney and Buttel 1985). Furthermore, though it is easiest to understand the structural transition in terms of individuals and families moving from family farms to wage labor, there are more abstract (but equally important) shifts that occur. Sometimes global trade patterns conceal the transition, as an entire nation of small farmers become displaced or marginalized, while urban populations come to depend on industrialized agriculture from Europe, North America, Japan and Australia (see Buttel et al. 1985). Furthermore, even the winners among the fewer and larger must share a larger portion of their farm profits with the companies that produce the technology, and they become dependent on those companies in a manner quite similar to the way that wage laborers depend on their employers (Kloppenburg 2004). Finally, the technological treadmill and its long-term consequences can have effects on the structure of agricultural research. In most parts of the world, agricultural research has been conducted by non-profit and government agencies. It has been thought to be in service to the public good, in large part because 100 years ago, the vast majority of the world’s population was engaged in farming. As the treadmill transition reduces the farming population to 2% of the whole or less, three mutually reinforcing drivers spur the privatization of research. First, as farm population declines, the political base for publicly supported research declines. With fewer farmers, there are fewer people to write elected officials in order to argue for policies

200

8 Social Impact and the Technology Treadmill

that favor the agricultural sector. Second, as farm population declines, productivity enhancing research comes increasingly to look like a subsidy to special interests, rather than a service for the public good. With a significant majority of the population in farming, policies that serve agricultural interests can be seen as serving the public good, but when farm producers make up less than 2% of the population, this interpretation becomes less plausible. Finally, as farms become fewer and larger, the costs of marketing to farmers are lower and the potential rewards are higher. Private venture capital is attracted into agricultural research in way that it was not when farmers were many and poor (Kenney 1986; Busch and co-authors 1991). Publicly funded agricultural research—once understood as benefiting a broad segment of the relatively less well off—now appears redundant at best. Government funded agricultural research can even be seen as a subsidy to the large companies developing agricultural technologies with the goal of profits in mind. This account summarizes a great deal of social science research and in doing so oversimplifies socioeconomic assessment of agrifood biotechnology’s social consequences. Yet even a simplified account of the technology treadmill corrects the impression that anti-biotechnology activists were working solely from an ideological distaste for markets or fear of corporate power. Economists who studied the 1980s farm crisis in the United States were divided about what should be done about it. Some shrugged their shoulders and said that it just proved there were too many farmers for an efficient agricultural economy, (Tweeten 1983, discussed later). Others saw it as a situation calling for government intervention, (Breimyer 1983). There was no debate about the treadmill’s economic causality, however. From an ethics perspective, an analysis that fails to address the bigger and fewer phenomenon in normative terms is thus seriously incomplete. Four points must be emphasized in completing the summary, however. First, the engine that is driving most of these changes is simple economic rationality. People who adopt or invest in the development of new technology do so because they think that they can benefit economically; people resist the technology because they think that resistance will benefit them more than simply adopting the new technology. Second, any production enhancing technology is likely to have these effects, and the impact of any specific product or class of products such as biotechnology will be diluted or intensified by that of other technologies—computerization, satellite imagery, mechanization—that may be coming on line at the same time. Third, my summary of the treadmill omits outcomes that both defenders and critics of GMOs would deem important. Consumers generally benefit (even if farmers do not) when production-enhancing technology is adopted (see Tweeten 1991). Furthermore, the widespread use of genetic engineering is clearly affecting the way that people think about everything in nature, including themselves (Nelkin and Lindee 2004). There is, in short, a cultural change on top of all this socio-economic change. The mechanisms that link a production technology to these secondary and tertiary consequences are even more obscure, more controversial and more difficult to trace. Some will be picked up in discussing the ethical significance of social consequences from agrifood biotechnology below, but others are simply not captured by the matrix organization of this chapter. Fourth, none of what has been said in this section need entail anything at

8.2 The Social Consequences of Food Biotechnology

201

all about whether the social consequences described are good or bad, fair or unfair, just or unjust. Clearly many of the authors who predicted or analyzed social consequences had opinions on these questions, but one must have a reasonably clear picture of what makes for right and wrong, for fairness and justice, before such questions can be approached with even a modicum of philosophical rigor or conceptual clarity.

8.3 Theories of Justice This section reprints earlier editions’ overview of the main themes that emerge in philosophers’ attempts theorize the elements of social justice. As noted in the Introduction, the first edition of the book was written for an audience that could not have been expected to have facility with the basic terminology of political theory. An appreciation of the ethical debate over social consequences presupposes some familiarity with the terms of political theory. This section is retained in the current edition as a primer for readers who want one. The main themes have already appeared in previous chapters. A utilitarian or consequentialist approach to social justice evaluates social changes in terms of whether they tend to produce an attractive ratio of benefit to cost for all affected parties. Rights based theories evaluate social change as acceptable when they take place under circumstances where rights are respected and enforced, and as questionable—possibly unacceptable—when they do not. These two alternatives emerged in one form or another in each of three previous chapters, as did the related idea that it is fair procedures that make for justice, without regard to outcomes. The importance of virtue arguments was raised in connection with ethics in production and feminism, discussed in Chap. 7. Marxism and structural injustice are treated as distinct philosophical issues, though it would not be implausible to suggest that they reflect an alternative approach to the theory of justice.

8.3.1 Utilitarianism and Utilitarian Theories of Justice Previous chapters have explored the contrast between expected value treatments of food safety and the problem of consent, utilitarian and other sentience views on animal welfare and the expected-value approach to environmental risk. In each case, some variant of utilitarian philosophy is evident. Utilitarianism is the moral and political philosophy usually associated with the English philosophers Jeremy Bentham (1748–1832) and John Stuart Mill (1806–1873). Bentham and Mill advocated for a single, fundamental principle for evaluating an individual’s action, a public policy or law, and even a broad social change. It was “that principle which approves or disapproves of every action whatsoever, according to the tendency which it appears to have to augment or diminish the happiness of the party whose interest is in question,” (Bentham 1789, 2). Known alternately as “the principle of utility,” “the greatest

202

8 Social Impact and the Technology Treadmill

happiness principle,” and “the utilitarian maxim,” the rule is usually generalized to consider the greatest good for the greatest number of affected parties, all things considered. Mill reserved the term justice for “certain classes of moral rules which concern the essentials of human well-being more nearly, and are therefore of more absolute obligation, than any other rules for the guidance of life,” (Mill 1861, 58). The principle of utility needs a great deal of further specification before it can be used as a decision rule for policymaking, but even in its general form it entails a number of philosophically significant commitments: • The justice of an action or social change is determined by its effect on the welfare (e.g. health, wealth or well-being) of individuals. • The effects of an action or social change on multiple individuals are to be summed or aggregated. • Rights, norms and legal codes are relevant to the morality of an action or social change only insofar as they are instruments for bringing about consequences for individuals. • Actions or policies are justified when they achieve a maximum (or at least optimum) production of welfare, when compared to other alternatives (Sen 1987). These philosophical commitments have been debated extensively for over 200 years. Many social theorists have developed interpretations of the utilitarian approach that abandon or modify one or more of these assumptions. Some have argued that it is impossible to use the theory because individual welfare cannot be measured. Others have questioned the maximization rule. Whether one is committed to maximizing welfare or not, the practice of comparatively ranking multiple options tends to turn utilitarian moral evaluation into a procedure that seeks optimal or efficient distributions of benefit and cost. Working out the details of these modifications will tire all but the most patient readers, however, and a simple presentation is adequate to the task at hand. Social changes brought on by productivity enhancing technology have been generally thought consistent with the principle of utility. Those who advocated agricultural research at the turn of the century clearly thought that improvements in technology would benefit the farmers themselves (Rosenberg 1961). The treadmill concept shook this belief, but not the utilitarians’ favorable view of technical change. If the combination of benefits to food consumers (in the form of reduced food prices) plus benefits to the winners in technical change is sufficient to compensate for costs to the losers, the end-state redistribution of welfare (consumers and big farmers are better off, small farmers are worse off) is still consistent with the principle of utility. Several generations of agricultural economists applied utilitarian principles to the evaluation of technical change, and liked what they saw, (see Tweeten 1987 for an example; Thompson et al. 1994, 233–245 for a discussion). However, a number of side issues complicate the utilitarian calculation. In some cases, individuals choose the option that maximizes utility, given their own set of limited choices, but the result is that everyone is worse off, (Epstein 1996). This may arise because each individual is neglecting ‘externalities,’ that is, consequences— either costs or benefits—borne by persons or groups who are not decision makers or

8.3 Theories of Justice

203

parties to a contract. From a philosophical standpoint, one does not have a defensible utilitarian analysis until external impacts are accounted for. Since technical change affects not only humans and animals, but also subsequent generations extending into the future, a complete analysis of external costs may be difficult to achieve. Nevertheless, those who apply a utilitarian analysis of social justice tend to be favorably disposed toward technical change. This is true whether one assesses technological change in the broadest sense (see Rosenberg and Birdzell 1986), or whether one applies the theory specifically to agriculture and food biotechnology, (see Huttner, Miller and Lemaux 1995). History teaches that technology seldom delivers all the benefits that are promised, and that costs are often higher than expected. Nonetheless, when costs and benefits are averaged over winners and losers and over time, it is difficult to argue with progressive tendencies of technical change, evaluated in utilitarian terms (see Rosenberg 1992 for an extended argument; Tenner 1996 for a more qualified endorsement). The utilitarian approach suggests that we should just accept the results of technological innovation, subject (perhaps) to some minor modification to address problems of collective action and externalities. Of all philosophical theories of justice, the utilitarian view comes closest to providing a rationale for traditional views of technological progress.

8.3.2 Justice and Rights Justice talk is pervasive in debates over the social consequences of agricultural technology. Labels for genetically engineered foods are justified in terms of a consumer “right to know.” Harms to animals or even ecosystems are said to be unjust because they violate rights held by these entities. In many disputes over public policy and technical change, rights arguments appear in direct rebuttal to utilitarian arguments. Often the point is to reject the utilitarian practice of aggregating or summing benefits and costs. Harms that violate the rights of an individual are thought to be so severe that no amount of compensating benefit to others can justify them. More generally, if someone (or some thing) has a right to X, whether X be information, property, economic opportunity or life itself, they may make a justified claim to X. This claim imposes duties to others and on the entire society who must either deliver X, or must at least not interfere with the rightsholder’s pursuit or disposal of X. These duties “trump” or override other cost/benefit considerations (Fineberg 1980; Donnelly 1989). Technical change would be justified on a rights view so long is it did not violate any individual’s rights. Note that this principle has the potential to be both more and less exacting than a utilitarian approach. It is more exacting in that even a single rights violation makes technical change unacceptable. Even if only one party’s rights are violated by a technical change, the change is unjust. It is less exacting in that changes need not pass an efficiency test, nor would one worry about collective action dilemmas, so long as each individual is making choices protected by rights. Of course, anyone can claim a right; the question is when are these claims justified? What rights

204

8 Social Impact and the Technology Treadmill

do people (or animals) really have? Two broad strategies for answering this question bear directly on the problem of technical change. They are distinguished from one another through the difference between negative rights, or rights that require only that others forbear (e.g. not perform) harmful acts, and positive rights, or rights that require one to undertake action on behalf of others. Libertarian theories recognize negative rights only. Broader rights theories include both negative and positive rights. These broader theories will be referred to here as ‘egalitarian.’ Libertarian Theories. Libertarians approach the question of which rights to recognize by assuming that the most desirable state is one of perfect and complete liberty. However, if everyone is at complete liberty, everyone is also at risk, for people who are totally free are free to harm one another. Hence it is rational to accept principles that restrict liberty at exactly the point that an exercise of liberty would be harmful to someone else. This reciprocal restriction of liberty means that one has a negative duty with respect to harming others, that is, a duty not to do things that harm others. Other people are justified in claiming that one must forbear such harmful actions, hence they have a moral right that is violated when harmful acts are performed. Libertarian rights protect the life and personal security of people, their liberties of conscience, movement and speech, and their free use of their property, so long as that use does not harm others, (see Thompson et al. 1994, 39–44). Libertarian protection of property rights provides the strongest philosophical argument for free-market economic principles. It is always wrong to interfere in someone’s use or exchange of property, unless of course that use constitutes a harmful act. This means that it would be wrong to interfere even in collective action dilemmas where individuals use property in ways that are contrary to their own interests. Although they may be making themselves worse off collectively, no individual’s act violates the life, liberty or property of another. On the other hand, libertarian theories also provide the strongest philosophical arguments for intervening to prevent externalities. If a person’s use of technology harms another, through pollution or exposure to environmental risk, for example, it is wrong, irrespective of whether it provides social benefits that compensate for those harms, (Machan 1984). Egalitarian Theories. Many of the rights claimed by individuals in advanced societies require more than abstinence or non-interference by others. If one has a right to education, someone must do the educating when this right is claimed. If one has a right to information, someone must provide it. If one has a right to employment, someone must offer a job. These positive rights expand the scope of rights arguments considerably, and they also increase the likelihood that there will be conflicting rights claims. Clearly if there are positive rights that require the entire society to set up schools, for example, these rights will require taxation that, on the face of it, violates individuals’ negative rights to control the use of their property. Rights theorists who admit positive rights are thus deeply concerned with the problem of limiting the expansion of rights claims and with reconciling conflicts among rights. Most approaches do this by placing positive rights in a hierarchy, so that claims to basic needs such as minimal health care, food and income opportunities are met for everyone. Once such basic rights have been guaranteed, it may be possible to

8.3 Theories of Justice

205

expand the scope of rights claims to include literacy, higher education, or perhaps even recreational opportunity, (Shue 1980). Positive rights arguments provide the most plausible way to interpret the ethical significance of structural changes brought about by technical innovation. The treadmill phenomenon described above has had mixed results for human opportunity. On the one hand, it has created opportunities for work outside of agriculture, and is the cornerstone of liberal societies that aim to guarantee a wide variety of positive rights to healthcare, education and opportunity for their citizens. On the other hand, the transitions described by sociologists are changing the agricultural production sector so that fewer people control decision-making. The autonomy of individuals may be eroding at the same time that the universe of food choices is expanding, (Busch and co-authors 1991, 191–203; Burkhardt 1992). If people have a positive right to have control over their lives and destinies in a strong sense, the decline of rural communities in which many (if not most) people had the opportunity to work for themselves, rather than for wages, may be seen as an inherently regressive social trend.

8.3.3 Justice and Virtue Both utility and rights are historically recent innovations when viewed in the 2500 year time frame of philosophical thinking. The view that a society is just to the extent that it provides a structure of interpersonal relationships, incentives and reinforcements to virtue is a more traditional way of conceptualizing justice. Philosopher Alisdair MacIntyre launched a revival of virtue theory with his book After Virtue (1984). He offers the following as a “partial and tentative definition of a virtue. A virtue is an acquired human quality the possession and exercise of which tends to enable us to achieve those goods which are internal to practices and the lack of which effectively prevents us from achieving any such goods,” (MacIntyre 1984, 191) By ‘practice,’ he means. any coherent and complex form of socially established cooperative human activity through which goods internal to that form of activity are realized in the course of trying to achieve those standards of excellence which are appropriate to, and partially definitive of, that form of activity, with the result that human powers to achieve excellence, and human conceptions of the ends and goods involve, are systematically extended. Tic-tac-toe is not an example of a practice in this sense, nor is throwing a football with skill; but the game off football is, and so is chess. Bricklaying is not a practice; architecture is. Planting turnips is not a practice, farming is. (MacIntyre 1984, 187).

MacIntyre criticized the moral traditions that discuss moral character only as an instrument for maximizing utility or respecting rights. This characteristically modern way of thinking about character and virtue inverts the proper form of the relationship, as seen by a virtue theorist (MacIntyre 1984, 108–120). The virtues we associate with good moral character are not tested by whether they encourage social utility or respect for rights. Virtues emerge out of the practices that represent the deepest moral commitments of a community. It is only when these moral commitments are

206

8 Social Impact and the Technology Treadmill

understood that it becomes possible to talk about the morality of social utility or rights. Virtue theory presents at least three ways in which technical change might be thought ethically problematic. First, to the extent that technical change is linked to social rationalization and to increasing sway of economically formalized interpersonal relationships, it may contribute to a general decline in the virtues. Second, to the extent that technology makes the performance of tasks routine and unreflective it contributes to the loss of human practices. Third, to the extent that traditional agrarian societies and family farms represent repositories of human practice and its virtues, technical change in the food system is particularly inimical to an ethics of virtue. How would a virtue theorist address these themes? Rationalization and Commodification. Some of the most sweeping objections to biotechnology are based on the view that once sacred spaces are being given over to the economic sphere. Things, processes or activities that were never even thought of as being capable of being traded, bought or sold are now being “commodified,” or turned into goods that can be owned or exchanged at a price. On this view, it is objectionable to even think of life and life processes as “having value,” in the sense used by utilitarians, or as being claimed as a property right. Even applying the moral categories of cost and benefit to these hitherto untraded, uncommercialized qualities or dimensions of life is itself morally despicable (Nelson 1994; Kimbrell 1993). In truth, it is difficult to pinpoint the moral force of these objections. Perhaps it would be more straightforward to characterize them simply as religious views in the sense discussed in Chap. 11. The view that modern society is becoming dangerously subject to legal and customary norms of commercial exchange, individual satisfaction and rigidly structured rules and codes does not appear to require an explicitly religious foundation, however. Surely many people are tempted by this sort of thinking on occasion, and surely some who are strongly committed to it base would describe their views in the language of community and virtue, rather than religion. Though all technology and modernization are part of this threat, biotechnology can be viewed as particularly significant in virtue of its capacity to bring an entirely new domain of objects into the realm of commodity exchange. The Loss of Practice. Albert Borgmann has argued that one of the great creeping threats of technology is that it turns practices that define and give meaning to human life into automated or rote routines. Cooking can be a practice in which a person strives for excellence, balancing nutrition with budget and aesthetics, or it can simply be a means to an end, something that should be accomplished as efficiently as possible. One irony of modern food technology is that it allows those who see cooking simply as a means for being fed to realize many of the nutritional, economic and aesthetic benefits that would, in earlier times, have been reserved to those who excelled in cooking as a practice. Borgmann clearly appreciates the tradeoffs that new technologies involve, and he would certainly not neglect nutritional, economic and aesthetic benefits associated with new food technologies. Nevertheless, he does believe that the cumulative affect of such technologies is that people

8.3 Theories of Justice

207

cease to occupy themselves with practices, at all. In doing so, they become shallow and base (Borgmann 1983). There can be little doubt that biotechnology can be employed in ways that erode practice, though it is also likely that biotechnology is itself a form of practice for the scientists who undertake the work. However, it may also be possible to deploy biotechnology in the service of practice, just as technology such as silicon rods has improved the practice of fly-fishing, rather than eroding it. This is an important ethical argument, but not one that cuts deeply against the development of food biotechnology. It is only in conjunction with either a commodification argument or an agrarian argument that loss of practice could be of more than cautionary significance. Agrarian Virtue. MacIntyre’s view is particularly important to questions in the food system, and it is significant that he uses farming to illustrate his notion of a practice. As was argued in Chap. 7, agrarian societies traditionally conceptualized their morality in terms of personal loyalties and virtues. The agrarian transition that has been brought about by technical change has created a world in which people interact with counterparts that are far more distant in space and time, reducing the importance of personal loyalties. Relationships are specified more by economic transactions or by claims of legal and political rights than by family or community roles. Indeed, any given individual in modern society may occupy many roles throughout their life, so much so that role morality and virtue can no longer support an adequate account of social justice. If one believes that technical change has led to the erosion of agrarian societies, and one believes that these societies were better suited to the production of virtuous citizens than are industrial societies, it is possible to generate a broad and sweeping argument against those changes in agricultural technology that militate against the continuation of family farming. This argument differs from utilitarian or rights arguments that evaluate the ethics of agrarian transition in terms of technology’s effects on individual farmers and their dependents. What matters morally is not that these individuals are harmed, nor that their rights are violated by excluding them from key decision making opportunities. The wrongness of this change consists in the fact that future generations will lack the virtues, indeed the very idea of virtue, that emerge naturally out of agrarian communities.

8.4 Structural Injustice and Structural Criticisms Rights theories and share a moral ontology with utilitarianism. As discussed in Chap. 7’s section on “Environmental Values in Consumption,” they presume that ethical theories speak to agents on the precipice of choosing how they will act. Although they differ in the advice that they give to agents, all of these approaches command actors to consider the options before them, and to choose which course to follow based on the highlighted moral considerations. A different philosophical tradition holds that the sources of injustice reside less in agency than in the structure

208

8 Social Impact and the Technology Treadmill

of rights, privileges, obligations and incentives in which agents operate. This need not mitigate agents’ responsibility for injustice entirely, but it does suggest that we should look beyond the standpoint of some person or group when offering an analysis of injustice. The tension between structure and agency animates interaction (or lack of it) between social scientist who study food systems and small cadre of philosophers who work in the area. By 1980, the critical social theory practiced in rural studies emphasized structural explanations of oppression, injustice and regressive trends in rural areas. However, a renewed interest in social movements and the efficacy of resistance to oppression sparked a new interest in agency, (Friedland 2008). In the meantime, philosophers were moving in the opposite direction. Philosophical work on structural injustice has experienced significant growth among writers commenting on the food system since the last revision to this book was completed. Food justice has emerged as a topic with a large and rapidly growing literature. Philosophical contributions to this literature stress the role of power in food systems, but following the influential work of Michel Foucault (1926–1984), and the feminist political theory of Iris Marion Young (1949–2006) they do not understand power as something under the control or direction of agents. These theoretical trends meet in the middle when they identify the source of injustice in food systems as the lack of power. People marginalized in virtue of race, gender, poverty or historic patterns of colonization are unable to control the production and distribution of food in their respective communities. These people are often farmers themselves; their political and economic subjugation to larger forces penetrates deeply into their lived experience and their community ties. Thus food justice extends far beyond food security or hunger, pointing toward the way in which crucial aspects of a person or community’s existence is dominated by forces they have no meaningful ability to affect, (Patel 2007; Gottleib and Joshi 2010; Alkon and Agyeman 2011). Owing especially to the philosophical writings of Iris Marion Young, many contemporary analysts of food justice describe it as a form of structural injustice. Teea Kortemäki identifies four characteristic features of structural injustice as it occurs in food systems: 1. 2. 3. 4.

Not traceable to individual harms or practices by individual actors; Results from complex interactions, making it hard to say where injustice begins; Accrues over the long term, occurring as process rather than as an outcome; Results from the accumulated outcome of actions that are in themselves normal and not blameworthy, (Kortemäki 2019).

Young’s work has become especially influential among theorists who adopt a feminist perspective. As noted in earlier editions, feminist philosophy shares principles and approach with philosophical studies that emphasize the perspective or experience of racial or ethnic minorities and colonized peoples. There is thus a conceptual link between feminism narrowly construed as philosophy that arose in response to social movements dedicated to empowerment of women, and a broader interpretation of feminism that sees it as encompassing some philosophical components in post-colonialism, gay and lesbian studies as well as black studies, African

8.4 Structural Injustice and Structural Criticisms

209

or Middle Eastern studies, Hispanic studies, or Asian studies. To some degree, all these approaches and intellectual movements have challenged epistemologies of the modern period (1550–1900). In particular, they have noted how European science tended to emphasize sharp conceptual boundaries, dichotomous logic and programs of reductionism in the sciences, (Harding 2015). These intellectual practices have contributed to social values that see women as radically different from (and generally inferior to) men, as well as to scientific values intolerant of ambiguity in data or systems of classification, approaches that may have neglected elements of ecological, historical or social context in their approach to various phenomena. Feminism has emphasized gender differences within broader contexts of continuity, and has tended to valorize, rather than denigrate, difference. Biologists such as Evelyn Fox Keller or Donna Haraway have suggested that women scientists may come to their subject matter differently from men may be more tolerant of apparent contradictions and more ready to accept the possibility that phenomena in nature are themselves ambiguous or continuous, (Keller 1983; Haraway 1991). Young argues that this requires every citizen to adopt a perspective of responsibility for justice. While the three theories of justice discussed above suggest that individuals commit acts of injustice for which they should be held responsible, they have little to say about the impersonal mechanisms that reproduce structural injustice. Young invites her readers to adopt a forward-looking stance in which all citizens accept responsibility for the social reforms needed to address structural injustice (Young 2010). A number of authors have drawn on feminism in studies of biotechnology. Judy Wajcman’s widely cited book Feminism Confronts Technology established a precedent for thinking that the feminist approach would be particularly fruitful in examining a domain such as technology, where men are primary decision makers. Her review of reproductive technologies stresses biomedical applications of biotechnology, but is conceptually broad enough to encompass agrifood biotechnology, as well (Wajcman 1991). Vandana Shiva has often claimed that the interests of traditionally marginalized groups including women, peasants and people of color are at risk in the commercialization of transgenic crops. Two of her early efforts made explicit links to feminist political thought (Mies and Shiva 1993; Shiva and Moser 1995). However, it may be more typical for those who take a feminist approach to follow the route taken by Haraway (1997) and by Bowring (2003) where agrifood and biomedical applications are not really distinguished in making broadly positive (in the case of Haraway) or broadly negative (in the case of Bowring) judgments about this new domain of technology. Other authors make few explicit references to feminism per se, yet apply styles of analysis that seem to draw heavily on feminist traditions. Annette Burfoot and Jennifer Pudrier describe efforts to collect and preserve plant, animal and human germplasm as expressions of European colonization and a male fascination with control, but make no explicit appeal to a feminist approach, (Burfoot and Pudrier 2005) The primary relevance of feminism in the present context is that feminist approaches have broadened the philosophical basis for challenging the assumption that utility and rights frame the philosophical terms of debate for a theory of social justice. Feminists are providing a new vocabulary in which to articulate moral claims

210

8 Social Impact and the Technology Treadmill

about agrifood biotechnology’s moral significance. They are by doctrine and personal inclination less likely to specify clear principles that could be applied to issues in agrifood biotechnology in algorithmic fashion. As noted at the beginning of the chapter, making reference to Marx can prompt some scientists to dismiss arguments out of hand. Nonetheless, many of the structural critiques leveled by feminists have precedents in Marxist sociology. Robert Tucker (1972) argues that ethical commitments are implicit in Marx’s early philosophical writings, and that his later political writings simply work out these principles of self-realization for the emerging Europe that he saw. Marx could be read as an egalitarian arguing for the positive right to dignity, though there are elements of his thought that seem utilitarian, such as his emphasis on the economic consequences of technical change. What would separate him from the agricultural economists who see technological change as progressive is first a different view of the harm done to those who lose their homes in agrarian transitions, and second a long-term view which sees these transitions eventually coming to ruin in a general collapse. In addition, Marxist thought anticipates feminist thought in problematizing the relationship between material conditions of life and the form of consciousness or subjectivity In feminist treatments, false consciousness has been supplanted by an analysis of how deeply seated ignorance of the food system has become a structural feature of the commodity orientation of contemporary food systems, (Rawlinson 2019). Jack Kloppenburg, Jr.’s penetrating sociological study of food biotechnology makes a convincing case for the claim that early twentieth century changes in seed technology (especially the development of hybrid varieties of maize) came about so that seed companies could appropriate a larger share of the value added in crop production than they would have had with open pollinated varieties. Farmers may save seed from open pollinated varieties to replant in succeeding years, but must purchase new seed corn for hybrid varieties every year. Kloppenburg argues that this pattern of technical change is an instance of the Marxist pattern: the capitalist uses technology to gain a larger share of the value, and gains this share at the expense of labor, (Kloppenburg 2004). In the standard treatment, a new technology lowers costs and eventually dominates the industry (e.g. the treadmill). Those who work in the industry (including small entrepreneurs who cannot continue to produce for lower returns) are forced to accept wages offered by the owners of technology (e.g. capitalists). Consistent with Kortemäki’s characterization of structural injustice, the social impacts studied by Kloppenburg are not easily analyzed in terms of utilitarian, rights based or virtue based theories of justice. One key problem is that they seem to be a causal consequence of social infrastructure, rather than being the direct result of any action by an individual or organization. As such, it is difficult to identify the agent that could have been said to have acted unjustly in bringing these consequences about. What is more, there is a pattern to consequences resulting from the organization of industrial societies. Many of worst consequences fall disproportionately on women and members of racial and ethnic minorities. Although Kloppenburg does not stress this point, it is this patterned distribution of harmful outcomes that provides an important ethical rationale for seeing them as instances of injustice. Again, Young’s

8.4 Structural Injustice and Structural Criticisms

211

conception of responsibility for justice is formulated to address this problem by calling on citizens to redress injustices they had no role in creating through political action.

8.5 Social Consequences for Small and Family Farms Willard Cochrane’s technology treadmill updates Marx’s analysis of technical change and applies it agriculture. Furthermore, the treadmill discussion grounds concerns associated with structural injustice in an economic model with explanatory power. The introduction of a continuing stream of productivity enhancing technology has a general tendency to shift the structure of industrialized agriculture toward fewer and larger farms, to reorient returns on food production toward capital from land and labor, and to limit the scope and flexibility of decision making for primary producers. Although more detailed empirical specification of these general trends might be controversial in its own right, the point of this section is to examine the ethical significance of these general impacts. Thus the question considered from a number of philosophical perspectives: Is there an ethical justification for resisting the transition from smaller (and family oriented) to larger (and industrially managed) farms? Alternatively, is there an ethical justification for promoting it?

8.5.1 Family Farms: Utilitarian Arguments Economists have struggled mightily with this problem for decades. A strict application of utilitarian welfare economics implies that the ethical significance of impact on family farms must be measured in terms of stress (both financial and emotional) placed on farm families, and on their long-term income capacity (Hussen 1979). The United States Department of Agriculture was applying such criteria to the evaluation of technical change in U.S. agriculture as early as 1940 (USDA 1940). It is also conceptually possible to include other, more esoteric forms of welfare value in the calculation. In explicitly applying a utilitarian framework, Luther Tweeten argues that welfare costs to family farms are outweighed by the benefits of production enhancing technology to consumers, but he thinks that policies aiming to protect family farms from such forces are nevertheless valid in virtue of family farms historical value, (Tweeten 1983). Although Tweeten does not say how he measured it, apparently historical value was able to offset the value of lowering the cost of food for consumers, tipping the balance in favor of family farms. One might also note the aesthetic or symbolic value of family farms, but comparative ranking of these values will be speculative, at best. Gary Comstock observes that family farms have emotional value for many people. “Since family farms are “ours,” since they are objects of love, and since they are now sources of considerable anguish, we ought to rescue them,” (Comstock 1987, 402– 3). Although he does not present emotional value within the context of making a utilitarian argument, the anguish of which he speaks is a good candidate for standard

212

8 Social Impact and the Technology Treadmill

utilitarian analysis. Just as resource economists have produced ways of evaluating the recreational and existence value of wild nature, why not use similar techniques to assess the value of family farms? It would be possible to generate a discussion of the utilitarian approach to social consequences that rivals that of Chap. 6’s analysis of environmental impact, describing the moral significance of each category of value, and discussing how it might compare other forms of cost and benefit. As Jeffrey Burkhardt concluded in one of the first published discussions on the ethical implications of social change from biotechnology, it is exceedingly unlikely that this approach will produce a convincing argument against any product of biotechnology, absent serious health or environmental risk. The more potent ethical criticism derives not from the claim that the social costs of small farm stress outweigh the benefits of biotechnology, but that the way this change is coming about is not fair, (Burkhardt 1991, 320–324).

8.5.2 Family Farms: Rights and Fairness The fairness theme is capable of generating two related arguments against technical changes that militate against small or family farms. One is that the process of technical change is unfair because small farmers’ (or others’) interests are not adequately represented. The second is that a social structure composed largely of small family farms is inherently fairer than one of fewer and larger farms. The first argument is philosophically straightforward, though highly controversial. Clearly technical changes have the capacity to substantially alter the nature of people’s opportunities, the value of their property and their prospects for prosperity, but what rights are being violated? There are two strategies for answering this question. One claims that the rights of individual farmers are violated, the other claims that the family-fun farms play a key role in legitimating the entire system of rights within industrial democracies. Innovations may encroach upon individual rights in virtue of the coercive way that they constrain decision-making. Kloppenburg notes that biotechnology such as herbicide tolerant seed limits farmer choice. If you use one company’s seed, you must also use their herbicide. Kloppenburg predicts that companies will use genetic engineering to integrate the entire farm production process, linking seed to an entire package of chemical inputs and processing technologies. This would, he argues compromise farmer decision making and choice (Kloppenburg 1984). However, it is hard to frame an argument that convincingly shows how farmers are deprived of any rights here. The old technologies are still available; farmers still have a right to use them. What they do not have is a right to both the old technologies and to the economic returns promised with the new ones. This, however, is an unexceptional situation, and not one that promises to suggest important moral objections to biotechnology. Perhaps consumer rights are denied or disrespected. As Comstock notes, people love family farms, and may wish to preserve them by favoring family farms in their market behavior (see also Hunter 1992; von Duijn 1995). Citizens have few direct

8.5 Social Consequences for Small and Family Farms

213

measures to affect technical changes through economic markets, and those that are available (such as boycotts) tend to be highly ineffective, (see Smith 1990). Nevertheless, consumers would not be in a position of being deprived of their rights unless biotechnology companies systematically attempted to prevent them from finding out about the source and origins of their food. Some evidence suggests that this is indeed part of the political and economic agenda of the food biotechnology sector, and such behavior is not only morally indefensible, it promises to erode public trust as well. This is thus an important but fairly narrow basis on which to formulate a rights-based objection to agrifood biotechnology. Other arguments charge that technology makes sweeping challenges to democratic rights. Langdon Winner argues that technical changes have social effects that are quite like changes in the legal or constitutional structure of society. Citizens of a democracy would not tolerate such sweeping changes coming about through governmental action without due process, but scientists and business leaders seem to able to bring about wrenching social change through a process that is totally isolated from public influence and participation. Such actions amount to an almost total usurpation of the most fundamental democratic rights, (Winner 1983). Winner’s general argument surfaced in biotechnology debates over the “4th criterion,” a proposal to regulate technology based on social impact (Lacy and Busch 1991). This is an argument that deserves to be taken seriously, but it is also an argument with such far ranging political consequences that it deserves to be at the heart of political debate on humanity’s technological future, not consigned merely to the debate over agrifood biotechnology. Whatever philosophical merits the argument has, it proved singularly ineffective in the rBST debate in the United States, at least. The Executive Branch concluded a review of literature on the social consequences of rBST with a telling sentence: “At no time in the past has the U.S. Federal Government prevented a technology from being adopted on the basis of socio-economic consequences.” (U.S. Executive Branch 1994, 35–6). Using the 4th criterion to regulate biotechnology would almost certainly have broader unintended consequences than biotechnology itself. It is thus not surprising that this approach has met with skepticism. The alternative strategy argues that family farming occupies a pivotal role in the system of rights. The view that a society of family farms represents an almost ideal instantiation of fundamental democratic rights has a long history, though not as old as some would claim. The link between small farms and democracy is often attributed to Thomas Jefferson (1743–1826), but historian A. Whitney Griswold (1906–1963) largely invented this alleged feature of Jefferson’s thought in his 1948 book Farming and Democracy (see Wunderlich 1984). Whatever its historical pedigree, the argument that family farms are crucial for democracy has figured in populist politics for a century. It became a staple of U.S. farm policy analysis around 1950 (see Brewster and Wunderlich 1961, pp. 200–3). Harold Briemyer (1914–2001) may have offered the most persuasive version of this argument in his book Individual Freedom and the Economic Organization of Agriculture (Briemyer 1965), while Jim Hightower was its most prolific spokesperson in the years preceding the introduction of agrifood biotechnology, (see Hightower 1976).

214

8 Social Impact and the Technology Treadmill

The general idea is that the transition described by Marx does indeed pose a moral problem for capitalism. Neither Breimyer nor Hightower would be so impolitic as to attribute the argument to Marx, but that is where it belongs philosophically. In brief: If capitalism systematically consigns labor to a situation of wage servitude, it cannot be considered morally legitimate. However, both Breimyer and Hightower think that wage labor jobs are perfectly acceptable so long as workers have an option. Farming, small-scale entry level farming that is, was to be that option. The argument here is that an economic structure including both wage labor jobs and also the opportunity to enter or leave family farming at will is ipso facto an ethically just structure of economic opportunity rights. Take away the opportunity to be one’s own boss on a farm and capitalism becomes coercive. If workers have no choice other than to accept going wage rates, capitalism is unjust. This version of agrarian populism would never succeed except in places where land is relatively available, but as the century turned, the capital and knowledge requirements for operating a farm alone cast doubt on farming’s capacity to stand as redoubt against the vicissitudes of the wage labor trap. If capitalism has this moral failing, agrifood biotechnology is not its singular undoing.

8.5.3 Family Farms and Moral Virtue An initial case for linking family farms to claims about virtue and character was sketched in Chap. 7. A further construction of the ethical virtues of farming can be drawn from the writings of American essayist Wendell Berry. Berry’s novels, poems and essays celebrate traditional farm life, and describe the virtues and character traits that are necessary for successful farming. Berry places the virtue of stewardship within a mutually reinforcing ecology of virtues that also include citizenship, industriousness, community, and family. Like Griswold, Berry bases his discussion of citizenship upon a questionable interpretation of Jefferson’s praise of farmers. Berry claims that Jefferson observed the effect of factory life on the character of the working class and concluded that wage laborers would be less reliable citizens than farmers. The specialization required by factory work made both workers and owners oblivious to the broader consequences of their actions. The ecological knowledge implied by a farmer’s stewardship practices, by contrast, prepares farmers to be more mindful of the unanticipated consequences of their actions. For this reason, according to Berry, farmers are more valuable as citizens. Berry also argues that industrialization undermines the moral meaning of work. Properly, work is both the formation and expression of personal identity. The hard work that is necessary for the traditional farm life has the effect of giving the farmer a well-developed sense of self, an identity that attaches naturally and harmoniously to a set of interests that arise from work. The factory pattern of life, by contrast, encourages people to identify with leisure activities, and to acquire interests that are not related to their identity or self-expression in any essential way. Berry’s understanding of work is ecological, a point that becomes clear when it is interpreted in

8.5 Social Consequences for Small and Family Farms

215

light of his vision of community. Farmers depend not only upon each other, but upon the tradesmen and merchants of the rural town. These are particular, non-universal dependencies that establish strong moral bonds to specific individuals. A farmer is in community with people whose lives are linked by the work activities that form their personalities and identities. In such places, Berry argues, community becomes meaningful as an ethical concept (Berry 1977). What is true for the community also holds for the family in Wendell Berry’s ecology of the virtues. Traditional farm life assigns tasks to each member of the family, so that husbands do the plowing and planting, wives tend to butter making and baking, children tend chickens and elders make quilts, jams, tools and tend to other farm needs. Each member of the family can see the importance of their work life to the overall survival and prosperity of the family. The family, in turn, is the source of production that sustains each member. Children learn that actions have consequences. Self-interest is again turned toward the virtue of family loyalty. In industrialized economies, by contrast, the relationship between work and prosperity is mediated by money. Family life requires cash that must be earned outside the home. Jobs are held to support the family, but the family itself no longer exists to perform work. Those who don’t work—children and retired elders—do not form part of the integrated, self-sustaining production that defines family identity on the farm. As a result, the family becomes defined as a consumption unit, and family members’ appreciation of virtue in productive work begins to fade (Berry 1977). The commitment to gender-determined roles in my summary of Berry’s view will prove offensive to feminists, but to some extent, the virtue argument for thinking that family farms are significant has been taken up and reinforced by feminists. For example, Deane Curtin’s book Chinnagounder’s Challenge draws on agrarian philosophy to argue for a new conception of ecological citizenship. Curtin sees the decline of virtues that encouraged stewardship of natural resources and community solidarity resulting from an imposition of utilitarian philosophy in rural communities. He interprets this quite literally, stressing policies that John Stuart Mill, author of Utilitarianism (1861), implemented in his role as an executive of the British East India Company. However, Curtin’s general philosophical framework is explicitly feminist and post-colonial. He situates the entire argument not in the virtue theory of Alisdair MacIntyre or the agrarianism of Wendell Berry, but in the importance of perspectives and voices that were silenced by doctrinaire applications of neo-liberal political thought (Curtin 1999). Neither Curtin nor Berry has much to say about agrifood biotechnology, and biotechnology could be at most one of many technological forces undoing the ecology of virtue in industrialized families. Furthermore, the entire argument on which the link to virtue is premised has implications that are troubling. Similar arguments would be made to oppose the rights of women, or to assert the rights of traditional families over those of single parent households, not to mention homosexual relationships (see Thompson 2000). These comments are not to dismiss the important themes that Berry introduces, but it is clear that much more work is needed to work out the implications of virtue ethics for food biotechnology.

216

8 Social Impact and the Technology Treadmill

8.6 Conclusion In one respect, all philosophical theories of social justice share a similar problem. Each offers a plausible diagnosis of a very large and far-reaching set of social problems. Yet precisely because these problems are so far-reaching, the claim that innovations in agrifood biotechnology are responsible for injustice becomes implausible. Collectively these distinct philosophical approaches to social impacts on small farmers describe some of the most serious ethical challenges to any agricultural technology. This, of course, includes agrifood biotechnology, but it is less clear that gene technologies have unique features that justify singling them out as perpetrators of social injustice. With the dawning of the twenty-first century it has become apparent that scientific and technological developments have the capacity to reshape society in sweeping and unexpected ways. Langdon Winner is only one recent political theorist who has argued for public action to wrest some measure of control over technical change from the market-based forces described first by Marx and then by Cochrane. René von Schomberg’s insightful papers (1993,1995) on science and policy applied the then-recent work of Beck (1992), to an analysis of the problem. Arie Rip has become associated with a broad approach he calls “constructive technology assessment,” in which scientists a members of the public interact extensively to plan and mediate conflicts, (Rip et al. 1995). Andrew Feenberg’s writings can also be mentioned in this connection, as can, of course, Hans Jonas himself. Indeed, Europe has supported large research initiatives in responsible research and innovation (RRI) since the second addition of this book in an attempt to get some sort of handle on these problems, (Hartley and coauthors 2016; Limson 2018). A detailed discussion of RRI lies beyond the scope of this book. Two hypotheses on the relationship between the risk-based approach to ethics outline in all three editions of this book and RRI conclude this chapter. First, my assumption in writing the first edition was that biologically trained scientists engaged in an RRI-like anticipatory study of their own research would benefit from the organizational scheme of the risk-based approach, as well as some minimal familiarization with standard ethical concepts. They should, for example, be able to understand the way that a traditional welfare-focused interpretation of risk as expected-value establishes very different burdens of proof than the informed consent criteria they were being required apply in human subjects protocols. It is doubtful that most RRI projects have provided scientifically-trained personnel with a capacity to perform reflective evaluations with even this minimal level of ethical expertise, (see Ribeiro et al. 2017). But perhaps I was wrong to make this assumption. As noted in the Introduction to this book, I have come to understand that very few practicing scientists will assume the responsibility to read a book-length treatment. What does that say for their willingness to undertake effective RRI? This points toward the second set of unanswered questions. Scientists and biotechnology companies must be justifiably frustrated by activists attempt to lay one of the most fundamental moral and social problems of the late twentieth century at their door. Scientists must indeed participate more actively in the debate over technology

8.6 Conclusion

217

and our future, but is it fair to hold biotechnology hostage to that debate? Furthermore, this is a question that impinges no less on developing countries and the structure of science (discussed just below), as it does on the small farm debate and the 1980s farm crisis in the United States. It will be revisited yet again in Chap. 12. In a narrower sense, the lesson to learn is that agrifood biotechnologies have been excused from the need to consider the cultural issue of family farms, the history of agrarian change, and the arguments from virtue. Biotechnology is not uniquely threatening to family farms and agrarian issues, though among the cluster of technologies that have brought on decades of change in these social forms, biotechnology is a uniquely attractive target of criticism. These deep and important moral issues continue to be ignored by many of the philosophers who have moved into food ethics over the last decade. Losing this opportunity to contemplate the moral significance of farming is unwise, but the shift to food justice has weakened the sense agrifood biotechnology had anything to do with structural injustice, in the first place. Philosophers who follow Iris Young are much more likely to view a racially marginalized working mother living in public housing as a victim of structural injustice than a farmer bankrupted by the technology treadmill. At the same time, it is equally foolish to conclude that simple resistance to technology or to the multi-national corporations that promote justifies roadblocks to the benefits that food biotechnology can bring.

References Alkon, A.H., and J.A. Agyeman (eds.). 2011. Cultivating food justice: Race, class and sustainability. Cambridge, MA: The MIT Press. Beck, U. 1992) [1986]. Risk Society: Towards a New Modernity. M. Ritter (trns) Sage Publications, London. Bentham, J. 1789) [republished 1948] An Introduction to the Principles of Morals and Legislation. Hafner Press, New York. Berardi, G.M., and C.C. Geisler (eds.). 1984. The social consequences and challenges of new agricultural technologies. Boulder CO: Westview Press. Berry, W. 1977. The unsettling of America: Culture and agriculture. San Francisco: Sierra Club Books. Borgmann, A. 1983. Technology and the character of contemporary life. Chicago: University of Chicago Press. Bowring, F. 2003. Science, seeds and cyborgs: Biotechnology and the appropriation of life. London: Verso Press. Breimyer, H. 1965. Individual Freedom And The Economic Organization Of Agriculture. Urbana: University of Illinois Press. Breimyer, H. 1983. Food for people and profit: An alternative interpretation. The farm and food system in transition: Emerging policy issues FS 6, Cooperative Extension Service, Michigan State University, East Lansing, MI. Brewster, J., and G. Wunderlich. 1961. Farm size, capital and tenure requirements. In Adjustments in agriculture—A national basebook, ed. C.F. Christian, 196–228. Ames: Iowa State University Press. Browne, W.P. 1987. Bovine growth hormone and the politics of uncertainty: Fear and loathing in a transitional agriculture. Agriculture and Human Values 4 (1): 75–80.

218

8 Social Impact and the Technology Treadmill

Browne, W.P., J.R. Skees, L.E. Swanson, P.B. Thompson, and L.J. Unnevehr. 1992. Sacred cows and hot potatoes: Agrarian myths in agricultural policy. Boulder CO: Westview Press. Burfoot, A., and J. Poudrier. 2005. Biotechnology as modern museums of civilization. In Biotechnology unglued: Science, society and social cohesion, ed. M.D. Mehta, 133–160. Vancouver: UBC Press. Burkhardt, J. 1991. Biotechnology, ethics and the structure of agriculture. In Ethics and agriculture: An anthology on current issues in world context, ed. C. Blatz, 317–330. Moscow: University of Idaho Press. Burkhardt, J. 1992. Ethics and technical change: The case of BST. Technology in Society 14: 221–243. Busch L., W.B. Lacy, J. Burkhardt, and L.R. Lacy. 1991. Plants, power and profit: Social, economic, and ethical consequences of the new biotechnologies. Cambridge, MA and Oxford, UK: Basil Blackwell. Buttel, F.H. 1986. Agricultural research and farm structural change: Bovine growth hormone and beyond. Agriculture and Human Values 3: 88–98. Buttel, F.H. 2000. The recombinant BGH controversy in the United States: Toward a new consumption politics of food? Agriculture and Human Values 17: 5–20. Buttel, F.H., and R. Barker. 1985. Emerging agricultural technologies, public policy and implications for Third World agriculture. American Journal of Agricultural Economics 67: 1170–1175. Buttel, F.H., M. Kenney, and J. Kloppenburg Jr. 1985. From green revolution to biorevolution: Some observations on the changing technological bases of economic transformation in the Third World. Economic Development and Cultural Change 34: 31–55. Cochrane, W. 1979. The development of American agriculture: A historical analysis. Minneapolis: University of Minnesota Press. Comstock, G. 1987. Conclusion: Moral arguments for the family farm. In Is there a moral obligation to save the family farm?, ed. G. Comstock, 399–418. Ames: Iowa State University Press. Crosby, A.W. 1986. Ecological imperialism: The biological expansion of Europe, 900–1900. Cambridge, UK: Cambridge University Press. Curtin, D.W. 1999. Chinnagounder’s challenge: The question of ecological citizenship. Bloomington, IN: Indiana University Press. Donnelly, J. 1989. Universal human rights in theory and practice. Ithaca, NY: Cornell University Press. Edge, D.O. 1973. Technological metaphor. In Meaning and control, ed. D.O. Edge and J.N. Wolfe, 31–59. London: Tavistock Publications. Epstein, R.A. 1996. Norms: Social and legal. The Good Society 61 (1): 3–7. Fineberg, J. 1980. Rights, justice and the bounds of liberty: Essays in social justice. Princeton, NJ: Princeton University Press. Friedland, W.H. 2008. Agency in the agrifood system. In The fight over food: Producers, consumers and activists challenge the global food system, ed. W. Wright and G. Middendorf, 45–68. University Park PA: Penn State University Press. Gottleib, R., and A. Joshi. 2010. Food justice. Cambridge, MA: The MIT Press. Griswold, A.W. 1948. Farming and democracy. New Haven, CT: Yale University Press. Haraway, D.J. 1991. Simians, cyborgs and women: The reinvention of nature. London: Free Association. Haraway, D.J. 1997) Modest_Witness@Second_Millennium. FemaleMan_Meets_OncoMouse™: Feminism and Technoscience. New York: Routledge. Harding, S.G. 2015. Objectivity and diversity: Another logic of scientific research. Chicago: University of Chicago Press. Hartley, S., F. Gillund, L. van Hove, and F. Wickson. 2016. Essential features of responsible governance of agricultural biotechnology. PLoS Biology 14 (5): e1002453. Hightower, J. 1976. Eat your heart out. New York: Crown Books. Hugill, P.J. 1993. World trade since 1431: Geography, technology, and capitalism. Baltimore: Johns Hopkins University Press.

References

219

Hunter, D. 1992. To live free as natives, free of fear: what citizens should require from animal biotechnology. In: Animal biotechnology: opportunities and challenges, ed. J.F. MacDonald. National Agricultural Biotechnology Council, Ithaca, NY, 33–140. https://ecommons.cornell. edu/bitstream/handle/1813/49749/nabc4_17_Hunter.pdf?sequence=1&isAllowed=y. Accessed 26 Feb 2020 Hussen, A.M. 1979. Assessment of the economic and social impacts of agricultural technology: A case study. Western Journal of Agricultural Economics 4 (2): 17–31. Huttner, S.L., H.I. Miller, and P.G. Lemaux. 1995. US agricultural biotechnology: Status and prospects. Technological Forecasting and Social Change 50: 24–39. Iowa Public Broadcasting. 2013. 1980s farm crisis, market to market: films for the classroom. https://www.iptv.org/mtom/classroom/module/13999/farm-crisis. Accessed 25 July 2019 Jeansonne, G. 1974. The automobile and American morality. Journal of Popular Culture 8: 125–131. Jonas, H. 1984. The imperative of responsibility: The search for ethics in a technological age. Chicago: U Chicago Press. Kalter, R.J. 1984. Production cost: Commercial potential and the economic implications of administering bovine growth hormone. In Proceedings of the Cornell nutrition conference for feed manufacturers. Cornell University, Ithaca, NY. Kalter, R.J. 1985. The new biotech agriculture: Unforeseen economic consequences. Issues in Science and Technology 13: 125–133. Kalter, R.J., R. Milligan, W. Lesser, W. Magrath, L. Tauer, and D. Bauman. 1985. Biotechnology and the dairy industry: Production costs, commercial potential and the economic impact of the bovine growth hormone. Agricultural Economics Research Bulletin 85–20. Department of Agricultural Economics, Cornell University, Ithaca, NY. Keller, E.F. 1983. A feeling for the organism: The life and work of Barbara McClintock. San Francisco: WH Freeman. Kenney, M. 1986. Biotechnology: The university-industrial complex. New Haven, CT: Yale University Press. Kenney, M., and F. Buttel. 1985. Biotechnology: Prospects and dilemmas for Third World development. Development and Change 16: 61–91. Kimbrell, A. 1993. The human body shop. New York: HarperCollins. Kloppenburg, J.R., Jr. 1984. The social impacts of biogenetic technology in agriculture: Past and future. In The social consequences and challenges of new agricultural technologies, ed. G.M. Berardi and C.C. Geisler, 291–323. Boulder, CO: Westview Press. Kloppenburg, J.R., Jr. 2004. First the seed: The political economy of plant biotechnology. Madison: University of Wisconsin Press. Kortemäki, T. 2019. Nobody’s fault? Structural injustice, food and climate change. In Food, environment and climate change: Justice at the intersections, ed. E. Gilson and S. Kenehan, 47–62. London: Rowman and Littlefield. Lacy, W., and L. Busch. 1991. The fourth criterion: Social and economic impacts of agricultural biotechnology. In Agricultural biotechnology at the crossroads: Biological, social and institutional concerns, ed. J.F. MacDonald, 153–168. Ithaca, NY: National Agricultural Biotechnology Council. Landes, D.S. 1983. Revolution in time: Clocks and the making of the modern world. Cambridge, MA: Harvard University Press. Limson, J. 2018. Putting responsible research and innovation into practice: A case study for biotechnology research, exploring impacts and RRI learning outcomes of public engagement for science students. Synthese 198: 1–26. https://doi.org/10.1007/s11229-018-02063-y. Machan, T. 1984. Pollution and political theory. In Earthbound: New introductory essays in environmental ethics, ed. T. Regan, 74–106. New York: Random House. MacIntyre, A. 1984. After Virtue. Notre Dame, IN: In University of Notre Dame Press. Marx, K. 1867 [1978]. Capital. In The Marx-Engels reader, 2nd ed., ed. R. Taylor. New York: WW Norton. Mies, M., and V. Shiva (eds.). 1993. Ecofeminism. London: Zed Books.

220

8 Social Impact and the Technology Treadmill

Mill, J.S. 1861 [1979]) Utilitarianism. G. Sher (ed) Hackett Publishing Co, Indianapolis, IN. Mumford, L. 1934. Technics and civilization. New York: Harcourt Brace. Nelkin, D., and M.S. Lindee. 2004. The DNA Mystique: The Gene as a Cultural Icon, 2nd ed. Ann Arbor: University of Michigan Press. Nelson, J.R. 1994. On the new frontiers of genetics and religion. Grand Rapids, MI: William B Eerdmans Publishing. Patel, Raj. 2007. Stuffed and starved: The hidden battle for the world food system. Brooklyn, NY: Melville House. Rawlinson, M.C. 2019. The climate of food: Justice, truth, and structural change. In Food, environment and climate change: Justice at the intersections, ed. E. Gilson and S. Kenehan, 91–118. London: Rowman and Littlefield. Ribeiro, B.E., R.D. Smith, and K. Millar. 2017. A mobilising concept? Unpacking academic representations of responsible research and innovation. Science and Engineering Ethics 23: 81–103. Rip, A., T.J. Misa, and J. Schot (eds.). 1995. Managing technology in society: The approach of constructive technology assessment. New York: St Martin’s Press. Rosenberg, C.E. 1961. No other gods: On science and american social thought. Baltimore, MD: Johns Hopkins University Press. Rosenberg, N. 1992. Technology and the wealth of nations. Stanford, CA: Stanford University Press. Rosenberg, N., and L.E. Birdzell Jr. 1986. How the west grew rich: The economic transformation of the industrial world. New York: Basic Books. Ruttan, V. 1991. Moral responsibility in agricultural research. In Beyond the large farm: Ethics and research goals for agriculture, ed. P.B. Thompson and B.A. Stout, 107–124. Boulder, CO: Westview Press. Schurman, R., and W.A. Munro. 2010. Fighting for the future of food: Activists versus agribusiness in the struggle over biotechnology. Minneapolis: University of Minnesota Press. Schmitz, A., and D. Seckler. 1970. Mechanized agriculture and social welfare: The case of the tomato harvester. American Journal of Agricultural Economics 54: 569–577. Sen, A.K. 1987. On ethics and economics. London: Basil Blackwell. Shiva, V., and I. Moser (eds.). 1995. Biopolitics: A feminist reader on biotechnology. London: Zed Books. Shue, H. 1980. Basic rights. Princeton, NJ: Princeton University Press. Smith, N.C. 1990. Morality and the market: Consumer pressure for corporate accountability. London: Routledge Publishing Co. Tauer, L.W. 1992. Impact of BST on small versus large dairy farms. In Bovine somatotropin & emerging issues: An assessment, ed. M. Hallberg, 207–217. Boulder, CO: Westview Press. Tenner, E. 1996. Why things bite back: Technology and the revenge of unintended consequences. New York: Alfred A Knopf. Thompson, P.B. 2000. Discourse ethics for agricultural biotechnology: Its limits and its inevitability. Science and Engineering Ethics. 6: 275–278. Thompson, P.B., R.J. Matthews, and E. vanRavenswaay. 1994. Ethics. Macmillan, New York: Agriculture and public policy. Tucker, R. 1972. Philosophy and myth in Karl Marx. Cambridge, UK: Cambridge University Press. Tweeten, L. 1983. Food for people and profit: Ethics and capitalism, the farm and food system in transition: Emerging policy issues FS 5, Cooperative Extension Service. East Lansing, MI: Michigan State University. Tweeten, L. 1987. Has the family farm been treated unjustly. In Is there a moral obligation to save the family farm?, ed. G. Comstock, 212–232. Ames: Iowa State University Press. Tweeten, L. 1991. The costs and benefits of bGH will be distributed fairly. The Journal of Agricultural and Environmental Ethics 4: 108–120. USDA (United States Department of Agriculture). 1940. [1984] Technology on the farm. In The social consequences and challenges of new agricultural technologies, ed. G.M. Berardi and C.C. Geisler, 51–61. CO: Westview Press, Boulder.

References

221

U S Executive Branch. 1994. Use of bovine somatotropin BST in the United States: Its potential effects, A study conducted by the executive branch of the federal government. Washington, DC, January. https://www.biotech.wisc.edu/docs/default-source/outreach-documents/use-of-bov ine-somatropin.pdf?sfvrsn=999ec28_0. Accessed 26 Feb 2020 von Duijn, P. 1995. Biotechnology between science push and consumer pull: Who steers? In Contested technology: Ethics, risk and public debate, ed. R. von Schomberg, 155–175. Tilburg: International Centre for Human and Public Affairs. von Schomberg, R. 1993. Controversies and political decision making. In Science, politics and morality: Sceintific uncertainty and decision making, ed. R. von Schomberg, 7–26. Dordrecht: Kluwer Academic. von Schomberg, R. 1995. The erosion of our valuespheres: The ways in which society copes with scientific, moral and ethical uncertainty. In Contested technology: Ethics risk and public debate, ed. R. von Schomberg, 13–28. Tilburg: International Centre for Human and Public Affairs. Wajcman, J. 1991. Feminism confronts technology. Cambridge, MA: Polity Press. Winner, L. 1983. Techné and politeia: The technological constitution of a society. In Philosophy and technology, ed. P.T. Durbin and F. Rapp, 97–111. Dordrecht: D Reidel. Wittfogel, K. 1958. Oriental despotism, a comparative study of total power. New Haven, CT: Yale University Press. Wittfogel, K.A. 1962. The Marxist view of China (Part 1). The China Quarterly 11: 1–20. Wunderlich, G. 1984. Comment on Shrader-Frechette. Agriculture and Human Values 13: 29–30. Yonkers, R.D., J.W. Richardson, R.D. Knutson, and B.B. Buxton. 1986. Accomplishing adjustment in the dairy industry during technological change: The case of bovine growth hormone, Texas Agricultural Experiment Station bulletin TA-21797. College Station, TX: Texas A&M University. Young, I.M. 2010. Responsibility for Justice. New York: Oxford University Press.

Chapter 9

Can Agrifood Biotechnology Help the Poor?

Abstract This chapter completes the review of socioeconomic risks from food and agricultural gene technologies begun in Chap. 8. Here, the focus is on challenges to the claim that gene technologies make or will make substantial contributions to the welfare of poor and marginalized people, especially in the less industrialized regions of Africa, Asia and Latin America. The institutional organization of science is central to the debate. Public laboratories and experiment stations contributed important critics allege that past innovations but patents, changes in funding patterns and other features of gene technology limit the future prospects of non-for-profit innovations in the food system. As such, an ethical analysis of biotechnology’s ability to help the poor must engage issues in the organization and incentives driving the research. These include the capacity and willingness of commercial enterprises to serve needs of the poor. Keywords Food security · Food sovereignty · Agricultural development · Global food needs · Agricultural research · Green revolution Defenders of gene technologies in food and agriculture cite the twin grounds of meeting global food needs and helping the poor. Critics are skeptical. They may portray the argument as ruse for large companies to gain entry (and eventual control) into sectors of the world’s agriculture where they currently have limited markets for their products. They may view current forms of smallholder agriculture as superior to the industrialized model that dominates in the United States, Europe and Australia. Either rationale means that introducing any aspect of industrial production systems will be greeted cynically. For many reasons, the impact of biotechnology on less-industrialized agriculture is debated hotly, and appeals to ethical principles are frequent. Rachel Schurman and William Munro find that moral principles were especially important in this quarter of the GMO debate. Proponents of the technology felt a moral responsibility to increase the global capacity for food production, while opponents believed that biotechnology was part of a larger corporate strategy to take over the global food system, (Schurman and Munro 2010). No area other of debate

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_9

223

224

9 Can Agrifood Biotechnology Help the Poor?

over the consequences of rDNA techniques for modifying plant and animal genomes draws upon ethics so clearly. In this chapter, I will take some pains to document the anticipatory critique of gene technology that was generated within the agricultural establishment, that is, by scholars employed in agricultural research institutions conducting research that was supported by traditional funders. I hope to correct the impression that agricultural researchers were uniformly aligned in favor of rapid and largely ungoverned implementation of gene technologies. I see this as relevant to contemporary debates in two respects. First, as philosophers have taken up food ethics in the 21st century, they have overlooked a key issue that is crucial for answering this chapter’s title question, namely, that poor farmers are not simply a subclass of the urban poor. Early critiques were far more sensitive to the plight of smallholders in less developed countries, in part because the debates were conducted by people who had much more intimate knowledge about the practice of agriculture. Second (and less favorably to the agriculture establishment), these internal critiques were and continue to be drowned out by less well-informed voices. As such, agricultural scientists have gravitated to the bad arguments reviewed in Chap. 2, ignoring problems for which they should have assumed responsibility. Although the ethical appeals are explicit, the issues are actually intricate. Disambiguating industrial and less industrial of peasant-based production systems is the first level of complexity. In many parts of the world that get described as “developing”, highly mechanized and chemically intensive production systems occupy large acreages, yet exist side by side with smallholder farming with few purchased inputs such as fertilizers or motorized machinery. In this situation, things that help the economy are not necessarily the same things that would help the poor. There is also the question of who does the helping. Prior to the advent of rDNA techniques, a large complex of not-for-profit research institutes dedicated to agricultural and nutritional improvement had agency. A contrary model sees for-profit innovators as more likely to deliver benefits to the poor. Assuming that gene technology has applications that could be beneficial in less industrialized economies, it remains unclear who should give, and who should receive.

9.1 The Ethics of Agricultural Research Many readers will have some familiarity with the Green Revolution and its charismatic figurehead, Norman Borlaug. As has been widely reported, Borlaug (and others) developed dwarf varieties of wheat and rice that grew well in the tropical climates typical of less industrialized regions and that respond well to added fertilization. Borlaug was a plant breeder by trade, and the Green Revolution varieties of wheat and rice were achieved by modifying plant genomes through a tedious process of crossing and backcrossing until the breeder achieves a plant that both has the desired characteristics, and that “breeds true,” (e.g. reliably passes those characteristics on to its progeny). When coupled with fertilization, Green Revolution

9.1 The Ethics of Agricultural Research

225

varieties allowed farmers to achieve an impressive increase in the amount of grain they harvested. Borlaug received the Nobel Peace Prize for this accomplishment, and even critics admit that Green Revolution varieties saved hundreds of thousands of lives between 1970 and 2000, (see Ehrlich and Ehrlich 2009). Fewer know that critical voices within agricultural research organizations pointed out harmful impacts from Green Revolution efforts even as these varieties were first appearing in farmers’ fields, (Falcon 1970; Griffen 1974; Wade 1974). The philosophical critique that ties condemnation of the Green Revolution to the decades-later writings of Vandana Shiva misunderstands the institutional setting for Borlaug’s work and underestimates his contemporaries’ ability to reflect on their work in ethical terms. Borlaug worked under the auspices of two research organizations, the International Maize and Wheat Improvement Center (CIMMYT), based in Mexico, and the International Rice Research Institute (IRRI), headquartered in the Philippines. Both are members of a consortium known as the Consultative Group on International Agricultural Research (CGIAR). The non-profit research agencies in CGIAR derive funding from government agencies (such as the U.S. Agency for International Development [USAID]), from international organizations (such as the United Nations Food and Agricultural Organization (FAO) or the World Bank) and from philanthropic groups (such as the Rockefeller Foundation or the Gates Foundation). They operate on a model of agricultural research developed between the American Civil War and World War II by nationally organized and publicly funded organizations. The United States Agricultural Research Service (ARS) or the French Institut national de la recherche agronomique (INRA) exemplify this tradition as do laboratories operated by the British, the Dutch and the Germans that were intended to support farming domestically as well as in far flung colonial empires. These researchoriented organizations have always been connected to agricultural universities. In the United States, the Morrill Act of 1862, the Hatch Act of 1887 and the Smith-Lever Act of 1914 authorized public funding for research, teaching and delivery of results to working farmers in each state. The prestige of this network reached its peak in the 1960s and 1970s, (Busch and Lacy 1983). The scientists and administrators in it constituted the agricultural research establishment at that time. The research establishment mindset went something like this: Agricultural research is designed to benefit farmers. It is morally justified by two complementary arguments. First, farmers are morally worthy beneficiaries, and second because throughout history, applied agricultural science has driven an overall decrease in the price of food. Each of these tenets deserves more detailed analysis, but for convenience let us refer to their conjunction as the establishment mentality (EM). Notice that EM does not necessarily imply that agricultural research should be done in the public sector. For-profit research could also fulfill these two criteria, at least in theory. Though as will be discussed presently, public sector scientists had reason to presume that not-for-profit laboratories and research institutes would be the primary locus for agricultural research, this did not necessarily imply that profit-driven research was incapable of achieving the twin goals of EM. In fact, it seems likely that most of them

226

9 Can Agrifood Biotechnology Help the Poor?

thought of for-profit research in seed or chemical companies as occurring in partnership with their own research in government institutes and agricultural universities, (Busch and Lacy 1983). The farmer-benefit side of EM owes a debt to Abraham Lincoln, who delivered a stump speech at an agricultural fair in Wisconsin while campaigning for the U.S. Presidency in 1859. Lincoln argued that government should pay for agricultural research because improvements that “bring the soil up to its full potential,” benefit farmers, and farmers were not only the most numerous group in the population, they were also among the most economically disadvantaged, (Lincoln 1859). Lincoln’s appeal to the number of people who benefit echoes utilitarianism, while his focus on poverty is egalitarian. Lincoln was, of course, a politician making promises to a specific constituency here, but in the context of mid-19th century America, his argument had real moral plausibility. Lincoln delivered on these promises, as well, by establishing the U.S. Department of Agriculture (USDA) and signing legislation that funded the nation’s system of agricultural universities, (Rasmussen, n.d.). Neither of Lincoln’s assumptions hold true for the United States or Europe today, where the farming population hovers around 1% of the total and most farmers are decidedly middle class, (Browne et al. 1992). Yet when one considers nations whose economic development remains low as measured by Gross Domestic Product (GDP), one also finds that between 40 and 80% of their population are farmers. Lincoln’s half of the EM argument still makes perfect sense in much of the 21st century world. Although the agricultural establishment was highly focused on helping farmers, the Establishment Mentality has a food consumer component as well. The view is that when the relative cost of food goes down, people can spend a portion of their income on other things. Since everyone eats, this benefits everyone. It would be difficult to imagine how one could do better with respect to the “greatest number” part of the utilitarian maxims injunction to whatever does the greatest good for the greatest number. At the same time, people with low incomes spend a greater proportion of their income on food than wealthy people. Among all the policy interventions one might make, support for agricultural research has great potential for meeting the test posed by John Rawls’ Difference Principle: do that which most benefits the worst off group. So like Lincoln’s farmer half of the EM, the consumer half is recommended by both utilitarian and egalitarian moral theory. The mentality of the agricultural establishment was imbued with moral confidence in the justifiability of their quest. The validity of the underlying perspective notwithstanding, this made them less critical of their mission and methods than they might have been, (Zimdahl 2006). As noted already, the EM applied equally to innovations that came out private industry and to the research findings that came from universities, foundation-funded international research institutes and other organizations supported by government funds or charitable donations. Indeed, the history of agricultural technology follows a series of entrepreneurs who built fortunes of their discoveries: Eli Whiney and the cotton gin, Cyrus McCormick and mechanical harvesters, Luther Burbank and modern seed varieties. The Establishment Mentality tended to underplay the significance of this history. The 1980s and 1990s were a time when the easy moral confidence of the agricultural research establishment was ripe for criticism. As discussed

9.1 The Ethics of Agricultural Research

227

later in the chapter, rural social scientists began to cite Willard Cochrane’s analysis of the treadmill phenomenon (see Chap. 8), which showed that farmers as a whole did not benefit from efficiency-improving innovation. Furthermore, as food purchases come to require a smaller share of the household budge, benefits to consumers in the form of lower food prices become less compelling.

9.2 Ethics and Agricultural Development Norman Borlaug’s hopes for biotechnology were framed against the background of the research establishment’s shared assumptions. Although I believe the argument he put forward in defense of genetic engineering was flawed (see Chap. 2), by the time of his advocacy for biotechnology he had achieved a lifetime of advocacy on behalf of farmers, and he had been heard at the highest levels of government. Borlaug was included in a group advising U.S. President Ronald Reagan on aid to Poland following the success of Solidarno´sc´ (the Polish worker’s union that ultimately ended the rule of Poland’s Communist Party). He is reputed to have said “What Poland needs is a new potato.”1 The story testifies to Borlaug’s intense loyalty toward farmers: he put their needs foremost at every opportunity. Borlaug was insisting that agricultural development assistance should be included in Poland’s aid package. At the same time, the mere suggestion that a new crop variety would be the answer for a society trying to recover its political and economic agency after years of Soviet domination also reveals an overweening confidence in the technological fix for social ills, (see Scott 2018 for a thorough discussion of the technological fix in the context of biotechnology). However, in addition to the twin tenets of Establishment Mentality, enthusiasm for agricultural research has additional roots in the context of international development. Succinctly stated, the use of government dollars to promote economic development beyond international borders has its roots in the United States Marshall Plan to speed economic recovery in Europe after the Second World War. The general idea is that strategic administration of financial and technical assistance could substantially accelerate in Europe, and if it works there, why not try in the nations that were emerging from the shadow of colonialism during the 1940s, 50s and 60s? The Marshall Plan itself is now acknowledged to have had both humanitarian and military-political objectives, (Kunz 1997), and the same is certainly true of the effort to extend similar policies to what was then referred to as “the Third World”. As I have summarized elsewhere, the expression “Green Revolution,” was coined in the context of opposing the red revolution that was putatively being fomented by the United States’ Cold War nemesis, the Soviet Union, (Thompson 2015b).

1 This story circulated at Texas A&M when Borlaug and I were both on the faculty there. I have not

found any confirmation of it, and it may be apocryphal. Even so, it is illustrative of Borlaug’s mind set in both its positive and negative aspects.

228

9 Can Agrifood Biotechnology Help the Poor?

International development (e.g. industrial, social and economic development in former colonies) has a complex theory and an equally complex history. For present purposes, it is enough to stress two themes. First, economic growth as measured by a country’s GDP came to be the prime indicator of development. Second, poverty alleviation and assistance to the needy has always been important. These goals sometimes, but not always, go hand in hand. Although offering lifesaving or life changing help to people in extreme poverty is morally compelling, it does not always translate into an increase in GDP. However, development theorists gradually came to embrace programs of technical assistance to agriculture because they were thought to be excellent candidates for resolving the tensions between these two goals. The moral argument was essentially the same as Lincoln’s. Influential advances in development economics throughout the 1960s provided a reason to think that it would actually work, (see Thompson 1992, pp. 84–86). In short, giving aid in the form of technical assistance to farmers would help them, and since they were both numerous and disproportionately poor, the humanitarian goals of assistance would be satisfied. At the same time, being able to sell agricultural commodities in global markets would add directly to the nation’s GDP, and would actually spark further growth in other economic sectors, (Johnston and Mellor 1961). The need for not-for-profit research organizations was not made explicit in these arguments, but it was the water that leaders of the agricultural research establishment were swimming in. Agricultural scientists began their love affair with biotechnology in the 1980s when the EM was at its zenith. The capacity for internal debate also expanded in the 1980s. Working with William Lacy, Lawrence Busch (1945–2020) established the Center for Agricultural Research Policy (CARP) at the University of Kentucky. Their book Science, Agriculture and the Politics of Research raised cognizance of the way that publicly funded agricultural research was becoming subservient to a for-profit innovation agenda, (Busch and Lacy 1983). Frederic Buttel (1948–2005) also drew upon science studies in a long string of papers expressing concern that privatization of agricultural research would prevent small farmers from enjoying the benefits of gene technology. Buttel and Busch argued that agricultural scientists’ self-confidence to the contrary, the agricultural establishment was at a crossroads, and would face an era of considerably reduced financial and political support in the future. They make frequent references to the expensive nature of gene technology in their 1988 article questioning whether the political willingness to fund this work at public research organizations could be sustained, (Buttel and Busch 1988). As the 21st century approaches its 3rd decade, it is clear that they were right. My book, The Spirit of the Soil: Agriculture and Environmental Ethics, built upon the critiques of the 1980s when it was originally published in the mid 1990s. I argued that mainstream agricultural researchers operated from a “productionist paradigm” that frustrated reflective evaluations of yield-enhancing research technology, and that they tended to regard environmental impacts as targets for the next technological fix, (Thompson 2017). With Busch, Buttel and Lacy, I was regarded as a bit of a nuisance by leaders in the agricultural research establishment, (see Moore 2001), yet we did not really question the normative commitments of EM as much as we regarded ourselves as helping the establishment do a better job of living up to the true meaning of its creed.

9.2 Ethics and Agricultural Development

229

Like many of the scientists and administrators in universities, government agencies and the CGIAR system, we presumed that the public, non-profit side of agricultural research had the lion’s share of the responsibility for developing a research agenda that did benefit farmers and the poor. In sum, much can be said in condemnation of the performance of the agricultural research establishment, especially during the decades after World War II. Harsh assessments of the Green Revolution stand at the forefront. Nevertheless, it is important to recognize that there was an ongoing debate within the establishment that issued from shared ethical premises: commitment to helping farmers, on the one hand, and to egalitarian poverty alleviation, on the other. The balance of this chapter will document ways in which agrifood biotechnology played a role in the undoing of EM, but first it is essential to take brief notice of two other factors that shaped the political and economic environment in which research intended to promote agricultural development would be undertaken. One is the rise of neoliberalism, while the other is a decline in the public’s understanding of agriculture. With neoliberalism, we broach yet another large topic with a massive and internally inconsistent literature. For simplicity, it is important to keep the focus narrowly on agricultural development. Neoliberal economic philosophy was an attack on the philosophical legitimacy of humanitarian goals in agricultural development policy. It stressed the role of agriculture in contributing to GDP and downplayed the importance of farming to meet subsustance needs. This attack was coupled to a theoretical critique of domestic policies that aid farmers through subsidy or price supports. What is more, the debt crisis of the 1970s and 1980s placed governments still emerging from the legacy of colonialism in a position of needing to generate hard currency through foreign exchange. In a nutshell, farmers were strongly encouraged to plant crops that could be sold in international markets, and discouraged from planting crops or raising livestock that would be consumed locally. Under this new regime, the EM’s tenet of helping farmers retained its pretense; after all, farmers were being assisted in making a shift to crops like coffee, cotton or cocoa, and they would benefit economically when these high value crops are sold in international markets. At the same time, these strategies put more and more people into dependency on the commodity agricultures of Europe and North America for the basic grains the people need to eat, (McMichael 2012). While this is an oversimplification, it must suffice in the present context. Buttel, Busch and Lacy were challenging the accuracy of the establishment mindset (EM) from a perspective that remained hopeful that science-led projects could actually benefit poor farmers in Africa, Asia and Latin America, but the policy shifts that were already under way were working against this hope. From a philosophical perspective, however, another kind of mindset was emerging among people who had little experience with or understanding of agriculture. This is a set of presumptions that associate hunger with the food banks and homeless shelters of the industrialized world’s urban centers. Aiding the hungry is morally compelling, on this view, but helping them means either literally feeding them or giving them the means to go out and buy something to eat. It does not mean engaging in an expensive research activity to develop higher yielding seeds or tillage practices that conserve water or retain soil tilth. The word tilth is not even in the electronic dictionaries that

230

9 Can Agrifood Biotechnology Help the Poor?

people operating from this mindset use to write analyses of hunger or to make appeals for more effective modes of giving assistance. Tilth refers to the condition of tilled soil, especially its capacity to hold water and support root growth. Concept-blindness to tilth is symptomatic of a larger moral problem. Even when aid is conceived as something beyond a handout, it is not tied to farming but to the development of skills that would make a poor person more employable in an urban, industrial economy. Ironically, this mindset also presumes that when smallholders have not been upset by a catastrophic flood, earthquake or invasion of locusts, everything about their farming system is just fine. There is really nothing that needs technical improvement. The focus, then, of international development and ethically motivated assistance is not on helping farmers, as the Lincolnian side of EM had it. The imperative is to get food into urban centers and to make sure that it is accessible by the poor people who live there. As I have argued at more length elsewhere, the ethical analyses of humanitarian obligations debated among philosophers ignore the institutional and technological infrastructure of international aid, (Thompson 1992). Even the recent turn toward “effective altruism,” looks narrowly at how much an organization spends on its own internal activity (staff salaries, fundraising efforts, etc.) in comparison to how much is spent in programming. Advocates of this approach argue that charitable contributions should be judged according to their efficiency in converting donated monies to actual help for the poor, (Singer 2015). Though entirely reasonable in its own right, the standpoint of charitable effort (as Jane Addams put it) obscures features of a needy person’s situation that are crucial to the actions they undertake in negotiating the challenges of survival under difficult conditions, (Addams 1915). The difference between a poor farmer and a poor person scrounging for food in an urban area is ethically crucial. The view that lack of monetary income is the definitive feature of poverty (at least in so far as humanitarian assistance is concerned) has led ethicists to overlook challenges that poor people face in their capacity as farmers, (see Thompson 2015b, Chap. 4). This extended set-up to the chapter’s answer to the title question, “Can agrifood biotechnology help the poor?” is nearly complete. Agricultural insiders such as Norman Borlaug were speaking from a mindset comprising the (to them) obvious ethical importance of helping farmers, and the presumption that doing so would help other people in poverty as a side effect. This provided an excellent rationale for emphasizing technical programs in agricultural science that had the promise of increasing the yields of agricultural crops. This implicit faith in what I have called “the producitonist paradigm” was being shaken but not broken by research which showed that innovations actually fuel farm bankruptcy (the topic of Chap. 8), while research demonstrating the growing environmental costs of the Green Revolution was viewed as an invitation for yet another technological fix. At the same time, on the ground realities in many places where large numbers of poor people actually are farmers were taking a turn for the worse at precisely the same time that wellmeaning egalitarians in the industrialized world were losing touch with on the ground agricultural realities altogether. Affirmations of biotechnology’s potency for helping poor people presumed both moral and empirical background assumptions that were

9.2 Ethics and Agricultural Development

231

questionable, to be sure, but the broader public was in no position to evaluate those assumptions, in any case. The stage was set for a debate that was as irresolvable as it was acrimonious.

9.3 Social Consequences in Peasant Agriculture In plain truth, much of the ethical analysis for social consequences experienced by family farmers in the industrialized world applies equally to resource-poor farmers in developing countries. Long before any GMOs were in the field, Fred Buttel predicted that biotechnology would have unfavorable impact on the rural poor in Africa, Asia and Latin America, while benefiting relatively better-off farmers in those regions. Farms will become larger and fewer. (Buttel et al. 1983; Buttel and Barker 1985; Buttel 1995). Buttel’s work with two Ph.D. students, Martin Kenney and Jack Kloppenburg, Jr. laid the groundwork for more popularly oriented work by activists Henk Hobbelink (1991) and Vandana Shiva (1993, 1995). However, Buttel recognized that the moral significance of transition from agrarian to industrial economy is different in Africa and in many parts of Latin America or Asia than in the United States. More people, both in absolute numbers and as a percentage of the population, are affected. Those who are affected are worse off than European or American farmers, to begin with, and are more vulnerable to displacement. They lack the alternative opportunities for employment that exist in more diversified economies, and many live in countries where social services do not provide an adequate safety net for the poorest of the poor. When food biotechnology displaces labor from agriculture (as it might, for example, if it hastened the advent of herbicides to replace hand weeding), it harms the land-less laborer, the poorest of the poor in the world’s poorest societies. The human cost of agrarian transition in the industrialized world is measured in terms of financial and emotional stress, with occasionally more tragic consequences (see Hendrickson 1987). In parts of the world where more than 20% of the population are farmers, it is measured in exposure, disease, malnutrition and death from the diseases of food deprivation. As noted, although academic philosophers have had relatively little interest in the decline of small farms over the 20th century, they have been much more attentive to the intellectual and moral challenges posed by unequal economic development on a global scale. While one must look to agricultural economists or rural sociologists for a utilitarian or rights based analysis of the family farm issue, some of the most prominent philosophers of recent years have written detailed analyses of hunger and development. Peter Singer constructed an influential moral argument for humanitarian assistance. Singer’s utilitarian formulation holds that if giving aid to keep someone from starving has significantly greater benefit to the recipient of aid than cost to the donor, the donor is obligated to give, (Singer 1972, 1977).2 In addition, Singer has argued against a consumption-oriented food ethic that recommends eating locally produced food. He argues that purchasing foods produced by resourced-challenged farmers abroad provides more total benefit than buying from middle-class farmers

232

9 Can Agrifood Biotechnology Help the Poor?

in industrially diverse economies, (Singer and Mason 2007). Neither of these arguments reflect the costs or benefits of transition from agrarian to industrial social organization. A transition from one farming technology to another having limited moral significance in contemporary Europe, North America and Austria might have far more serious consequences in places where many people still farm at a nearsubsistence level. If so, an equally straightforward utilitarian argument might be developed. If biotechnology accelerates the fewer and larger trend in the developing world, the suffering of those who lose their ability to farm outweighs any benefit to those who make more from farming. Whether a reasonable expectation of farmer bankruptcy and farm loss can be laid at the doorstep of food biotechnology and genetic engineering research is exceedingly difficult to say. There are at least as many people predicting benefits to resource-poor farmers (see Persley 1990; Beachy 1991; Chappell 1996; Wambugu 1999; Mackey 2003) as costs, but counting the number of authors on each side of the issue is a poor way to decide the issue. When the first edition of this book was published in 1997, the literature on social consequences for developing countries included precious little in the way of detailed ex ante studies on the implementation and ultimate adoption of food biotechnology or its products. Perhaps the nature and impact of biotechnology in developing countries was so speculative in the 1980s and early 1990s that useful empirical and theoretical work was impossible, and perhaps studies are currently underway that will rectify the situation. When the second edition appeared in 2007, more evidence was available, (See Pardey 2001; Pray and Naseem 2003; Buttel and Hirata 2003), but as the data grows, the effect of gene technology becomes more equivocal. Bt cotton is alleged to have caused bankruptcy and farmer suicide in India. A review of this issue is beyond the scope of my analysis, but it is clear that Indian cotton farmers took on debt in order to finance GM seed purchases, and that the Indian financial offers little relief to farmers who cannot service their debt, (Stone 2002; Kloor 2014). One detailed analysis of the socioeconomic impact of Bt cotton plantings found farmer access to irrigation is the most critical feature in determining whether a smallholder benefits or suffers from adopting the GMO, (Guiterrez et al. 2015). A meta-analysis of Indian cotton plantings identifies even more complexity, concluding that Bt’s positive impact on yields and profitability is real but modest when compared to other changes that were occurring in farmer practice. The authors conclude, “It now appears that Bt cotton’s primary impact on Indian agriculture will be its role in this rising capital-intensiveness rather than any enduring agronomic benefits,” (Kranthi and Stone 2020). As the causal connections come to be seen as more complex, criticism of agrifood biotechnology’s social impact in dominantly agrarian economies lose some of its punch. They lack the clarity of Robert Kelter’s prediction of rBST’s impact on dairy farmers (Kalter 1985, see Chap. 3 for a discussion). There are overlapping reasons for this. The time-consuming difficulty of research that takes a detailed look a farm management decision making explains why few studies tell us how smallholders are cope with new crop varieties. Dan Hicks has argued that disciplinary and methodological variations in the standard for evidential value penetrate deeply into social science studies that have attempted to document both harmful and beneficial

9.3 Social Consequences in Peasant Agriculture

233

outcomes at the farm level, (Hicks 2015). Although the technology treadmill idea is still influential, the belief that one can single out one innovation and accurately predict its impact now seems somewhat naïve. Nevertheless, ethically motivated concern over social transition from an economy with many small farms to one of larger farms was a motivating factor for many early critics. The Rural Advancement Foundation International (RAFI) is a civil society group that promotes the interests of small farmers both through projects that provide direct aid, and through advocacy. Throughout the 1980s and 1990s, RAFI and aligned individuals or groups published reports on agricultural technology’s adverse impact. Erosion of farmer control over seeds and other key resources was of concern, (see Shand 1991; Manicad 1996). In 2001, some of the most vocal advocates at RAFI formed the ETC Group, which continues to have a strong voice in questioning the impact of agricultural innovations in less industrialized economies. Now distinct from the activities of the ETC Group, RAFI still exists, but it has not had a high profile in debates over biotechnology. The rationale for continuing critiques is, in fact, the same as that discussed at some length in Chap. 8, though, as noted above, the Lincolnian argument for helping small farmers attains greater urgency in areas where farmers are both poor and numerous. These early advocates for smallholders in Africa, Asia and Latin America were up against the agricultural establishment, which was moving ahead with biotechnologybased research programs at universities, national laboratories and the CGIAR system. The ideological difference between these contesting parties came down to a profound disagreement over agricultural technology’s impact on poor farmers. One aspect of the disagreement was grounded in the technology treadmill (discussed in Chap. 8) while classic disagreement between capitalist and socialist economic policy was another. Note that the treadmill argument suggests that slow adopting farmers are less economically competitive. They are thereby forced into foreclosure, fueling the “fewer & larger” cycle of agrarian transition. As such, ending farmer dependence on commodity markets was a favored response among rural advocates. On the one hand, biotechnology was drawn into this debate because the agricultural establishment described it in terms of increases in efficiency that would feed into the treadmill. On the other hand, small farm advocates noticed something that establishmentarians such as Borlaug did not: The agents behind on-the-ground biotechnologies were not universities or international centers like CYMMIT or IRRI; they were large agricultural chemical companies such as Monsanto, DuPont and Ciba-Geigy (later Syngenta). The philosophical dimensions of the treadmill argument are substantive, but subtle. The case against powerful international corporations was easy for anyone to understand. In summary, Borlaug and other members of the agricultural establishment assumed that both private firms and publicly funded organizations would be deploying the techniques of gene transfer in the next generation of seeds and livestock breeds. Their philosophical commitments were those of the establishment mentality (EM). Though not without flaws, the EM made it plausible to think that small farmers would have free access to new varieties that would both increase their yields and would mitigate some of the unwanted environmental effects of Green Revolution

234

9 Can Agrifood Biotechnology Help the Poor?

varieties. A well-constructed argument against this way of linking biotechnology and EM, required two things: First, the socio-technical theory of treadmill effects, made even more treacherous in a time when governments were pushing small farmers into crops sold in international markets; Second, a normative or ethical commitment to small farms and resource poor farmers. Since critics and members of the established shared the commitment to farmers, there was no need to defend it in the early discourse. All this was playing out among people who were informed about agriculture; they cared deeply about its future course. At the same time, philosophers, social scientists and other humanitarian activists were building a discourse that did not involve any moral commitment to farmers, though it did provide an ethical rationale for helping them as long as they are poor. This set the stage for a turn of events worthy of Hegel’s phrase “the cunning of history,” though in this instance the outcomes have hardly been progressive.

9.4 World Feeders and Ethical Consumers This section provides a succinct account of a four-decade transition in thinking on the ethics of food (see Thompson 2015b for more detail). The transition has three crucial moments. First, a “feed-the-world” trope now dominates arguments for biotechnology promulgated by the agricultural establishment. The claim that this technology could help farmers is not prevalent in this discourse. Second, a food ethics movement advocates making food purchases based on the consequences for people, animals and environments. Third, food sovereignty has emerged as a new language for making pro-farmer arguments. Although incipient in the early 2000s, all three of these developments have become more important since the last revision of this book was completed. Helen Zoe Veit’s history of the United States Food Administration (USFA) and successor programs argues that relief efforts associated with the two world wars of the 20th century gave rise to the thought that industrial agriculture could (and should) feed the world. President Woodrow Wilson (1856–1924) established the USFA by executive order in 1917. The idea was to help the United States’ European allies during a time when their own agricultures had been devastated by war. The USFA’s activity included a massive public relations campaign intended to convince Americans that the personal sacrifices involved in doing this were morally praiseworthy, if not obligatory. Veit marks a shift in the moral foundations for the agricultural sector toward obligations directed toward humanity at large, as opposed to self-interest and regional solidarity. The trend continued through the 20th century, finding expression in the formation of “Food for Peace,” initiative launched by President Dwight D. Eisenhower (1890–1969), (Veit 2013). Veit’s history documents and explains how industrial farmers have justified any unwanted side effects of their production systems through a cosmopolitan morality of benefit to the world at large. The philosophical point here must be stressed but not overstated. The Lincolnian rationale for agricultural research discussed above gave scientists a good feeling

9.4 World Feeders and Ethical Consumers

235

about what they were doing, but since many of them had come from farming households, they were helping friends and family back home. If Veit is correct, the good feeling becomes internationalized over the course of the 20th century, and this marks a shift in the moral foundations for agriculture and its supporting institutions, including agricultural science. This is an important point from a philosophical perspective because it articulates a rationale for helping farmers produce more that goes considerably beyond farmers themselves. Agriculture comes to be seen as having a duty to feed the world, and the chief custodians of this duty are the scientists who will provide the next generation of agricultural technology. This duty could be understood as an amendment to EM (as described above) that requires a substantially different philosophical rationale. Yet it is doubtful that very many scientists experienced any intellectual tension between their confidence that they were helping farmers, and their duties to feed the world. The point is that by the 1980s, farmers and scientists both saw their collective endeavor as justified by an imperative to feed the entire planet’s human population. However, as an ethical principle that did not require justifying the peculiar needs or status of farmers, the imperative was also handy for defending biotechnology to an audience that was no longer convinced that farmers had any kind of special moral standing. There are problems with the argument, of course. If agriculture writ large has this responsibility, there is no obvious reason to prefer an agricultural system that supports 20%, 40% or even 80% of a nation’s population, as opposed to one with 1% (or less) in farming. Observations like this were highly substantive among farm sector insiders, though insiders were split on how they viewed its ethical significance. Indifference to the socioeconomic structure of agriculture would have been especially opposed to the analysis being promulgated by RAFI, Food First and other smallholder advocates. However, in 2007 and 2008, the world experienced a sharp uptick in food prices, leading to riots and political unrest in urban centers of many countries with large farming populations, (see Clapp and Cohen 2009). The food crisis sparked a flurry of debate that gave biotechnology advocates an opportunity to argue that increased yields were needed. They could then argue that opponents of GMOs were callously keeping food out of the mouths of hungry people, (Collier 2008; Paarlberg 2009; see Schurman and Munro 2010, p. 180 for further documentation). Viewed from an ethical perspective, it is important to recognize that helping hungry people in cities might actually harm the rural poor, especially when the help comes in the form of food aid from the fields of better-off farmers in food-exporting countries. Donated food has long been known to hurt farmers’ ability sell their produce in local markets, (see Thompson 2010, for a more detailed discussion). Nevertheless, the “feed the world” rationale for biotechnology came to dominate in public discourse. It is arguably the most persuasive ethical rationale in favor of gene technologies even today. Smallholder activists were not sitting on their hands in the 1980s, 1990s and 2000s. Among their signature efforts were arrangements intended to give poor farmers a larger share of what the end consumer finally pays for the products they have grown. Certification schemes organized under the banner of “fair trade” will be familiar to readers. Though not without controversy, fair trade labels are intended to help

236

9 Can Agrifood Biotechnology Help the Poor?

small farmers who were producing export crops, especially in commodities like coffee, cocoa and cotton, (see Downie 2007 for a brief discussion from The New York Times). Does the public’s interest in buying free trade contradict the claim that commitments to the moral significance of farmers was declining between 1970 and 2020? Arguably not. For one thing, very little in promotional materials for fair trade products bases the appeal on the moral significance of helping farmers, as distinct from people who are merely poor. Indeed, one of the challenges of free trade has been the tensions between schemes that improve a farmer’s competitiveness and those that include additional values, such as environmental protection or fair wages throughout the supply chain, (Jaffee and Howard 2016). More broadly, fair trade is just one component of the movement toward ethical consumption of food. The idea that one should make food consumption decisions that are consistent with one’s moral commitments was a prominent part of Peter Singer’s argument in the book Animal Liberation, (1975). Singer was, of course, focused somewhat narrowly on the suffering of livestock in this early treatment. As readers will surely be aware, books, films, videos and short articles exposing harms to both humans and non-humans appear with regular frequency. Michael Pollan’s book The Omnivore’s Dilemma framed his overview of food system practices as the ruminations of someone pondering how to eat ethically, (Pollan 2006). Singer himself answered with a more comprehensive study attempting to integrate his vegetarianoriented concern for animal welfare with other considerations, including the welfare of low wage employees in the food industry and resource poor farmers, (Singer and Mason 2007). Adjusting one’s food purchases according to an estimate of the harm associated with their production defines food ethics for many. Fair trade or concern for farmers is but one component in this picture. In treatments by philosophers, in particular, one would be hard pressed to find an argument that attributes distinction to the occupation of the farmer. Farmers are treated more like owners of capital, who are more likely to be the perpetrators of moral offense, rather than the victims. Following Singer’s lead, concern for animals seems to rise to the top level of concern in philosophical food ethics. The combined impact of world-feeding advocacy and consumption-oriented food ethics is a cultural setting where people are not immediately receptive to the underlying moral claims of the social consequence critiques mounted in the early years of biotechnology. In this kind of political environment, it has proven wise to downplay the concerns that motivated RAFI and many early critics of agrifood biotechnology, and to rely on arguments that raise questions about the safety, regulation and environmental impact of gene technology. None of these questions depends on the Lincolnian pro-farmer attitudes that were foundational to the creation of agricultural science as well as to the not-for-profit organizations that were the custodians of Lincoln’s moral commitments throughout the 20th century. It is, to be overly concise, much easier to scare the pants off people. Given this situation, it has been necessary to reconstruct pro-farmer moral arguments. Food sovereignty is the primary vehicle for this reconstruction, but it is not without problems. The paradigmatic statement of food sovereignty is the Nyéléni Declaration of 2007. It was developed at a conference of civil non-governmental organizations

9.4 World Feeders and Ethical Consumers

237

(NGOs) working on behalf of smallholding farmers, and has since become associated with the work of ViaCampesina, a consortium of NGOs that continues this work. The document includes a number of themes. It represents a generalized form of resistance to development activities that promote industrialization, especially in agriculture. It interprets sustainability less as global agriculture’s ability to feed the entire human population than as the ability of peasant farming communities to persevere in the face of globalization and pressures to adopt modern farming methods. A more detailed discussion of the philosophical implications of the Nyéléni Declaration are beyond the scope of this chapter, (see Werkheiser 2016; Timmermann et al. 2017). In the present context, it is important to stress how food sovereignty was intended as an alternative to food security as a development goal. Development specialists have long stressed food security, which could be defined colloquially as “knowing where your next meal is coming from.” A frequently cited definition from the World Food Summit of 1996 states that food security exists “when all people at all times have access to sufficient, safe, nutritious food to maintain a healthy and active life”, (Anonymous 2015). USDA identifies degrees of food security that range from anxiety about household food access or compromised variety or nutritional quality of household diets to disruption of eating patterns and reduced food consumption, (ERS 2018). What is crucial here is that these definitions emphasize food access, and do not confer significance on food production. In contrast to this emphasis on access, food sovereignty is intended to provide support for a community’s relationship to its farmers and their chosen methods of food production. Food sovereignty might be harmed even as food security is improved. A low cost grocery store might provide community members better food access at a lower price, but it could also drive the farmers in the region out of production, altogether. Food security, not food sovereignty, continues to be the goal of the United Nations Sustainable Development Goal 2 Zero Hunger, promulgated in 2012, (UN, n.d.). The Nyéléni orientation to food sovereignty reinstitutes a moral commitment to farmers. This is not the case, however, in all characterizations of food sovereignty. In other cases, food sovereignty is discussed in the language of control over one’s food system, (see, for example, Alkon and Mares 2012; Block et al. 2012). In laying the stress on control, rather than farmers, food sovereignty is opened to the possibility that a given community might express sovereignty over their local food system by any of several means. More generally, the language of control connects well with the idea that what matters ethically is resisting the control of others, and especially control by globally organized corporate actors. While there may indeed be compelling ethical and political reasons for doing this, it nonetheless reduces the resonance between food sovereignty as an ethical goal and the rationale for assisting farmers. Even more broadly, this language has elicited pushback from people who interpret it as resistance to all forms of globalized trade in foods. No European nation is currently capable of producing all the food consumed by their population, and none are trying to achieve it.

238

9 Can Agrifood Biotechnology Help the Poor?

What does all this mean for biotechnology? There are four possibilities: A. Biotechnology will harm people in developing countries through the “fewer and larger” mechanism of agrarian transition, documented in studies of the Green Revolution. B. Biotechnology will harm people in the developing world through the mechanism of global trade. It will increase the gap between the efficiency of industrialized agriculture and resource poor farmers, tipping the economic advantage toward larger farmers. C. Biotechnology harms people in the developing world primarily through the mechanism of intellectual property. D. Biotechnology has effects on the agricultural research system that have reduced its capacity for innovations that help resource-poor farmers in less industrialized regions.

Early critics of biotechnology saw a clear link between A and C, but their key moral commitments reflect their presumption that doing good for farmers is doing good simpliciter. Their ethical values are reflected in A, rather than C. However, as debate over food policy progressed in light of crop shortages in 2007 and 2008, the world-feeders laid more and more stress on efficiency arguments. The worldfeeding argument has had appeal for humanitarians who envision hunger as an urban phenomenon of insecure food access. What is more, the COVID-19 pandemic of 2020 sparked calls for shortening food supply chains, even in highly industrialized countries. These developments point toward B as the primary focus of debate, but the moral significance of losing a competitive advantage is difficult to defend in a neoliberal context. Defenders of peasant farming have seen programs that would send food from the industrialized world to poor people in Latin America, Africa or Asia as an assault on poor farmers, even if such programs might do good for the urban poor. In the international context, food sovereignty has been formulated as a defense of small-holding farms as an institution, a rationale that draws its inspiration from A. Food ethicists who interpret food sovereignty as a defense of consumer preferences or the access for the urban poor misunderstand the basic argument. The third argument anticipates themes that will be taken up in Chap. 9 (see Anonymous 1996 for a version of the argument). One critical dimension of C-type arguments holds that gene technology constitutes an illegitimate taking of improvements introduced into crops by generations of peasant farmers living in regions of genetic diversity, (Juma 1988; Sharma 2003). For these arguments, it is the innovative step undertaken by smallholders that establishes the ethical claim. The fact that they are poor may be viewed as incidental. A discussion of D-type arguments closes this chapter. Before moving on to consider impacts on science itself, it is worth noticing that the assumptions of the first two arguments produce an ironic tension. A-type arguments assert that resource-poor farmers will be harmed if their countries get seeds and vaccines from rDNA technology; B-type arguments assert that they will be harmed to the extent that they are forced to do without them. The moral foundations of both arguments are similar. On one viewpoint, it is the poverty of farmers that matters ethically. Undercutting middle class farms would not be an occasion for moral distress. Advocates of food sovereignty have struggled mightily to reinstitute arguments that were familiar elements of agrarian philosophy, and they have had some success. Perhaps the best option, however, is to see this tension as expressing

9.4 World Feeders and Ethical Consumers

239

a tragic truth: poor farmers are going to be harmed if they get biotechnology, and they will also be harmed if they are forced to do without it.

9.5 Social Consequences for the Conduct of Science Sheldon Krimsky’s Biotechnics and Society (1991) and Busch, Lacy, Burkhardt and Lacy’s Plants, Power and Profit (1991) both predicted that some of the most serious social consequences from food biotechnology would be experienced within the community of science itself. They predicted that commercialization of science would divert research away from basic research as well as from research aimed at publicly beneficial, but less profitable subjects. They predicted that the conduct of science itself would be hurt by burdensome licensing and IPR secrecy procedures, and by restrictions on the disclosure of proprietary information. They predicted that corporations would gain ownership of the products of biotechnology without paying a fair share of the costs for research and development. These developments are ethically significant because, as argued above, Norman Borlaug’s advocacy for biotechnology was predicated on the assumption that agricultural science would continue to function at least as well as it had throughout his career. To an extent, all of their predictions have been realized, though perhaps not to an extent that an impressionable reader of these books might have expected. The 1997 text of Food Biotechnology in Ethical Perspective continued with the following observation: At least two institutes at Texas A&M solicit annual fees from food industry firms for which these companies get nothing more than the right to “get close” to university scientists, as the director of one such institute puts it. As Director of an ethics center, I have yet to sense a desire for companies to “get close,” or to get early, privileged access to research results. To the extent that availability of funds inevitably influences what research is done, it is impossible to deny that research choices at Texas A&M are more responsive to market forces than they have ever been before.

Many voices were added to the list of those expressing concerns in the ensuing decade. In 2000, The Atlantic Monthly, a large circulation U.S. news and opinion magazine, ran a cover story entitled “The Kept University.” The authors did not single out agrifood applications of gene transfer and were more interested in drugs and medical biotechnology. Nevertheless, the article brought the growing alliance between universities and private industry to widespread public attention. The authors argued that this trend would compromise not only the direction, but also the results of university research, (Press and Washburn 2000). Larry Busch devoted the last ten of his research activity to a sustained critique of the effects that for-profit innovation has on university science and education. Busch saw this impact as operating at many scales. It was not simply the causal mechanisms associated with treadmill-like impacts on small farmers. Busch argued that the entire culture of the educational sector was being transformed, and that faculty and administrators were losing touch with the idea that universities operate on the

240

9 Can Agrifood Biotechnology Help the Poor?

basis of an assumed fiduciary responsibility to further the public good, (Busch 2017). The ethical significance of change within the institutions of science can be analyzed along at least three distinct lines. I will call them the aesthetic purity argument, the social function argument, and the public trust argument.

9.5.1 The Scientific Purity Argument One might argue that science, like art or sport, has an internal purpose that can only withstand so much pollution from extraneous sources. The internal purpose of science is pursuit of truth. According to this view, the social context of science is largely irrelevant to its essence, which is to employ observation, deduction and experimental procedures in the discovery of nature’s laws and in the development and verification of logically coherent theory. Scientists must of course have buildings and equipment, just as they must eat and breathe, but the social and economic forces that impinge upon the conduct of research have no more effect on its essence than do the mental fatigue or bodily ailments that eventually force any individual scientist to quit the laboratory for sleep and relief. This image of science, though challenged of late, is fairly standard throughout 20th century philosophy of science (see Brodbeck 1953). However, biotechnology is not ruining science in the way that it might be alleged to be ruining farming, for not only is science a deeply technological practice, the ability to use rDNA techniques in the activity of science takes great skill and art. Nevertheless, the commercialization of science might ruin its capacity to exist as a practice that gives meaning and focus to the lives of scientists. It might do this by substituting externally profit-driven goals for the internal goals defined by pursuit of truth. The potential for wealth production might divert scientists from the essence of science. To the extent that one sees science as a practice, internally determined and characterized by its essence, it is reasonable to interpret this turn of events as a form of corruption and a loss of virtue for scientists (Goldworth 1991). When the aesthetic dimension is placed at the forefront, it becomes possible to bemoan the loss of scientific purity, just as one might mourn changes that have taken place in art or sport (see Ruscio 1994). One might, then, argue that scientists should be pursuing gene technologies because that is where their curiosity and the practice of science has taken them. It is worth stressing that this is assuredly not the way that Krimsky or Busch interpret the moral significance of social consequences for science. Both take science to be socially embedded, and deeply influenced by its dependence on funding sources, be they public monies or venture capital. In addition, philosophers of science have moved steadily away from this vision of science since 1997 (Kitcher 2003; Elliott 2017). Even if protecting the purity of science for its own sake has merit, it is not an argument that engages the reasons why Borlaug or other members of the agricultural science establishment defended biotechnology.

9.5 Social Consequences for the Conduct of Science

241

9.5.2 The Social Function Argument John Stuart Mill offered a defense of strong academic freedom for scientists in his essay “On Liberty.” Mill argued that scientists should be free from interference in pursuit of whatever interested them because, given the unpredictability of the applications of science, total freedom of thought is the best path toward realizing social benefit from science (Mill 1859). Mill did not think that there needed to be an argument to establish the root claim that science produces social benefits. There have been many reformulations of this utilitarian argument for letting scientists follow their curiosity in the intervening century and a half. Measuring the social returns to research has become a minor industry among economists, and Robert Evenson (2002) has done research that specifically ties this theme to agrifood biotechnology. As is often the case with social science, this research seems to presume a broadly utilitarian framework without articulating any philosophical commitment to utilitarianism explicitly. However, Philip Kitcher’s book Science, Truth and Democracy makes just such a philosophical argument. Kitcher integrates some fairly conventional philosophy of science with a discussion of the philosophical critique levied by Herbert Marcuse (1898–1979), Theodor Adorno (1903–1969) and Max Horkheimer (1895–1973), the Critical Theory school. These Marxist theorists had argued that by being situated in capitalist societies, scientific ideals of truth and method had become distorted. Kitcher rejects the Critical Theorists’ criticism of scientific method, but accepts the argument that capitalism tends to have a distorting effect on the kinds of questions scientists ask, and on the kinds of research they eventually undertake. He then moves on to develop a theory of what science should do, what the research agenda ought to be, given the norms of democratic societies. Here, he argues that citizens in a democracy will want to support those scientific projects that are most likely to improve their quality of life. It is the ultimate consequences for human welfare that should determine the agenda for scientific research (Kitcher 2003). It would seem that Kitcher’s view comes close to the mindset that I have attributed to establishmentarians among agricultural scientists (e.g. EM). One would need to add some additional argumentation establishing the moral importance of agriculture, and of the value that agriculture’s practitioners lend to other social institutions. There is a robust philosophical tradition of such argumentation that extends back to ancient Greece, though I have not provided a detailed discussion of it here. However, as described above, agricultural scientists and the administrative staff at agricultural universities and research institutes (including the CGIAR centers) would have shared the sentiments articulated by Abraham Lincoln during the middle of the 19th century. They believe that science has a social function, and that scientists freedom to pursue whatever seems most interesting and exciting at the time helps scientists fulfill that function. They also believed something not discussed by Kitcher, namely, that the social function of agricultural science is to help farmers. But many of the critics discussed throughout this book—Krimsky, Busch, Buttel, Kloppenbug, Shiva—were claiming that the intrusion of the profit motive was

242

9 Can Agrifood Biotechnology Help the Poor?

distorting agricultural science’s ability to fulfill a social function, however it is understood. Kitcher presumes that capitalist societies tend to deviate from their guiding norms in favor of research that is profitable for capitalists. Research would be skewed to the kinds of questions wealthy people ask, and they can be presumed to ask questions about how they can become wealthier still. As such, he suggests a thought experiment in which citizens vote for the kinds of science they want. He is not advocating voting as a serious decision mechanism for research policy, merely using this idea to test how the ethical content of the utilitarian’s goal of maximizing welfare for the population as a whole. He sees two main ethical problems with the voting ideal, one being that people cannot be expected to have enough scientific sophistication to predict which lines of inquiry really are to their benefit. The other is that people may have immoral preferences for research; they may support research that reinforces their illegitimate preferences. Although the first problem is probably the most relevant to agrifood biotechnology, it is the latter question that gets a detailed treatment by Kitcher. Examples of medical biotechnology and genetics get a fair amount of attention. In short, Kitcher is concerned that racial prejudice or faulty views on the links between genetics and moral conduct will skew the voting, (Kitcher 2003). Kitcher’s response to this problem (again not focused on agrifood biotechnology or the GM debate) is to suggest something like a “citizen jury” in which people are given access to various expert perspectives on the likely prospects of science, (Kitcher 2003). Citizen juries have been convened to discuss agrifood biotechnology, with mixed results. In some cases, citizens have rejected gene technologies even after independent analysis of the expert testimony in their defense as found it persuasive, (Pimbert et al. 2001). However, in Australia analysts of several science communication projects (including citizen juries) found that they were biased toward acceptance of biotechnology (Schibeci and Harwood 2007). Other analysts have highlighted the difficulties in recruiting participants that could claim to function as legitimate representatives of the public at large, (Lezaun and Soneryd 2007). Chapter 12 discusses some of the philosophical issues that arise in connection with communication and governance of gene technologies in the food sector. The point to note in the present context is that Kitcher’s version of the social function argument embroils the institutions of science in efforts to increase the public’s voice in steering the direction of scientific research. As mentioned earlier, Lawrence Busch and William Lacy argued that food and agricultural science became structurally tied to commercial interests well before the advent of biotechnology. These ties produced an institutional structure that was conservative and tradition bound in its choice of research problems, just the opposite of what Mill and Kitcher envision (Busch and Lacy 1983). The predictions in Plants, Power and Profit are an extension and application of that earlier work. If their empirical analysis is correct, then utilitarianism would support the same conclusion as the aesthetic purity argument, but for very different reasons. Science should remain somewhat distant from commercial influence because, so-called free market economics to the contrary, commercial influences do not align science with public benefit. Yet neither Busch and co-authors nor Krimsky appeal directly to the social

9.5 Social Consequences for the Conduct of Science

243

function argument in criticizing the social consequences of food biotechnology for science itself. Their arguments instead appeal to the importance of public trust.

9.5.3 The Public Trust Argument If scientists working in research organizations have accepted public funds to pay their salaries and those of their graduate students, to provide physical facilities, and perhaps even to purchase equipment, is it fair that the results of their research should be controlled by private industries that may have contributed only a fraction of the total investment? This rhetorical question insinuates the moral principle that what has been paid for with public funds belongs to the public. To divert public property toward private use violates an ethical principle that should need no argument. As Busch, Lacy, Burkhardt and Lacy put the case, “society may pay twice: once for the research and again for its benefits and products,” (1991, p. 196). But it should be noted that history, sociology and English professors collect royalties on the books and poems that they publish (and in some few cases, the amounts are not trivial), while no one raises an eyebrow. Despite the authority with which Krimsky and Busch, Lacy, Burkhardt and Lacy advance this critique, there are murky questions in research ethics here that deserve a wider and more considered hearing. Divided loyalties and conflicts of interest betray the public trust in another sense, as well. According to Krimsky, the most significant social consequence of change within scientific institutions is “the disappearance of a critical mass of elite, independent and commercially unaffected scientists to whom we turn for vision and guidance when we are confounded by technological choices,” (Krimsky 1991, 79). We can interpret the public trust as a social contract, just as John Locke (1632–1704) or Jean-Jacques Roussea (1712–1778) understood it. Food biotechnology, however, has a role in the contemporary social contract that differs from the science of Mill’s day. The scientists of the 17th, 18th and early 19th centuries were upper class gentlemen tinkering in their spare time. Science is now seen to be essential to the protection of life and health. It can help identify threats to individual or environmental health that would have been written off as “acts of God,” in earlier times. Science is also a source of threats to health, as the preceding chapters have documented. Technoscience is both a threat and a guarantor against threats. To those who fear the commercialization of science through biotechnology, the problem of public trust is a case of asking the fox to guard the henhouse. Pertinent to the theme of this chapter, the question, “Can agrifood biotechnology help the poor?” is now situated within the context of the institutional organization of agricultural science. It is less whether the technology could help the poor under some idealized or imagined scenario, or even whether gene technology will help the poor, given the convoluted nexus of disadvantages and power inequities that constitute their repression. The question really becomes, “Can scientists be trusted to make helping the poor a morally binding component of their research?” This is in one sense an easier burden to bear, as they are bound to have good intentions far more often than

244

9 Can Agrifood Biotechnology Help the Poor?

they could be bound to produce good results. At the same time, observers would be justified in faulting scientists, research groups or institutes whose short term interests in obtaining funding, patents, or professional rewards distort their decision-making. What is more, failure to give reflective consideration to the ethical questions surveyed throughout this book is grounds for withholding trust. This way of construing the relationship between science and the public anticipates ethical issues that will be taken up in Chap. 12. They cut across every area in which food biotechnology might be thought to have unintended consequences and depend as much on public attitudes as they do on the institutional structure of science. As little as the public might care about the institutional effects of biotechnology within the agricultural science establishment, they may well be among the most far reaching. These moral issues are being raised in connection to the way that universities and public research organizations are changing their funding relationships with the food industry, and to the changing importance of intellectual property. There can be little doubt that biotechnology precipitated many of these changes, as scientists established equity positions in private firms, and universities sought to establish more capable intellectual property offices throughout the 1980s (Teitelman 1989), but similar things happened throughout other sectors of science. Many of the social changes on the structure of agricultural science now appear to be tied to the Reagan/Thatcher era, and to the end of the cold war. The rise of agrifood biotechnology was less a causal factor in this transition than a happenstance coincidence with events that had nothing to do methods of gene transfer. It may be time to inspect the infrastructure of our research organizations and to think about repairing any damage, but food biotechnology and some revised relationship between public and private sector research will be the norm.

9.6 Conclusion I have argued that failures of the Green Revolution notwithstanding, the institution of agricultural science entered the 1980s with a significant capacity to use gene technology to benefit resource-poor farmers in less-industrialized economies. Indeed, failures within Green Revolution initiatives spawned a decade of social learning and debate within the agricultural development community, placing the research organizations committed to serving these farmers in a better position to do so. At the same time, two factors mitigated against the utilization of this internal capacity for reflective socioeconomic risk assessment. One was the private sector’s colonization of universities and research institutes. The other was an ethos within applied biology that I have associated with positivism, (Thompson 2017). Whether my analysis is correct or not, (see Wolf 2019), agricultural scientists with training in biology, physics and chemistry cede both authority and funding to social scientists quite reluctantly. One cannot be confident that international research organizations would have realized their capacity for more socially reflective implementation of technology, if only the private sector had not poisoned the well.

9.6 Conclusion

245

Activist opposition to GMOs proved to be a two-edged sword in this milieu. On the one hand, civil society organizations gave a through spanking to the agricultural research establishment, especially with respect to the fate of agrifood gene technology in Europe and Africa. This certainly made the biophysically trained science community more receptive to the social scientists and philosophers who were trying to encourage a more effective way to launch yield-enhancing technology. On the other hand, the critics were prone to broad claims, some of which were just factually inaccurate. What is more, they adopted a worldview, “…incapable of seeing any conditions under which agricultural biotechnology might be useful,” (Schurman and Munro 2010, p. 192). This led industry scientists to despair of any constructive dialog, and many public sector researchers followed them. At the same time, social scientists and philosophers who took up the study of GMOs after 2000 did not familiarize themselves with the critical literature generated within agricultural research institutions. They tended to take the critics claims at face value, and are still producing studies that demonstrate little understanding of or sympathy for the challenges faced by resource-poor farmers. When agricultural scientists argue that gene technology can help resource poor farmers in less industrialized economies, they are likely presuming that non-profit research institutes will be doing the work. Early critiques of agrifood biotechnology questioned whether that assumption would be valid. In addition to the points noted above, it is also important to note an argument that generally comes from the political right. Strong regulatory regimes for biotechnology make it very difficult for nonprofit institutes to release engineered crops or livestock breeds because the costs of generating data and providing the legal expertise to move a variety through the regulatory process often exceed the costs of basic research. This observation has been made in arguments for lightening the regulatory requirements for gene-edited crops, (Kupferschmidt 2018). If regulatory hurdles prevent publicly funded researchers from delivering beneficial crops to the poor, there must be persuasive risk-based reasons for them. When the circumstances of smallholders who farm in dominantly agrarian economies are considered in their totality, there are reasons to question whether the establishment mindset (EM) prevailing within agricultural research is really prepared to do them good. This chapter has discussed how domestic policies within emerging economies have pushed small farmers into the production of commodity crops for sale in international markets. A discussion of whether that is, on balance, good or bad for them is beyond the scope of this volume. The rise of food sovereignty movements advocating on behalf of smallholders provides at least prima facie reasons for thinking that becoming entangled in global commodity markets is not beneficial to the poor. But this argument is only a condemnation of gene technology if one assumes that gene technology will only be used on commercially valuable commodity crops. That assumption is precisely what a robust, public sector agricultural research establishment might have been expected to counter. And so the argument circles back again. Could agrifood biotechnology help the poor? The answer to that question is assuredly, yes, but when we follow the risk assessment injunction in include the

246

9 Can Agrifood Biotechnology Help the Poor?

likelihood that an outcome will prevail into our judgment, the answer is less favorable. There are uses of gene technology that could provide smallholders with varieties better adapted to their growing conditions, especially in an era when greenhouse gas emissions are causing instability in those conditions. I have not given the question of how biotechnologies could be tailored to help smallholders its due in this chapter. Part of the answer lies in the concept of lower cost intermediate technology, while another part is to embed a technology release within an infrastructure that helps farmers avoid the complex agronomic and financial failures that accompanied the introduction of Bt cotton in India. The proliferation of small seed companies and the lack of effective oversight created a “black market,” in seeds—including transgenics—where farmers believe they have accessed GMO traits in seeds that are not certified and that may have very low quality control. Even seeds sold as certified varieties may be counterfeit, (Herring and Kandlikar 2009). Someone might argue that indigenous farmers already have all the resources they need, and that help from a predominantly white science establishment in the industrialized world is only one more permutation of the colonizing mentality that has repressed the poor for so long, (see Whyte 2018). The chapter has gone on too long to engage that argument fully. It must suffice to say that, like criticisms of the Green Revolution, legitimate worries that Western scientists were not respecting the intelligence or accomplishments of peasant farmers were taken on board in agricultural development circles several decades ago. I do not assert that the system was working well in 1980, much less that it works well today, but scholars who ignore the critiques of the 1970s, 1980s and 1990s are depriving themselves of an opportunity to learn from the past. What is more, many resource poor farmers could, indeed, benefit from better tools. The reasons why gene technology might not be the best way to help the poor have more to do with the social institutionalization of agricultural science and its critics than with the nature of gene technology itself.

References Addams, J. 1915 [2002]. Democracy and Social Ethics. Urbana: University of Illinois Press. Alkon, A.H., and T.M. Mares. 2012. Food sovereignty in US food movements: Radical visions and neoliberal constraints. Agriculture and Human Values 29: 347–359. Anonymous. 1996. Intellectual property rights, community rights and biodiversity. Bija-the-Seed 15:254–30. Anonymous. 2015. Food security. Disabled World. https://www.disabled-world.com/fitness/nutrit ion/foodsecurity/. Accessed 31 July 2019. Beachy, R. 1991. The very structure of scientific research does not mitigate against developing products to help the environment, the poor and the hungry. Journal of Agricultural and Environmental Ethics 4: 159–165. Block, D.R., N. Chávez, E. Allen, and D. Ramirez. 2012. Food sovereignty, urban food access, and food activism: Contemplating the connections through examples from Chicago. Agriculture and Human Values 29: 203–215. Brodbeck, M. 1953. The nature of philosophy of science. In Readings in the Philosophy of Science, ed. H. Feigl and M. Brodbeck, 3–7. New York: Appleton-Century-Crofts Inc.

References

247

Browne, W.P., J.R. Skees, L.E. Swanson, P.B. Thompson, and L.J. Unnevehr. 1992. Sacred Cows and Hot Potatoes: Agrarian Myths in Agricultural Policy. Boulder CO: Westview Press. Busch, L. 2017. Knowledge for Sale: The Neoliberal Takeover of Higher Education. Cambridge, MA: MIT Press. Busch, L.M., and W.B. Lacy. 1983. Science, Agriculture, and the Politics of Research. Boulder, CO: Westview Press. Busch, L., W.B. Lacy, J. Burkhardt, and L.R. Lacy. 1991. Plants, Power and Profit: Social, Economic, and Ethical Consequences of the New Biotechnologies. Cambridge, MA and Oxford: Basil Blackwell. Buttel, F.H. 1995. The global impacts of agricultural biotechnology: A post-green revolution perspective. In Issues in Agricultural Bioethics, ed. T.B. Mepham, G.A. Tucker and J. Wiseman, 345–360. Nottingham, UK: University of Nottingham Press. Buttel, F.H., J. Kloppenburg Jr., M. Kenney, and J.T. Cowan. 1983. Genetic engineering and the restructuring of agricultural research. The Rural Sociologist 3: 132–144. Buttel, F.H., and R. Barker. 1985. Emerging agricultural technologies, public policy and implications for Third World agriculture. American Journal of Agricultural Economics 67: 1170–1175. Buttel, F.H., and L. Busch. 1988. The public agricultural research system at the crossroads. Agricultural History 62: 303–324. Buttel, F.H., and A. Hirata. 2003. The Gene Revolution in Global Perspective: A Reconsideration of the Global Adoption and Diffusion of GM Crop Varieties, 1996–2002. Paper No 9, PATS Staff Paper Series, Cooperative Extension, University of Wisconsin-Madison. Chappell, J. 1996. Plant biotechnology comes of age—again. Molecular Breeding 2: 1–6. Clapp, J., and M.J. Cohen. 2009. The food crisis and global governance, In The Global Food Crisis: Governance Challenges and Opportunities, ed. J. Clapp and M.J. Cohen. Waterloo, ON: Wilfrid Laurier University Press. Collier, P. 2008. The politics of hunger: How illusion and greed fan the food crisis. Foreign Affairs 87 (6): 67–79. Downie, A.C. 2007. Free trade in bloom. The New York Times, October 7. https://www.nytimes. com/2007/10/02/business/worldbusiness/02trade.html. Accessed 31 July 2019. Ehrlich, P.R., and A.H. Ehrlich. 2009. The population bomb revisited. The Electronic Journal of Sustainable Development 1: 63–71. Elliott, K.C. 2017. A Tapestry of Values: An Introduction to Values in Science. New York: Oxford University Press. Evenson, R.E. 2002. From the green revolution to the gene revolution. In Economic and Social Issues in Agricultural Biotechnology, eds. R.E. Evenson, V. Santaniello, and D. Zilberman, 1–16. New York: CABI Publishing. ERS (Economic Research Service). 2018. Definitions of Food Security, U.S. Department of Agriculture. https://www.ers.usda.gov/topics/food-nutrition-assistance/food-security-in-theus/definitions-of-food-security/. Accessed 31 July 2019. Falcon, W.P. 1970. The green revolution: Generations of problems. American Journal of Agricultural Economics 52: 698–710. Goldworth, A. 1991. Publicly funded scientific researchers are not entitled to profit from their discoveries. Journal of Agricultural and Environmental Ethics 4: 192–199. Gutierrez, A.P., L. Ponti, H.R. Herren, J. Baumgärtner, and P.E. Kenmore. 2015. Deconstructing Indian cotton: Weather, yields, and suicides. Environmental Sciences Europe 27 (1): 12. https:// doi.org/10.1186/s12302-015-0043-8. Griffen, K. 1974. The Political Economy of Agrarian Change. London: Macmillan. Hendrickson, P. 1987. Those who are no longer with us. In Is There a Moral Obligation to Save the Family Farm?, ed. G. Comstock, 47. Ames: Iowa State University Press. Herring, R.J., and M. Kandlikar. 2009. Illicit seeds: intellectual property and the underground proliferation of agricultural biotechnologies. In The Politics of Intellectual Property: Contestation over the Ownership, Use, and Control of Knowledge and Information, eds. S. Haunss and K.C. Shadlen, 56–79. Cheltenham, UK: Edward Elgar.

248

9 Can Agrifood Biotechnology Help the Poor?

Hicks, D.J. 2015. Epistemological depth in a GM crops controversy. Studies in the History and Philosophy of Science. A 50: 1–12. Hobbelink, H. 1991. Biotechnology and the Future of World Agriculture. London: Zed Books. Jaffee, D., and P.H. Howard. 2016. Who’s the fairest of them all? The fractured landscape of US fair trade certification. Agriculture and Human Values 33: 813–826. Johnston, B.F., and J.W. Mellor. 1961. The role of agriculture in economic development. The American Economic Review 51: 566–593. Juma, C. 1988. The Gene Hunters: Biotechnology and the Scramble for Seeds. Princeton, NJ: Princeton University Press. Juma, C. 2014. The global risks of rejecting biotechnology. https://geneticliteracyproject.org/2014/ 12/09/global-risks-of-rejecting-agricultural-biotechnology/. Accessed 6 March 2020. Kalter, R.J. 1985. The new biotech agriculture: Unforeseen economic consequences. Issues in Science and Technology 13: 125–133. Kloor, K. 2014. The GMO-suicide myth. Issues in Science and Technology 30 (2): 65–78. Kranthi, K.R., and G.D. Stone. 2020. Long-term impacts of Bt cotton in India. Nature Plants 6: 188–196. Kitcher, P. 2003. Science, Truth and Democracy. New York: Oxford University Press. Krimsky, S. 1991. Biotechnics in Society: The Rise of Industrial Genetics. New York: Praeger Publishers. Kupferschmidt, K. 2018. EU verdict on CRISPR crops dismays scientists. Science 361: 435–436. Kunz, D.B. 1997. The Marshall Plan reconsidered: A complex of motives. Foreign Affairs 76: 162–170. Lezaun, J., and L. Soneryd. 2007. Consulting citizens: Technologies of elicitation and the mobility of publics. Public Understanding of Science 16: 279–297. Lincoln, A. 1859 [2018]. Address before the Wisconsin state agricultural society, Abraham Lincoln Online. https://www.abrahamlincolnonline.org/lincoln/speeches/fair.htm. Accessed 29 July 2019. Mackey, M. 2003. The application of biotechnology to nutrition: An overview. Journal of the American College of Nutrition 21: 157S-160S. Manicad, G. 1996. Biodiversity conservation and development: The collaboration of formal and non-formal institutions. Biotechnology and Development Monitor 26: 15–17. https://biotech-mon itor.nl/2604.htm. Accessed 2 March 2020. McMichael, P. 2012. Development and Social Change: A Global Perspective, 5th ed. Thousand Oaks, CA: Sage Publications. Mill, J. S. 1859 [1961]. On liberty. In The Philosophy of John Stuart Mill, ed. M. Cohen, 185–319. New York: The Modern Library. Moore, S.A. 2001. Technology and Place: Sustainable Architecture and the Blueprint Farm. Austin: University of Texas Press. Pardey, P.G. 2001. The Future of Food: Biotechnology Markets and Policies in an International Setting. Washington, DC: International Food Policy Research Institute. Paarlberg, R. 2009. Starved for Science: How Biotechnology Is Being Kept Out of Africa. Cambridge, MA: Harvard University Press. Persley, G. 1990. Beyond Mendel’s Garden; Biotechnology in the Service of World Agriculture. Wallingford, UK: CAB International. Pimbert, M., T. Wakeford, and P.V. Satheesh. 2001. Citizens’ juries on GMOs and farming futures in India. LEISA Magazine 17 (December): 27–30. Pollan, M. 2006. The Omnivore’s Dilemma: A Natural History of Four Meals. New York: Penguin Books. Pray, C.E., and A. Naseem. 2003. Biotechnology R&D: Policy Options to Ensure Access and Benefits for the Poor. ESA Working Paper No. 03–08. https://www.fao.org/tempref/docrep/fao/ 007/ae041e/ae041e00.pdf. Accessed 3 March 2020.

References

249

Press, E., and J. Washburn. 2000. The Kept University. The Atlantic Monthly 285(3):39–54. https://www.theatlantic.com/magazine/archive/2000/03/the-kept-university/306629/. Accessed 6 March 2020. Rassmusson, W.D. n.d. Lincoln’s agricultural legacy, National Agricultural Library, USDA. https:// www.nal.usda.gov/lincolns-agricultural-legacy. Accessed 29 July 2019. Ruscio, Kenneth P. 1994. Policy cultures: The case of science policy in the United States. Science, Technology and Human Values. 19: 205–222. Schibeci, R., and J. Harwood. 2007. Stimulating authentic community involvement in biotechnology policy in Australia. Public Understanding of Science 16: 245–255. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists versus Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press. Scott, N.D. 2018.Food, Genetic Engineering and Philosophy of Technology. New York: Springer. Shand, H. 1991. There is a conflict between intellectual property rights and the rights of farmers in developing countries. Journal of Agricultural and Environmental Ethics 4: 131–142. Sharma, D. 2003. GM Food and Hunger: A View from the South. New Delhi: Forum for Biotechnology and Food Security. Shiva, V. 1993. Monocultures of the Mind: Perspectives on Biodiversity and Biotechnology. London: Zed Books. Shiva, V. 1995. Biotechnological development and the conservation of biodiversity. In Biopolitics: A Feminist Reader on Biotechnology, ed. V. Shiva and I. Moser, 193–213. London: Zed Books. Singer, P. 1972. Famine, affluence and morality. Philosophy and Public Affairs 1: 229–243. Singer, P. 1975. Animal Liberation: A New Ethic for Our Treatment of Animals. New York: Avon Books. Singer, P. 1977. Reconsidering the famine relief argument. In Food Policy: The Responsibility of the United States in the Life and Death Choices, ed. P.G. Brown and H. Shue, 36–53. New York: Free Press. Singer, P. 2015. The Most Good You Can Do: How Effective Altruism Is Changing Ideas about Living Ethically. New Haven, CT: Yale University Press. Singer, P., and J. Mason. 2007. The Ethics of What We Eat. Emmaus, PA: Rodale Press. Stone, G.D. 2002. Biotechnology and suicide in India. Anthropology News 43(5): 5. https://pages. wustl.edu/files/pages/imce/stone/stone_biotechnology_and_suicide_in_india.pdf. Accessed 15 July 2020. Teitelman, R. 1989. Gene Dreams: Wall Street, Academia and the Rise of Biotechnology. New York: Basic Books. Thompson, P.B. 1992. The Ethics of Aid and Trade: U.S. Food Policy, Foreign Competition and the Social Contract. New York: Cambridge University Press. Timmermann, C., G.F. Félix, and P. Tittonell. 2017. Food sovereignty and consumer sovereignty: Two antagonistic goals? Agroecology and Sustainable Food Systems 42(3):274–298. Thompson, P.B. 2010. Food aid and the famine relief argument (Brief return)”. The Journal of Agricultural and Environmental Ethics 23: 209–227. Thompson, P.B. 2015a. From Field to Fork: Food Ethics for Everyone. New York: Oxford University Press. Thompson, P.B. 2015b. From world hunger to food sovereignty: Food ethics and human development. Journal of Global Ethics 11: 336–350. Thompson, P.B. 2017. The Spirit of the Soil: Agriculture and Environmental Ethics, 2nd ed. New York: Routledge. UN (United Nations). n.d. Food security and nutrition and sustainable agriculture. Sustainable Development Goals. https://sustainabledevelopment.un.org/topics/foodagriculture. Accessed 31 July 2019. Veit, H.Z. 2013. Modern Food, Moral Food: Self-Control, Science, and the Rise Of Modern American Eating in the Early Twentieth Century. Chapel Hill: University of North Carolina Press. Wade, N. 1974. Green revolution (I): A just technology, often unjust in use. Science 186: 1093–1096. Wambugu, F. 1999. Why Africa needs agricultural biotech. Nature 400: 15–16.

250

9 Can Agrifood Biotechnology Help the Poor?

Werkheiser, I. 2016. Individual and community identity in food sovereignty: the possibilities and pitfalls of translating a rural social movement. In The Routledge Handbook of Food Ethics, ed. M. Rawlinson and C. Ward, 393–403. Routledge. Whyte, K.P. 2018. Indigenous science (fiction) for the Anthropocene: Ancestral dystopias and fantasies of climate change crises. Environment and Planning E: Nature and Space 1 (1–2): 224–242. Wolf, C. 2019. Displacing the Productionist Paradigm: A Comment on Paul Thompson’s Spirit of the Soil. Ethics, Policy & Environment 22: 235–242. Zimdahl, R.L. 2006. Agriculture’s Ethical Horizon. San Diego, CA: Academic Press.

Chapter 10

Conceptions of Property and the Biotechnology Debate

Abstract The chapter provides an analytic framework for applying classic philosophical theories of property and the distribution of property rights in the context of emerging technology. Instrumental theories of property view property as a convention that should be evaluated according to the purposes it serves. Ontological theories of property claim that holding and exchanging items of property is a natural or intrinsic feature of the human condition. The early debate over so-called Terminator seeds is used to link key philosophical questions to real disputes in policy and practice. The Terminator case illustrates distinctions between property in tangible goods (such as seeds) and intellectual property, as well as the relationship between these forms of property and the risk-based approach that is the focus of earlier chapters. The chapter reviews a sample of the literature on contested property claims in products of gene technology, and discusses how authors draw selectively on concepts from different philosophical traditions. The chapter also identifies logical flaws in many arguments, both for and against the application of intellectual property rights to GMOs and other products of gene technology. In the end, I argue that philosophical theories of property can be enlisted both to support and to criticize current practices. The chapter does not provide a conclusive standard for deciding the legitimacy of property claims in genes, sequences and gene products. Keywords Patents · Terminator · Property rights · Natural law · Utilitarianism · Libertarianism · Political theory Disputes over the ethical justifiability of ownership claims in genes, seeds and processes central to the manipulation of life forms are crucial elements in the debate over agrifood biotechnology and GMOs. Property rights play a crucial role in defining commercial and political relationships among human beings. Changes in these relationships inevitably introduce uncertainty in social affairs, and often alter the balance of competing interests among buyers and sellers, producers and consumers. Gene technologies have altered property relations in two different ways. First, legal definitions of property have been changed through legislative and judicial action, and these changes have initiated further change in contractual agreements that specify

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_10

251

252

10 Conceptions of Property and the Biotechnology Debate

rights in specific cases. For example, patenting of genes and gene sequences enabled the creation of technology licenses that stipulate restrictions on a farmer’s right to save and replant seed, (Carstenson 2005). A cascade of issues follow, including legal and cultural issues in the enforcement of new legal codes, and their interaction with other laws, such as antitrust. Second, technology itself can introduce change into the structure of customary rights, often by altering the ease with which one person can exclude another person from access to a good. The most straightforward example is a dam, which gives upstream users an ability to exclude downstream users from access to water. Such technologies precipitate both moral and legal disputes that can last for generations. Gene technology has changed the customary structure of ownership in the genetically-based traits of agricultural plants and animals by allowing innovators to alienate these traits from the seeds or breeding populations in which they were previously embedded. To complicate matters even more, the 1980s and 1990s saw a surge of interest in the philosophical foundations of property claims. This work had only begun to have influence on the philosophy of intellectual property when the first edition appeared, but the literature has continued to grow, (see Drahos 2016). The first edition of this book took a simplifying approach. Consistent with the overall intent of the book, the goal of this chapter was to equip readers with a primer on the ethics of property claims, and then examine how these ideas might be applied to biotechnology. The resulting analysis shows how intersecting strands of philosophical theories on property produce a landscape of concepts that point in many different directions. In short, my analysis demonstrates that there are many ways to approach property rights in seeds, in living organisms and in genetic constructs, all of which have a sound basis in the history of European law and political thought. The new edition retains this approach. Philosophical theories of property provide both general and explicit statements of the rationale for deciding legal and moral questions about the status of ownership claims. Debates over property rights in biotechnology began with specific legislative proposals such as the US Animal Patent Act of 1986, and by filing of patent applications for DNA sequences and processes in the early 1990s. While these debates make occasional appeal to philosophical theories of property, moral claims entangle with questions about filing requirements, tests for efficacy, and the rules for licensing and defending patents. Prior to the publication of the first edition of this book, discussions of intellectual property related to biotechnology and genetics tended to review legal mechanisms and to omit discussion of underlying ethical issues (see, for example, Lechtenberg and Schmid 1991; Murashige 1994). Shortly after the first edition appeared in 1997, authors having a considerable background in the law of intellectual property contributed new treatments, though medical, rather than agricultural, biotechnology was generally their focus (see Eisenberg 2003; Barton 2004). Scholarly publications on patenting in agricultural and food biotechnologies exploded at about the same time that the second edition was appearing in 1997, (see Somsen 2007). Interest from scholars gradually turned toward the implications of the Agreement on Trade-Related Aspects of Intellectual Property Rights (the TRIPS Agreement, a part of the WTO) where patents for pharmaceuticals and for GMOs alike stimulated debate, (Strauss 2009;

10 Conceptions of Property and the Biotechnology Debate

253

Prasad et al. 2012). A detailed discussion of this literature would only detract from the primary philosophical orientation of the chapter, however. A very brief context setting discussion has been added to the new edition, preceding the philosophical discussion of property rights.

10.1 Property Rights in Genetic Information: The Context Early critiques of biotechnology’s potential for helping resource poor farmers in less industrialized economies focused on the control of seed traits. Calestous Juma, Pat Roy Mooney and Vandana Shiva, were collectively responsible for initiating a large and growing literature on the moral significance of intellectual property rights (IPRs) on genes, rDNA processes, and whole organisms for developing countries. Calestous Juma (1953–2017) was a native of Kenya and member of the faculty at Harvard University. His 1988 book, The Gene Hunters: Biotechnology and the Scramble for Seeds brought the concept of biopiracy to widespread attention. Juma discussed how agricultural scientists had traveled the world searching for plants with valuable characteristics. They typically found them in farmers’ fields, where local growers were well aware of the value of their seeds. Local growers were not aware that plant breeders could move these traits into new varieties that would be sold commercially, and they typically gave away or sold seeds at the price they might have gotten from someone who simply wanted to make a stew. Juma argued that this practice was immoral, effectively a form of theft, and predicted that it would become a larger issue in an era or rDNA enabled gene transfer, (Juma 1988). Juma’s work brought the importance of maintaining some control over genetic resources to the attention of the international development community, sparking an era of reform in the arrangements for collection of germplasm. He was a sharp critic of the way that Western scientists and seed companies exploit the knowledge of resource poor farmers, but he was not against the development of new plant varieties as such. Neither did he object to the use of gene technology in modifying the genomes of crops that resource-poor farmers would grow. In 2014, Juma published a paper entitled “Global Risks of Rejecting Agricultural Biotechnology” on the Genetic Literacy Website where he endorsed the basic premises of the world-feeding argument (see Chap. 2), writing that, “It is important to realize that developing countries face a separate set of risks from industrialized countries,” (Juma 2014). Canadian author Pat Roy Mooney (not to be confused with University of Kentucky sociologist Patrick Mooney) was a founding member of RAFI (see Chap. 9) and a longtime advocate for smallholding farmers. Working with Food First in the 1980s, Mooney argued that the genetic potential in seeds is a public good, rather than a resource available for capture by private capital, (Mooney 1979). His book with Cary Fowler, Shattering: Food, Politics and the Loss of Genetic Diversity was successful in recruiting many scientists and public officials to the cause. Fowler, also a founding member of RAFI, became the Executive Director of the Svalbard Global Seed Vault, an organization dedicated to the preservation of plant genetic

254

10 Conceptions of Property and the Biotechnology Debate

resources and committed to retaining their status as a publicly available resource. Since 2000, Mooney has dedicated his effort to the production of documentary films with an environmentalist and anti-corporate message. Shiva is, of course, the best known of all anti-biotechnology activists. In the mid 1990s, Shiva’s Research Foundation for Science, Technology and Natural Resource Policy in New Delhi listed twenty-four publications on capture of genetic resources, along with its magazine, Bija—the Seed. Since the publication of the first edition, much of the ongoing work on genetic resources was taken over by the CUTS Centre for International Trade, Economics and Environment in Jaipur, India, which now has offices in several parts of the world, including Washington, DC. Shiva herself has continued to make very broad claims linking biotechnology to a host of ills, (Shiva 2016). A rhetoric of theft and illegitimate taking permeates much of her activism, suggesting that usurpation of property is still a central theme in her work. Perhaps Shiva can be best seen as the person who brought critiques mounted by others to widespread attention. Her 1993 book Monocultures of the Mind links declining biodiversity to biotechnology companies’ pursuit of profit. She followed Juma in defending Indian smallholders’ prior rights in control of a biological resource, especially in connection with a U.S. scientists’ attempt to obtain a patent on functional properties of the neem tree, (Shiva 1997). A later book, Stolen Harvest, repeats these claims but applies the metaphor of theft even more broadly, (Shiva 2000). Prior to the work of Juma, Shiva and Mooney, developed world researchers collected germplasm from centers of diversity that lie in developing countries with little thought to the propriety of doing so. Sometimes germplasm was collected from the wild, but often simply by buying at local markets where beans, potatoes and grain are sold for food. Scientists took the germplasm back to laboratories of the developed world where plant breeders used it to develop improved varieties. This is how the commercial seed industry evolved from the trail-and-error observations of farmers. However, under this model nothing prevented a farmer from buying improved seed one year, growing it out and selling the seed again in the following year. By the late 19th century, seed companies marketed their branded varieties to farmers as much based on the purity and quality of the seed as on the specific traits of the variety. The idea of “plant breeders’ rights” arose to acknowledge and incentivize the innovative step in improved varieties. Plant breeders’ rights were enshrined in U.S. law with the Plant Variety Protection Act (PVPA) of 1970. These rights did afford plant breeders with a form of intellectual property, specifically, the right to prevent the re-selling of seed from the varieties that they had registered. Plant breeders’ rights do not prevent other breeders from building on an improved variety, nor do they prevent farmers from saving seed for their own use. Under PVPA and similar laws to protect plant breeders’ rights, farmers could replant seed from their crops, but could not resell it for use as seed without obtaining a license from the owner of a varietal registration. Furthermore, germplasm (including that of improved varieties) could be used freely to develop new varieties, which could be then be marketed to farmers. In other words, a one breeder could build on the work of another, and could obtain a new variety registration by showing that the new breed had agronomic traits not found in the earlier one. Under this regime, a plant breeder could capture value from

10.1 Property Rights in Genetic Information: The Context

255

plants grown in far-flung regions by crossbreeding them with varieties already in use domestically. Juma, Mooney and Shiva each mounted similar critiques of this practice, arguing that compensation was owed for germplasm incorporated into the improved variety. This argument becomes complex in its details: in some versions, compensation is owed to farmers, who surely did improve germplasm over the course of many generations, while in other versions compensation was claimed by governments in the regions were germplasm was being collected. The rise of biotechnology in food production occurred at a time when political leadership in less industrialized countries were becoming cognizant of the value of their genetic resources, both for agronomic and for pharmaceutical uses. The moral claims being argued by Juma, Mooney and Shiva were not specific to recombinant methods of transformation. They applied equally to traditional methods of plant breeding. Nevertheless, these new claims of ownership had a significant impact on the way that gene technologies were viewed in the context of international agricultural development. As noted, indigenous farmers, their governments or the whole society might claim ownership in genetic resources, though the moral arguments to support each claim would necessarily differ. Other arguments purport to show that no property claims on germplasm are defensible; hence, people in agriculturally less-developed countries need not respect the plant variety registrations and patents awarded in the developed world. These are logically incompatible claims, of course, and one difficulty in applying philosophical rigor to this politically heated controversy is that advocates seem willing to toss out virtually any argument, hoping that it will work. On the other side, trade representatives and representatives of developed world biotechnology companies have been disinclined to view these claims as resting on any serious argument at all, preferring to rely on economic power and the privilege they currently enjoy under the status quo. The international property rights dispute is not an example of ideal philosophical discourse, to say the least. Two philosophical threads might be untangled from this morass of issues, however. One is to examine how various ways of defining and defending claims to property bear on the international property rights issue. That is the main focus of this chapter. The second thread concerns the social impact on the poor. It is evident that advocates of resource-poor smallholders are of the opinion that IPR’s deny them their due rights. On the one hand it may be that these farmers have property rights of their own, even if they have an informal basis or have not been recognized in the courts of the industrialized world. Philosophically, this argument devolves back to the question of how one would establish legitimate property claims over the germplasm. On the other hand, one might think that IPR’s will harm developing country farmers by depriving them of something other than a property right, such as robbing them of some important economic opportunity in the future. The latter possibility is clearly real, for if biotechnology companies develop more productive seeds and place them on developing country markets, the logic of the technology treadmill dictates that those who adopt the new seeds early will benefit, and those who are too slow to adopt them may never get the chance. IPR’s figure prominently in this argument, for it is IPR’s that prohibit entrepreneurial farmers from growing up a handful of

256

10 Conceptions of Property and the Biotechnology Debate

purchased seed and sharing it at no or low cost with the entire village. If better seeds become purchased inputs, the pattern of harm is, once again, fewer and larger farms. Again, if it is marginal, resource poor farmers who are being put out of farming, the consequences of the treadmill may be serious indeed, but this is repeating an argument that has already been made before. The abuse of property rights raises yet another set of issues. Farmer-grown, farmerdeveloped crops are called land races. Articles in Bija—the Seed claimed that farmers will lose the right to freely plant seed from land races or other publicly available varieties (Anonymous 1996). This is thoroughly inconsistent with any of the moral foundations for IPR’s. The legal codes that establish IPR’s in industrially developed countries specifically protect any existing uses of the raw materials from which new seed varieties or plants are derived. Nevertheless, it would be incorrect to conclude, as Western specialists often do, that there are no moral issues here. Legal codes are not always administered fairly in the industrialized world, and in countries where social hierarchy and local power count for much, the situation will be worse. Chapter 9 notes how counterfeit seeds and illegal reproduction of seed has exposed Indian farmers to grave risks from crop failure. A strong system of property rights would protect farmers from such abuses. As such, one must be cautious in laying such failures at the feet of patents or other forms of IPR. As in Shiva’s writings IPR’s are also associated with a line of reasoning that emphasizes the profit-oriented nature of agrifood biotechnology. Devinder Sharma’s 2003 pamphlet GM Food and Hunger: A View from the South provides a good example of the argument. Sharma covers a lot of ground in only forty pages, but a succinct summary of his argument runs as follows: Developed country scientists and biotechnology companies have promoted agrifood biotechnology as a response to hunger, but profit is their sole motivation. Sharma seems to think that the case against biotechnology is proven when the motivations of its developers become clear. Intellectual property figures prominently in the evidence that he assembles to make that case, (Sharma 2003). Sharma’s emphasis on motives is philosophically important because many scholars of IPRs would discount the significance of motive entirely. Greed can drive behavior that would be classified as unethical by virtually any standard, so stressing motivation becomes a way to broaden the argument against profit-seeking technology. Sharma is certainly correct to claim that unambiguously immoral activities have caused many of the actual harms that small farmers in India have experienced in the wake of biotechnology. The extent to which legitimate actors (such as biotechnology firms) have moral responsibilities to anticipate illegal and immoral activities by third parties is a topic for future research on the ethics of emerging technology. Sharma’s work illustrates how claims about property intersect with claims about human behavior, providing an indirect link between the arguments discussed in this chapter and the risk-based framework that structures the book as a whole. IPRs are thus relevant to risk assessment, even if many of the philosophical arguments for defending or criticizing property claims appear to operate in an entirely different intellectual space.

10.2 Ownership in Genetics: A Contested Subject

257

10.2 Ownership in Genetics: A Contested Subject The trajectory of philosophical debates over patenting or other ways of claiming ownership in genes or processes for transforming genomes is convoluted. Competing perspectives emerged early on in the era of biotechnology in connection with the massive sequencing effort of the human genome project. On occasion, arguments that originated out of opposition to the patenting of human genes as spilled over into the debate on agrifood biotechnology. Sigrid Sterckx (1997) argued that a clause in European patent law proscribing patents for immoral inventions applies to products of biotechnology. Philosophers known for their contributions in bioethics have occasionally weighed in on topics of relevance to agriculture and food systems. David Magnus (2002) argued that biotechnology is a means for expropriating traditional farmers’ contributions to genetic resources, and that sanctioning this expropriation with patents awarded to scientists and biotechnology firms is a form of injustice. Lori Andrews (2002) has issued a call for an entirely new way of thinking about the ethical rationale for intellectual property that would address issues in medical and agrifood biotechnology alike. The noted bioethicist Baruch Brody (1943–2018) also reviewed debates over intellectual property in genes and gene processes, (Brody 1989, 2010). All of these studies note ways in which conventional notions of intellectual property fail to jibe with expectations about morally justifiable attitudes to life processes and living things. More recently, Justin Biddle and David Resnik have reinvigorated this line of work by philosophers with new studies, (Biddle 2014, 2016; Resnik 2016). Readers wishing a comprehensive introduction to current debates on property and agrifood biotechnology will want to consult these sources. However, much of the work by philosophers has been motivated by debates over biomedical biotechnology, rather than applications to agriculture and food. Needless to say, this medically oriented work is insensitive to the role of alternative IPRs (such as plant breeders’ rights or registered livestock breeds). Furthermore, when the medical bioethics of intellectual property claims is included, the literature becomes so voluminous that even a cursory review is impossible. The robust debate over the so-called Terminator gene provides a helpful entrée into some of the issues in the realm of food and agriculture. “Terminator” was the facetious term that critics of biotechnology used to describe a family of gene constructs intended to make seeds sterile. Advocates of the technology prefer to call them “genetic use restriction technology” or GURTs. That is, a gene construct performs the function of constraining (or at least limiting) certain uses of seeds. In the Terminator case, the use that was restricted was seed saving by farmers. The practice of saving seeds to plant in the future would be frustrated because these seeds would not establish a robust crop. As with rBST, one can often discern how an author views the case by the terminology used. The boosters called it bovine somatotropin, while the knockers said bovine growth hormone. In order to make an attempt at neutrality, I will henceforth alternate between “Terminator” and “GURTs”. The first and most famous GURT was protected by a patent awarded jointly to the United States Department of Agriculture (USDA) and the Delta and Pine Land Co. in

258

10 Conceptions of Property and the Biotechnology Debate

1998. Delta and Pine was subsequently purchased by agricultural biotechnology giant Monsanto, making the Terminator case emblematic not only of several key issues relating to intellectual property, but also of Monsanto’s clumsy handling of public relations (see, Specter 2000; Charles 2001). Terminator was brilliantly exploited by activist critics of biotechnology, who drew support from academic authors who stressed the irony of a life science company’s attempt to produce seeds that would not reproduce (Berlan and Lewontin 1998; Crouch 1998; Shah 2001; ETC. Group 2002). There are at least four substantive ethical issues raised by GURTs. First, there is the patent itself. The patents on Terminator constructs are patents on gene sequences; hence, there is the basic question of whether genes should be “ownable” at all. Second, there is the way that GURTs effectively make genetic traits “ownable” through a physical, technological means, as distinct from a legal institution, (such as a patent). Historically, farmers can replant seeds from the crops they grow year after year. They purchase seeds once, then genetic traits that are present in the germplasm of the crops they grow will be passed in the next generation of seeds, which can be saved and planted again. Seeds containing GURTs produce a crop bearing seeds that will not germinate, meaning that farmers must buy new seeds every year. Thus GURTs “take” an effective property right on the continuing genetic potential of the crop germplasm from farmers and “give” it to seed companies. If farmers want the improved genetic potential of the crop, they must buy it over and over again. This physical transformation in the control and salability of genetic traits can occur without regard to whether the GURT itself is patentable. The ethical question: is this way of technologically altering the traditional property relationship between farmers and seed companies ethically justifiable? The third ethical issue comes back to biopiracy. As mentioned above, critics of agricultural science have long argued that the developers of land races (or their heirs) have a moral property right in the genetic traits of these crops, even if international law has failed to invest this right with legal force (Mooney 1979; Fowler and Mooney 1990). Though virtually any form of scientific plant development might fall prey to the biopiracy critique, RAFI argued that that Terminator seeds and biopiracy were both part of a larger corporate strategy to exert control over genetic resources, (Shand 1998; Robinson 2010). Vandana Shiva also identified Terminator seed as particularly egregious example of biopiracy, (Shiva 1997, 2000). Finally, do Terminator genes pose unacceptable risks to human health or to the environment? Although this question should certainly be posed for any application of biotechnology, some have apparently envisioned especially catastrophic risks in connection with GURTs: I would like to mention a major environmental risk associated with Terminator, concerning more than one billion poor people whose main food source is based on replanting second generation seeds. The introduction of death genes in crops such as rice or wheat would have a great impact on the fate of millions of people: considering them non-target organisms, the negative impact of Terminator raises to unacceptable levels. (Giovanetti 2001)

10.2 Ownership in Genetics: A Contested Subject

259

This author may believe that Terminator genes will be fatal to people who eat them, though it is more likely that she has envisioned Terminator genes becoming established in food crops beyond the commercial varieties in which they have been intentionally introduced. In fact, both scenarios are equally unrealistic. Plants containing GURTs are far less likely to have environmental impact of any kind than are all other plants precisely because GURT technology dramatically reduces the plant’s reproductive fitness. That is its intended effect, and some have endorsed the use of Terminator type genes as a means to limit the risk of unintended gene flow from transgenic plants (Muir 2001). This does not exhaust the sense in which Terminator seeds might pose risks, however. Clearly, farmers who save the seed progeny of Terminator crops expecting them to perform comparably to the parent will observe a devastating crop failure in the following year. What is more, even normal crops lacking this particular gene construct can be affected by Terminator pollen if planted in the vicinity of Terminator crops. Farmers in the United States, Europe, Australia and other areas where production is commercialized could suffer economic losses, though it is unlikely that they would attempt to save Terminator seed with the intention of replanting it. However, smallholders who save seed or farm in locales where they would be vulnerable to pollen drift might lack the institutional supports that create a barrier between economic loss and starvation. They could suffer a crop failure that translates into a human catastrophe (Pinstrup-Anderson and Schiøler 2000). This risk provides a sufficient ground for opposing the development of Terminator seeds in staple food crops, especially in poor countries where farmers are saving seed for subsistence needs. In fact, no Terminator seeds have ever been released for use by farmers, (Genetic Literacy Project, n.d.). It is worth taking a few pains to emphasize the distinctness of these four arguments. The last concern, that Terminator seeds could be the cause of a local or regional food crisis, is a powerful risk-based argument against the technology. The risk argument provides a persuasive reason to override the claim of seed developers to commercialize and profit from their innovation—a claim that normally accompanies the issuance of a patent. These risks justify bans on GURTs that dramatically reduce the fertility of seeds, absent powerful assurance that negative effects can be contained. Such bans would clearly be justified in settings where smallholder crop failures would be followed immediately by localized famine, but might also apply more broadly. Unless one could show that distribution of GURT protected seed has been carefully controlled, the potential of unintended consequences of a localized but catastrophic nature cannot be dismissed. Monsanto announced that it would not pursue commercialization of Terminator seeds on just these grounds, (Glover 2007). However, it is also crucial to see that this is a risk argument, unrelated to the link between GURTs and intellectual property. A review of debates over Terminator from the late 1990s until 2010 shows would demonstrate how muddled these concerns became in the minds of biotechnology critics. Some apparently believed that all GMOs were GURTs, and the belief that the biotechnology industry was pursing the commercialization of Terminator persisted (Herring 2007). One encounters people who still believe that this is the case today,

260

10 Conceptions of Property and the Biotechnology Debate

(see Yusef 2010 for a summary of the controversy). The other three arguments, however, are more central to the focus of this chapter: On what ethical grounds can property rights in genes or gene processes be sustained in the first place? Questioning the patentability of gene sequences appears to address that question directly, but in focusing the concern on patents and patentability, there is a chance that we may be sidetracked by legal arcana of patent law. There are, in fact, many ways to establish ownership of a good, and the second set of questions focused on technologically created control make this point abundantly clear. If genes or genetic traits are not the sort of things that it is ethically justifiable to own, why should it matter whether the exclusion and control we associate with patents is achieved through technological means? At the same time, it is important to see that while the first two questions involve the ethical or legal basis on which we might claim that someone can legitimately claim to own a gene or a genetic trait on ethical grounds, the biopiracy question is significantly different. On the face of it, this question seems to involve not whether genetic traits can be forms of property, but who has the ethical right to claim them as property. If one argues that corporations and scientists in wealthy countries are unjustly appropriating the property of farmers who created land races, one would appear to have accepted the legitimacy of property rights in genes and genetic traits already. The question is not whether genetic traits can be owned, but who owns them, and under what conditions can they be transferred. The alternative would be to argue that genes and genetic traits are public goods that are unethically and inappropriately placed in private hands when biotechnology companies expropriate them for commercial purposes. However, if this were the view, then the farmers and descendants of farmers who develop and conserve land races would not be in position to claim ownership or demand compensation for the loss of their property. The biopiracy literature seldom recognizes the distinction between these two lines of argument. Indeed the lack of consistency (much less subtlety) in the position of those involved in debate over property rights and agricultural biotechnology is one of its most frustrating features. The balance of this chapter provides a schematic overview of arguments used to establish the ethical grounds intellectual property rights (IPRs), then examines how each of these argument forms might be applied in the case of plant and animal transformation using gene technologies.

10.3 The Theory of Property Philosophers produced an extensive new literature on property and property rights during the last quarter of the 20th century (see Becker 1992). Even a representative summary of this literature is impossible, but most authors distinguish two central ethical questions, as well as two philosophical approaches to the development of theoretically adequate replies (for example Ryan 1984; Goldman 1987). The two ethical questions:

10.3 The Theory of Property

261

1. What counts as property? That is, how are we to understand the concept of property and which sorts of things can and cannot be classified as property, given a particular moral conception of property? 2. Who owns what? How are assignments of ownership to be made? How is the general distribution of property within society to be justified? Recombinant organisms or sequences raise philosophical questions that transcend the categories of risk-based technological ethics in part because they appear to challenge accepted ways of answering the first question. As we have seen, some critics of biotechnology have suggested that it will have a disproportionate negative impact on the poor in developing countries. As noted, the risks of planting Terminator seeds can be distinguished from the question of whether they should be covered by IPRs. However, the Terminator case also illustrates how difficult maintaining such a distinction can be. A schema for classifying the justifications given for property rights can help by showing how health, environmental or socioeconomic risks from the use of a technology can be incorporated into the justification or legitimation of IPRs. More philosophically, separating these two questions illuminates how the first question may or may not be a normative one, depending on one’s point of view. The matter of what can and cannot be property might simply be a matter of fact, determined by empirically observable characteristics of the good in question, or it might simply be a matter of legal convention. The current systems for establishing and enforcing IPRs exist as legal facts. Legal positivists insist on purely descriptive language in analyzing legal concepts and would regard the definition question simply as a matter of ascertaining how property rights are defined and administered in any given society. However, the definition question can also be asked in a purely normative vein: What sort of thing is it morally or ethically legitimate to regard as property? Human beings, for example, were held as chattel property in slave societies. One strategy for opposing slavery has been to argue that regarding humans as property is itself morally wrong, even while the practice of chattel slavery was legal. On this view, the concept of property cannot be applied to human beings without committing a moral wrong. In contrast to legal positivism, this philosophical position on the morality of slavery interprets the question of property status normatively. Debates over slavery may seem far afield from agricultural biotechnology, but they are more relevant than one might expect. Prior to the Human Genome Project (HGP), the ownership of human beings had not been thought to have much to do with patentability. Now, the US Fourteenth Amendment banning slavery has been interpreted to exclude human beings from otherwise applicable aspects of patent law. In 2000, U.S. President Bill Clinton and U.K. Prime Minister Tony Blair issued a joint statement promising that the results of the HGP would be “freely available” citing again proscriptions that were introduced into property law in connection with the end of slavery. There are, thus, historically important ethical considerations regarding property rights that are logically independent of the technical legal apparatus developed to facilitate patents. These considerations establish ethical constraints upon patent law that can become legally binding.

262

10 Conceptions of Property and the Biotechnology Debate

The schema for organizing ethical theories of property consists of nested distinctions. At one level, approaches to property that view it as a convention for achieving further ends differ from approaches that determine the property status of a good in terms of its characteristic features. I will refer to justifications of the first type as instrumental theories of property, and to rationales that pick out characteristic features as ontological theories of property. Instrumental theories treat property as a social, linguistic or legal construct validated in terms of its capacity to produce or secure other ethical goals. The alternative approach treats the property status of an entity as an ontological question, as a feature of some definitive trait or trait that render it fit to be owned and exchanged. That is, whether a good or thing can be claimed as an item of property is thought to depend upon whether it has (or lacks) key qualities and characteristics, on being a thing of a particular kind. Further distinctions are operative within each of these categories. Instrumental theorists can view property rights either as tools for safeguarding political liberty, or as a social apparatus justified in terms of its costs and benefits. Several examples of the ontological approach will be mentioned, but two, natural law and labor theory, will be singled out for discussion.

10.4 Instrumental Conceptions of Property One way to understand property is to see it as a social construction, a mutually agreedupon convention, or a social institution. On this view, property consists simply in the fact that we abide by rules or patterns of conduct in our use or disposal of certain goods. The central ethical question then becomes, are those rules or patterns justified? An instrumental approach to property presumes that these otherwise arbitrary social conventions are validated to the extent that they prove useful in producing or securing some more fundamental kind of good. The philosophical literature identifies at least three types of good that property rights might be thought to produce, protect or secure. One is liberty, a second is social utility or value and a third is social stability. This third line of argument will not be developed in the present context because social instability arising from disputes over IPRs in genes and genetic sequences can be analyzed in terms of social utility. Whether focused on liberty or utility, the instrumental approach to property rights requires an argument to show how property rights serve as tools for securing the more fundamental good. This argument itself has two components. First, there must be some account of the more basic ends (be they liberties or social benefits) that property is thought to protect, to further or otherwise to produce. Second, there must be some account of the link between socially recognized and legally enforced property rights and the more basic end that they are thought to serve in instrumental fashion. Those who have seen liberty as the fundamental good furthered by property rights are libertarians, while those who see utility, happiness, satisfaction or some other use value as the fundamental goods are utilitarians.

10.4 Instrumental Conceptions of Property

263

10.4.1 Libertarian Theory Property rights might be instruments for protecting civil liberties to the extent that freedom of action, freedom of expression and freedom of exchange depend upon the institution of property rights for their effective exercise. A person may feel constrained in his or her ability to produce or enjoy some goods if that person cannot be assured some degree of control over the use of the goods. Many liberties depend upon an individual’s ability to have certain morally important goods at that individual’s disposal, and if the protection of such liberties is thought to be a valid social norm, then recognition of the corresponding property rights will follow. Libertarian political theorists (see Chap. 8) have argued that personal liberties are the most basic political good. In a morally ideal world, people are totally free and unconstrained, but in the real world, we give up our freedom to harm or interfere with others in exchange for the assurance that they will not harm or interfere with us. Thus the fundamental liberties are civil rights, such as a right to assemble, free speech and a right of non–interference in personal affairs (Nozick 1974). Libertarianism is often associated with an ontological argument given by John Locke that takes up the question of slavery, or ownership of human beings. In brief, if property rights are applied to human beings, that will have significant implications for political and economic freedom. Locke’s argument is discussed in more detail later in the chapter. The point here is simply that if property rights are conventions intended to protect freedom, legal systems permitting slavery contradict the basic purpose of property rights. This element in the intellectual history of property rights also explains why libertarians are adamant defenders of the view that social benefits should not override the protection of human liberty. They are especially reluctant to accept the view of the utilitarians, discussed below, that a calculation of social benefits is relevant to the question of when liberties should be protected.

10.4.2 Utilitarian Theory Utilitarian or value–based views are far more predominant in discussions of biotechnology (see OTA 1989). Here, property rights are justified only when they facilitate the creation and allocation of social utility. In the most common philosophical theories, individual preferences are the standard of utility. That is, one good has value in virtue of its being preferred over other goods by individual human beings. For utilitarians, rights of any kind are justified by the fact that they tend to promote the satisfaction of individual preferences throughout the population at large. Property rights are no exception. Thus legal codes governing the use, control or exchange of goods should be evaluated in light of whether the population as a whole experiences greater satisfaction with them or without them. The ‘non–obviousness’ clause in patent law is an example of utilitarian reasoning used to deny a property claim. For the utilitarian, creation of property rights is justified only when they increase net

264

10 Conceptions of Property and the Biotechnology Debate

social value. Allowing someone to claim ownership of ideas or design principles that would be obvious to most people cognizant of general practice in a discipline or trade would only create obstacles to the dissemination of technology and the creation of value. The utilitarian will not sanction all and every appeal for property status, but only those that promise to increase utility.

10.5 Ontological Conceptions of Property Ontology is the division of philosophy that formulates categories or theories for what sorts of things there are (e.g. physical objects, ideas, relations, mathematical objects). Philosophical ontology attempts to account for the general differences in what is (for example, the difference between a physical object and an idea, between a historically existing person and a fictional character). Simply obtaining an internally consistent account of these differences is difficult. An ontological theory of property accounts for what is or is not accorded the status of property by attempting to describe characteristics or traits that make a particular thing ownable, or capable of being a possession. These criteria may refer to specific traits that are either possessed or lacked by the object in question, or they may be purely relational, referring to a relation between the object owned and its owner, or to relations obtaining among a number of people or things. One way to build an ontological theory of property is to start by listing the sorts of things that are treated as property in any given social setting, that is, to treat the ontology of property simply as a project of description. The fact that something is regarded as an item of property is evidence that it can be regarded as an item of property. One might proceed further by asking whether new or unusual goods are in some way analogous or similar to those things that are already regarded as property. Alternatively, it is possible to begin with criteria that stipulate or appeal to a moral, theological, aesthetic or pragmatic standard and to use these standards to establish further criteria for determining the general sorts of things that can be justifiably understood as possessions or holdings, the broadest class of things that can legitimately be said to be owned. Ethical questions about whether a specific object should or should not be classified as property are determined by applying the criteria in individual cases.

10.5.1 Natural Law Theory Natural law is a comprehensive approach to questions in ethics and political theory. In one sense, a natural law is a law of nature, the sort of regularity in nature that is the traditional object of scientific observation and experiment. In the sense relevant to property rights, however, natural laws are ‘principles of objectively right conduct, the rightness of which is immanent in human nature or the nature of things,’

10.5 Ontological Conceptions of Property

265

(MacCormick 1987, p. 275). Over its several hundred year history, natural law theory has embraced patterns of argument derived from many of the approaches that have been discussed in previous chapters, differing from them (if at all) only in claiming an objective ontological status for its fundamental principles. However, the objectivism of natural law theory is less relevant to the current chapter than is an approach to a general definition of property that is not captured in other approaches, and especially in instrumental approaches. As noted, natural law theory presumes that what is natural is, in a deep sense, what is right. The idea that property is a component of natural law has been influential in European history. Such a belief is particularly plausible when one’s concept of nature includes a benevolent, but also judgmental God, who has designed the fixtures of the earthly realms in accordance with His plan. Given such a theology, a natural theory of property may include an attempt to ascertain God’s intentions as revealed in the characteristics of things commonly regarded as items of property. These would certainly include personal effects such as clothing, or common goods routinely bought, sold or transferred by gift. The natural law tradition can be interpreted as starting with these common items of property and attempting to discern characteristics that could be applied to other cases, including genes and gene processes. Rivalry, for example, refers to whether it is possible for more than one person to use or consume the good without diminishing the amount of good available for others. Goods such as canned food and clean water are rival; goods such as street lighting and national defense are non–rival. A second natural characteristic is how easy it is to exclude others from using or consuming a good. Canned foods are relatively excludable in that one may lock them up, preventing their appropriation and use by others. By contrast, it may be difficult to exclude people from access to water or street lighting. A third natural characteristic is alienability: the ability to separate a good from someone’s person for purposes that include exchange. Because the U.S. Declaration of Independence begins with praise of rights to life, liberty and pursuit of happiness, Americans have come to think of ‘inalienable rights,’ as something like ‘supremely important rights’ but Webster’s Third New International Dictionary defines ‘alienability’ simply and unambiguously as ‘the capability of being transferred to other ownership’. The right to use a good such as land or water, for example, can be transferred by sale or gift. Other rights cannot be transferred from one person to another without being vitiated or rendered null: my right to life or religious liberty cannot be meaningfully transferred to someone else, for example. Each person must have his or her own inalienable (inherently non–transferable) rights secured in a just society. This is arguably Thomas Jefferson’s exact point when he wrote The Declaration of Independence, and the reason why he did not include the right to property as an inalienable right (Wills 1978). Natural facts about alienability, excludability and rivalry provide one way to decide whether something can be claimed as property. Goods that are naturally rival, excludable and alienable are easily defensible as items of property. Goods that are highly non–rival and non–excludable are not natural candidates for property (Thompson et al. 1994, pp. 202). These three traits leave considerable gray area where the relative rivalry, excludability and alienability of goods do not provide the basis for

266

10 Conceptions of Property and the Biotechnology Debate

a secure judgment. In such cases, a fourth element of natural law theory may emerge which treats all of nature as a heritage to be shared equally by all human beings. John Locke was a staunch defender of property rights for the emerging seventeenth– century English middle class, but even he recognized that all must share nature. Such a principle for deciding property claims would accept that highly rival, excludable and alienable goods are ‘fit’ to become property, but would decide the gray cases in favor of a non–property or common–property determination. Chief Justice Warren Burger (1907–1995) wrote the majority opinion for the US Supreme Court decision in Diamond vs. Chakrabarty, which appeals to such a view implicitly. The Court held that Ananda Mohan Chakrabarty (1938–2020) deserved a patent for his bacterium because it was his own handiwork, and not ‘a manifestation of nature, free to all men and reserved exclusively to none,’ (US Supreme Court 1980, italics added).

10.5.2 The Labor Theory of Property However, ontological theories need not appeal to natural law. In theory almost any trait might be stipulated as a criterion. In an early paper on biotechnology and property rights Ned Hettinger proposes criteria that would challenge the property status of any living thing and would categorically rule out all sentient life forms (Hettinger 1995). Hettinger appears to claim that the simple property of being a living thing disqualifies a plant or an animal as a candidate for being someone’s property. Interpreted literally, it would mean that the lettuce and tomato sandwich in my lunch bucket really isn’t mine, after all. It is likely that he intends something weaker, such as the claim that patent holders should not be able to extend their IPR to living (sentient?) organisms bearing the genes covered by a patent. The strong version of Hettinger’s criterion is unlikely to win wide acceptance, denying as it does property status to domesticated animals and challenging well–established chattel property rights. Nevertheless, it serves to illustrate how alternative criteria might be proposed. One might consider explicitly theological criteria, or criteria that test for autonomy or rationality as alternative developments of the ontological strategy. One important alternative is the labor theory of property, also derived from John Locke. A labor theory of property holds that a person’s productive work is the basis for a property claim. People are entitled to claim what they make or create as their own. The mere act of discovery does not establish a property claim, but the appropriation of the discovered good to some further purpose does imply some element of labor. As long as previous property claims upon the appropriated good are discharged fairly, the work that a person does in picking up, transforming or safeguarding the appropriated good establishes a property claim. In standard applications, ownership of goods produced while in the employ of another person or organization are, subject to prior negotiations, transferred from the laborer to the employer as a consequence of the wage or salary contract. A labor theory of property is not to be confused with the labor theory of value developed by early economists, including Adam Smith (1723–1790). The labor

10.5 Ontological Conceptions of Property

267

theory of value made dubious claims linking the legitimate value (or justifiable price) of a good to the value of the labor expended in producing it. Rejection of this theory was a critical feature in the move from classical economic theories, such as those of Smith and Marx, and the neoclassical approach that dominated the 20th century. However, claiming that the ownership entitlement is established by the investment of labor in the good’s appropriation, creation, manufacture or development entails nothing about its economic value. If value is determined by exchange, as neoclassical economists assume, it is clearly possible to invest substantial amounts of labor into items that are of no value whatsoever. A labor theory of property would nevertheless support the claim that such valueless items are the property of their manufacturer irrespective of whether they have exchange value or social utility.

10.6 Linking Instrumental and Ontological Theories of Property Most theorists of property have had little interest in developing philosophically pure approaches to their subject matter. Instead, different approaches to property are often mixed. When a particular set of legal criteria for property can be shown to satisfy several different philosophical approaches, that is a mark in favor those criteria. There are two important strategies for linking these argument forms. One ties natural law’s emphasis on rivalry, excludability and alienability to the utilitarian focus on creating social utility, while the second ties the central criterion of labor theory to the libertarian interest in protecting human freedoms. Linking natural law and social utility. Rivalry, excludability and alienability dramatically affect the costs and benefits of any individual or group’s attempt to control the use of a given good. Attempting to control a highly non–excludable good is costly, and little benefit can be derived from trying to control a highly non–rival one. To the extent that this general pattern holds, relative degrees of rivalry, excludability and alienability will track the particular configuration of social rules that tends to promote optimal social utility, the best ratio of social benefit to social cost. However, a systematic departure from the pattern becomes crucial to the debate over intellectual property rights. Ideas and innovations have the potential to create social value, but since they are non–rival and poorly excludable, advantages to the creator or innovator are nullified when all share in the benefits of the innovation. Lacking a socially constructed and legally enforced right to the innovation, the only way for an innovator to profit from ideas is to keep them secret. Utilitarian conceptions of property were the impetus for widespread development of patent offices in the eighteenth and nineteenth centuries. It should not, then, be surprising that biotechnology and intellectual property are often discussed in terms of a utilitarian or value–based approach. The creation of social value is the intellectual rationale for utility patents, especially in the United States. Hence, demonstrating the need for incentives to develop and disseminate biotechnologies has emerged as the key burden of proof in

268

10 Conceptions of Property and the Biotechnology Debate

patent oriented debates (see Lesser 1989). The natural law account stressing rivalry, exclusion cost and alienability will coincide with the utilitarian aim to promote social utility in many cases. Linking labor and liberty. As noted, libertarians have strong arguments for protecting existing property rights, but weak arguments for saying where these rights came from in the first place. The argument linking labor to liberty remedies this problem and is straightforward. A system of property rights that failed to recognize a person’s property right in their own labor would compromise liberty by consigning people to effective servitude. There is little point in insisting on human liberty if the products of a person’s labor can be arbitrarily appropriated without consent or compensation. This argument requires a slight modification of the labor theory of property developed by John Locke, because it stresses not the self-ownership of each person, but each person’s initial ownership of their own labor. While Locke thinks that one’s self-ownership should not be regarded as an alienable good, the product of one’s labor clearly is transferable to another person. In fact, leading libertarian theorists do stress the claim that that recognition of property rights in labor is necessary for the protection of liberty in just this way. Labor needs to be salable or alienable, in order to make it possible for someone to work for wages. Clearly, one way to earn a living is to produce things (buggy whips or bushels of corn) that can be sold to others, but many (if not most) people in contemporary society sell their labor. They agree to work mowing lawns or making buggy whips. The wage they are paid reflects the local market for labor, the wage or salary for which comparably productive people are willing to work. To deny people the opportunity to enter into contracts with others either for the buggy whips they have made or directly for their work in the form of wages would restrict individual liberty, (Hospers 1971; Paine 1991).

10.7 Property and Agrifood Biotechnology: Ontological Approaches How do each of the several conceptions of property point in different (though not necessarily contradictory) directions when applied to questions in biotechnology? One reason why it is difficult to say anything definitive about property rights for biotechnology is that each of the conceptions of property developed over the centuries are now subject to forms of interpretation that differ substantially from those of the past. In the present context, it is less useful to strive for conceptual purity than to see how key concepts might be interpreted and combined to form a rationale for evaluating biotechnology. One clarifying task can be achieved by stressing how ontological claims can be differentiated from utilitarian ones. However, the ubiquity of utilitarian argumentation in the biotechnology debate has a polluting effect that makes it difficult to perceive how non–instrumental criteria might be applied.

10.7 Property and Agrifood Biotechnology: Ontological Approaches

269

However, another reason for the difficulty of making definitive judgments is that the products and processes of biotechnology are themselves very different. At first, genetically altered organisms were at the center of debate, with considerably more emphasis on animals than plants. A further controversy emerged over US National Institutes of Health (NIH) filing of patent claims on various and sundry fragments of cDNA. While this action was widely criticized at first, the action was defended on the ground that legal procedures in the United States entail that failing to file effectively eliminates the opportunity for NIH to claim rights in the future, while leaving open opportunities for private companies to do so. What is philosophically interesting in the case is the likelihood that criticism of NIH reflects a widely held opinion among the scientific community that the sequences under consideration should be understood as discoveries, rather than as inventions (Anderson 1991; White 1994). Without implying anything about how patent officers or courts might adjudicate patentability under existing law, it will be illustrative to consider how each conception of property might be applied both to whole organisms and to fragments of genetic code. Natural criteria for property survive into the present in a form significantly altered from their application in natural law. In the first instance, the theological warrant for property has all but vanished, with theological arguments offered most commonly to limit, rather than promote, the application of property claims. Thus, the new strategy rejects Locke’s original judgment that all things, including human beings, are covered by property rights. The current practice applies ontological criteria to make the normative claim that some things should never be considered as property, at all. Human beings are the paradigm example of a non–property good. From this starting point, at least two rather different strategies for applying natural criteria are available. One stresses analogy to the human case, the other stresses rivalry, excludability and alienability. The application of labor criteria is more straightforward. Since scientists are people, don’t they, too, own the products of their labor?

10.7.1 Ruling Out Ownership of Human Genes One way to arrive at the conclusion that human beings cannot legitimately be understood as property, even as property reflexively owned (in Locke’s sense), is to argue that the concept of property implies a status of subservience that is inconsistent with certain natural facts about human beings. This view starts from the observation that humans are free and autonomous agents, acting in pursuit of rationally chosen interests. Given this orientation, regarding oneself as one’s own property might be self–contradictory, since one could consider the potential use or sale of oneself as a potential means for realizing those interests. Many readers will recognize the form of this argument as Kantian. While it might still be possible to exchange labor for other goods on a Kantian view, the autonomous agent that is at the core of the Kantian conception of the person could not be considered as a means to any rationally chosen end because entertaining such an idea would be to regard oneself as (potentially)something other than a rational being. As already noted, arguments of

270

10 Conceptions of Property and the Biotechnology Debate

this sort have surfaced with respect to property claims over the human genome, but how is this relevant to agrifood biotechnology? Recent attempts to extend this notion of personhood to nonhuman animals entail that ownership of any subject of a life, to use the phrase favored by Tom Regan (see Chap. 5), cannot be justified on ethical grounds. As sketched by Hettinger in an early draft of his 1995 paper, this view extends to any transgenic animals that also possess requisite moral characteristics such as consciousness and a consistent mental identity over time, (Hettinger 1992). However, the pseudo Kantian argument applies much more readily to individual animals and to human beings than to the products of genetic engineering. It is, after all, individual human beings who possess autonomy, rather than the species as a whole, much less a segment of code from the human genome. The argument could be applied to a case in which an individual’s rights were compromised by experiments that extract or derive genetic technologies from samples of that individuals’ DNA. Here, the individual in question might have been treated like property to the extent that others want to claim ownership of something uniquely derived from his or her body. Rebecca Skloot’s study of cells derived from Henrietta Lacks (1920–1951) illustrates the problem, (Skloot 2010). The case is not without ambiguity, however, for body products such as whole blood and semen are bought and sold in many countries, including the United States. A Kantian modification of natural law that would rule against products of genetic engineering would appear to have an even stronger application to these more routine cases. Furthermore, the extension of this argument to plants or animals depends upon the controversial extension of Kantian arguments to non–human organisms. Nonetheless, something like this argument appears to surface in the thinking of many people who oppose intellectual property rights in genes and gene processes. Michael Fox, for example, expressed the view that ‘the patenting of animals reflects a human arrogance towards other living creatures that is contrary to the concept of the inherent sanctity of every unique being and the recognition of the ecological and spiritual interconnectedness of all life,” (USHR 1988, pp. 64–65). Andrew Kimbrell believes that allowing patenting of plants or animals opens the door to patenting of human genes. He describes a ‘two decade long slippery slope’ in which apparently narrow decisions on the property status of genes and gene processes have laid the groundwork for what he regards as objectionable claims, based largely on their applicability to human genetic materials, or, he would argue, actual human beings, (Kimbrell 1993, pp. 188–202). To some extent, Fox’s and Kimbrell’s views are derived from considerations discussed in Chap. 5 on animals, or in Chap. 10 on religious beliefs. In either case, however, it is not clear that their opinions reflect arguments that uniquely address the moral status of intellectual property. Instead, Fox, Kimbrell and other critics in this vein seem to apply an argument form takes the wrongness of transgenic technology as a premise. From this, they conclude that recognizing property rights related to transgenic is also wrong. This may be a respectable form of argumentation, but it is important to see that more fundamental views about the ethics of gene technology do the real moral work here. The putative conclusion— that property rights in products of gene technology are wrong—is not in itself deeply interesting.

10.7 Property and Agrifood Biotechnology: Ontological Approaches

271

10.7.2 Rivalry and Excludability An interpretation of natural property criteria that stresses properties of rivalry and excludability offers norms that are more applicable to biotechnology. In this view, the property rights would be recognized to the extent that natural features of excludability and rivalry are present. Ownership would be limited to that which could easily be controlled by virtue of its physical characteristics and property rights would primarily protect against common forms of theft. Such a view favors chattel property rights, or ownership of a specific individual, but provides strong grounds for rejecting all intellectual property rights. Biotechnology might even be used to engineer rivalry and excludability into certain organisms, by introducing and eliminating traits that affect reproduction or uses that deviate from intended purposes, as the discussion of Terminator and GURTs at the beginning of the chapter illustrates. An GURT might, for example, increase the rivalry of a hen that lays golden eggs by engineering traits that would preclude her being used for fried chicken. Such strategies would not, however, protect others from reverse engineering any organisms they legitimately could acquire. As such, they do not protect intellectual property, as that term is typically understood. Indeed, the Terminator case is precisely a case of this general kind. Terminator seeds are rival in way that ordinary seeds are not. Ordinary seeds can be grown for food or they can be grown to produce more seeds to plant next year. Although one cannot, of course, both eat and replant a particular seed, the crop in the field seen from the farmer’s perspective can be allocated to either or both of these purposes. They are non-rival, though in either case the seed is consumed in use. The Terminator crop, in contrast, cannot be used as next year’s seed. Growing for food and growing for seed have become rival uses, and farmers must decide which purpose they have in mind before they buy seed. Note, however, that while GURTs are very effective at protecting seed from being used to produce even more seed, they are absolutely useless from the standpoint of protecting one’s investment from a competing seed company. Without the additional protection of a patent, competitors can reverse engineer the GURT and incorporate the genetic trait into a crop variety of their own. The analysis thus far suggests that truly intellectual property is not justified by the natural law standard, but that physical or technological transformations of rivalry or exclusion cost might well be. Adding the criterion of alienability into the mix may favor property rights in genes and sequences, since it is the technology of rDNA that makes these items alienable from the cells and molecules in which they occur naturally. Nevertheless, this interpretation of natural law is particularly important for the biotechnology debate because it provides the most obvious foundation for those who wish to ground property in something natural and to question biotechnology and genetic engineering in particular, in light of its alleged unnatural character. As already noted, these extensions of natural law are very much at odds with many current practices, but either Kantian or rivalry/excludability interpretations are more plausible than classical views that relied heavily on theology.

272

10 Conceptions of Property and the Biotechnology Debate

10.7.3 Do Scientists Own Their Labor? If natural law provides the strongest argument against property rights in genes and gene processes, perhaps labor criteria establish the strongest and most plausible claim for property rights in biotechnology. There can be no denying that transgenic organisms and even fragments of code become available to us as a result of a great deal of labor. This labor is both intellectual and physical, though perhaps not as physically onerous as that involved in clearing and improving land. If labor establishes a claim upon a parcel of land, it should also establish a claim upon the fruits of biotechnology research. There are, however, important qualifications. In Locke’s examples, labor establishes a property claim through working land or gathering apples. These activities are different from intellectual discovery and design in several important respects. They are processes of physical production and consumption. They have tangible goods as their object, and most importantly, land and apples already have characteristics of rivalry and excludability. In these examples, labor can be seen as a process of alienating these goods from their natural surroundings. Once alienability has been added to their natural rivalry and excludability, property claims can be readily justified in natural law terms. One cannot, therefore, say that Locke thought of his labor criterion as distinct from natural law criteria. Given the fact that intellectual goods fare less well on natural law grounds, it is less clear that a property right to a discovery, particularly an intellectual discovery, or an idea can be justified by the labor criterion. There are, however, several different shadings that can be given to the argument. One may answer the query as to whether intellectual discovery involves the right kind of labor in different ways—producing diametrically opposing results. Locke’s view of labor can be interpreted strictly in terms of alienation of rival and excludable goods from nature. Since ideas, designs, and, (more relevant to our purposes), genes or gene processes are neither rival nor excludable, no intellectual property claims in them are ethically justified. The alternative way to read Locke is that any failure to recognize a claim of ownership in the product of any labor is an interference in personal liberty; hence, all intellectual property claims based on the labor of the researcher are justified. In between, it is possible to argue that the relevant sense of labor involves transformation, not simply alienation, of goods existing in nature. This view might permit the argument that a transgenic organism or a process for isolating or manipulating genes will be defensible as property, while the fragment of code will not. The key to such an argument is the claim that something has been produced in making a transgenic organism, while something has merely been discovered in identifying the sequence. This claim is itself subject to nuance and alternative interpretations. Physicists, for example, must produce quarks in order to discover them. Is the situation similar for genetic sequences? Labor and natural law criteria provide philosophically powerful insights into the way that we think of what is ownable and what is not, but the central concepts in these philosophical approaches are open–ended. They are subject to many and subtly different interpretations. Therefore, they provide the basis for extended philosophical

10.7 Property and Agrifood Biotechnology: Ontological Approaches

273

and moral argument about intellectual property rights in biotechnology, rather than definitive answers. Readers looking for such answers may disappointed, but in truth all that can be done is to sketch some strategies for applying the basic principles of labor and natural law theory to biotechnology. This may be useful as an aid to following the thinking and argument of those who engage in this debate, and to forming an opinion, but it is just incorrect to imply that these approaches to property theory truly decide the issue one way, rather than another.

10.8 Property and Agrifood Biotechnology: Utilitarian Approaches A frankly asserted and undefended utilitarianism frames much of the debate over biotechnology and patents. The lack of appreciation for other ways of rationalizing property rights leaves little room for compromise or appreciation of alternative approaches. However, it is important to note that there are other ways to interpret the consequentialist, social choice feature of utilitarianism. Even staying within the assumption that property rights are conventions legitimated by their consequences or outcome, libertarian philosophy justifies these contractual agreements in terms of their impact on liberty. What is more, it is possible to reinterpret elements of the labor and Kantian account in libertarian terms. In both cases, the justification of intellectual property rights for genes, gene sequences, and genetic traits or processes has less to do with what biotechnology is (e.g. its ontological status) than with the social effects of recognizing property rights to the goods in question. What is more, these outcome-based criteria can be applied to the legal institution of property rights (akin to rule-utilitarianism) or to each individual case (act utilitarianism). Recognizing property rights to a key sequence or gene transfer process for an agronomically important crop like maize or soybeans will have vastly different consequences than making a similar judgment for a minor crop such as rutabaga, for example. Viewed in strictly ethical terms, these quantitative differences might be crucial to whether one would want to recognize intellectual property claims on either utilitarian or libertarian grounds. Although utilitarians tend to assume that it is the general institution of patenting that they are defending, there are reasons to support a case-by-case application of the utilitarian injunction to maximize social utility. Discussed above, the utilitarian argument justifies intellectual property rights by claiming that they create incentives to invent and share products that are socially beneficial. The persuasiveness of the argument depends upon whether this claim is true. Here, it matters a great deal whether the market structure for seed and production of a given crop is diverse and competitive, or small, highly integrated and non–competitive. Will the intellectual property right increase or stifle competition in that industry? The question cannot be answered in general terms. The consequences will depend both on the intellectual good in question and on the market structure at a given point in time.

274

10 Conceptions of Property and the Biotechnology Debate

It might be thought that a different question could be answered, though. Can it be said that, on average or for the most part, recognizing a general system of intellectual property rights will tend to promote the social good? Advocates of intellectual property rights have thought that it would. A 1991 review of intellectual property rights in agricultural biotechnology concluded with the statement that broader recognition of such rights would promote agricultural research, and that ‘This increase in research and greater interactions among scientists will rebound to the gain of society,’ (Lechtenburg and Schmid 1991, p. 105). Baruch Brody concludes his review of ethical arguments on animal patents by observing that ‘the claim… that a patenting system promotes beneficial consequences by providing an incentive to create useful inventions… is the most widely used argument by proponents of patenting transgenic animals,’ (Brody 1989, p. 151). He also notes that the structure of this argument makes it ‘difficult to assess in particular cases, for it is often hard to tell how desirable will the outcomes be and how likely they are to occur’. Yet he concludes that the general experience of developed countries with patents provides a reason for finding substantial moral support for patenting transgenic animals, a conclusion that would, one presumes, extend to all forms of food biotechnology. More than thirty years after Brody made that assessment there is reason to be less sanguine. Before products were in the field, it seemed reasonable to think that strengthening of intellectual property rights for biotechnology helped seed companies and food biotechnology firms attract venture capital (Berghorst 1991) and gave large chemical and pharmaceutical companies more confidence in developing new products in face of regulatory uncertainties, (Young 1991). In hindsight, these presumption rested on a narrow basis of research. Whether the growth of intellectual property rights has truly benefited food and agricultural research appears to be very much a matter of perspective. Those who have been successful in obtaining patents have benefited, often indirectly from enhanced status and better internal funding from their home institutions rather than from the patent’s earnings. Yet for others, research has become almost prohibitively costly (Overhauser 1994; Sederoff and Meagher 1995), and most universities found that an intellectual property office that returns more than 5% or 10% on its own operating costs is doing quite well (Haussler 1996). These concerns about gene patenting’s impact on the research process are reiterated for gene editing, (Egelie and coauthors Egelie et al. 2018). In addition, patents are occasionally sought not to commercialize a discovery, but to prevent competitors from pursuing a competing line of research, (Malshe 2018). Finally, the legal costs of both protecting one’s patents and of negotiating the thicket of overlapping patents in some areas of biotechnology have proven to be so substantial that they have brought the incentivizing assumptions of the utilitarian argument into question (Galini 2017). Over time, the enthusiasm for patents in agrifood biotechnology has worn thin. One of the key issues has been freedom to operate, a technical dimension of patent law limiting both future research and dissemination of inventions that would reasonably be expected to infringe upon an existing patent. The problem was dramatized by the discovery that Ingo Potrykus’s approach to vitamin A enriched rice (so-called Golden Rice) was potentially in violation of over seventy patents. At first it appeared that not only would Potrykus need all these permissions to procede with Golden

10.8 Property and Agrifood Biotechnology: Utilitarian Approaches

275

Rice, but the non-profit organization for which he worked could also be sued by each of these patent holders for failing to obtain permission simply to conduct the initial research. Some sources accused Potrykus of entering an agreement with AstraZeneca Corporation to develop Golden Rice largely to shed this liability (RAFI 2000). Can anything be concluded about the ethics of utilitarian arguments for intellectual property rights to food biotechnology? First, one must admit that the general form of the utilitarian argument for intellectual property rights has a long and distinguished pedigree and that utilitarian arguments have indeed been very influential in setting public policy. Second, there is a distinct possibility of reasonable disagreement about whether the empirical evidence supports the extension of this distinguished tradition to food biotechnology. Lacking sounder empirical support than appears to be available, Brody’s conclusion favoring intellectual property rights on utilitarian grounds seems too strong. The jury is still very much out on this question, and the longer the jury stays out, the less compelling the utilitarian argument sounds. The uncertainty that pervades assessment of the actual consequences of intellectual property conventions provides a basis for taking alternative ethical arguments all the more seriously.

10.9 Property and Agrifood Biotechnology: Libertarian Approaches The general claim of the libertarian view is that if an individual’s labor originates the property right, then appropriation without consent violates that individual’s civil liberties. As with the utilitarian view, the first step in assessing whether the libertarian argument is applicable to research in agrifood biotechnology is to ask whether the claim is, in fact, true. Here, the questions reviewed above in connection with a scientists’ ownership of their labor come up once again. In particular, though the intellectual researcher is as entitled to own the immediate fruits of his or her labor as any day laborer is, this entitlement does not establish the terms on which publication or dissemination will take place. Such terms are not difficult to divine for excludable, rival and alienable goods, because any uncompensated use of the goods deprives the laborer of a competing use and clearly compromises the laborer’s liberty. Intellectual goods—ideas—are different. As long as the researcher decides to keep his own counsel, they are perfectly excludable. No one else knows they exist. Once public, however, they are notoriously non-excludable and non–rival. The problem is not so much with the vulnerability of researchers’ liberty as with the conditions under which it would be reasonable for them to consent to general publication. In the laissez–faire system of non–interference rights proposed by libertarians, ordinary labor contracts represent a negotiated settlement between the labor and the capitalist or final owner of the alienable good, and the voluntarily accepted wage represents the compensation necessary to secure consent. Such contracts might well differ with respect to intellectual goods, however. The intellectual laborer knows that

276

10 Conceptions of Property and the Biotechnology Debate

upon publication, the intellectual good is both non–rival and non–excludable; hence, he or she must negotiate not merely with one eventual owner or user of the good, but with every person in the society who is likely to use the good prior to publication. Presuming that such negotiations could be carried out, people in the society are likely to agree to such terms, since such an agreement may be the only way that they will get to use the good at all. People will not, however, agree to rights and licenses controlling knowledge that is easily obtained. One person might pay for knowledge about a short cut to the airport, for example, but it is unlikely that everyone in society would be willing to recognize the exclusive right of any individual to such knowledge. Judgments about the novelty of the relevant knowledge will therefore become part of the negotiations. Such negotiations are likely to prove time consuming and expensive, however, and it can easily be imagined how a system much like patent law would arise to standardize the problem of assigning rights and licenses. Viewed as a socially negotiated procedure, a patent system solves the problem of negotiating with everyone prior to publication and provides the intellectual laborer with the option of seeking compensation when the negotiated rights are violated. The broad implication of libertarian theory is, thus, a strong property right in biotechnology on the part of the individual innovator and a socially negotiated system of making contracts for the exchange and use of innovations. Novelty, however, is one of the criteria used in making patent decisions and it is worth noting well that it rests more securely on libertarian/consent standards than on utilitarian ones. Libertarian or consent–based approaches to intellectual property, thus, end in providing a more straightforward prima facie justification for intellectual property rights than any other view. There are, however, two additional qualifications. First, it may be questioned whether any idea in food biotechnology (or science) generally can truly be traced to the labor of a single individual or even to a specific group of individuals. Science is a social process and scientific ideas are plausibly conceptualized as the result of a long and very public course of development. To the extent that discovery and scientific innovation are conducted in the public sphere, the basis for the libertarian rationale begins to evaporate. Second, although the system of utility patents and its stress on novelty provides one account of the conditions under which people might consent to a system of intellectual property, it is hardly the last word on this subject. Researchers might consent to publication of their ideas for little more than enhanced status and public recognition, for example. (Indeed, that would appear to be the standard that has governed science for several centuries.) The suggestions made in this chapter describe how one might begin to work out a consent–based system of intellectual property rights, but they do not provide a full accounting of the philosophical issues that would be relevant in the final analysis.

10.10 Against Property Rights in Food Biotechnology

277

10.10 Against Property Rights in Food Biotechnology The above sections show that the philosophical case for recognizing intellectual property rights in genes, sequences and genetic processes is mixed, with no decisive arguments either way. Yet there is a fairly large literature opposing intellectual property rights in GMOs, gene processes and seeds produced using gene technology. A number of books have already been written on the subject and the literature has grown steadily since the first edition of this book. As noted in Chap. 9, many presumptions on which criticism of intellectual property rights have been based are simply wrong. The furor over the Neem tree provided a dramatic example that anticipated many of the ambiguities associated with Terminator seeds. Critics claimed that patents granted to W.R. Grace for an insecticidal extract from the Neem plant would preclude Indian farmers from their traditional use of Neem (Anonymous 1995). Similar claims have been made about attempts to patent substances used by folk healers in India (Anonymous 1996). Shiva’s writings on biopiracy have continued this tradition, making India a hotbed of opposition to biotechnology (Barooah 1999). The critics’ arguments are plagued by ambiguity, however. Patents never provide a legal basis for challenging uses of a substance, good or product that were established before the patent was applied for. As such, the idea that patent laws could be used to provide a legal basis for companies to prevent a farmer (in India or elsewhere) from doing what they have always done is just factually incorrect. This is not to say that intimidation and extortion in the misuse of patents never happen, only that there is no legal basis for it. Representative versions of several arguments against intellectual property rights will be reviewed in this final section of the chapter. All either fail to make their case, mostly by begging the central question: are intellectual property rights for food biotechnology ethically justifiable in principle? In many cases, they beg the question because they fall victim to one of three ambiguous concepts. Many fail to distinguish between the consequences of the technology and the consequences of intellectual property rights. Others stumble over the very ideas of ownership and the commodity form. As Milligan and Lesser (1989) noted in their review of the US debate over animal patents, most critics fail to address the central question because they use arguments that oppose the technology to oppose property rights in the technology. Intellectual property rights can be awarded for products (take chemical weapons, for example) that we may hope will never be used. In fact, owning property rights in such technologies may be instrumental to controlling and limiting their use. There may thus be legal grounds for recognizing a property right even when there are also compelling arguments to restrict a technology’s use. As noted above, patents can be used to prevent use. Arguing that a technology is risky, harmful or downright evil is thus not in itself an argument against patents or other forms of intellectual property rights material to the technology. Nor is it clear that food and agricultural biotechnology would be stopped or even substantially slowed by an absence of such rights. It is, after all, the seeds and food products that will eventually determine the profitability of food biotechnology, and intellectual property rights are at best tangentially related

278

10 Conceptions of Property and the Biotechnology Debate

to that. Milligan and Lesser acknowledge the legitimacy and importance of these issues, yet question whether these arguments bear on the question of ownership. Here is a quick inventory of critics who conflate the case against using a technology with reasons to disallow property rights. Cary Fowler and Pat Mooney identify a series of potential risks to genetic diversity and smallholder wellbeing that might be associated with any agricultural technology that promotes monoculture and industrialization (Fowler and Mooney 1990). Jose de Souza Silva describes the long history of unequal exchange between North and South in his critique of IPRs for products of gene technology (de Souza Silva 1995); he does not, however, indicate how changes in the recognition of intellectual property claims to genes or genetic resources will affect the trajectory of that history. Vandana Shiva notes the importance of maintaining both genetic and cultural diversity (Shiva 1997, 2000), but aggressive pursuit of developing country markets for agricultural inputs such as seed, fertilizer and mechanical farm equipment can proceed in the absense of intellectual property rights protecting these technologies. Just as Milligan and Lesser note, these comments seem more directed at the consequences of using the technology than at the consequences of protecting it with intellectual property rights. In fact it really seems to matter little whether the industrialization that these critics fear comes about as a result of one well–capitalized firm promoting a single technology for which it holds an intellectual property right, or several well–capitalized firms promoting similar technologies for which none of them hold intellectual property rights. Yet if nothing is done to regulate technology or to control the economic power of capital, that is what the difference between recognizing IPRs and not recognizing them comes down to. With such regulation of socio-economic impacts in place, the consequences for environment and distribution of wealth are likely to be very different, even when intellectual property rights for gene technologies do exist (or at least as far as we can tell). As their own arguments show, industrialization has occurred in the past without the need for intellectual property rights. There is, however, a deep basis for this confusion. As the Terminator case illustrates, changes in technology, and especially biotechnology, themselves affect the excludability, rivalry and alienability of goods. Ideally, a farmer might want a crop that grows only in the microclimate of his or her particular farm. That would be perfect natural excludability, but most farmers enjoy some degree of excludability from the fact that only farms having similar soil or climate characteristics can produce the crop. Farmers working in this group of farms have a collective natural property right in the crop. When a biotechnologist (or even a traditional plant breeder) transforms seed so that a crop can survive in colder climates, or can resist the pests and diseases of the tropics, this change reduces the natural exclusivity built into the characteristics of the crop. Genetic engineering makes even more obvious and dramatic changes in alienability. In learning how to move genes from crop to crop, the genetic engineer turns what was once an inalienable trait of the crop (the gene) into an alienable good (hence a candidate for ownership) itself. If one takes a strong natural law view of property, it is possible to interpret these technological changes as changes in property rights. This raises an interesting and underappreciated issue for the philosophy of technology. Yet note that even if one

10.10 Against Property Rights in Food Biotechnology

279

wanted to pursue this argument, it would have little to do with the putative topic of this chapter, namely intellectual property rights. The question of whether it is ethical to transform the physical characteristics of excludability, rivalry and alienability is at best marginally related to whether a social convention to protect ideas, discoveries or designs should be established. Similarly, whether the social and legal consequences of the technology itself are acceptable is only vaguely related to whether genes and gene processes are sufficiently like the other kinds of intellectual goods that are protected by intellectual property rights to warrant extension of the social convention to them. Property Claims versus Property Conventions As noted early in this chapter, the fundamental ethical question is whether genes, segments of code or gene processes can be owned, consistent with our ethical beliefs and traditions for recognizing a given good as an ownable form of property. Even if this fundamental ethical question is answered in the affirmative, there are a host of additional questions about the social and legal means for protecting this property right. The frequent pattern of critics is to make sweeping statements opposing the legitimacy of intellectual property rights for biotechnology, then to support these claims with arguments that refer only to the defects of a specific system for protecting them. Henk Hobbelink does this, for example, when he writes that the ‘who owns it’ question is the most profound question in the development of agricultural biotechnology, then launches into a series of problems with US-style patents (Hobbelink 1995, pp. 230–231). Hope Shand, Vandana Shiva and even David Magnus commit a similar error in citing the biopiracy debate as point against intellectual property rights in genes or genetic traits (Shand 1991; Shiva 1997; Magnus 2002). In fact, the central point presumes that intellectual property rights exist, and that the problem consists in the failure of international patent systems to recognize the contributions (e.g. the prior ownership) of indigenous farmers in developing those rights. The underlying fairness question that critics of biopiracy raise is an important ethical question. It is, in fact, also a legal question. Given sufficient resources, representatives of farmers in developing countries might well be able to challenge patents awarded to scientists and seed companies through the courts (Feinsilver 1995). That they are not likely to do this has less to do with intellectual property as such and more to do with general ethical questions associated with the disparity between the access of the rich and the poor to legal services. It is important to see the poor must be able to make a moral claim to ownership of plant genetics for the biopiracy argument to go through. Furthermore, this claim is not limited by a smallholder’s ownership of single specimens of some plant variety; it applies to any plant having the relevant gene sequence. It is therefore an intellectual property right. When critics deny the validity of IPRs in genetic resources, they undercut the claim that they are trying to advance on behalf of the poor.

280

10 Conceptions of Property and the Biotechnology Debate

10.11 Ownership and Commodities A further theme in critics’ arguments is that ownership of life is itself wrong. To the extent that the theme devolves from religious beliefs, the argument will be taken up in Chap. 11. A number of critics have developed secularized versions of the argument, however. The strongest versions are directed toward a widely shared sentiment that patenting of human genetic materials is morally abhorrent (Macer 1992b, 1994). Theologian Ted Peters relates this reaction to a shared horror felt in response to the manipulation of Nazi eugenicists, (Peters 1994). Mark Hanson provides an overview of the way that this link between ownership of genes or gene process and improper commercial exploitation of life has been developed in medical bioethics. While he finds that vagueness and ambiguity pervade most approaches to the question, he concludes that patents do represent “an encroachment of commodification on our understanding of traits and organisms,” (Hanson 2002, p. 172). If these arguments make claims relevant to the key ethical questions of this chapter, they must say more than doing certain things (like practicing eugenics or genocide) is wrong. They must claim (and then justify) that owning (or claiming to own) genes, genomes, etc., is wrong. Andrew Kimbrell does make this claim pointedly in his antibiotechnology polemic (Kimbrell 1993). Kimbrell introduces the point by reviewing the legal actions surrounding the case of John Moore. Two scientist/physicians who had been treating Moore for leukemia successfully commercialized a line of cell tissues drawn originally from Moore. Kimbrell finds it bizarre that the courts have held that the creation of this commodity is perfectly acceptable, (though Moore himself enjoys no right of ownership in the cell line). Kimbrell argues that Moore’s rights were compromised not by the failure to obtain consent or to share the income from Moore’s spleen cells, but by the fact of commercialization itself. Individual rights, Kimbrell says, would be similarly compromised by experiments that extract or derive genetic technologies from samples of an individual’s DNA (Kimbrell 1993, pp. 206-210). Similar arguments against the commercialization of cell lines are made in Rebbeca Skloot’s book The Immortal Life of Henrietta Lacks (2010). However, Kimbrell and Skloot are discussing medical cases. The cell lines are derived from human beings with an unambiguous claim to moral standing, and it is likely that Peters also had biomedical gene technologies in mind. The reader might wonder why such issues would even come up in a book devoted to food and agricultural biotechnology. There are two reasons. First, it is theoretically possible that food researchers will find good reasons to use so–called human DNA in food products and processes. Second, it is possible to make an analogous argument with respect to all forms of life, not just humans. We saw this in the argument made by Hettinger (1995), and Hanson’s claims are similarly general, (Hanson, 2002) David Loy also makes sweeping claims about the wrongness of taking the attitude of ownership toward any form of life from a Buddhist perspective, (Loy 2003). The claim then is that ownership of genes itself compromises both human dignity and the dignity of life, irrespective of what one does vis-à-vis real human beings (like offering compensation or securing consent). This argument pivots to associate the

10.11 Ownership and Commodities

281

attitude of ownership with the other forms of unwanted consequence that have been the main topic of this book. Clearly, one way to flesh out this argument is to give it a religious meaning, and, again, that theme will be taken up in Chap. 11. But could this claim be made based on purely secular considerations? One way to do so associates (or conflates, depending on one’s point of view) ownership with domination and control. As noted at several junctures throughout the previous chapters, criticism of the view that nature is available for human domination and control has been a prominent theme in recent environmental ethics. Other chapters review the merits of the argument itself, but one way to see problems with property rights is to see these rights as permitting total license on the part of the rights holder. However, one would be hard pressed to find authors, much less legal documents, that defend property rights in terms of human domination and control. Ordinary chattel property rights provide the counter example. Livestock are arguably one of the oldest and most universal forms of ownable goods. Sheep, cattle and goats are, in normal circumstances, highly rival and excludable goods, and their ready alienability allows them to serve as a symbol of wealth itself in some societies. Clearly there is variability from society to society about what an owner may permissibly do with livestock, but there are few societies (and no industrial democracies among them) where the owners of livestock may do literally whatever they please with their animals. Both legal and customary sanctions against cruel and neglectful treatment are commonplace. Clearly there are verbal conventions where people use the word ‘own’ to indicate a morally objectionable form of domination and control (e.g. ‘I own you; you’re mine’), but to equate simply all forms of ownership with these morally objectionable senses is equivocation of the worst possible kind. Something more subtle may be going on with respect to the arguments that refer specifically to human genes. How do we handle the permissibility or impermissibility of human slavery? As already noted several times, one philosophical tradition has it simply that human beings should never be regarded as ‘ownable’ potential forms of property, hence slavery is never acceptable. Yet it is not clear what even this strong view implies for human genes. Kimbrell (like many) fails to avoid twin part–whole fallacies that are confusing by virtue of their complex logical relation to one another. A division fallacy occurs in inferring that because the whole (e.g. the individual human) cannot be owned, the parts (e.g. organs, blood, cells and DNA) cannot be owned. A composition fallacy occurs in inferring that because individuals cannot be owned, the whole species (represented by the genome) cannot be owned. One may indeed want to argue that genes or genomes cannot be owned, but the property status of individual human beings does not have logical bearing on the property status of parts of human beings (e.g. their DNA) or the wholes (the genome) of which humans are a part. Again, it is worth remembering that blood and semen are viewed as items that can legitimately be bought and sold. These arguments that take up problems of ownership as they relate to human genes or to life processes seem to be raising profound issues, but they are currently raising them in a fundamentally confused and ineffective manner. Perhaps better versions of these arguments will be forthcoming, and I repeat for the last time that specifically religious versions of the arguments will be discussed below in Chapter 11. For the

282

10 Conceptions of Property and the Biotechnology Debate

present, however, the arguments seem to hang on ambiguities in the word ‘ownership’ as well as on the logical ambiguity of part–whole relationships. Until better versions of these arguments are available, the case of the critics must be regarded as unproven. Moral philosophy provides multiple ways to interpret the ethical status of intellectual property claims for biotechnology, however, and given that the case for recognizing these property rights cannot be regarded as entirely proven, even faulty arguments have psychological and political force. The next round of debate over intellectual property can advance our understanding of the issues by being more specific in avoiding all three of these ambiguities and in focusing more intently on the problems of intellectual property as such.

10.12 Conclusion The chapter does not present an ethical argument for either accepting or rejecting IPRs for agriculture and food biotechnology. Instead, it develops a schema for philosophical theories of property, and examines hypotheses on how each could apply. The schema introduces intellectual discipline into the debate over property rights in agrifood biotechnology. My general conclusion is that the traditions of philosophical theory and argumentation provide reasons to support intellectual property rights (but not necessarily patents) in gene sequences as well as the key processes for identifying gene function and transferring DNA in agricultural plants and animals. They also provide some reasons against extending our traditions for recognizing goods as own-able, and then assigning property rights to individuals or organizations. In short, philosophy does not settle the issues, but it could improve argumentation and mutual understanding among disputants. Even within the comparatively narrow tradition of Western political thought, there are so many ways to conceptualize and legitimize property claims that any decisive argument is unlikely to be forthcoming. One point of ethical critique emerges: The dominance of utilitarian-style arguments to justify IPRs for the genetics of crop and livestock species reflects an ethical blind spot. Authors who write in favor of extending patent laws to gene technologies often write as if this is the only type of argument that counts. As such, they fail to understand the burdens of proof that they must meet in order to make cogent responses to critics who challenge the instrumentalism and welfare-maximizing assumptions of the utilitarian approach. They also may neglect some strong arguments in their favor, as a blend of libertarianism and “fruit of scientific labor” kind of argument would entail strong a strong right to control intellectual property. In any case, the ethical vocabulary in which IPRs for gene technologies could be broadened significantly through a simple inventory of the way that property rights have been critiqued and defended throughout history. Critics of IPRs are more cognizant of broader ethical traditions. Nevertheless, I doubt that intellectual rigor is actually what many of the contestants desire. My analysis of arguments in the literature through the 2007 revision of this book concludes that critiques who have opposed IPRs in gene technologies have committed one of

10.12 Conclusion

283

several logical errors. First, they may have developed an argument that is plausible for gene technologies applied in a biomedical context, where issues of privacy and the history of eugenics are relevant. Yet, they have neglected obvious reasons why the application of gene technologies to plants and livestock species are different. Second, many critics cite risks or unintended harms that either have or might accrue from use of a GMO, but there are many dangerous technologies that are protected by IPRs. Since patents can be used to prevent the use of an invention, we should not presume that the risks of use and the ethics of property claims are so tightly linked. Finally, criticisms that associate attitudes of arrogance or domination with claims of ownership are usually so broadly stated that it is difficult to see how they could count as an argument against some form of IPR for agrifood gene technologies. Although I have not attempted a thorough update of scholarship on patents in agrifood biotechnology, the criticisms discussed in this chapter are inconclusive, at best. Yet as with the early debate over Terminator seeds, claims about property rights and farmers’ control (or lack thereof) have strong emotional resonance. The suggestion that commodification—or intensification of ownership regimes—is fundamental to the moral faults of industrial agriculture has not waned in the intervening decades (see, for example, Akram-Lodhi, 2015; Rawlinson, 2019). As some recent analysts of the GMO debate suggest, analytic clarity may be less important in driving the rhetoric of the debate than a more conventional analysis of whose ox will be gored when the discourse turns out one way, rather than another, (Heilscher and coauthors, 2016). Needless to say, this is not the way that philosophers have traditionally approached argumentation, but it does provide an explanation for the lack of rigor and clarity in debates over IPRs for gene technology. Finally, when critics do not conflate reasons to be mindful of risks with reasons to grant or withhold IPRs, their arguments often claim that living things, life itself and, by extension, genes and gene processes are simply not the sort of things that can be owned. Their arguments overlap to a large degree with those who claim that manipulating genes is itself wrong, irrespective of the risks or benefits of doing so. These arguments depend on religious and metaphysical viewpoints that are taken up in the succeeding chapter.

References Akram-Lodhi, A.H. 2015. Accelerating towards food sovereignty. Third World Quarterly 36: 563– 583. Anderson, C. 1991. US patent application stirs up gene hunters. Nature 353: 485–486. Andrews, L. 2002. Patents, plants and people: The need for a new ethical paradigm. In Engineering the Farm: Ethical and Social Aspects of Agricultural Biotechnology, ed. B. Bailey and M. Lappé, 67–80. Washington, DC: Island Press. Anonymous. 1995. A GATT for Genes. Bija-the-Seed 14 (December): 1–4. Anonymous. 1996. Intellectual property rights, community rights and biodiversity. Bija-the-Seed 15: 25–30. Barooah, M. (1999) Intellectual property rights: Issues and options before India. Currents in Science, Technology and Ethics 2(2): 1, 6–8, 12.

284

10 Conceptions of Property and the Biotechnology Debate

Barton, J. 2004. TRIPS and the global pharmaceutical market. Health Affairs May/June 23 (30): 1–9. Becker, L.C. 1992. Too much property. Philosophy & Public Affairs 21: 196–206. Berghorst, D.T. 1991. Investment and the new technologies. In New Technologies and the Future of Food and Nutrition, ed. G.E. Gaull and R.A. Goldberg, 123–126. New York: John Wiley and Sons. Berlan, J.-P. and R. Lewontin. (1998) Cashing in on life: Operation Terminator. LeMonde Diplomatique, Accessed March 11, 2020 at http://mondediplo.com/1998/12/02gen. Biddle, J.B. 2014. Can patents prohibit research? On the social epistemology of patenting and licensing in science. Studies in History and Philosophy of Science Part A 45: 14–23. Biddle, J.B. 2016. Intellectual property rights and global climate change: Toward Resolving an Apparent Dilemma. Ethics, Policy & Environment 19: 301–319. Brody, B.A. 1989. An evaluation of the ethical arguments commonly raised against the patenting of transgenic animals. In Animal Patents: The Legal, Economic and Social Issues, ed. W. Lesser. New York: Stockton Press. Brody, B.A. 2010. Intellectual property, state sovereignty, and biotechnology. Kennedy Institute of Ethics Journal 20: 51–73. Carstensen, P. 2005. Post-sale restraints via patent licensing: A seedcentric perspective. Fordham Intellectual Property, Media & Entertainment Law Journal 16: 1053–1080. Charles, D. 2001. Lords of the Harvest: Biotech, Big Money, and the Future Of Food. Cambridge, MA: Perseus Books Group. Crouch, M. (1998) How the Terminator terminates, RAFI, Accessed March 11, 2020 at https://iatp. org/sites/default/files/How_the_Terminator_Terminates.htm. Drahos, P. 2016. A Philosophy of Intellectual Property. New York: Routledge. Egelie, K.J., S.P. Strand, B. Johansen, and B.K. Myskja. 2018. The ethics of access to patented biotech research tools from universities and other research institutions. Nature Biotechnology 36: 495–499. ETC Group. 2002. Terminate Terminator. Ottawa, ON: The Etc Group. Eisenberg, R. 2003. Patent swords and shields. Science 299: 1018–9. Feinsilver, J.M. (1995) Cuban biotechnology: The strategic success and commercial limits of a first world approach to development. In: N.P Peritore and A.K. Galve-Peritore (eds) Biotechnology in Latin America: Politics, Impacts and Risk. Scholarly Resources, Wilmington, DE, pp 97–126. Fowler, C., and P.R. Mooney. 1990. Shattering: Food, Politics, and The Loss of Genetic Diversity. Tucson: University of Arizona Press. Gallini, N. 2017. Do patents work? Thickets, trolls and antibiotic resistance. Canadian Journal of Economics/Revue canadienne d’économique 50: 893–926. Genetic Literacy Project. (n.d.) What’s the controversy over ‘terminator’ seeds? Genetic Literacy Project, Accessed March 11, 2020 at https://gmo.geneticliteracyproject.org/FAQ/whats-contro versy-gmos-terminator-seeds/. Giovanetti, M. (2001) Environmental impact of transgenic crops: potential risks associated with Terminator gene technology. Non-Technical Abstracts, Biotechnology, Science and Modern Agriculture: a New Industry at the Dawn of the Century Conference, June 2001, Ravello, Italy. Glover, D. 2007. Monsanto and smallholder farmers: A case study in CSR. Third World Quarterly 28: 851–867. Goldman, A.H. 1987. Ethical issues in proprietary restrictions on research results. Science, Technology and Human Values 12: 22–38. Hanson, M.J. (2002) Patenting genes and life: Improper commodification? In: D. Magnus, A. Caplan and G. McGee (eds) Who Owns Life? Prometheus Books, Amherst, NY, pp. 161.-174. Haussler, W. (1996) Negotiating intellectual property agreements, presented at Third Asia Pacific Conference on Agricultural Biotechnology, Prachuab Kirikhan, Thailand, November 12. Heilscher, S., I. Pies, V. Valentinov, and L. Chatolova. 2016. Rationalizing the GMO debate: The ordonomic approach to addressing agricultural myths. International Journal of Environmental Research and Public Health 13: 476–486. Herring, R.J. 2007. Stealth seeds: Bioproperty, biosafety, biopolitics. The Journal of Development Studies 43: 130–157.

References

285

Hettinger, N. (1992) Owning varieties of life. Discussion Paper 92–10. Center for Biotechnology Policy and Ethics, College Station, TX. Hettinger, N. 1995. Patenting life: Biotechnology, intellectual property, and environmental ethics. Boston College Environmental Affairs Law Review 22: 267–275. Hobbelink, H. 1991. Biotechnology and the Future of World Agriculture. London: Zed Books. Hobbelink, H. 1995. Biotechnology and the future of agriculture. In Biopolitics: A Feminist and Ecological Reader on Biotechnology, ed. V. Shiva and I. Moser, 226–33. London: Zed Books. Hospers, J. 1971. Libertarianism: A Political Philosophy for Tomorrow. Los Angeles: Nash Publishing. Juma, C. 1988. The Gene Hunters: Biotechnology and the Scramble for Seeds. Princeton, NJ: Princeton University Press. Juma, C. (2014) The global risks of rejecting agricultural biotechnology, Accessed July 29, 2020 at: https://geneticliteracyproject.org/2014/12/09/global-risks-of-rejecting-agricultural-biotechno logy/. Kimbrell, A. 1993. The Human Body Shop. New York: HarperCollins. Lechtenberg, V.I., and A.A. Schmid. 1991. Intellectual property rights. In Agricultural Biotechnology: Issues and Choices, ed. B.R. Baumgardt and M.A. Martin. IN: Purdue Agricultural Experiment Station, West Lafayette. Lesser, W. 1989. Animal Patents: The Legal. Economic and Social Issues: Stockton Press, New York. Loy, D. (2003) Remaking the world or remaking ourselves? Buddhist reflections on technology. In: P.D. Hershock, M. Stepaniants and R.T.Ames (eds) Technology and Cultural Values on the Edge of the Third Millennium. University of Hawaii Press, Honolulu, pp 176–187. MacCormick, N. 1987. Law. In The Blackwell Encyclopedia of Political Thought, ed. D. Miller. Oxford, UK: Basil Blackwell. Macer, D.R.J. 1992. Attitudes to genetic engineering: Japanese and international comparisons. Christchurch, NZ: Eubios Ethics Institute. Macer, D.R.J. 1994. Bioethics for the People by the People. Christchurch, NZ: Eubios Ethics Institute. Magnus, D. 2002. Intellectual Property and agricultural biotechnology: Bioprospecting or biopiracy? In Who Owns Life?, ed. D. Magnus, A. Caplan, and G. McGee, 265–276. Amherst, NY: Prometheus Books. Malshe, D. 2018. Patent Pools. Routledge, New York: Competition Law and Biotechnology. Milligan, R., and W. Lesser. 1989. Implications for agriculture. In Animal Patents: The Legal, Economic and Social Issues, ed. W.H. Lesser, 100–14. New York: Stockton Press. Mooney, P.R. (1979) Seeds of the Earth: A Private or Public Resource? Inter Pares for the Canadian Council for International Cooperation and the International Coalition for Development Action (London), Ottawa. Muir, W. (2001) RE: USDA says Yes to Terminator, AgBioWorld, April 6, 2001. Accessed March 11, 2020 at http://www.agbioworld.org/newsletter_wm/index.php?caseid=archive&newsid=1141. Murashige, K.H. 1994. Intellectual property and genetic testing. In The Genetic Frontier: Ethics, Law and Policy, ed. M.S. Frankel and A. Teich, 181–98. Washington, DC: American Association for the Advancement of Science. Nozick, R. 1974. Anarchy. State and Utopia: Basic Books, New York. Overhauser, J. 1994. Intellectual property and genetic testing: a scientist’s perspective. In The Genetic Frontier: Ethics, Law and Policy, ed. M.F. Frankel and A. Teich, 209–214. Washington, DC: American Association for the Advancement of Science. OTA (U.S. Congress, Office of Technology Assessment). (1989) New Developments in Biotechnology:Patenting Life–Special Report, OTA-BA-370: U.S. Government PrintingOffice, Washington, DC. Paine, L.S. 1991. Trade secrets and the justification of intellectual property: A comment on Hettinger. Philosophy & Public Affairs 290: 247–263. Peters, T.F. 1984. Creation, consummation and the ethical imagination. In Cry of the Environment: Rebuilding the Christian Creation Tradition, ed. P.N. Joranson and K. Butigan, 401–29. Santa Fe: Bear and Co.

286

10 Conceptions of Property and the Biotechnology Debate

Pinstrup-Andersen, P., and E. Schiøler. 2000. Seeds of Contention: World Hunger and the Global Controversy over GM Crops. Baltimore: Johns Hopkins University Press. Prasad, R., U.S. Bagde, and A. Varma. 2012. An overview of intellectual property rights in relation to agricultural biotechnology. African Journal of Biotechnology 11: 13476–13752. RAFI (Rural Advancement Fund International). (2000) Golden rice and Trojan trade reps: A case study in the public sector’s mismanagement of intellectual RAFI Communique September/October Accessed March 13, 2020 at: http://www.etcgroup.org/files/publication/305/01/com_goldenrice. pdf. Rawlinson, M.F. 2019. The climate of food: Justice, truth and structural change. In Food, Environment, and Climate Change: Justice at the Intersections Rowman and Littlefield, ed. E. Gilson and S. Kenehan, 91–118. MD: Lanham. Resnik, D.B. 2016. Scientific Realism and the Patent System. Journal for General Philosophy of Science 47: 69–77. Robinson, D.F. 2010. Locating biopiracy: Geographically and culturally situated knowledges. Environment and Planning A 42: 38–56. Ryan, Alan. 1984. Property and Political Theory. New York: Basil Blackwell. Shah, A. (2001) Terminator technology, Global Issues: Genetically Engineered Food Accessed March 11, 2020 at http://www.globalissues.org/EnvIssues/GEFood/Terminator.asp. Sederoff, R., and L. Meagher. 1995. Access to intellectual property in biotechnology: constraints on the research enterprise. In Genes for the Future: Discovery, Ownership, AccessNational Agricultural Biotechnology Council, ed. J.F. MacDonald. NY: Ithaca. Shand, H. 1991. There is a conflict between intellectual property rights and the rights of farmers in developing countries. Journal of Agricultural and Environmental Ethics 4: 131–142. Shand, H. 1998. Agricultural biodiversity, biopiracy and food security. Canadian Journal of Development Studies/Revue canadienne d’études du développement 19: 161–183. Sharma, D. 2003. GM Food and Hunger: A View from the South. New Delhi: Forum for Biotechnology and Food Security. Shiva, V. 1993. Monocultures of the Mind: Perspectives on Biodiversity and Biotechnology. London: Palgrave Macmillan. Shiva, V. 1997. Biopiracy: The Plunder of Nature and Knowledge. Boston: South End Press. Shiva, V. 2000. Stolen Harvest: The Hijacking of the Global Food Supply. Cambridge, MA: South End Press. Shiva, V. (ed.). 2016. Seed Sovereignty, Food Security: Women in the Vanguard of the Fight against GMOs and Corporate Agriculture. Berkeley, CA: North Atlantic Books. Skloot, R. 2010. The Immortal Life of Henrietta Lacks. New York: Crown. Somsen, H. (ed.). 2007. The Regulatory Challenge of Biotechnology: Human Genetics. Food and Patents: Edward Elgar Publishing, Cheltenham, UK. de Sousa, S.J. 1995. Plant intellectual property rights: The rise of nature as a commodity. In Biotechnology in Latin America: Politics, Impacts, and Risks, ed. N.P. Peritore and A.K. Galve-Peritore. DE: Scholarly Resources, Wilmington. Specter, M. 2000. The pharmageddon riddle.”. The New Yorker 76 (13): 58–71. Sterckx, S. 1997. Biotechnology. Ashgate, Brookfield, VT: Patents and Morality. Strauss, D.M. 2009. The application of TRIPS to GMOs: international intellectual property rights and biotechnology.”. Stanford Journal of International Law 45: 287–320. Thompson, P.B., R.J. Matthews, and E. vanRavenswaay. 1994. Ethics. Macmillan, New York: Agriculture and Public Policy. United States Supreme Court (1980) Diamond v Chakrabarty. 447 US 303. White, T.J. 1994. Intellectual property and genetic testing: A commentary. In The Genetic Frontier: Ethics, Law and Policy, ed. M.S. Frankel and A. Teich, 199–208. Washington DC: American Association for the Advancement of Science. Wills, G. 1978. Inventing America: Jefferson’s Declaration of Independence. New York: Doubleday. Young, F.E. 1991. Regulatory challenges and biotechnology advances. In New Technologies and the Future of Food and Nutrition, ed. G.E. Gaull and R.A. Goldberg, 101–104. New York: John Wiley and Sons. Yusef, M. 2010. Ethical issues in the use of terminator seed technology. African Journal of Biotechnology 9: 8901–8904.

Chapter 11

Religiously Metaphysical Arguments Against Agrifood Biotechnology

Abstract Metaphysical claims assert categories and categorical systems for the broadest and most general characterizations of reality and experience. The chapter discusses the nature of metaphysical claims and the role of religious or theological doctrines in lending support to them. Early debates over gene technology emphasized metaphysical and religious topics and questioned whether established doctrines in mainstream religious traditions were compatible with applications of genetic engineering. The chapter surveys those debates, with emphasis to their significance for agrifood biotechnologies, and situates them within the context of other technologies that have been alleged to pose religious or metaphysical challenges. Those who write in favor of gene technology from a religious perspective have not based their arguments on metaphysical claims. Hence, religiously metaphysical arguments tend to take a critical stance toward gene technology. However, metaphysical arguments appear to have decreased in frequency and significance since earlier editions of this book. Keywords Playing god · Sanctity of life · Bioethics · Intrinsic objections Every new development in biotechnology has seemingly spawned a spasm of philosophical reflection on the possibility that there is something fundamentally wrong about it. There was an outcry when in vitro fertilization was first used to produce a human baby (Nugent 2018). Years earlier, when artificial insemination was first tried, the people involved in the experiment were sworn to secrecy, (Yuko 2016). Following the establishment of bioethics programs in many medical schools throughout the 1980 s, innovations that involve reproduction at either the genotypic or phenotypic level have been subject careful scrutiny. While many ethical concerns have sat comfortably under the umbrella of risk analysis, others claim the new tools or techniques are inherently or intrinsically wrong, irrespective of their potential for adversely affecting human health, animal welfare, the social order or the environment. This chapter explores objections to food and agricultural gene technologies that identify the use of rDNA to modify plant or animal genomes as inherently or intrinsically wrong. The chapter begins with three sections that characterize the nature of

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_11

287

288

11 Religiously Metaphysical Arguments …

these arguments. They are metaphysical in the sense that they challenge the basic conceptual categories that most biologists use (often unconsciously) to explain the mechanisms of molecular biology. Some, but not all, of these challenges arise out of religiously-motivated ways of understanding the basic forces that control the shape of events. The chapter continues with three sections that discuss metaphysical or religiously based statements on the ethics of agrifood biotechnology. The first examines statements made by religious authorities speaking on behalf of established churches, synagogues or other religious organizations. The second continues themes from Chap. 10, discussing religiously based arguments against property rights in genes or gene processes. The last reviews the statements of academic theologians. The chapter concludes with a discussion of the ethical significance of these religiously motivated views.

11.1 Intrinsically Wrong The arguments discussed in this chapter hold that gene technology is wrong irrespective of its consequences. This implies that the considerations brought forward in a risk-based approach will never be decisive. To put it differently, examining the risks and benefits of genetic modification is irrelevant to the ethics of biotechnology, given the philosophical orientation of people who rely on this type of argument. Among the rationales offered to support such views, are that genes, genomes or perhaps simply life itself is sacred, or that these genetic tools take us into a territory where human beings should never go. Sometimes language like this means that the risks of gene technology are especially grave, so much so that they outweigh all projections of benefit. This type of claim can be incorporated into a risk-based approach in the form of risk management that forbids acceptance of such grave risks. Similarly, the logic of the risk-based approach accommodates hazards and exposure mechanisms that appeal to supernatural powers. Threats of eternal damnation or the wraths of vengeful gods are conceptually understandable as the outcome of immorality or acting with evil intent. Indeed, Blaise Pascal (1623–1662) formulated influential theological arguments for practicing the Christian faith in just such risk-based terms, (Hájek 2018). However, the philosophical claims one would make in applying a risk-based approach this way have a different character than previous chapters have discussed. What is more, one might reject the very idea of weighing risks, benefits and prospective trade-offs that might occur in the wake of gene transfer or modification as inherently corrupt. In these arguments, we are invited to see that act of intervening in nature’s methods for mixing genes or reproducing organisms is morally problematic in-itself. It cannot be redeemed by compelling benefits, making the exercise of evaluating its risks not only moot, but also indicative of a moral flaw. For some observers of contemporary biotechnologies, these are concerns that most fully qualify as originating on ethical grounds. Rachel Schurman and William Munro trace the origins of the social movement against GMOs to people who, in the 1980s, believed that gene technology possessed

11.1 Intrinsically Wrong

289

unprecedented potential for evil. Cautioned by the excesses of the early 20th century eugenics movement and related racist ideologies that had sanctioned Nazi doctor’s experimentation on Jews and the U.S. Tuskegee experiments on blacks, some of these early activists believed that the very idea of genetic manipulation was indicative of hubris, excess, greed and human domination of nature, (Schurman and Munro 2010). These early critics did not make a sharp distinction between human biotechnology and the application of recombinant DNA tools to agricultural plants and animals. Veterinarians and livestock producers used reproductive technologies without controversy before researchers in human medicine tried them. It is plausible to imagine a slippery slope where food and agriculture serve as a proving ground for actions that will eventually be applied to human beings. In the 1980s, methods to modify plant and animal genomes began to be discussed at the same time that sequencing the human genome was being debated in bioethics. I will not engage in speculation as to whether this new social context influenced the public’s reception for food and agricultural biotechnologies. However, it will prove especially helpful for readers who work in agriculture to remind themselves that the metaphors of boundary crossing and forbidden knowledge have a long history in the context of technical changes that affect the reproduction of human beings. More broadly, virtually all philosophers and religious thinkers recognize a distinction between reasoning that evaluates an action in light of its possible effects (including both impacts on welfare and rights violations), and reasoning that finds something internal to a practice to be the basis for ethical evaluation. Perhaps the clearest example of the difference can be seen when people act from evil intentions, from the desire to do somethingf wrong. Something ethically problematic about genuine wickedness remains, even when an evildoer chances upon acts that have morally praiseworthy outcomes. Notice that this would hold even if an uninvolved party were able to perform an objective risk analysis of the evildoer’s plans, concluding that intentions aside, they were very likely to result in extremely beneficial outcomes. Said another way, malevolence is not excusable just because hoped-for evils either did not or were unlikely to be realized. There is, in fact, an important subclass of anti-GMO arguments that turn upon the claim that developers have evil intentions, or at least that they are motivated exclusively by profit seeking (which comes to the same thing in the minds of some critics). However, the arguments discussed in this chapter seek to locate the ethical problem with gene transfer in the very act of performing a genetic modification, rather than in some character flaw in the agent. As hinted in the chapter title, religion can provide rationales that have this feature. Religiously based dietary norms, for example, proscribe the consumption of certain foods or prescribe conditions for preparing or serving a meal. The test for the ethical correctness of these norms does not reside in their implications for personal or environmental health. Yet members of the relevant faith tradition accept the legitimacy of these norms as prescriptions intrinsic to the traditions, rituals and practices that define the faith. Arguments of this sort are made with some frequency in discussions of gene technology, but there have been shifts in the regularity with which they are made in the years since the first edition of this chapter was published in 1997. There does seem

290

11 Religiously Metaphysical Arguments …

to be a declining interest in them since the first edition of this book was written in the mid-1990s. Official and semi-official commissions convened to review gene-based technologies (such as the U.S presidentially established bioethics commissions or the U.K.’s Nuffield Council) never fail to include some discussion of religious concerns as well as other reasoning that supports a blanket injunction against gene manipulation. However, university-based experts on technological ethics have grown increasingly impatient with such broadly stated arguments over the last thirty years. One U.S. study on synthetic biology notes that the religious concerns expressed when gene technologies were being discussed in the 1980s are restated in connection with synthetic biology. The report observes that these concerns are never taken to have policy relevance. The report’s authors conclude that unless religious reasons can be correlated with some risk to human or environmental health, they are incompatible with the establishment clause in the U.S. Constitution, which prevents the Federal Government from acts that promote one religious perspective over another, (Presidential Commission 2010). It is less clear that gene editing, gene drives and other advances in biotechnology will be subjected to the kind of debate reviewed in the balance of this chapter, especially if their applications are confined to the agrifood sector.

11.2 Situating Religious and Metaphysical Claims First, some qualifiers. Many who base their concerns and objections to food biotechnology on religious grounds are applying arguments that are typical of a risk-based approach to technological ethics. They cite risks to environment, animals or social stability that have a technological origin, and base their ethical arguments either on principles that articulate what is seen as an unfavorable balance of benefit to harm, or by appealing to principles of participation, rights and consent. Practitioners of a given faith often articulate their personal duty to embrace moral principles in religious terms. This can make an ethical duty seem more profound or vital, but there is an important sense in which the addition of religious motivation for acting ethically does not change the content of ethical claims that might be made on secular grounds. The mechanics of securing consumer consent for the sale of foods modified through genetic engineering, for example, are the same whether political or religious reasons ground the duty. Consent criteria become no easier (or more difficult) to satisfy when they are based on religion, and the question of who should bear the cost of labeling policies that protect consent is not changed. Perhaps those who feel religiously enjoined to reject consequentialist or trade-off arguments will be more tireless in the opposition to utilitarian reasoning than others will, but the philosophical point at issue is not substantially changed (Deane-Drummond 1995). Given this, many statements on genetic engineering from religious organizations or religiously inclined people repeat topics covered by the preceding chapters of this book. The main task of this chapter is to examine two kinds of religious argument that contribute a distinct line of reasoning to the debate over food biotechnology and

11.2 Situating Religious and Metaphysical Claims

291

genetic engineering. They will typically involve the claim that when one manipulates an organism using rDNA techniques, one does something that is itself wrong, without regard to the immediate consequences. Academic philosophers often refer to this as an “intrinsic” objection, (Streiffer and Heidemann 2005; Kaebnick 2009). Intrinsic reasons are sometimes associated with rights-based or deontological systems, while intrinsic values are described as those that are not subject to value-for-value tradeoffs. In either case, from the perspective of a risk-oriented ethics, the philosophical debate will concern the decision rule. An intrinsic objection, then, can be understood as a rule that tells the decision maker not to allow or consider options of a given kind, irrespective of their hazards or the likelihood that hazards will be realized. This chapter will examine comprehensive or far-reaching arguments that purport to show the inherent immorality of all biotechnology. Next, some definitions. Unlike many of the other arguments discussed in this book, the arguments discussed in this chapter draw directly on particular religious and metaphysical beliefs. I am using the words religious and metaphysical here to make a distinction pertinent to the discussion that follows, and I am not committed to these definitional points beyond the present context. A metaphysical belief is a broad framing belief about the nature of reality. Not all metaphysical beliefs are religious in the sense of resting on or reflecting a religious doctrine or practice. The core sense that I will give to the term religious in this chapter refers to doctrines that are held and promulgated by one or more of the world’s religions—Islam, Buddhism, Christianity, Hinduism, Judaism, and so on. My particular emphasis will be to characterize beliefs that do not function in mainstream biological explanations of genes and gene processes as religious in nature, and there may well be some such beliefs that draw little support from any organized religion or spiritual tradition. However, some of the beliefs and doctrines most typical of the world’s religions are clearly metaphysical in nature. Belief in the existence or non-existence of gods, spirits or ghosts having (or lacking) powers to intervene in human affairs are metaphysical. Beliefs about the meaning of death, life after death, reincarnation, heaven, hell, or purgatory are metaphysical beliefs. When beliefs about right and wrong action are combined with beliefs about the intentions or wishes of a creator god, or intertwined with beliefs about sin, karma or the fate of the soul they become both religious and metaphysical. To reiterate, metaphysical beliefs are ubiquitous; everyone has some, but we do not all have the same ones. For example, many of us, at least, agree that food animals are sentient and can feel pain, though we might not agree on other beliefs about the nature of cognitive experience or spirituality. Do animals have an immortal soul? Unlike a discussion of animal pain, further discussion of this question invites contributions from those religious traditions where talk of souls and immortality has been important (indeed, constitutively important). It is it is possible to conduct an informed conversation about the ethics of food biotechnology’s impact on animal health and sentience. Yet, the most troublesome and conceptually difficult objections to genetically modified animals make the claim that modification of animal genomes is itself wrong, irrespective of secondary costs and benefits. The report Patenting Life prepared for the United States Congress refers to them as “metaphysical and theological arguments”

292

11 Religiously Metaphysical Arguments …

(OTA 1988), and my word use in this chapter attempts to follow the rough distinctions implied thereby. Whether limited to beliefs about animals or extended to cover plants, microorganisms or genetic constructs themselves, metaphysical and religious arguments are among the most potent objections to food biotechnology. Arguments based on technological ethics merely qualify the direction and application of biotechnology. Metaphysical and theological arguments might provide grounds to prohibit it altogether.

11.3 Metaphysical and Religious Arguments: An Exhibitive Introduction A moral argument makes a metaphysical claim when the rightness or wrongness of an action is based directly on features ascribed to the ultimate reality or nature of a being, a creature, a natural construct or, indeed, being itself. For example, Chap. 10 reviewed ontological criteria for deciding whether a good can be considered as someone’s property. While it is difficult to imagine a moral judgment utterly lacking in metaphysical implications, the arguments that are the focus of this chapter have some distinguishing characteristics. They involve direct claims stating that something of moral significance is (or is not) the case. They neither contextualize these claims, nor develop further criteria for verification through logic and empirical evidence or legitimation through a social process. As noted already, the current consensus among philosophers of science is that all sciences presuppose some metaphysical beliefs, but empirically inclined philosophers would also argue that these beliefs are highly warranted in virtue of the coherence and predictive utility that is associated with the ensemble of beliefs constituting a scientific worldview. It is thus likely that every moral argument relies on metaphysical beliefs. However, the arguments that have been analyzed in the previous chapters of the book do not depend on the accuracy of some notoriously controversial metaphysical beliefs, beliefs such as whether or not there is a God, and of even whether there is an external world beyond our senses. In ordinary or vernacular speech, people may formulate moral claims in a similarly direct, contextless and dogmatic style. They may simply be cutting corners and speaking loosely, or they may presume that their audience shares the relevant beliefs. As such, getting to the heart of a religiously metaphysical argument can require some analysis. For example, the scientifically inclined author of an early treatise on ethics and genetic engineering offers the following as an introduction to the discussion of human genetics: Man is an animal in that he grows, reproduces and evolves like all other organisms. However, unlike other animals, man possesses a mind which manifests itself in cultural phenomena, thereby posing problems for an effective study of his genetic inheritance. First and foremost, unlike mice, fruit-flies, and other animals used for gathering experimental data, man cannot be mated at the discretion of the scientist for the sake of determining the genetic make up of the offspring. (Santos 1981, 8)

11.3 Metaphysical and Religious Arguments: An Exhibitive Introduction

293

The implicit sexism aside, this passage by M.A. Santos makes at least three statements with important moral implications whose basis could only be metaphysical, given the balance of his treatment. 1. Human beings possess minds. One way of reading Santos attributes the possession of mind to the human species rather than individual human beings. This is questionable, but a charitable interpretation of his remarks yields this first, more reasonable metaphysical claim. 2. Other animals do not possess minds. Santos actually states that other animals do not possess minds that manifest themselves in the form of culture. In some contexts, this would be an ethically crucial distinction, but the less subtle interpretation seems more consistent with the author’s apparent intent. He probably should not be read as implying that animals are not sentient, however. These two statements are offered in support of the third: 3. Man cannot be mated at the discretion of the scientist. Again, the actual statement is qualified, but Santos seems to mean this as an absolute moral proscription of discretionary mating of human beings for purposes of scientific research. All three of these statements could be unpacked at a length that would test the patience of all but the most philosophically inclined reader. Statements 1 and 2 are open to query into the nature of minds, but Santos’ point seems to be emphasizing some crucial distinction between the minds of human beings and minds of other animals. This distinction is the basis for his main moral claim, which is a moral prohibition against breeding humans for laboratory research. What makes this into a religiously metaphysical argument? Consider an alternative argument that would have established the case against laboratory breeding of humans by saying it violates fundamental human rights. A number of philosophical or religious views proffered to clarify and defend the meaning and legitimacy of human rights could support such an argument. They include principles resting on consent, social consensus, long-run social utility, or rational consistency, as well as the view that rights have a foundation in religious faith or in God’s plan. The metaphysical connotation of Santos’ prose arises in part from its dogmatic character: it just is the case that scientists “cannot” (as opposed to should not) mate humans. Written in 1981, Santos’ text should not be too heavily criticized for its insensitivity to more recent evidence that many non-human species exhibit mindedness, but in the 21st century, the second statement appears dogmatic, too. The metaphysical tone is reinforced by the obscurity of alleged links between the proscription and two other metaphysical claims. Why does the possession or nonpossession of a mind bear on the permissibility of breeding? No reasons are forthcoming, leaving the reader with the impression that the author just “sees” the world this way, in much the same way that we simply “see” the world as spatially extended, or “see” time as irreversible. Furthermore, specific Judeo-Christian doctrines seem to lie beneath the surface, especially passages in Genesis which authorize humans to kill animals. Religiously metaphysical arguments may make perfect sense to someone who shares the author’s intuitions, but usually offer little to those of us who do not.

294

11 Religiously Metaphysical Arguments …

Although the argument above does not address food biotechnology, it provides an example of metaphysical claims in support of a moral prescription. What is more, metaphysical and theological arguments against food biotechnology are often just special cases of arguments against genetic engineering tout court. If it is simply wrong to cross species boundaries, it is wrong no matter what species are involved. As Schurman and Munro suggest, advocates such as Jeremy Rifkin and Andrew Kimbrell may have derived a metaphysically-based view that genetic engineering is wrong first by considering cases that involve human genes and then inferring that food biotechnology might revive the eugenics movements associated with the horrors of the early 20th century. Such an argument would require additional premises that would bridge the chasm between humans and other animals (not to mention plants) that we see in the quotation from Santos. Once such metaphysical views are articulated, they entail the wrongness of genetic engineering for plants and animals, as well as for humans on logical grounds. Since none of these metaphysical claims, (including those that support the worldview of the typical molecular biologist), can be discussed outside the context of a philosophical conversation, facile dismissals of religiously metaphysical arguments are unwarranted.

11.4 Analyzing the Religiously Metaphysical Case Against Genetic Engineering Since the earliest days of molecular genetics, virtually everyone with any cognizance of this science expected a visceral reaction from religious conservatives. Clearly, it would possible to think that the entire science of recombinant DNA is just wrong, with manipulation of plant, animal and human genomes simply being the most egregious violation of a basic and irreducible moral principle. The wrongness implied is often expressed as “playing God,” the phrase chosen as the title for June Goodfield’s 1977 book on the Asilomar conference. However, it is surprising how elliptical and non-specific statements of this basic moral proscription tend to be. Karen Lebaqz, a professor of Christian Ethics at the Pacific School of Religion reviewed a United States Presidential Commission’s discussion of the possible meanings of “playing God,” (U.S. President’s Commission 1982). The Commission concluded that fears of playing God are actually fears of the consequences from gene technology. This interpretation rejects the most straightforward interpretation, namely that, as Lebaqz puts it, “the nature of the knowledge involved is taken to generate a prohibition,” (Lebaqz 1984). Lebaqz herself stops short of endorsing such a prohibition, stating instead that the problem resides in an unquestioned commitment to the rational analytic methods typical of technological ethics. The 1982 Commission and Lebaqz both exemplify one approach to the “playing God,” claim. They recognize the phrase “playing God,” as both metaphysical and as religious in the senses developed above, and they respond to the obscurity of the phrase by offering an interpretation of its meaning in the context of genetic

11.4 Analyzing the Religiously Metaphysical Case Against Genetic Engineering

295

engineering. These interpretations open the door to further philosophical analyses and debate. The 1982 Presidential Commission elects to interpret the phrase as a use of religiously metaphysical language to assert what is, in the final analysis, a claim that can be adequately addressed using the expanded framework of risk analysis that has guided the discussion in chapters three through nine of this book. Lebaqz disputes the Commission’s interpretation by asserting the phrase is actually intended as a criticism of philosophical methods in technological ethics (e.g. risk assessment), or at least to uncritical adoption of such methods. As I read Labaqz, she is both asserting that the religious language needs to be addressed on its own terms, and offering an interpretation of it that would shift the ethical analysis to focus on the hubris with which the technology is being pursued. This shift is also an avenue for engagement with secular ethics, albeit one that emphasizes virtue and vice as the primary moral categories for analysis, (see Sandler 2004 for a secular version of this argument). The reinterpretation of religious language is an important philosophical task in its own right. It permits translation and communicative engagement across cultural boundaries, and is a way of demonstrating the respect for religious thought and its institutions that Labaqz calls for in her critique of the Presidential Commission’s report. At the same time, interpretations that translate a claim into the secularized vocabularies of utilitarianism, neo-Kantian rights theory or neo-Aristotelean virtue ethics may not do full justice to Lebaqz’s claim that something about the very nature of gene technology generates the rationale for its prohibition. Closer examination of some instances where religiously metaphysical appeals are used to express disapproval of gene technologies provides the basis for amending secularized reinterpretations of religious language with another alternative. Even the most visible public opponents of gene technology tend to promote ethical prescriptions more through innuendo than through direct statement. A 1987 article by Andrew Kimbrell and Jeremy Rifkin separates ethical considerations from social and ecological risks, but when Kimbrell and Rifkin go on to specify what ethical considerations might be, what they produce is a list of questions: What is wrong with a cow the size of an elephant, or a sheep the size of a horse, or “glowing” tobacco plants? Is there any meaning in the morphology of animals or plants, both internally and externally? Should we alter nature or mutate, perhaps permanently, the forms and shapes of the biotic community so that they better conform to our agricultural or industrial needs? Do plants and animals have any right to be treated as sufficient “ends” in themselves, and not merely as “means” in a system of production? What are the ethical implications of the likely proposal to engineer plant or animal genetic material into humans? Finally, who is to decide these issues: Congress? Scientists? Corporations? Theologians? The public? Federal agencies? (Kimbrell and Rifkin 1987, 126).

Although it is easy to guess how Kimbrell and Rifkin would answer these questions, they are not explicit in formulating a statement that genetic engineering is categorically wrong on moral or metaphysical grounds. This pattern continues even in their more polemical works. There is an important point of rhetoric to note in this connection. Aristotle examined how leaving a key premise or two unstated serves to persuade people in many

296

11 Religiously Metaphysical Arguments …

social contexts. He made a distinction between philosophical ethics, where logical completeness is important for proper evaluation of an argument, and other argumentative contexts, such as legal proceedings or a political debate. Aristotle states that in the latter case, a speaker should rely on what is commonly believed, rather than what stands the test of rigorous (e.g. philosophical) examination. There are both shallow and deep rationales for this advice. Metaphysical claims can be omitted because they appeal to “what everybody knows,” even if what everybody knows is wrong. If one’s goal is simply to manipulate opinion, then calling attention to assumptions that might prove difficult to defend will only defeat one’s purpose. However, Aristotle scholars argue that as a philosopher, he could not have endorsed such an unscrupulous view. Rather, these assumed commitments function to maintain the social fabric enabling collective action, deeds or decisions that could be said to reflect “the will of the people.” Other critics follow the rhetorical strategy of listing unanswered questions. Writing some fifteen years before he became the Chairman of President George W. Bush’s Bioethics Council, Leon Kass took issue with suggestion that a genetically engineered organism is just a composition of matter (hence patentable), by asking “What about other living organisms—goldfish, bald eagles, horses? What about human beings? Just compositions of matter? Here are deep philosophical questions to which the court has given little thought…” (Kass 1985 149-50) Baruch Brody noted that those who proffer metaphysical and theological arguments “themselves recognize that they need to do a lot more work to articulate the inchoate concerns they feel,” (Brody 1989, 142). Brody notes that religious denominational statements and religious study groups decry the philosophical reductionism and materialism that they see in contemporary molecular biology. Yet they are unable (or unwilling) to turn this revulsion toward calling for an explicit ban on genetic engineering. When pressed, they revert to the concerns of technological ethics that need no religious basis, (Brody 1989, 142-3). In reinterpreting religious claims, one fleshes out the unstated assumptions by attempting to make an explicit statement of the metaphysical claim. There are other candidates besides those offered by the 1982 Commission, Lebaqz or Brody. Barriers to cross-species reproduction that exist in nature represent a divine or metaphysical order that is not to be breached; interference in the natural mechanisms of reproduction is morally wrong because it violates God’s will; movement of genes disrespects the current generations’ debt to the ancestors. It is not difficult to find common people who believe such things. Anyone who cares to strike-up a conversation with strangers on airplane or among members of the local church will be able to add anecdotal testimony to the existence of this belief among the general populace. The early Hoban and Kendall survey on public attitudes to genetic engineering in the United States showed that many adults cite religious motivation for concern about genetic engineering, (Hoban and Kendall 1993). Yet anecdotes and statistical surveys are even more “inchoate,” than Brody’s expert witnesses. The Reverend Wesley GranbergMichaelson is typical. He will say that “The Judeo-Christian view says that… there are limitations on what we can do,” (quoted in Brody 1989, 144), but he does not explicitly say that genetic engineering and molecular genetics research violate those

11.4 Analyzing the Religiously Metaphysical Case Against Genetic Engineering

297

limits. There is thus a profound sense in which philosophical ethics reaches its limits in confronting these arguments. However, a few things can be said. First, it seems likely that many of the lay informants that cite religious values in their concerns about genetic engineering also hold metaphysical beliefs that flatly contradict fundamental tenets of evolutionary biology. Like creationists, they must regard biology’s metaphysical claims as merely speculative, as nothing more than tools for organizing work, at best, and as false and probably dangerous beliefs at worst. It is tempting to dismiss these views as philosophically naive. Bernard Rollin is willing to write off all religious opposition to biotechnology as being committed to a view of nature that is entirely unsupported by modern science (Rollin 1986). Yet we should remember that were metaphysical disputes to be decided by vote, the party of evolutionary biology would almost certainly be a minority one. The potential for anti-scientific radicalism among the religiously conservative is unknown, and the implicit attitude of the scientific community seems to be, ‘Let sleeping dogs lie.’ This may be good advice from a political perspective, but surely, an ethically responsible perspective on food biotechnology must find it at least a bit duplicitous. If we have an ethical responsibility to communicate respectfully with the wider public (the subject of the next chapter), it will be necessary to find some way of engaging religiously conservative beliefs in a respectful manner. Second, the likely view of this silent majority notwithstanding, it is possible to formulate a religious argument against human intervention in reproductive processes without contradicting the basic metaphysical claims of evolutionary biology. Such an argument might take several forms. One example comes from Lane Lester and James Hefley’s argument in support of the conclusion that “cloning is not God’s way.” Lester and Hefley offer a religiously conservative argument that rejects the importance of any conflict between religion and biology, and that supports their conclusion by substituting a religiously metaphysical claim that is logically compatible with evolution and molecular genetics. They argue that the biblical objection to genetic manipulation is “the lack of a normal family background,” (Lester and Hefley 1980, 60). Citing Genesis 1:27–28, they conclude, “Clearly God intended society to be built on the two-parent family,” (Ibid.). It is a matter for speculation how these authors would view food biotechnology, but their argument stands as an example of how one might formulate a religious argument without contradicting the factual claims of biologists. Although I do not endorse this pattern of reasoning, it exemplifies a metaphysical form that is logically compatible with a reductive understanding of molecular genetics. One example is enough to show that the form is logically possible. Finally, if ambiguous statements that gesture vaguely at religiously metaphysical beliefs are indeed part of the social fabric, there are secularly ethical reasons not to insist on a translation that allows a more conventionally philosophical analysis of the arguments to proceed. First, insistence upon translation into the languages of welfare, rights and duties or virtues and vices disrespects those who sincerely believe that a mystical force or supernatural entity lies at the root of their concern. Not only will such translations fail to express what believers actually mean to say, they have

298

11 Religiously Metaphysical Arguments …

the potential to threaten the communicative basis on which further conversations could proceed. This points us to a second point. Adverting to Aristotle’s Rhetoric, the public morality that supports and legitimizes political action may not conform to the standards of philosophical rigor. “Political action” should, in this context, be understood to include not only regulation but also the structure of market relations that allows for the deployment of gene technology in agricultural production and food processing.

11.5 Religious Statements on Genetic Technology Through 2005 Given this orientation, it is appropriate to review some statements on gene technology made through the official channels of religious organizations. Most religious groups that have made official statements on genetic engineering are primarily concerned with the possibility of human eugenics, but they express this concern in language that applies to food biotechnology, as well. The United Methodist Church (1992), for example, adopted a resolution on genetic science at its General Conference, which reads in part: Failure to accept limits by rejecting or ignoring accountability to God and interdependency with the whole of creation is the essence of sin. Therefore, the question is not can we perform all prodigious works of research and technology, but should we? (p 2)

The resolution goes on to endorse genetic technologies in general, and to qualify their application largely along lines that conform to technological ethics, though it does oppose patents on organisms based upon, “the sanctity of God’s creation and God’s ownership of life” (p 5). The Methodist declaration was preceded by a ten-page report from a special task force on genetic science. Significantly, almost one third of this report deals with food biotechnology, including discussion of the impact of recombinant bovine somatotropin, herbicide tolerant crops and release of genetically engineered plants into the environment (United Methodist Church 1991). The task force treated these issues in much the same way that they have been treated in Chaps. 4 through 7 of this book, which is to say that they noted no special metaphysical circumstances or principles that would produce an evaluation of the products in question that differs from evaluations based entirely on secular grounds. The task force report offered an endorsement of agricultural genetic engineering, subject to two qualifications. First, the report stressed public input into the planning and distribution of benefits from food biotechnology. Second, the task force urged that food biotechnology promote the sustainability of family farms, natural resources and rural communities (United Methodist Church 1991, 122). The Methodists have continued to make recommendations on gene technology, with a 2016 revision of the 1992 resolution that calls for public involvement in policy and planning for agricultural biotechnology and careful attention to environmental and food safety risks (United Methodist Church, 2016).

11.5 Religious Statements on Genetic Technology Through 2005

299

The July 1989 recommendation adopted by the Central Committee of the World Council of Churches is more focused on medical applications of biotechnology. It calls for a prohibition of testing for sex selection, proposes a ban on experiments involving the human germ line and makes several other statements on human reproduction. The recommendations oppose patenting of animal life forms, and call for “swift adoption of strict adoption of strict international controls on the release of genetically engineered organisms into the environment,” (World Council of Churches 1990) No specific theological rationale for these recommendations is reported. The United Church of Christ (UCC) offered a statement that was broadly supportive of food biotechnology, stating that “Genetic engineering gives us new ways to relieve suffering and increase food production.” The UCC concludes, “We support the application of genetic engineering to agriculture, forestry, mining and pollution control, provided there is adequate regulation and public participation in evaluating new uses,” (United Church of Christ 1990) The 1984 report on genetic technology from the National Council of Churches of Christ also notes a number of the ethical concerns discussed in the preceding eight chapters, and places the ethical critique within the context of religious faith. The report states that scientific findings and theories, “neither annul, displace, nor validate the belief in divine creation.” (NCCC 1984, 22) The report continues, A high testimony to the value of each created human life and of all humanity was, and remains, the act of Incarnation. This is one of the foundation stones of the Christian faith. The life that was blessed by being created in the image of God was confirmed and ratified by the becoming-human of the eternal Word of God in Jesus the Christ. In Jesus Christ human kind is re-created and renewed. This rejuvenation supplies force for the Christian witness to the original goodness and value of human life. Life is the created gift of God: that conviction can be further enhanced in this world and made eternal by God’s action in Christ. For these reasons, each and all human life is to be held in high respect. Traditionally, then, Christian theology regards the effect on human life as the primary theological criterion for making ethical judgments about genetic science (NCCC 1984, 22–23).

The report thus marries a moral principle that could serve as the basis for a thoroughly secular technological ethics (that the effect on human life is the primary criterion for making ethical judgments) to a metaphysical statement about God’s role in creation and about the divine status of Jesus. The report continues with a series of statements that rest on this “primary theological criterion.” Immediately following, for example, the report states: We know that we and all human beings should be responsible for unborn generations of humanity. The human gene pool—that is, the totality of genetic material available for reproduction—is in danger of corrupting its offspring through imprudent, excessively risky genetic modifications (NCCC 1984, 23).

Panelists warn of risks and stipulate moral principles for weighing risk and benefit that are entirely consistent with the secular analysis presented in other chapters of this book. Although this analysis is interlaced with statements such as “Life is holy because God is holy,” (NCCC 1984, 23), the effect of these metaphysical statements is to reinforce the moral authority of the panel’s concern for effects on human life.

300

11 Religiously Metaphysical Arguments …

Nowhere does the report state that modification of genes could be wrong except in virtue of its impact on the quality of human life. This admittedly limited survey of statements on gene technology from church organizations suggests two broad philosophical routes to such religiously based statements (pro or con) on genetic engineering. One proposes metaphysical statements that bear directly on the morality of genetic modification, irrespective of its technological consequences, but the other follows the form of philosophical analysis plotted throughout this book. The difference is that a particular philosophical position regarding technological ethics is articulated in traditionally religious language. Clearly, a number of theologians writing on public policy issues reject the philosophical foundations of utilitarian philosophy. It follows that they would discount the moral importance of the consequentialist or trade-off arguments that have been reviewed in previous chapters. The literature on abortion and on the use of nuclear weapons for deterrence exhibits this pattern of reasoning far more evidently than the literature on biotechnology (see Finnis 1987). Analyzing a number of religious statements on genetic engineering, Audrey Chapman, then of the American Association for the Advancement of Science, notes that they “do not engage in systematic and extended theological reflection as a basis for drawing ethical and policy guidelines,” (quoted in Nelson 1994, 183). Chapman also notes that the statements are often vague and unspecific even when they do attempt to make pronouncements. The Church of Scotland’s Society, Religion and Technology Project is especially remarkable given religious organizations’ usual tendency to superficiality and incompleteness. A permanent and professionally staffed activity of the Church beginning in 1970, the Project formed a working group on ethical issues associated with genetic engineering of non-human species that published its 337-page report under the title Engineering Genesis in 1998. This report does not represent a doctrinal statement on the part of the Church of Scotland. Rather it serves as an example of an alternative approach that delimits a number of issues and indicates how the Christian faith tradition can be brought to bear upon them. Most of Engineering Genesis deals either with characterizing the scientific subject matter (which includes non-agricultural applications such as the development of animal models for medical research) or with considerations that have been characterized here as general matters of technological ethics, that is, questions of risk, consent and the procedures and institutions that govern technology and research. The report does single out a number of religious and metaphysical objections to biotechnology, which are characterized as “intrinsic” ethical arguments. Five types of intrinsic argument are subjected to detailed discussion in Engineering Genesis (discussed below): (1) playing God; (2) natural or unnatural? (3) relationships; (4) trans-species gene transfer; and (5) the status of animals, plants and microorganisms (Bruce and Bruce 1998). An earlier attempt to raise the level of religious debate about genetic technology was coordinated by the Institute of Religion at the Texas Medical Center between 1990 and 1992. The effort coordinated working groups on genetics issues at several sites. Although the focus of this effort was heavily oriented toward medicine, genetic counseling and genome mapping, one study group formulated a survey that asked

11.5 Religious Statements on Genetic Technology Through 2005

301

respondents to represent their faith orientation to questions on the acceptability of genetic engineering applied to plants and animals. The anonymous Jewish respondent stated, “Cross species fertilization is prohibited,” though he or she continues with the statement “genetic crossbreeding may be less of a problem… Submicroscopic actions are often not culpable even when macroscopic activities yielding the same results are.” The respondent qualifies this, however, by saying that mystical Jewish communities “would be far more troubled with the new species or trans-species creation no matter how human beings created them,” (Seydel and coauthors 1992, 35) An anonymous Protestant respondent to the survey notes that “Within Protestantism generally there are no objections to selective breeding, cross species fertilization, or the creation of “new” species,” (quoted in Seydel and coauthors 1992, 60) The respondent goes on to note a number of specific product-related concerns that are typical of technological ethics. The Episcopal respondent answered all questions relating to plants and animals with the following statement: We are stewards rather than manipulators of God’s creation. Something initially beneficial later can lead to unforeseen detrimental effects, i.e. green revolution; development of new species and loss of older ones and subsequent disease/failure of “new species.” Ecological balance could be changed. All knowledge can be misused, (quoted in Seydel and coauthors 1992, 68). It seems reasonable to interpret these statements as a generally favorable response with respect to the religious acceptability of food biotechnology.

The prominent theologian J. Robert Nelson (1920–2004) was the convener of the Houston conferences. He collected and annotated material (but not principal addresses) collected at the conferences in an unusual book entitled On the New Frontiers of Genetics and Religion (Nelson 1994). The book also reprints a number of personal statements on new reproductive technologies written from different faith perspectives, along with selection of ecumenical and denominational statements on genetic engineering. It is thus an excellent starting point for readers wishing more information on the matters reviewed above in this section. However, the book is of even more limited relevance to food biotechnology than are religious statements in general because Nelson edited the results of the Houston conference to concentrate narrowly on medical technologies and medical genetics, omitting all direct discussion of food and environmental issues. Nelson may have believed that disquiet evident in some comments on agrifood biotechnology would distract from what he took to be more serious issues. The religious community’s tendency to apply the blue pencil toward each other’s expressed views is a continuing problem for those who would wish to derive some insight into the faith basis for beliefs about food biotechnology.

11.6 The Religious Case Against Property Rights in Genes: 1985–1995 Religious arguments were prominent in U.S. debates over the extension of intellectual property rights (IPR’s) in genes, gene sequences or in whole genomes, especially

302

11 Religiously Metaphysical Arguments …

during the decade when IPR’s were in their most intensive period of Congressional review and political discussion, making the debate over IPR’s a useful test bed for probing religious attitudes to biotechnology.. For example, several religious leaders testified against IPR’s in transgenic animals prior to passage of the Transgenic Animal Patent Reform Act in 1989. Testifying on behalf of the National Council of Churches, the Reverend Wesley Granberg-Michaelson cited a biblical responsibility to preserve the integrity of creation. In calling for a moratorium on animal patents, he described patenting of transgenic animals as “an unprecedented shift in humanity’s relationship to the God-given natural environment,” (U.S. House of Representatives 1988, 201). Rabbi Michel Berenbanm joined in calling for a moratorium, resting his case on the distinction between “what constitutes life and what is merely an inert manufactured commodity,” (U.S. House of Representatives 1988, 202). The issue of patents and ownership again became the flashpoint for religious opposition to biotechnology in 1995. Religious leaders from Roman Catholic, Jewish, Muslim, Hindu and both mainline and evangelical Protestant churches in the United States issued a statement opposing patenting human and animal genes. The story made the front page of the New York Times on May 13, where Richard Land of the Southern Baptist Convention was quoted saying, “This issue is going to dwarf the pro-life debate within a few years.” The statement itself did not include a rationale, stating only that the signatories “oppose the patenting of human and animal life forms,” but subsequent stories on the statement indicated that for most signatories, opposition was based on the belief that the basic units of life are sacred and demeaned by patenting. Catholic Bishop William Friend (1931–2015) was quoted in the New York Times saying that alteration of human genes “compromises the incomparable dignity of the human species.” Friend and many Catholics would accept the patenting of plant and animal genes (Andrews 1995). While these statements reveal the seriousness of religious opposition to at least some forms of genetic technology, the arguments reported in the religious press are vague. They were opposed by faculty from religious or theological institutions such as Leroy Walters, Theodore Peters and Ronald Cole-Turner, each of whom argued that although new genetic technologies should provoke broad societal reflection on moral issues, reactionary opposition to genetic engineering must not be allowed to prevent or postpone research and development that could produce compelling benefit to humans. If intellectual property rights in genes and genomes would hasten such benefit, these theologically oriented authors support them. These arguments on behalf of IPR’s reintroduce a pattern of philosophical argumentation discussed in Chap. 10, and testify, again, to the way that technological ethics bleeds over into the religious and metaphysical realm. The vagueness of religious language reported in the press presents a challenge to philosophical analysis. We may speculate that one of two lines of reasoning undergirds religious opposition to patents. One, that manipulating genes is itself wrong, will be discussed at greater length below. Another viewpoint might countenance the manipulation of genes for purely humanitarian purposes, but balks at the institutionalization of IPR’s in virtue of the seeming commercialization of life that this entails. In this view, the boundary that is threatened is not the species boundary between human

11.6 The Religious Case …

303

beings and other animals, but the boundary between sacred and profane domains of human practice. Delineation of the sacred represents a crucial moral boundary for virtually any religious cosmology or view of nature. To call a place or an activity sacred is to distinguish it from ordinary places or activities, and to denote the importance of respecting special rules, or of adopting a proper attitude with respect to the place or activity. In many religious traditions, but especially within the Judeo-Christian tradition, marking a place or activity as sacred denotes the impropriety of ordinary commerce within or with respect to it. Mark 11:15–17 and Luke 19:45–46 describe how Jesus of Nazareth (4 BCE–30/33) drove money changers from the temple in Jerusalem. In Luke, Jesus chastises the priests for profaning the temple, turning it into “a den of thieves.” These passages from the Christian Gospel emphasize the inappropriateness of commerce within sacred places, and underline the distinction between the sacred and ordinary commercial activity. The way that any religious group understands the boundary between sacred and profane (or ordinary) activities is, of course, contingent upon religious beliefs and traditions that vary not only from culture to culture, but also from time to time and place to place even with a coherent religious tradition. Yet we should not be surprised when religious believers classify human activities that involve food and fertility among the sacred. The act of “saying Grace,” or blessing the meal just before it is consumed is a form of re-sacralizing foods that may have been bought and sold prior to human consumption. Even more significantly, ensuring fertility in all of its manifestations—human, livestock, plant and soil—is a common theme of sacraments in all religious traditions. Twentieth century theological traditions display considerable ambivalence toward fertility-based metaphysics, wanting no part of religious beliefs that sanction such practices as idolatry and even human sacrifice in order to promote fecundity. Nevertheless, it would be surprising if the most basic processes of reproduction for humans, animals, plants and microbes were not associated with some residual feelings of sanctity. Religious objection to IPR’s on genes, sequences and genomes might therefore be built on an argument with two metaphysical premises. First, it would be necessary to proscribe commercial transactions within the domain of the sacred, that is, to interpret the sacred as violated or transgressed either by specific transactions or (more plausibly) when sacred activities become pervasively characterized by the buying and selling of alienable goods. Second, it would be necessary to designate the reproduction of human, animal, plant or microbial organisms as a sacred process. It would be logically possible to designate only some subset of these organisms as having sacred significance; hence supporting the judgment that commercialization of plant reproduction is morally acceptable, while IPR’s for humans or animals are not. Although the details of these two premises would need to be specified through theological arguments specific to a given religious tradition, it is quite plausible to think that such arguments could be offered and found compelling by the faithful. Such an argument would not necessarily proscribe the application of genetic engineering in the improvement of plants or livestock, or in developing techniques for food processing, nor need it preclude the normal commercial exchange of food and fiber

304

11 Religiously Metaphysical Arguments …

commodities produced using food biotechnology. It might, however, support the conclusion that IPRs in genes and gene processes constitute an abrogation of sacred boundaries, and produce a strong religious case against property rights.

11.7 Academic Theology and Genetic Engineering Theology within the setting of scholarship and university research is a discipline unto itself, with obvious connections to philosophy, but with a distinct literature and its own technical concepts, traditions and scholarly language. Scholarly academic theology often diverges from the religious beliefs of the laity, of the active clergy, and even from the doctrines of organized churches. While a layperson might wonder about the ethics of food biotechnology because of an immediate concern about the propriety of eating a given food, or buying a given seed, the question of genetic engineering emerges for academic theologians as one piece in a larger set of questions about the relationship between science and religion. At bottom, these questions do (or should) inform or impinge upon the more practically focused beliefs of the layperson, the parish priest, or the church functionary, but the pattern in the scholarly literature is to pursue much larger metaphysical themes. A 1986 book entitled God and the New Biology struggles with the implications of the reductionist turn in biology, and with the theological implications of statistical explanations, (Peacock 1986). The index does not include entries for ‘biotechnology,’ or ‘genetic engineering,’ because these topics do not arise. Two examples of academic theology, Wolfhart Pannenberg’s Toward a Theology of Nature (1992) and Langdon Gilkey’s Nature, Reality and the Sacred (1993) represent very different approaches to their subject matter undertaken by two theologians with a lifetime of work on the subject. Where Gilkey (1919–2004) softens tensions between religion and science, calling for an understanding of the sacred that does not challenge the metaphysical presuppositions of the sciences, Pannenburg (19282014) attacks the metaphysical presumptions of science head on, provoking their advocates to modify (or more rigorously defend) all metaphysical beliefs, whether based on religion or science. Neither of these distinguished theologians examines the implications of their theology for genetic modification, much less food biotechnology. Deriving such implications from their more general views would itself be a task of theological scholarship. As noted, academic theologians have been supportive with respect to the ethical acceptability of genetic research and its biomedical applications. Georgetown University’s Leroy Walters served on the National Institutes of Health’s Recombinant DNA Advisory Committee for many years, but his approach to bioethics is not explicitly theological (see Walters 1978). Walters is an example of a theologian who works largely within the sphere what has here been called standard technological ethics. Ron Cole-Turner’s The New Genesis is a short but systematic enquiry into the theological implications of recent work in molecular genetics. Turner accepts factual allegations of molecular biology at face value, and displays equal interest in the importance of

11.7 Academic Theology and Genetic Engineering

305

these facts for theological doctrines of creation and for the ethical implications of genetic technologies in biomedical applications. For the latter questions, he supplies an ethical analysis that is entirely consistent with technological ethics: concern for human dignity, for rights and consent, but no religiously based proscription of genetic engineering for purposes that benefit human beings (Cole-Turner 1993). Ted Peters is another theologian active in discussions of medical biotechnology. As described in a 1984 paper, his ethical approach is explicitly theological, based on “proleptic eschatology.” This means that Peters advocates an approach to enquiry that begins with a vision a future, harmonious global community, based on the lessons revealed in the life of Christ, then defines ethical action as that which will realize or bring about that vision, (Peters 1984). Peters has been active in organizing symposia on the Human Genome Project, and has worked to keep religious groups and denominations open to the beneficial applications of biotechnology, primarily with respect to human health, but regarding food and agriculture, as well. A recent paper applies his approach to the debate over stem cell research. There he takes conservative theologians to task, arguing that their unresponsiveness to the vision offered by the medical research community and to the theological arguments offered by more liberal theologians (such as himself) is indicative of an intellectual closure that is inconsistent with the Christian religious tradition. Though Peters has not addressed food and agricultural biotechnology specifically, his approach would appear to be broadly accepting of it on theological grounds. Andrew Linzey is one theologian who has maintained a steady focus on nonhuman issues. Linzey has produced a complex theological argument for animal rights that (1) rebuts theological views (such as those of Santos and the NCCC) that see humans as unique, that (2) appeals to a particular conception of Christ’s lesson for humanity, and that 3) interprets animal suffering as both morally and theologically meaningful. He asks and (unlike many) answers the key theological question: “What does it mean for humans to exercise a priestly role of redemption? Quite simply: it concerns the releasing of creation from futility, from suffering and pain, and worthlessness,” (Linzey 1995, 55). This theology produces a religiously based case for animal liberation. Linzey concludes his book Animal Theology with a polemical chapter that equates genetic engineering of animals with the sin of human slavery. Yet this chapter does not claim that genetic engineering is wrong when applied to plants, nor does it depend upon metaphysical beliefs about the sanctity of species or of reproductive processes. Linzey’s opposition to genetic engineering is entirely a consequence of his views on the need for radical revision of the relationship between humans and animals. Genetic engineering comes in for his wrath not because it appears to constitute any novel threat not posed by animal experimentation, meat-eating or hunting (three other activities discussed) but because it represents a new instance of the same old threats. In Linzey’s view genetic engineering requires that we see animals as exploitable for human uses. He likens this to discredited metaphysical and theological arguments that were once put forward in defense of human slavery, (Linzey 1990). Since genetic manipulation is, in Linzey’s view, premised on this morally

306

11 Religiously Metaphysical Arguments …

indefensible attitude toward both human and non-human animals, it is itself indefensible. Precisely because Linzey does not base his concerns on species boundaries that must not be crossed or the moral significance of genes and genetic processes, his view is a poor model for the kind of religious beliefs that are (speculatively) at the root of lay concerns. It seems unlikely that many of the lay informants who have expressed qualms about genetic engineering of animals would go as far as Linzey in reformulating humanity’s relationship with the animal world. In the end, however, Linzey’s argument becomes curiously muddled. On the one hand, he wants to claim that genetic engineering itself merely extends a theologically indefensible attitude toward animals. On the other, he saves his most explicit rhetoric for a critique of property rights in transgenic animals. He writes, “No human being can be justified in claiming absolute ownership of animals for the simple reason that God alone owns creation,” (Linzey 1995, 148, italics in the original), but while Linzey seems to have special animosity toward IPR’s for animals, his position entails far more radical changes for ordinary chattel property rights than for IPR’s and genetic engineering. If creating transgenic animals is itself wrong, why focus the invective on IPR’s? Perhaps the answer is that thinking of genetically engineered animals in terms of property is simply the most extreme example of the attitude Linzey wishes to decry. In any case, it is difficult to infer any basis for discussing the ethics of transgenic animals outside of a system of property rights from Linzey’s rhetoric. The most plausible way to read him is as proposing a categorical rejection of genetic engineering applied to animals. Yet gene technologies aimed a preventing genetic disease would appear to further his goal of releasing creation from futility, suffering and pain. The 1997 edition of Food Biotechnology in Ethical Perspective noted the paucity of theological literature dealing with food or agricultural biotechnology prior to 1997. A 1995 bibliographic survey of literature entitled Ecology, Justice and Christian Faith surveyed more than 500 books and articles in the scholarly journals of religion. The compilers listed only five publications that take up biotechnology. Three of these are books that address environmental risks of food biotechnology from a perspective of religiously-based technological ethics (Barbour 1993; Shinn 1991; Meyer and Meyer 1991). A fourth is an article by Deiter Hessel that takes the same approach. Hessel argues that biotechnology would be permissible only if it is carried out by those who view the human vocation as one of living in harmony with nature, rather than as extending the Baconian project of human power over nature. While these authors identify themselves as theologically oriented, the arguments they deploy do not rely on theological views of God’s intentions regarding species boundaries. Those who adopt the eco-centric worldview would apply the norms of technological ethics discussed in Chap. 7 for theological reasons (Hessel 1993). The last bibliographic entry is a paper by Richard Chambers entitled “Plant Breeders’ Rights and the Integrity of Creation.” In making a fairly strong statement against genetic engineering, Chambers rejects the pragmatism implicit in denominational statements on genetic engineering, calling instead for a theological basis. This paper, prepared under the auspices of the World Council of Churches (WCC), references another WCC document as pointing toward the sought for theological foundations,

11.7 Academic Theology and Genetic Engineering

307

but Chambers does not believe these foundations are clearly or adequately articulated in the document, (Chambers 1988). The WCC authors call for a theologically grounded view directed specifically toward agricultural and food applications of biotechnology, but in large measure, that call is not answered. Following the announcement of Dolly—the first cloned sheep—in early 1997, theological debate over biotechnology began to shift slightly in the direction of non-human applications. Though not specifically a work of academic theology, the previously mentioned Engineering Genesis provides a thorough discussion of food and agricultural genetic engineering that includes citations and argumentation typical of an academic study. The authors follow a standard philosophical convention of distinguishing between intrinsic and consequentialist arguments against agrifood biotechnology. While the latter involve risks, costs and benefits that must be reflected in the trade-off balancing or utility optimizing style of ethical reasoning typical of utilitarianism, intrinsic concerns are intended to represent reasons that should not be subject to this style of thought. The intrinsic concerns discussed in Engineering Genesis tend not to rely on rights arguments such as those discussed in previous chapters, and in fact cite a number of religious and metaphysical objections to genetic engineering as it might be applied to plants, animals and microbes. For example, the authors of Engineering Genesis characterize the phrase “playing God” as a call for humility and a theologically based proscription of human interference in God’s plan for humanity. The report counters this call by noting that, although the Christian tradition does indeed endorse humility, it also endorses human creativity in the manipulation and transformation of nature in the service of God’s will. A similar line of argument is applied to the suggestion that genetic engineering might be “unnatural”: Christian traditions provide sources for constraining human beings abuse of the natural world, but also for seeing scientists’ work with genetic engineering as consistent with duties to serve as stewards of nature, (Bruce and Bruce 1998). The pattern of argument in Engineering Genesis is to note the diversity and complexity of Christian theology, and to indicate that grounds can be found both for opposing and for endorsing genetic engineering in agriculture and food. The overview of intrinsic concerns is followed by a lengthy discussion of issues relating to animal welfare, social justice, environmental impact and the safety and autonomy of consumers—a discussion not unlike the one undertaken in the first nine chapters of this book. The upshot seems to be a perspective on agrifood biotechnology that is, in the final analysis, accepting and cautiously optimistic, while also noting the seriousness with which standard issues of technological ethics must be addressed as products are developed. If this is correct, then the primary significance of noting intrinsic concerns in Engineering Genesis is to endorse the need to engage those who express these views in respectful Christian debate, and to ensure that they have ample opportunity as individuals to live within the dictates of their personal faith. As food and agricultural biotechnologies sit upon the threshold of new applications made possible through gene editing, there is some evidence that the enthusiasm for theologically based critiques is declining. In the United States, the 2010 Presidential report on synthetic biology and the 2016 National Academies report on gene drives both mention the potential for religiously based opposition, but move quickly to

308

11 Religiously Metaphysical Arguments …

sidestep the arguments, following the previously discussed strategy of the 1982 effort (Presidential Commission, 2010; NASEM, 2016). The Nuffield Council’s 2016 report on gene editing does not even consider religiously based ethical issues (Nuffield Council 2016). Although J. Craig Ventor’s announcement of an artificially sequenced microorganism did precipitate a flurry of theologico-ethical reflections, the prospects for gene editing using CRISPrCas9 have not thus far sparked a similar outburst. A full treatment of synthetic genomes is beyond the scope of the present chapter, but many of the religiously-based concerns noted for earlier generations of agrifood biotechnology are being repeated and revisited.. Ted Peters is back with a new paper calling (once again) for collaboration between theologians and secular ethicists. His focus, as in past years, is primarily on applications in human medicine, rather than food, (Peters 2019). The upshot, then, is that theologians and theologically oriented scholarship has yet to undertake metaphysically based concerns about agrifood biotechnology in a direct and straightforward manner. Those who take a more conservative theological perspective seem focused on biomedical applications, and to the extent that species boundaries are the concern, it is the boundary between the human species and all others that ought not to be crossed. More liberally minded theologians seem to be anxious to place some distance between themselves and their conservative colleagues. The mainline churches’ opposition to animal patenting that occurred in the 1980s and early 1990s appears to be something of an embarrassment to theological liberals, who now appear reluctant to say anything critical about applications of genetic techniques in the food and agricultural sector. If there is a middle ground, it is occupied by those such as Linzey, whose opposition to biotechnology has little to do with genetic manipulation as such, or the Church of Scotland’s Engineering Genesis group, who emphasize openness to a diversity of theological viewpoints.

11.8 The Ethical Implications of Religious Views Religious claims seem to turn on normative evaluations of boundaries. In this respect, the arguments extend a longstanding tradition in ethical thought. Ethical judgments throughout history have placed emphasis upon group boundaries. At its most primitive, ethical responsibility may have been confined to intragroup loyalties. The ancient Greeks, inventors of Western philosophical traditions, defined ethical responsibility according to a hierarchy in which the stringency of obligation is reduced as one moves from the family to the city-state, from the city-state to Hellenic peoples, from Greeks to other humans, and from humans to the balance of nature (MacIntyre 1988). Boundaries that establish rights, privileges and obligations within a hierarchical scheme were implicit components of morality in virtually every human society. Members of the nobility would owe duties to one another that were not owed to commoners; men owed duties to other men that they did not owe to women. Practices that challenge implicit boundaries take on ominous significance, particularly for those who imagine morality to depend upon a tightly knit fabric of personal

11.8 The Ethical Implications of Religious Views

309

norms, philosophical and religious justifications, and social reinforcement. However, systematic ambiguity runs throughout the attempt to base opposition to biotechnology on the two implications of boundaries outlined above. On the one hand, biological boundaries imbue humans with a unique status; on the other hand, ethical boundaries establish constraints on conduct. Logically, these may be entirely separable issues. Raising questions about the biological uniqueness of humans does not, on the face of it, also imply a questioning of absolute constraints on conduct. In this, modern biology’s “threat” to social order is somewhat imaginary. However, it is possible that the two questions are theologically, psychologically, or socially linked in some way that has gone un-explicated (or overlooked in my scholarship). The burden of proof should fall upon religious critics who wish to make use of a boundary argument, (see Evans 2013). The concern, clearly, of the science community is that the faithful will draw the boundary so that all use of gene technology for research and product development is found unacceptable. In fact, however, religious critics have more typically opposed patent protection for genes, sequences and gene products. Even theologically less dramatic arguments might nonetheless have extreme consequences for food biotechnology. One might conclude that fertility and reproduction are sacred in ways that preclude genetic engineering, not just commercial ownership of genes through IPR’s. Here, as in Lebaqz’ reading, the emphasis would be on the moral limitations on intentional actions of human beings, not on God’s role in the establishment of natural order. The argument would be much like the one outlined above with respect to sanctity and property rights, save that it would identify any human interference in reproductive processes as a violation of the sacred, rather than simply commercially motivated actions. Such a characterization of the sacred would have to be based on sacred texts or other theological sources unique to the faith tradition in which the argument is developed. This is evidently not an argument that many theologically trained individuals are inclined to make, but it appears to be a philosophically coherent option for those who would turn inchoate concern into a strong prohibition of genetic engineering. It would advance our understanding of ethical issues respecting genetic engineering for those whose religious views are so inclined to attempt an explicit statement of such an argument, if only to clarify and sharpen the terms of debate. Finally, Baruch Brody’s observation on animal patents may be the most important point to take away from any consideration of religious objections. It is the inchoate character of religious opposition that is its most significant fact. Religious leaders had more than a century to adjust to Darwin, yet it seems they are being expected to accommodate important social and political implications of genetic engineering in less than a decade, and with considerably fewer resources for deliberation and debate. The point is that all parties should attempt such arguments as can be made, and to have some patience with those whose views are not fully formed. The fact that a potent theological objection has not been raised thus far is only the weakest sort or evidence that it will not be raised sometime in the future. Developers of food biotechnology cannot wait forever before going ahead with their products, but they can and should assist religious leaders and religious groups in formulating and articulating their views, however inimical to the products of food biotechnology they

310

11 Religiously Metaphysical Arguments …

might, at the outset, appear to be. Only in this way can the inchoate be verbalized, and only when verbalized can the objections of the religious be evaluated and met on informed and rational grounds.

References Andrews, E.L. (1995) Company News: Religious leaders prepare to fight patents on genes, The New York Times, May 13. Accessed Jan. 1, 2020 at https://www.nytimes.com/1995/05/13/us/com pany-news-religious-leaders-prepare-to-fight-patents-on-genes.html. Barbour, I. (1993) Ethics in an Age of Technology. The Gifford Lectures 1989–1991, vol 2. Harper Collins Publishers, New York. Brody, B.A. 1989. An evaluation of the ethical arguments commonly raised against the patenting of transgenic animals. In Animal Patents: The Legal, Economic and Social Issues, ed. W. Lesser, 141–156. New York: Stockton Press. Bruce, D., and A. Bruce (eds.). 1998. Engineering Genesis: The Ethics of Genetic Engineering in Non-Human Species. London: Earthscan Publications. Chambers, R.G. 1988. Plant breeders’ rights and the integrity of creation. In Creation and the Kingdom of God, ed. D. Gosling, G. Limouris, and F. Rajotte, 36–42. Geneva: World Council of Churches. Cole-Turner, R. 1993. The New Genesis: Theology and the Genetic Revolution. Knoxville, TN: Westminster/John Knox Press. Deane-Drummond, C.E. 1995. Genetic engineering for the environment: Ethical implications of the biotechnology revolution. The Heythrop Journal: A Quarterly Review of Philosophy and Theology 36: 307–327. Evans, J.H. 2013. “Teaching humanness” claims in bioethics and public policy. In Synthetic Biology and Morality: Artificial Life and the Bounds of Nature, ed. G. Kaebnick and T. Murray, 177–204. MA: The MIT Press, Cambrdige. Finnis, J. 1980. Natural Law and Natural Rights. New York: Oxford University Press. Gilkey, L. 1993. Nature. Augsberg Fortress, Minneapolis, MN: Reality and the Sacred. Hájek, A. (2018) Pascal’s wager, The Stanford Encyclopedia of Philosophy (Summer 2018 Edition), Edward N. Zalta (ed.), URL = . Hessel, D.T. 1993. Now that animals can be genetically engineered: Biotechnology in theologicalethical perspective. Theology and Public Policy 5 (Summer): 40–54. Hoban, T.J., and P. Kendall. 1993. Consumer Attitudes about Food Biotechnology. Raleigh, NC: North Carolina Cooperative Extension Service. Kaebnick, G.E. 2009. Should moral objections to synthetic biology affect public policy? Nature Biotechnology 27: 1106–1108. Kass, L. 1985. Toward a More Natural Science. New York: Free Press. Kimbrell, A., and J. Rifkin. 1987. Biotechnology—A proposal for regulatory reform. Notre Dame Journal of Law, Ethics and Public Policy 3: 117–143. Lebaqz, K. (1984) The ghosts are on the wall: A parable for manipulating life, In: R. Esbjornson (ed) The Manipulation of Life. Harper and Row, San Francisco, pp 22 41. Lester, L.P., and J.C. Hefley. 1980. Cloning: Miracle or Menace?. Wheaton, IL: Tyndale House Publishers. Linzey, A. 1990. Human and animal slavery: A theological critique of genetic engineering. In The Bio-Revolution: Cornucopia or Pandora’s Box, ed. P. Wheale and R. McNally, 175–188. London: Pluto Press. Linzey, A. 1995. Animal Theology. Urbana: University of Illinois Press.

References

311

MacIntyre, A. 1988. Whose Justice? Which Rationality?. Notre Dame, IN: University of Notre Dame Press. Meyer, A., and J. Meyer. 1991. Earthkeepers: Environmental Perspectives on Hunger. Poverty and Injustice: Herald Press, Scottdale, PA. NCCC (Panel on Bioethical Concerns, National Council of the Churches of Christ, USA). 1984. Genetic Engineering: Social and Ethical Consequences. New York: The Pilgrim Press. Nelson, J.R. 1994. On the New Frontiers of Genetics and Religion. Grand Rapids, MI: William B Eerdmans Publishing. NASEM (National Academies of Sciences, Engineering and Medicine). (2016) Gene Drives on the Horizon. National Academies Press, Washington, DC. Nuffield Council on Bioethics. (2016) Gene Editing: An Ethical Review. Nuffield Council, London. Accessed Jan. 1, 2020 at https://nuffieldbioethics.org/assets/pdfs/Genome-editing-an-ethical-rev iew.pdf. Nugent, C. (2018) What was it like to grow up as the world’s first ‘test tube baby’? Time Magazine Accessed Dec. 31, 2019 at https://time.com/5344145/louise-brown-test-tube-baby/. OTA (U.S. Congress Office of Technology Assessment). 1988. Patenting Life. Office of Technology Assessment, Government Printing Office, Washington, DC: US Congress. Pannenberg, W. (1993) Toward a Theology of Nature. T. Peters (ed) Westminster/John Knox Press, Louisville, KY. Peacock, A. 1986. God and the New Biology. San Francisco: Harper and Row. Peters, T.F. 1984. Creation, consummation and the ethical imagination. In Cry of the Environment: Rebuilding the Christian Creation Tradition, ed. P.N. Joranson and K. Butigan, 401–29. Santa Fe: Bear and Co. Peters, T.F. 2019. Flashing the yellow traffic light: Choices forced upon us by gene editing technologies. Theology and Science 17: 79–89. Presidential Commission for the Study of Bioethics. 2010. NEW DIRECTIONS: The Ethics of Synthetic Biology and Emerging Technologies. Washington, DC: Executive Branch of the US Government. Rollin, B. 1986. The Frankenstein thing. In Genetic Engineering of Animals: An Agricultural Perspective, ed. J.W. Evans and A. Hollaender, 285–298. New York: Plenum Press. Sandler, R. 2004. An aretaic objection to agricultural biotechnology. Journal of Agricultural and Environmental Ethics 17: 301–317. Santos, M.A. 1981. Genetics and Man’s Future: Legal. Charles C Thomas, Springfield, IL: Social and Moral Implications of Genetic Engineering. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists versus Agribusiness in the Struggle over Biotechnology. Minneapolis: University of Minnesota Press. Seydel, F.D., J. Fullarton, B. Freundel, R. Enquist, N. Cummings and A. Chapman. (1992) Religious and Theological Attitudes toward Genetic Intervention and Alteration of Life-forms. Working Paper for the Genetics, Religion and Ethics Conference, March 13–15, 1992, Houston, TX. Shinn, R.L. 1991. Forced Options: Social Decisions for the Twenty-First Century, 3rd ed. Cleveland: Pilgrim Press. Streiffer, R., and T. Hedemann. 2005. The political import of intrinsic objections to genetically engineered food. Journal of Agricultural and Environmental Ethics 18: 191–210. United Church of Christ. (1990) The Church and Genetic Engineering: A Proposed Pronouncement. 17th General Synod, Fort Worth, TX. United Methodist Church. 1991. United Methodist Church Genetic Science Task Force Report to the 1992 General Conference. General Board of Church & Society Daily Christian Advocate Advance Edition 1991: 113–123. United Methodist Church. (1992) Genetic Science Resolution. May. General Conference. United Methodist Church. (2016) New Developments in Genetic Science, 2016 Book of Resolutions, #3181 Board of Church and Society, United Methodist Church, Accessed March 13, 2020 at https://www.umcjustice.org/who-we-are/social-principles-and-resolutions/new-dev elopments-in-genetic-science-3181.

312

11 Religiously Metaphysical Arguments …

United States House of Representatives, Committee on the Judiciary. (1988) TRANSGENIC ANIMAL PATENT REFORM ACT. House Report 100–888. United States President’s Commission for the Study of Ethical Problems in Medicine and Biomedical and Behavioral Research. 1982. Splicing Life: The Social and Ethical Issues of Genetic Engineering with Human Beings. Washington, DC: US Government Printing Office. Walters, L. 1978. Contemporary Issues in Bioethics. Encino, CA: Dickenson Publishing Co. World Council of Churches. 1990. Report on Biotechnology. Geneva, Switzerland: Church and Society. Yuko, E. (2016) The first artificial insemination was an ethical nightmare, The Atlantic Monthly Accessed Dec. 31, 2019 at: https://www.theatlantic.com/health/archive/2016/01/first-artificialinsemination/423198/.

Chapter 12

Communication, Education and the Problem of Trust

Abstract This concluding chapter from previous editions makes recommendations that follow from the previous eleven chapters analyzing food safety, animal health, environmental and socioeconomic risks associated with agricultural and food biotechnology, as well as discussions of intellectual property rights and religious objections. Scientists and the biotechnology industry have failed to meet reasonable and justifiable expectations for an explanation and defense of their objectives in developing gene technologies for crops, livestock and food processing. Although a rationale for these applications of biotechnology exists, it has not been put forward in a manner that promotes a democratic and respectful dialog. Articulated in 1997, the chapter was a set of ethical recommendations for agricultural insiders. In retrospect, it serves as an indictment that may explain why the technology was resisted and early hopes for agrifood biotechnology remain unrealized. Looking forward, it is a contribution to the literature on public engagement with science. Keywords Public attitudes · Discourse ethics · Technological ethics · Risk assessment · Risk communication · The deficit model · The grammar of risk The ethical issues discussed throughout the previous chapters would exist without widespread public resistance to food biotechnology. The bases for ethical concern about the impact of genetic engineering on food animal well-being or for concern about social consequences depend upon the validity of key norms, the accuracy of key predictions, as well as the truth of key factual assertions. My metaethics holds that it is logically possible for facts to be true, predictions to be accurate and norms to be valid entirely apart from whether any human being, much less a significant number of human beings, appreciates, believes in or endorses their truth, accuracy or validity. However, it is, all too easy to confuse public outrage with ethics itself. Public knowledge of or attitudes toward biotechnology has not been an important focus of the analysis in previous chapters because the goal has been to examine food biotechnology from the perspective of technological ethics, without regard to the popularity of or political motivation for any given argument.

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_12

313

314

12 Communication, Education and the Problem of Trust

This chapter begins with a context-setting review of public opinion on biotechnology before moving to characterize the ethical significance of public opinion within theories of trust and responsibility. Annette Baier’s theory of trust is built on a feminist reading of David Hume’s approach to ethics, laying stress on emotional cues. In the balance of the chapter, I develop an ethical theory for risk communication. This theory is not significantly changed from the original 1997 edition of the book. Others analysts have reached similar conclusions in the intervening years, especially in connection with the deficit model of science communication. Aside from a few brief mentions, I have not attempted to incorporate these lines of argument into the treatment that follows. The chapter closes with a discussion of key fallacies in risk communication.

12.1 Public Opinion and the Ethics of Gene Technology It cannot have escaped the notice of any likely reader for this book that all forms of biotechnology have been the subject of enormous public debate. Public controversy over agricultural biotechnology was elevated even prior to the publication of the first edition of this book in 1997, but skyrocketed in the final years of the 20th century. Social scientists have attempted to measure and to analyze public attitudes toward biotechnology, though their measurements do not present a consistent picture of the basis or degree of concern. In the late 1990s, the Eurobarometer survey comparing attitudes across the European Union revealed a widespread public skepticism about food and agricultural biotechnology. Succeeding versions of the survey documented shifts in the degree of concern and in the national identity of those expressing the highest degree of concern (see Durant et al. 1998; Gaskell and Bauer 2001; Bauer and Gaskell 2002). In the United Kingdom, a group at the Institute of Food Research, Reading also conducted a series of studies on public attitudes toward gene technology. Their findings correlate concern with ethical issues to perception of risk (Sparks et al. 1994; Frewer et al. 1994; Sparks et al. 1995; Frewer and Shepherd 1995). Although it is widely believed that North Americans are far more accepting of biotechnology than Europeans, when the same survey research methods are used to sample both groups the differences are not so well marked (Wandersman and Hallman 1993; Priest 2000, 2001), However, in other respects there are striking differences between these populations. Marlis Buchman studied the 1992 Swiss referendum that established a constitutional amendment calling for the regulation of biotechnology, concluding that the referendum was supported primarily by people with a high level of education who occupied professional or salaried positions, (Buchman 1995). Contemporaneous surveys in the United States show just opposite result (Hoban and Kendall 1993; Hallman and Metcalfe 1994). One way to interpret the ethical significance of these studies is that informants have one or more of the ethical concerns discussed in the previous nine chapters when they report resistance to biotechnology. If so, the surveys simply document that people actually do find these ethical issues problematic. That finding lends political

12.1 Public Opinion and the Ethics of Gene Technology

315

urgency to ethical problems, but it does not change their philosophical character. Yet the ethical issues that arise when widespread public perceptions or attitudes toward risk diverge sharply from those of the scientists who are in a position to have more accurate factual knowledge of the likely outcomes are not the ethical issues discussed throughout the first 10 chapters of this book. The divergence between public and scientific attitudes toward risk suggests that something has gone deeply wrong at the junction between science and broader society. From the perspective of many scientists, the conclusion to draw from such studies is that the public is poorly informed and cannot make reasonable judgments. If so, public opposition is interfering with the conduct of research, especially in Europe, (see Rabino 1991, 1994). Scientists responded to this problem with calls for public education or better communication. Between 1988 and 2016, the National Agricultural Biotechnology Council, a consortium of North American universities and non-profit organizations conducting research on food and agricultural biotechnology, sponsored an annual meeting to solicit consensus recommendations to government, the private sector and to NABC members. Attendance at these meetings was dominated by university and government scientists with substantial representation from the administrative offices of biotechnology companies. Attendees also include a few representatives from non-governmental organizations (NGO’s) and farmers, but few members of the general public. Such a diverse group reaches consensus on relatively little, but every meeting concluded that there is a pressing need for better communication and for public education. Does this conflict at the interface between science and the public present special philosophical problems? Arguably, there are two. The first is that the public conceives of the risk problem differently from scientists. This is not to say that they differ with respect to their assessments of probability and outcome, but that they understand risk issues according to different philosophical parameters. The second is that communication and public education themselves create moral responsibilities not covered in previous chapters. These issues are interrelated. If the public interprets risk differently from scientists, communicating with the public through providing information on probability and consequence will do little good. I will argue that misconceived efforts at communication can do (and have done) considerable harm. The harm they have done is to erode public trust in science. Scholars in science communication were formulating their own theoretical response to the problems discussed in this chapter even as the second edition was being prepared for release in 2007. It revolves around “the deficit model,” of science communication: that scientists know something that public doesn’t, and that role for communication or education is to fill that deficit in the public’s understanding, (Hansen et al. 2003; Sturgis and Allum 2004; Chess and Johnson 2007; Brossard and Lewenstein 2010). There is considerable overlap in the arguments laid out in this chapter and the social science critiques of the deficit model. Indeed, the 1997 edition of the book was cited by some of the theorists who contributed to that critique (Frewer 1998; Kuzma and Besley 2008). My original text (retained in the current edition) draws on work by John Zimon (1925–2005) that prefigured key elements of the deficit model. Revisions to take recent literature into account would constitute

316

12 Communication, Education and the Problem of Trust

a total reconstruction of the argument developed in this chapter. Unlike the deficit model, this chapter brings ethical considerations to the forefront. As such, the conversation between the ethical analysis of communication efforts that follows and the now extensive literature on knowledge deficits must await another opportunity. In addition to reprising many themes from the previous ten chapters, then, this chapter synthesizes discrete analyses of three philosophical problems: risk, communication and trust. Each of these topics might be worthy of a booklength treatment in their own right. Here I will offer summary analysis of each topic in reverse order, concluding with a section on the problem of trust as it relates to food biotechnology. Clearly there are many individual scientists with very different approaches and attitudes toward these public concerns, and just as clearly “the public,” is an abstraction, for in reality there are many groups and voices that interact with scientists and scientific organizations. In the interests of efficiency, however, I will speak of “science” and “the public” as if they represented two coherent perspectives on the problems that have been surveyed in earlier chapters of the book.

12.2 The Problem of Trust Trust is a moral relationship. Two parties who trust one another have a moral expectation with regard to each other’s conduct. The trusting relationship is one in which one not only believes that another will act in the fashion that is expected, but also that the other regards the obligation or responsibility to act in this fashion as a moral or ethical obligation, a moral responsibility. Annette Baier argued that lack of trust was an underlying factor in risk: if a social actor has proven to be untrustworthy, it is entirely reasonable to associate their actions with risk, (Baier 1986). Baier’s extended study of the moral dimensions of trust is the theoretical underpinning for my analysis in this chapter. Baier says that we should see trust as a relationship in which the trusted accepts responsibility to act in the interests of someone who will be the beneficiary of their action. Moral trust is a rich form of the fiduciary relationship where the trustees place the beneficiary’s interests above their own. It may be construed narrowly, as, for example, when financial advisors understand their responsibility to clients as confined to their economic interests, or quite broadly, as may be common among close friends or family members. A philosophical account of trust should thus accommodate the different ways in which a trusting relationship can be shaped. Baier builds her analysis of trust on prior work by philosopher Thomas Scanlon, who identified four principles that govern relations of trust. M: One should not manipulate others by deliberately raising false expectations about how one will discharge the relationship. D: One must take due care not to allow others to form reasonable but false expectations about how one will discharge the relationship. L: One must take steps to prevent any loss that others would face through reliance on their reasonable expectation of what one will do in discharging the relationship.

12.2 The Problem of Trust

317

F: One must maintain fidelity to precisely what one has assured others will be done; while one may not do less, one need not do more. Indeed one should not do more if doing so would alter the other’s expectations in an unreasonable fashion. (Baier 1994, 134).

Baier adds that these principles should be understood as examples of common vulnerabilities of trust, rather than as defining principles that circumscribe trust. She gives examples of broken trust that do not violate any of Scanlon’s principles. Unique features not common to all instances of trust will characterize any given relationship of trust. “Trust comes in webs,” she writes, “not in single strands, and disrupting one strand often rips apart whole webs,” (Baier 1994, 149). The question is whether this model of trust illuminates anything about the public’s attitude toward food biotechnology. The answer begins to take shape with some of the ways that Baier qualifies the relationship. She notes that people often confuse the moral relationship of trust with power relationships and with ordinary promises. Trusting differs from promise-making because promises require one to keep faith with respect to the specific conduct or obligations that are created by the act of making a promise. It is a contractual obligation that creates specific duties envisioned at the time the promise is made. Trust, however, requires one person to act on another’s behalf in ways that could not have been anticipated in a promissory act (Baier 1994, 137). Neither researchers nor biotechnology companies make promises that explicitly state obligations to the public. Advertising strategies that deploy the rhetoric of “our promise” are too vague to satsify Baier’s model of promising. However, the distinction between trust and power is particularly relevant to the present topic. A number of the ethical issues analyzed in preceding chapters deal with unequal power relations between scientists, research organizations or the food industry and some other social group. Consumers want labels on foods derived from biotechnology. Farmers in developing countries are concerned that they will lose the right to use genetic resources that they have husbanded through generations of trial and error farming. Animal advocates find farm animals totally at the mercy of genetic engineers. In each of these cases, an inequality of power underlies the ethical problem, and the proposed response—required labels, restricting access to genetic resources, or animal rights—is justified in terms of claims made by vulnerable parties. Caron Chess, a contributor to the deficit model critique, described such tensions between the vulnerable public and the purveyors of biotechnology as problems in trust, (Chess 1996). Yet Baier’s analysis of trust suggests something different. It is certainly accurate to say that the vulnerable parties (food consumers, indigenous farmers, animals) do not trust those who hold or seek power. But were these parties in a relationship of trust from the outset? Trust relationships are maintained and strengthened when the trustee discharges responsibilities by acting in the interest of the vulnerable party. Yet in the cases discussed above parties will not trust one another even if the food industry or the scientific community does act in the interests of vulnerable parties. There has never been an opportunity to develop a genuine fiduciary relationship in the first place, and in the case of industry, the structural implications of caveat emptor place consumers in an inherently adversarial position. In fact, many persons, organizations or institutional actors cannot be construed as

318

12 Communication, Education and the Problem of Trust

trustees because they are in no position to undertake the actions that would address the concerns of vulnerable parties. Only governments are in the position to take effective action in many of these instances. This suggests that the changes called for are to create more equitable power relationships. Correcting the ethical problem has nothing to do with establishing trust, and everything to do with redistributing power. Baier also draws a distinction between trust and the more extensive and openended forms of interdependency that characterize spousal and parental relationships or other family and community ties. Unlike these relations that one has as a result of factors beyond one’s control, trust is a relationship that may be initiated or terminated, and it may be characterized by specific limitations of scope or duration that are known to both parties. As she puts it, “Trust is acceptance of vulnerability to harm that others could inflict, but which we judge that they will not in fact inflict,” (Baier 1994, 152). One would not accept total vulnerability, nor would one extend a relationship of trust indefinitely. There must be some point in time at which will be possible to reexamine a relationship of trust, and to revise it if necessary. “To trust is to give discretionary powers to the trusted, to let the trusted decide how, on a given matter, one’s welfare is best advanced, to delay the accounting for a while, to be willing to wait to see how the trusted has advanced one’s welfare,” (Baier 1994, 136). Baier’s discussion of trust illuminates the problem of science and its relations to society in two ways. First, it helps us recognize that there are some dimensions of this relationship that have much more to do with power than trust. Many individuals (probably a majority of citizens in most industrial democracies) who purchase their food in supermarkets or restaurants are in a state of utter dependency on their food providers. It may not always have been so, for one needs to cast one’s glance only a generation or two back to find a time when reliance on a highly centralized food system was a matter of choice, not necessity. Clearly there are ethical norms that the responsible parties of the key organizations should (and generally do) follow, but let us not deceive ourselves. We are long past the time where we could alter our dependence on the central food system without prohibitive cost. This is a relationship of power, not trust, (Baier 1986). Yet there is a second point to consider. Perhaps it is business more than science that is deeply implicated in the web of power relations that define the food system. The kind of research done by social scientists suggests that people do have confidence that scientists will guard the safety and abundance of their food and the integrity of the environment. Scientists trust society to give them the support and freedom to carry out this task. The general public may not be in the position of being able to examine and revise their faith in scientists, but the first three of Scanlon’s principles pick out some vulnerabilities in this relationship. Have scientists manipulated the public by promising too much? Have they taken due care to be sure that public does not form reasonable (but false) expectations, absent any malicious or manipulative intent by scientists? Have they taken steps to prevent or compensate for losses that trusting parties may have experienced as a result of what scientists have done? And what of it if it turns out that science comes up short on any of these three principles? Baier’s model suggests that we should divide the question by treating the economic relationship between the biotechnology industry and the public as an instance of power,

12.2 The Problem of Trust

319

reserving the trust relationship to the activity of science, as represented by not-forprofit researchers and scientific organizations such as the National Academies and Royal Societies, or publishers such as Nature and Science. A more thorough discussion of the scientific community’s performance is taken up below in the section on risk, but a preliminary discussion of each question is revealing. One of the most common criticisms of food biotechnology is that it has been oversold, that its proponents have, in their quest for dollars, created wholly unrealistic expectations (Teitelman 1989; Busch et al. 1991). Yet more seriously, even those scientists who themselves have criticized the moneymen and their overzealous colleagues have done little to dissuade members of the public from reaching unrealistic expectations. Thus even if violations of the M principle (manipulation) are exceptions, violations of D (exercise of due care) are the rule. Finally, much of the entire controversy over social consequences, over structural impacts on the size distribution of farms, over the increasing difficulty of family farming and the decline of rural communities is precisely targeted at L, responsibility to limit or prevent loss. Food and agricultural scientists, so their critics claim, at least, have not taken the required steps to prevent losses by those who have entrusted them with their welfare and well-being. There are really only two possible conclusions here. Either food and agricultural scientists have abused the trust of their farming constituency and the wider public, or they have not regarded their relationship with the public as one of trust with respect to the matters under contention. Significantly, discharging responsibilities with respect to Scanlon’s principle of manipulation (M) and his principle of due care (D) requires communication. M is a norm of communicative process: do not manipulate by insinuations that lead to false expectations. D is a communicative mandate: you must take due care to communicate in circumstances where people might, left to their own devices, form false expectations. Yet while standard approaches to science communication would find M to be entirely unexceptional, few would go so far as to mandate D. What and when are scientists obligated to communicate? There are two problems here. One is to understand the method and point of science communication; the second is to identify its ethical dimensions.

12.3 Communication and Public Understanding of Science John Zimon tackles why science communications generally go badly in his 1992 article “Not Knowing, Needing to Know and Wanting to Know.” Each of his title phrases encapsulates a philosophical approach to communicating with the public. In what he called “the deficiency model” of science communication, “not knowing,” is presumed to be the cardinal fault of the average non-scientist. The scientist knows something the layperson does not. Better help the public overcome this deficiency. Of course, Zimon notes, this proves to be a ridiculous model for it is impossible to know everything, and equally impossible to circumscribe a specific set of facts, theories or methods that characterize the “science” that the public does not know.

320

12 Communication, Education and the Problem of Trust

The rational choice model appears to address the problem raised in D because it presumes that people need to know certain key facts that bear on the likelihood of achieving their stated or apparent objectives. The problem here, Zimon claims, is that when we examine the everyday projects of ordinary people, we find that they do not lend themselves to the rational choice model. Put straight out, people do not frame their lives as a series of objectives for which they are seeking the most efficient means (Zimon 1992, 13–17). Ecologist Donald Ludwig made a similar argument anticipating 21st century debates over non-scientists ability to understand climate modeling. Even if economists, political scientists or psychologists can explain or predict human behavior using a rational choice model, people do not represent their own life situations to themselves in such an organized, means-ends fashion. Information that links means to ends in a probabilistic fashion may indeed bear on whether people will achieve their goals, but it will not be taken up until expressed in language that better accommodates how people frame short-term means-ends thinking in terms of roles and narratives (Ludwig 1993). Writing in the early 1990s, Zimon and Ludwig can be read today as developing a philosophy of science communication that connects ideas from rhetoric, anthropology, cognitive science and the philosophy of language in complex ways. One strand of the perspective Zimon and Ludwig develop stresses narrative over logic and mathematics. While the history of philosophy may have emphasized conceptualizations of rationality that rely upon mathematical abstraction and logic, Zimon is claiming that people make sense of the world through storytelling, through construction and dissemination of narratives that have a beginning, a climax and a dénouement, which provides an opportunity for underling the moral to drawn from the story. At the second decade of the 21st century, Zimon’s view is supported by studies on the cognitive processing of probabilistically formulated information (Stanovich 2011), and “fast and slow thinking,” (Kahneman 2011), as well as those that emphasize the role of narrative in a person’s ability to contemplate and evaluate alternative courses of action, (Schechtman 2011). Even a sketch the relevant connections among recent literatures that support and extend the arguments of Zimon and Ludwig would exceed the remit of a study focused on agrifood biotetchnology. The point is that we should not be surprised that the rational-choice model should fail miserably as a theory of science communication. Zimon offers what he called “the context model” as a more adequate approach. This model starts with the presumption that from the perspective of a layperson, formal scientific knowledge is incoherent in that it is encountered piecemeal and fragmented from the broad theoretical models that frame knowledge claims for experts. Scientific knowledge is inadequate in that, “The use that people make of formal knowledge in any particular situation depends on their needs of the moment and represents only one element in a complex and varied response,” (Zimon 1992, 18). People do accept this knowledge passively, but gauge its credibility according to factors that are fixed by the situation in which the knowledge is to be applied. The significance of this cannot be underestimated given late 20th century thinking on

12.3 Communication and Public Understanding of Science

321

scientific expertise. A thesis popularized by Paul Gross and Norman Levitt (1943– 2009) portrayed science as a necessarily complex pursuit of truth that will not be understandable even by other scholars applying methods developed in other disciplines. Non-specialists’ attempts to integrate or translate a disciplined discourse by examining its implications for action wind up being “anti-science,” (Gross and Levitt 1994). Norman Borlaug applied this view to the biotechnology controversy, characterizing criticisms as forms of “anti-science zealotry,” (2000, 2001). Contrary to the suggestion that critiques of science lead to a widespread skepticism of science as a process, one sociological study suggests that general attitudes towards scientific expertise are not particularly influential in shaping the public’s understanding of science. It is the context in which messaging occurs and the specific way that information affects a given person’s interests that determines whether people are likely to be skeptical about science communication, (Trachtmann and Perrucci 2000). Furthermore, though conflicting views among experts may reduce a layperson’s tendency to accept scientific knowledge at face value, the inconsistencies disappear as people apply their own values in selectively adopting or rejecting scientific knowledge claims, (Zimon 1992, 18–19). The context model demonstrates that those who would initiate communication efforts on behalf of biotechnology must realize at the outset that they cannot control the public’s receptivity or interpretation of any given message. The success or failure of a science communication effort must always be measured with respect to what people wanted to know before the effort was initiated. Unfortunately, the extensive public opinion research on food biotechnology does little to document what people want to know about it. The surveys indicate that people want to know whether genetic engineering is being used in the food they eat, but the particular significance that any given individual attaches to this information is unclear. In Susanna Hornig Priest’s research on biotechnology members of focus groups given the opportunity to express their reaction to news stories on biotechnology identify an interest in ethical issues, defined in much the way that they are presented in the first nine chapters of this book. However, although focus groups initiate discussion in terms of the ethical significance of social consequences, property rights, and so on, they are not especially interested in extensive clarification of the values dimensions of these issues. (The members of these focus groups, in other words, are not very likely to enjoy reading this book.) Although the discussion must be initiated in terms of ethical issues, laypersons quickly become curious about both unwanted and beneficial consequences of biotechnology (Hornig 1993). In short, people want to know the things that scientists can tell them, as opposed to what philosophers can tell them, but they would prefer that scientists have some ability to present their information in an ethics-oriented framework. Extending Hornig Priest’s study to a generalization about public understanding of or interest in biotechnology is a bit speculative, yet interpreting it in light of Zimon’s analysis suggests some interesting hypotheses. Since people participating in focus groups are removed from the hurly burly of everyday decision-making, the methodology itself presents an opportunity for less fragmented interest in biotechnology. Their interest, however, is still contextualized by human concerns. It extends both to

322

12 Communication, Education and the Problem of Trust

facts about biotechnology and to facts relevant to specific goals, but only as a result of communication initiated within the framework of ethics and values. At the same time, the desire for fact-based information as a response to ethical issues suggests that it would be inappropriate to have ethicists (or public relations officers) conducting communication efforts. If this analysis is correct (an empirical hypothesis, to be sure) a communications effort conducted in terms of spreading factual information about biotechnology goes nowhere, while a communications effort initiated with a reasonably sophisticated overview of ethical issues spawns curiosity about the facts.

12.4 The Ethics of Science Communication The work of Zimon and Priest helps define the pragmatic considerations that must frame a discussion of communication, but neither of these analysts take up communication about science as an ethical problem. For present purposes we can note three key norms for science communication. First, science communication should be truthful. This almost goes without saying, of course, since moral proscription of prevarication is one of most basic and widespread norms. Truth-telling is subtly difficult in science communication, where information must be translated out of technical and into ordinary language. It is easy to unintentionally mislead. Yet the basic moral claim here is straightforward; it is Scalan’s M, (do not manipulate), described above. Second, Scanlon’s principle of due care (D) states that there are situations in which communicators have a positive obligation to provide information. One part of this obligation is also non-controversial once it is laid out: scientists have a responsibility to inform the public about objective dangers and risks. It would, for example, be unconscionable for a scientist who has evidence that a particular product is dangerous to withhold that information, or even to publish it in a forum where it was likely to remain unnoticed by users of the product. This, too, is a difficult norm to operationalize because it requires scientists to make torturous judgments about when it is appropriate to bring a concern into the public realm. Scientists try out many hypotheses in the course of discovery, and the mere fact that some, if confirmed, would point toward a public hazard is insufficient reason to bring speculation forward. Yet the burden of proof for publicizing a hazard is certainly much lower than for accepting the hypothesis. Arguably, the burden gets weaker and weaker in proportion to the seriousness and irreversibility of the hazardous outcomes. Carl Cranor and Kristen Shrader-Frechette have analyzed this problem in terms of Type I and Type II statistical errors, arguing that it is often more important to act on a result that might be true than to avoid accepting a result that might be false (Shrader-Frechette 1991; Cranor 1993). The duty to inform become critical to one of the key events in the controversy over agrifood biotechnology when Arpad Pusztai made public allegations about the risks of genetic engineering based on some very preliminary studies of transformed potatoes. This incident took on many of the classic elements of a whistleblowing case: Pusztai was disciplined by his employer, who felt his move to publicize results was premature and alarmist, but became a

12.4 The Ethics of Science Communication

323

celebrity among opponents of GMO’s (see Krebs 2000; Randerson 2008). Pusztai’s case illustrates that even if the duty to inform in cases of risk to the public seems ethically non-controversial, determining the exact circumstances in which this duty becomes mandatory may be complex and controversial indeed. For present purposes, however, it is sufficient to note the problem and to point out that the need for striking a moral balance is quite clear, even if knowing how to reach it in practice is not. While no one disputes the duty to inform the public about risks and hazards, a broader reading of the due care principle (D), might entail that the public has a right to know whenever its values are threatened. For example, if genetic engineering of food might offend a person’s aesthetic sensibilities or religious values, scientists would have a responsibility to take due care in informing the public of this possibility. If social consequences would negatively affect small farms or rural communities, scientists would have a responsibility to inform not only those who are affected, but also those in the broader public who espouse values of solidarity with rural groups. Clearly there is a possibility of extending D beyond all reasonable scope here. Due care does not require informing every person who might possibly be offended by the results or products of research in food biotechnology. What may be more reasonable is to stipulate an auxiliary principle: that science as institution (and scientists as its representatives) has a responsibility to undertake public communication efforts that promote participation as a democratic ideal. The basic argument for this view of communication has already been sketched in Chap. 8 on the social consequences of agrifood biotechnology. There Langdon Winner’s conception of the technological constitution was summarized (Winner 1983). The idea that technical infrastructure has deep political significance is also supported by Philip Kitcher’s analysis of the relationship between science and democracy, discussed in Chap. 9, (Kitcher 2001). Either provides reasons why individuals and groups might feel that it is important to be involved in the earliest stages of decision making about biotechnology. Communication is fundamental to this problem: involvement implies notification that one’s interests will (potentially) be affected, as well as some understanding of scientific and technological possibilities. It is not feasible to make research decisions by ballot, so some form of mediated participation simply must suffice to satisfy the ideal of participation. At a minimum, this means that members of the public must have some vehicle for advising the science community of its concerns, and for requesting information and response. Political theorists have long argued that a free press can satisfy this need, but the public must also have some assurance that scientists are listening. There is a desperate need for two-way communication, and for some means of assuring parties that their messages are being heard. The analysis developed here (and published originally in 1997) should be read as continuous not only with subsequent work on the deficit model in science communication, but with even more recent studies in the philosophy of science that emphasize “inductive risk” (Douglas 2000, 2009; Elliott 2017). Describing and speculating on how scientists could discharge fiduciary responsibilities to the broader would take the discussion deeply into this more recent work. It must suffice to note that prior to the actual release of the first genetically engineered plant varieties in 1997, the record

324

12 Communication, Education and the Problem of Trust

of science and industry was quite poor. Christopher Plein’s research concludes that private industry successfully manipulated the U.S. debate to portray public concerns as opposition to job creation and economic growth, (Plein 1991). Susanna Hornig Priest’s work shows that industry also shapes U.S. newspaper coverage of biotechnology (though not always in the manner that they might intend), (Priest and Talbot 1994; Priest 2001). Brigitta Forsman and Stellan Welin describe how scientists and industry closed off public participation in a national ethics commission on biotechnology in Sweden, (Forsman and Welin 1995). Ad van Dommelen documents how public and private sector scientists worked to restrict public participation in a participatory risk assessment in Germany (van Dommelen 1995). Angela Griffiths has shown that Canadian biotechnology researchers failed to even consider how a series of government directives (issued with broad political support) to emphasize sustainable agriculture might be incorporated into their research planning (Griffiths 1996). On the other side of the ledger were precious few success stories. It would appear that the NABC’s conferences had some modest impact on biotechnology planning in the United States, for example, and when scientists turned out for the meeting, it was at least evidence that they were listening. But while meetings between the NABC’s founding in 1988 and the early 1990s were lively, attendance by anyone besides industry and university scientists or administrators declined rapidly. The meeting in 1996 (immediately prior to the first edition) attracted less than a hundred participants, despite being held in New Jersey, where many non-scientists could have participated at low personal cost. Attendance rose when European reactions to biotechnology hit the headlines, but waned again when the controversy subsided. By 2006 (immediately prior to the second edition), attendance was again less than a hundred. By 2016, when NABC shut its doors, the meeting had become a place where research administrators from universities and the biotechnology industry could converse with no fear of disruption by members of the public.

12.5 The Problem of Risk Risk issues present both a special case for the ethics of science communication, and they are particularly crucial to any discussion of food biotechnology. There is a pattern of give and take in risk debates that is widespread across policy issues and for which scientific evidence is expected to be decisive. The first element of the pattern is criticism of the data, conclusions, or methods that have been used in assembling the scientific evidence. Criticism of this sort is part of science itself. The second element is an inference to the effect that uncertainty in data, conclusions or methods entails risk to members of the public. This inference is not characteristic of scientific reasoning; scientific risk assessment does not conclude that an activity is dangerous simply because it is uncertain. This divergence between scientific and “ordinary” rationality is, perhaps, the first wedge between science and the public when risk issues are debated. The final element is an attack upon the motives or values of scientists themselves, who are portrayed as trying to conceal risks and uncertainties

12.5 The Problem of Risk

325

from public view (Thompson 1986). The upshot is a political environment in which scientists are alienated from those who profess to speak for the public, and from their perspective, justifiably so. In order to see why this circumstance arises it is necessary to revisit the distinction between expected-value approaches to risk and those that stress consent. As argued in Chap. 4, scientific research techniques are well suited to the measurement of certain key relationships between exposure to a given substance and the subsequent occurrence of harm. These relationships are important in food safety because high correlations between exposure and harm give cause for concern about the human health effects of exposure to the substance. Though important, the measurable relationships between exposure and harm create a misleading communications context when they are taken to define risk to the exclusion of qualitative characteristics. One has long heard the opinion that scientists study the reality of risk (see Starr et al. 1976; Ruckleshaus 1983), or that people who are concerned with other factors that are relevant to risk are dealing with mere perception; while only the scientists deal with reality (Cook et al. 2004). This view of risk is logically, linguistically and epistemologically insupportable (Thompson 1990, 2018), but what is important here is that it uses the language of perception and reality—ostensibly an objective, science-based distinction—to conceal an ethical value judgment. That is, the measurable correlations between exposure and harm are deemed real (which is to say, important), while other elements that may be very significant for assigning responsibility or determining whether a person or organization should be trusted are consigned to “perception”. Risk and reality are both politically potent notions. The judgment to emphasize measurable relationships is often justified; presuming that these relationships model the reality of risk is not. How can the ethics of risk be untangled? How can one determine when it is appropriate to interpret risk as the probability that hazards will materialize, and when it is appropriate to have a broader and more flexible way of understanding risk? A close examination of the way that the word ‘risk’ gets used by ordinary people in ordinary conversational contexts can provide a great deal of insight into these questions. “Risk” is a common English word. It cannot be appropriated as a technical term without inviting miscommunication. Careful listening to the way that the word “risk” functions in ordinary speech reveals a varied pattern of use. It is particularly important to notice that the word ‘risk’ is both a verb and a noun, and that there are adverbial and adjectival forms of the root, as well. In contrast, the standard scientific definition, which holds that risk is defined as a function of exposure and hazard, readily converts into an ordinary language expression stating the chance or probability that a given hazard will occur. Thus, “the risk of agrifood biotechnology” gets treated as being equivalent to “the probability that hazards will occur, given agrifood biotechnology.” Notice that while “the probability that hazards will occur” can easily be understood as indicating a certain state of affairs (e.g. it functions readily as a noun), it is not at all obvious how one would convert this noun-phrase into a verb, an adjective or an adverb. This means that word ‘risk’ and its related grammatical forms are capable of conveying much more information

326

12 Communication, Education and the Problem of Trust

in ordinary language (at the cost of being ambiguous) than the standard scientific definition of risk. As a verb, to risk is to do something, to take action. Risks do not just happen; they are undertaken or done by someone (or by some other subject capable of taking action, like an organization, a government or a corporation). This reflects a deep and philosophically important feature of ordinary language. Verbs such as ‘happen’ ‘cause’ or ‘occur’ can appear in grammatically correct sentences in which the subject of the sentence is an ordinary thing, a natural phenomenon, and not an agent capable of acting intentionally. Earthquakes happen. Tsunamis occur with a measurable frequency. Volcanic eruptions can cause damage. But note well that when one puts the word ‘risk’ in the verb spot, one generates nonsense: Earthquakes risk. Tsunamis risk with a measurable frequency. Volcanic eruptions can risk damage. These phrases sound odd to the ear precisely because the verb ‘risk’ cannot appear in a grammatically correct sentence unless the subject of that sentence is an agent, a being that we understand as capable of intentional action. So a person can risk, and an organization can risk. An animal might be able to risk, because we do speak of animals as agents capable of intentional action. But a tulip does not risk being eaten by squirrels, nor does an earthquake risk the damage it might cause. These grammatical points are admittedly subtle. Earthquakes pose risk, to be sure, but the insertion of the verb ‘pose’ shifts the context so that we are now using the word risk as a noun and as a noun, the word risk can and does frequently indicate a possible state of affairs. This suggests that there are at least two broadly distinguishable ways in which the word risk functions in ordinary language. One usage maps fairly closely with the standard scientific definitions. It is the event-predicting or ‘state of affairs’ naming sense of risk. The other sense is reflected when the word is used as a verb. Here, risks are acts undertaken by agents capable of acting intentionally. This is the act-classifying sense of risk, for the point of saying that someone or some group “risked something” is to pick out that action and notice something special or distinctive about it. I do not mean to suggest that these two senses of ‘risk’ are always easy to distinguish. When people run risks they could not have taken consciously, the tendency is also to shift the word ‘risk’ to its nominative form. So it is meaningful to say, “Jim risked his life by driving drunk.” Here, the suggestion is that Jim’s driving drunk was an intentional act, and also an act that is especially remarkable. But it would be odd to say “Jim risked his life by eating peas,” or “The Romans risked their lives by using lead pipes,” even though eating peas and using lead pipes are both intentional acts. In the case of Jim and his peas, the oddness is felt in that one waits for the other shoe to drop: “And why was eating peas so dangerous for Jim?” Would it relieve the tension if someone replied with the refrain we often hear from scientific risk assessors, “Well you know, there is no zero risk.”? (Answer: It certainly would not.) To say something like “Jim risked his life by eating peas,” is to imply something about Jim, peas or the context at hand that makes this particular case unlike the others where there is really nothing exceptional or worth noting about someone’s eating peas. As for the Romans, it would not be odd to say that the use of lead plumbing created a risk to their health, because we know what the Romans could not have known, i.e.

12.5 The Problem of Risk

327

that the chance of lead poisoning creates hazards to health. And unlike the above attempts to form risk-sentences with subjects like ‘earthquake’ or ‘volcanic eruption’ there is nothing grammatically incorrect in saying that that people risk things without knowing it: “She risked her life unknowingly by smoking cigarettes.” So although this act-classifying sense of risk does not always imply that a person has knowingly chosen to risk, it does imply that the act in question is an intentional one. We would not, for example, describe an epileptic seizure as “risking one’s life,” despite the clear indication that there is a significant probability of harm associated with seizures. The reason is that enduring a seizure is not an intentional act. The grammar of risk allows “Why do you risk your life by having a cigarette?” but not “Why do you risk your life by having a seizure?”. It is clear that the word ‘risk’ is also used in ordinary language to describe a trait of future events, namely, that if they occurred they might be harmful. We can and do talk about the risk of an earthquake, a tsunami or a volcanic eruption. If the word risk is used to describe this trait of events, or if it is used to refer to events having this trait to a strong degree, different event-predicting grammatical rules come into play. Since situations such as enduring a seizure are significantly correlated with some probability of harm, they clearly do count as forms of risk in this event-predicting sense. Indeed, there appear to be no situations that do not involve some degree of risk, at least when it is the event-predicting sense of risk that we have in mind, and when the conversational context is clearly in the event-predicting mode “There is no zero risk” is not at all an odd thing to say. Ironically, when grammatical rules for act-classifying are applied, an epileptic seizure is not a risk, but when rules for eventpredicting are applied, it is. The philosophical grammar that distinguishes these two senses of risk is admittedly obscure. An epileptic seizure is a risk to one’s life, but to have a seizure is not to risk one’s life. The differences between act-classifying and event-predicting uses of risk are not sharp enough to warrant the claim that there are two, fully distinct meanings. Nevertheless, the different uses of the word risk suggest opportunities for technical or formal specifications of the term risk that stress event-predicting grammar to the exclusion of act-classifying grammar (or vice versa). The expected value analysis of risk, discussed in Chaps. 4, 6 and 8, trades heavily on the event-predicting grammar typical of ordinary use. Although there are many ways to specify risk quantitatively, those that follow the expected value approach define risk (R) as a function of the probability and value (utility) of future events (Friedman and Savage 1948). Expected values are themselves computed as a function of value or utility associated with the event U(e), and the probability of the event’s occurrence P(e). There are several ways of representing risk as an expected value. One simple and intuitive function is R = P(e) X U (e) for all U (e) < 0 This concept of risk can be linked to decision-making through the expected utility theory of choice. Although there are several decision rules that can be applied to convert expected utility calculations into action (Rescher 1983), the simplest one

328

12 Communication, Education and the Problem of Trust

assumes that the objective of decision-making is to select the option with optimal expected utility. The option with the highest net expected utility, once costs and benefits are weighed, is the one that should be chosen. To the extent that scientists adopt the expected-value approach to understanding risk issues, they refer exclusively to the event-predicting grammar of risk, and they reduce the broad and flexible grammar of risk to a set of quantitative relationships fixed by probability and value of harm. There are many occasions on which this reductive practice is helpful for decision makers. I would not have endorsed the risk-based model if I did not think that was the case. Nevertheless, there are also communicative situtions in which persisting with the event-predicting sense of risk goes badly. The expected value analysis of risk places a great deal of emphasis upon quantifiable probabilities, plus it is easily linked to a theory of choice. These two factors make it very attractive as a conceptual approach for science-based public policy (for the argument see Kneese et al. 1983; Freeman and Portney 1989). The expected value analysis of risk also provides a rigorous and sophisticated development of the event-predicting applications of risk that we note in ordinary language. The rigor in the expected value analysis, however, is achieved at the expense of act-classifying shades of meaning that can be detected in the ordinary concept of risk. Correlations between exposure and harm are extremely important in setting policy for food safety and quality, but they do not exhaust the ethically significant aspects of risk policy. Three examples follow.

12.6 Human Action, Risk, and Responsibility As noted above, the expected value analysis of risk applies equally well to intentional actions and natural events. One can quantify the fatality risk of driving drunk, of undergoing a seizure, or of being caught in an earthquake. Simple comparison of the expected values makes these events appear morally commensurate, but they are not. We hold people responsible for their action when they drive drunk, but we do not hold people responsible for the consequences of enduring a seizure or an earthquake. The expected value analysis of risk provides no clue as to whether an agent would be held responsible for their actions, or correlatively, as to whether it would be responsible to act in a prescribed way. This underappreciated feature in the grammar of risk creates an opportunity for misleading communications, as well as for some morally troubling situations described below. Interest in more nuanced senses of the work risk has increased significantly since the first edition of this book appeared in 1997. As discussed at more length in Chap. 14, Sven Ove Hansson has identified five core senses to the word (Hansson 2018), but none of them would map well onto the sense that the word has when used as a verb. When used as a verb, risking is a sub-class of action. Only agents (or entities metaphorically categorized as agents) can risk something. In contrast, non-agential entities can cause or be causally responsible for events. When I have presented some of this work on the philosophical grammar of risk to scientific audiences, I have occasionally been met with the claim that the

12.6 Human Action, Risk, and Responsibility

329

word simply is not a verb. Indeed, a quantitative study of word usage found that risk occurs as a verb in scientific publications much less frequently than in non-scientific usage, (Hamilton et al. 2007). Scientists have been socialized into a practice that suppresses connotations of the word risk that are meaningful in ordinary contexts. It would appear that Hansson has followed them. Specifically, to write or speak as if the risk of a seizure or an earthquake is indeed commensurate with the risk of drunk driving allows the audience to hear a message stating a broad moral equivalence, as if these activities should be ethically evaluated in similar terms. But while seizures and earthquakes can cause bad outcomes, they cannot risk them. In contrast, the act of drunk driving risks a harmful outcome, even when it does not cause one, (see Thompson 1985 for an early discussion of this grammatical feature). This becomes relevant to agrifood biotechnology when a science communicator places risks from genetic transformation into a comparison with risks that are, like earthquakes and seizures, associated with natural hazards. An example would be food safety risks associated with microbial contamination. To see this point, it is critical to understand how the act-classifying and eventpredicting meanings of risk perform two distinguishable (but also overlapping) communicative functions. We do not classify the seizure or the earthquake as acts, but drunk driving is an act. The act classifying rules of grammar for risk are part of taxonomy for sorting actions into different kinds. Some actions are considered risks; others are not. Articulating the criteria for sorting would be a large philosophical and linguistic project in itself, but the examples given above seem to involve paradigm cases or ideal type classifications, so that judgments as to whether an act is a risk can be drawn by analogy. In our society, driving while drunk is a paradigmatic case of risk; driving while sober is not. It also seems that traditional familiarity with the act in question is a criterion. Using the new-fangled radio-controlled convection oven is a risk; boiling peas on the stove is not. Any number of communicative functions may be fulfilled by this distinction, but one in particular is critical for ethics: calling an action a risk is one way of noting that a person will be held responsible for the consequences. Secondarily, it is a way of urging caution, rather than a claim that significant probabilities of harm exist or have been measured. An idealized depiction of traditional tort law provides the clearest account of how classifying actions under the category of risk plays a role in making decisions and in assessing responsibility. Innovations in the case law of torts beginning in the 1970s introduced the expected value analysis into liability decisions (Schroeder 1986), so the following portrayal of torts should not be taken as a description of legal practice. Traditional torts in the United Kingdom and the United States are based on common law. The purpose is to assess whether the defendant wrongfully harmed the claimant bringing suit, and whether the defendant should be required to pay damages. The claimant may meet this burden of proof by showing first that the actions of the accused were unreasonable, then that they actually resulted in harm to the claimant. This two stage burden of proof is critical to understanding the ethics of risk as they relate to culturally determined categories of action. Simple demonstration of harm is not enough to warrant damage in traditional torts, for the defendant’s act is judged to impose risk only when it is something that a reasonable person would not do. If

330

12 Communication, Education and the Problem of Trust

a reasonable person would have regarded the act as unexceptional and proper, the claimant cannot meet the initial burden of proof. The principle implies a general recognition that harm can occur as a result of happenstance, freak events or so-called acts of God, even when the actions of a defendant are completely ordinary acts of the sort that reasonable people perform every day, (see Gardner 2015). None of this is implies that this idealized picture of tort law should serve as model for legal regulation of risks from agrifood biotechnology. The point here is to describe an ethics of risk that a) differs from the expected value model; and b) draws upon the act-classifying sense of the word ‘risk.’ Tort decisions involve compensation for damages, but only when the person whose actions cause damage has acted in an unreasonable way. Legal scholar John Gardner argues that in such cases the reasonable person standard should be understood to mean that the person’s actions were justifiable, (Gardner 2015). My argument here is that this picture of the reasonable person embodies an implicit division of human activity into at least two categories, one being normal, ordinary activities that do not impose extra burdens for deliberative evaluation or due care, and a second category of actions that do impose these burdens. There is, possibly, a third category of actions that are so clearly seen to be beyond the pale of reason as to be called ‘reckless.’ I am suggesting that the grammar of risk maps on to these rough categories in the following way: to describe an action as ‘risk’ is to state that it is in the second or third category, which is to say that the action becomes a candidate for further judgments about moral responsibility for harm. Even when the person who is harmed meets the dual burden of proof (a risky act and an occurrence of harm), the defendant has an opportunity to demonstrate exculpatory factors, and the list of potential exculpatory factors is extensive. They include, for example, whether the defendant acted knowingly and whether the claimant had complicity in undertaking the risky course of action. So to say that an act is ‘risky’ in this sense is not to complete an assessment of moral responsibility for harm, but a key point to note is that many unexceptional activities are not even candidates for ascriptions of responsibility for harm. They are not (in this sense) risks. The key concept in proving both the initial claim of risk and in providing excuses is that of the reasonable person. In the traditional process of establishing responsibility, there is a large class of actions that are not risks, simply because they are so broadly accepted, even though there are measurable (and perhaps even relatively high) numerical probabilities that they might result in harm. As is generally the practice in common law and ordinary moral judgment, criteria for deciding what a risk is and what is not are established by drawing analogies to precedents. In the law, these criteria are set forth in judicial opinions and become more deeply embedded into law the longer they endure, and the more broadly they are applied. Laws regulating agrifood biotechnology are statutory and administrative, so the traditional practice of torts may be a poor model for reflecting the kinds of regulatory decisions have been (or need to be) made with respect to risks. The point is not to advocate reliance upon traditional case law, but to show how this idealization of torts draws upon the act classifying grammar of risk in making a determination of responsibility.

12.6 Human Action, Risk, and Responsibility

331

From an ethical perspective, there are many reasons not to lose sight of the actclassifying sense of risk. First, it links harm (or the possibility of harm) with actions for which persons could be held legally or morally responsible, and it does so in a way that conforms broadly to culturally based understandings of rights, duties and human virtue. The expected value analysis, by contrast, stresses the sense in which every instance of harm falls into statistical patterns. One gains some management capacity by emphasizing expected value, but one impoverishes the conception of personal or group responsibility for risk at the same time. The acts of individual persons or corporate groups contribute to the statistical pattern that provides the basis for an expected value. However, they are not morally responsible for the pattern, and without a pattern there is no expected value analysis of risk. Statistical patterns are revealed by analyzing data that collates classes of events, including behavior by individuals and groups. To the extent that the probability of an event is associated with these statistics, it can be seen as dissociated from any single act. Indeed strict logic would see the inference from data about a population or class of behaviors to a statement concerning the risk of a single action as a division fallacy. Walter Sinnott-Armstrong applied this form of reasoning in an influential analysis of moral responsibility for climate forcing emissions. One individual’s emissions do not cause climate change, nor do they cause the collective impact of many individuals. Hence, Sinnott-Armstrong argues, how can one hold individuals responsible? (Sinnott-Armstrong 2005). When the connection to action is weakened, the most plausible normative view is that risks should be managed at the level of statistical populations through mechanisms such as insurance or regulation. The debate over causality, climate change and collective action notwithstanding, there are corporate agents that develop and release GMOs. That their actions are statistically comparable to the probability of hazards that are not the result of intentional action is irrelevant. In addition, the act-classifying sense of risk actually functions as one of the cognitive filters discussed in Chap. 2. There it was shown that it is not really feasible to think that every application of technology (much less every possible decision that people might make) could be evaluated on a case-by-case basis. Instead we rely on habits or “filters” to identify when we should actually try to consider costs and benefits in a conscious, deliberative fashion. The unexceptional actions that are classified as “not risky” are not subjected to this kind of evaluation, while those in the second group, those classified as “risky” are. It is possible that an expected-value or risk-benefit type of evaluation will lead to the judgment that the risk in question is well worth taking, but the point here is that the act-classifying standards implicit in our common-sense background beliefs perform as filters, identifying which actions need to be subjected to the kind of consequence-predicting evaluation characteristic of scientific risk assessment. Without filters of some kind, the whole exercise of comparing the expected-value of consequences devolves into incoherence: it is simply impossible to evaluate every possible course of action in a conscious and deliberative way. Other reasons to emphasize the act-classifying sense of risk, and other ways to connect this way of thinking about risk with cognitive filters that organize our allocation of deliberative resources are discussed in the succeeding sections of this chapter.

332

12 Communication, Education and the Problem of Trust

For now, a final reason can be noted: scientists who talk of risk from biotechnology only in terms of hazard and exposure deny the public an obvious opportunity to raise questions of agency and responsibility. This is an unfortunate way to shape the message from the standpoint of an ethics of communication. Not only does it introduce opportunities for misunderstanding (discussed below), but it frustrates communication on the issue that may well be of paramount importance to a layperson: Who is responsible? Whom must I trust?

12.7 Equivocation Problems and False Authority Equivocation upon distinct meanings of the same term is one of the most egregious and indisputably fallacious forms of logical error. Although equivocation fallacies are conspicuous when exposed, they are often far from obvious to the people who commit them. Equivocation has ethical implications when it is the source of error in judgment, or in communication. Equivocation can also play a role in the creation of false authority. When a judgment or standard justifiable on one interpretation of the term is imposed upon a situation in which the alternative interpretation would be more appropriate this may simply be a mistake in judgment. But when a body of knowledge appropriate to one way of interpreting the term begins to be systematically applied to situations where the alternative interpretation is more appropriate, the nature of the ethical problem takes on a political dimension. Those who possess and promote this (inappropriate) body of knowledge become viewed as having authoritative expertise. In fact, their expertise may be much more limited than they (or anyone else) seem to think. More serious ethical issues arise when equivocation is used as a deliberate vehicle of deception. The equivocation of interest here occurs when an act-classifying use of the word ‘risk’ would be the most appropriate way to approach a decision or a communication effort, but the event-predicting sense is substituted in its place. I believe (though this is not the place to argue) that many well-documented anomalies in the literature of risk-studies can be traced to exactly this kind of equivocation. For example, researchers have been documenting a divergence between expert and lay attitudes toward risk for many years. Paul Sovic, one of the leading figures in this work on risk, summarized much of this work in an article entitled “Trust, Emotion, Sex, Politics, and Science.” His title reflects his conclusion that certain socially relevant variables (such as gender) are strongly related to the divergence between expert and non-expert attitudes toward risk, but also that other patterns of divergence cannot be so readily explained, (Slovic 1999). My hypothesis is that the difference between the cognitive-filtering of act-classification and the outcome-optimization of event-predicting accounts for a significant part of the divergence that Slovic and his colleagues have observed in three decades of empirical research on attitudes to risk. My hypothesis is that the experts and lay respondents are not actually talking about the same thing. Furthermore, while the experts may be more correct than the lay public when it comes to well-specified and highly contextualized decisions (such as:

12.7 Equivocation Problems and False Authority

333

should we regulate international trade in beef in order to control the movement of pathogens?), people in the lay public are more rational than the experts in the global sense. They are working from a background set of cognitive filters that would have to be in effect in order for the means-end optimization of the expected-value framework to have any intellectual coherence at all. In the present context, further digression into the broader literature of risk studies would only move the argument even further from the ethics of agrifood biotechnology. Although simple errors of judgment and intentional deceptions occur in the discussion of agrifood biotechnology, false authority may be the most important ethical issue associated with equivocation on the act-classifying and the eventpredicting meanings of risk as the word is used in discussing food safety and environmental impact. Most people apply the concept of risk in ordinary decision making without being fully aware of the semantic content or logical structure of either actclassifying or event describing usage. The context of speech is usually sufficient to specify the meaning intended in any given speaker’s utterance. The problem of false authority arises in connection with agrifood biotechnology when the expected value analysis of risk is applied in such a way as to make otherwise reasonable judgments appear illogical, uninformed, and even irrational. One instance of the false authority fallacy occurs when actions for which individual or corporate agents can be held responsible are compared to natural events in order to derive standards for acceptable risk (see Starr 1969). Many naturally occurring substances are estimated to possess greater carcinogenicity than heavily banned additives and heavily regulated chemical residues (Ames 1983), and we can expect a similar circumstance to be true for products of biotechnology. What should we make of this fact? The expected value analysis of risk can be interpreted to imply that there are certain trade-offs between risk and benefit that are acceptable, without regard to the origin of the risks. The preceding discussion of responsibility shows that origins are sometimes important. Although it is clear that the dangers of natural carcinogens have been tolerated or endured by human populations, the expected value analysis of risk begs the question of why we should tolerate or endure similar levels of expected harm from human action. When responsibility is important, the permissibility of risk is determined by comparing the act to the standard range of things that human beings do, by considering the importance of the ends in view, and by examining the alternative ways of achieving those ends. In this context, the judgment that a risk is acceptable implies that there are overriding moral or prudential reasons for acting in an exceptional manner. Acceptability, in other words, implies an intentional attitude toward the act, not mere tolerance for passively enduring a state of affairs. There is a genuine philosophical issue here. It may indeed be a foolish waste of public resources to ensure against harms that are already far less likely to occur than harmful natural events. The important philosophical issue is not illuminated, however, when the expected value analysis is falsely applied to cases where human agency and responsibility for risk are clearly important. The problem of false authority relates to the role of science in the public’s ability to participate in democratic decision-making. There are always good scientific reasons

334

12 Communication, Education and the Problem of Trust

for adopting the expected value analysis of risk, and there are sometimes good ethical reasons, too. When the expected value analysis comes to exclude the multiple shades of meaning that are associated with risk in common speech, however, some of the most natural ways of raising serious issues about responsibility for action appear absurd. People who are applying the grammar of risk in very standard and traditional ways are made to appear as if they are making logically insupportable statements, and the ethical issues that would be raised by these standard and traditional ways of talking about risk are made to seem chimerical and irrational. The danger is that the appearance of irrationality will be dealt with by handing policy over to experts; only in this case, the criterion for being an expert lies primarily in possessing an impoverished understanding of risk.

12.8 Moral Reductionism and Political Exclusion Those scientists who do take the concerns discussed in this book seriously tend to understand them as separate issues, just as they have been presented here. They tend to think, for example, that it is possible to resolve concerns about food safety or environment without simultaneously doing anything about social consequences. To a large extent, these are seen as risk issues, with the risk understood as a function of hazard and exposure and risks in each category being determined by distinct causal mechanisms. With respect to food safety, the mechanisms are biochemical. With respect to environment, the mechanisms are ecological. The mechanisms for social consequences are economic or sociological, with relatively little biological base. Each of these mechanisms is triggered by functional characteristics of the plant or animal product, rather than by genes, hence none of these risks are unique or different in kind when products of biotechnology are compared to products of ordinary plant or animal breeding. What is more, following the analysis of the first nine chapters, each of these mechanisms is associated with a different set of ethical questions and a different kind of moral significance. For food safety it is the ethically unproblematic human health. For animals we are involved in a philosophically tricky extension of moral consideration to non-human species, but this is quite distinct from the moral evaluation of environmental impact. When we get to social consequences, we are back in the realm of traditional (but contentious) political theory. As these areas are distinct in terms of physical or social mechanisms, they are distinct in terms of the ethical values that make the unwanted consequences morally significant. For convenience, let us call this a purifying view of risk: muddled mechanisms are sorted out, and the appropriate moral concepts are matched to each. The public, however, may not even make a distinction between individual scientists, scientific research, scientific theory and the specific products of biotechnology. Just as they may have elements of act-classification, (reflecting uncertainty, intentionality or consent) in mind when they use the word ‘risk,’ members of the public may have amalgam of scientists, universities, theories, products and corporations in mind when they use the word “biotechnology.” As such, it is not surprising that members

12.8 Moral Reductionism and Political Exclusion

335

of the public do not tend to see the risks of biotechnology in as ordered and distinct a fashion as scientists do. In fact, non-scientists tend to skip from concerns about food safety to environment to animals to property rights to social consequences as if they saw these risks in rather the same terms. Public speech on risks of biotechnology does support the idea that risks are brought about by different mechanisms with independent probabilities. Rather than seeing products of biotechnology as entailing a series of distinct and potentially manageable risks, they just see risk, not broken down into scientifically supportable mechanisms. Similarly, they may not recognize the tidy philosophical distinctions that have been used to separate questions of animal welfare from questions of solidarity with small farms. Because the propensity is run together what the purificationists want to keep apart, let us call this view a hybridization of risk. Yet hybridizers do integrate all these various dimensions of risk in a rational fashion. Biotechnology is seen as risky in much the same way that renting a home or buying an appliance is risky. The landlord/salesman may charge you too much, or may have another way to cheat you with some hidden information. The home or appliance may break, and it may even cause injury. What’s more, your spouse or family may not like it, and you may not be able to get your money back. All these dimensions go together in the risk of renting a home or buying an appliance, and they all attach to the person of the landlord or salesman. When you buy a house, risks that derive from structural features of the building become rolled together with risks that derive form the nature of the transaction. They often register in one’s mind in terms of whether one’s counterpart in the transaction can be trusted. In a similar manner, all the risks of biotechnology go together in the public mind, and they all attach to the person of the scientist. If the scientist behaves in a manner that lends credibility to any of these threats, biotechnology will come to be seen as extremely risky indeed. This, too, is a reasonable and time-honored way to manage the risks that might be created through one’s dealings with others. A scientist (especially one who takes Zimon’s “not knowing” or deficit model approach to communication) is constantly trying to divert the public’s attention away from personalities and toward the facts, to the mechanisms that actually create risk. The public, meanwhile, is intently focused on the behavior of the scientist, looking for evidence that these are people who can be trusted. Tragically, the attempt to divert attention away from personalities to facts is seen in this context as untrustworthy behavior, as an attempt to dodge responsibility. Speaking for consumers, activist Ken Taylor (1938–1995) said as much in a 1991 address to the National Agricultural Biotechnology Council, (Taylor 1991). British survey research on public attitudes toward biotechnology provides empirical support for this generalization (Sparks et al. 1994, 20). It is ironic that one of scientists’ most deeply held values, the respect for facts, should be a key source of distrust. There is, however, a deeper ethical problem here that relates to public participation in decision making. Thus far, purification and hybridization have been presented as alternative approaches to risk. It must be admitted that the argument from hybridization may not seem particularly compelling to someone trained in the sciences (or

336

12 Communication, Education and the Problem of Trust

philosophy). However, to arbitrate the contest between purification and hybridization in light of scientific or logical rigor is to misunderstand the burden of proof that hybridization demands. Ethical problems in food biotechnology are ethical precisely because someone might be responsible for doing something about them, and because that responsibility might entail social, institutional or governmental enforcement. Reasonable people can reach the conclusions of the hybrid interpretation without violating any canons of common sense. If we assume (as I do) that reasonable people should not be arbitrarily excluded from debate over policy and enforcement, then adherence to moral purification cannot be arbitrarily chosen as the standard for participation in debate. As such it is incumbent on those who would reject the hybrid view to justify the use of purification as an exclusionary tactic in governance and public policy, or failing such justification to ameliorate exclusionary applications of power with more inclusive political procedures. Put another way, those who choose the hybrid interpretation will be excluded from the seminars and lectures of scientists and philosophers. When they find themselves excluded from political or economic power the situation not only becomes more serious, but the fact of exclusion reinforces and validates the inferences that gave rise to the hybrid interpretation in the first place. It is now important to understand the system of purification in both narrow and broad terms. Narrowly, it is a set of ethical concepts that allow us to partition the complex welter of inchoate ethical concerns into logically distinct categories. Purification also reflects a broader set of intellectual categories that serve as principles for organizing knowledge into disciplines, departments and areas of concentration and for organizing at least some governmental authorities into agencies, administrations, services, and offices. Environmental impact is studied by environmental scientists and regulated by environmental agencies. Animal welfare is studied by physiologists and ethologists and is regulated (if at all) by institutional care committees. Toxicology and pharmacology are the scientific provenance of food safety. U.S. agencies such as the Food and Drug Administration (FDA) and Food Safety Inspection Service (FSIS) have analogues in most nations. Social consequences are studied by economists and sociologists and ethics is part of philosophy, but government does not regulate in these areas. Most industrialized countries afford a role in government to each of the categories. This role bestows political authority on the purified ethical analysis of animal biotechnology. What is more, funding for research and policy action also follows the bureaucratic organizational lines that conform to a purified view of risks. The hybrid interpretation is excluded, and it is important to see how. Ethical criteria effect policy change only when policy makers can apply them in enforcing the law. As Bernard Rollin notes in his work on biotechnology, religious believers may feel a moral obligation to practice rituals central to their faith, but the use of political authority to enforce such practices is now rare. As a practical matter, political authority is mustered by convincing individuals in positions of power, be they monarchs, legislators, judges or bureaucrats, that the mandate under which they wield their power justifies or perhaps requires action. Only absolute monarchs, however, are defined as having unlimited mandates. The more usual case is represented by the Food and Drug Administration of the U.S. Government, which has

12.8 Moral Reductionism and Political Exclusion

337

a clear mandate to enforce criteria that relate to human health, but no authority at all to even consider social consequences. This means that issues relating to human health count both ethically and legally, while issues relating to social consequences are not subjected to a legally binding test. Under the United States Constitution and the current Federal Code, authority to deal with social consequences reverts to the United States Congress, an institution unlikely to act on this issue in the foreseeable future due to sharp political divisions that transcend the problems of the agriculture and food sectors of the U.S. economy. However, the simultaneous inclusion of food safety and exclusion of religion and social consequence from public policy in the U.S. gives legal force to purification, and deprives the alternative worldview of symmetrical legal standing. To summarize: FDA will regulate based on a pure, probabilistically based interpretation of safety. An understanding of safety in which social consequences have a bearing on one’s feeling of well-being, for example, will be ruled out without being taken seriously. Those who advocate such a view have no standing, and are, as a matter of fact, likely targets of ridicule. The system of purification is invested with political authority and power. At the same time, the system of purification reflects, to a large degree, the set of categories that define divisions of knowledge within academic and scientific research institutions. Each category of consequence, human, animal, ecosystem and social, would be the object of study by separate departments, disciplines or sub-disciplines. These departments are routinely (though far from universally) seen to be operating in logically distinct spheres. There is thus a double institutionalization of the system of categories produced by purification, first in government and second in the academic departments of the sciences, including the social sciences and to some extent the humanities. The significance of this double set of social institutions is subtle and complex. On the one hand, it may be interpreted as validation of the order produced by the purification itself, suggesting that similar patterns of purification have been replicated and reproduced in a variety of otherwise independent contexts. On the other hand, it may be that government and science have co-evolved so as to produce mutually consistent organizational divisions for addressing complex issues. The social histories implied by each of these two alternatives raise large and deep philosophical issues that must be set aside here, but so long as the second alternative is plausible, the anxiety that arises from seeing purification as a form of power seeking is not only warranted, but increased. Seen from one vantage point, the system of purification establishes a leviathan of science and government. The basic assumptions that partition knowledge also partition government power. Those who do not share the basic assumptions or who rely on ordinary language rather than technical definitions of concepts are outsiders. Their arguments have no standing and cannot be converted into policy by the agencies that have been established with limited mandates. Scientists and scientific organizations, in the meantime, have been placed at the center of the leviathan. They control the definitions that are used to translate regulatory mandates into operational terms. They do the research that will form the empirical basis for policy decisions. It is unfair to suggest (as critics have) that scientists have an interest in manipulating the results of that research, for the long term viability of the leviathan depends upon objective

338

12 Communication, Education and the Problem of Trust

research procedures. It is entirely fair, however, to say that scientists have an interest in maintaining the structure of the leviathan, for it assures their status and the continuing demand for their services. Both government and private industry need scientific institutions to perform the dual function of defining criteria and evaluating specific products or technologies, and this need establishes a market for science, both in the form of jobs and research funds. This means, however, that scientists have an interest in preserving the system of purification. Those who would propose alternative interpretations are, thus, enemies of science, not in any elevated philosophical sense, but in having adopted basic assumptions that fail to support the system of purification that links science, government and private industry in a mutually supportive network. It is important to not to overstate the case here. An individual scientist’s or lab’s interests may, of course, diverge from those of the leviathan. There have been many individual scientists who have found reasons to criticize prevailing attitudes on agrifood biotechnology, for example, though virtually all of them have done so within the rigid risk categories of the purification point of view. The political winds that influence policy at the highest levels of government may also diverge from those of mainstream science. The case of climate change science is a case in point: U.S. policy in Republican administrations administration has never embraced the consensus view on climate change. Nevertheless, the convergence of epistemic style—a commitment to the even-predicting way of understanding risk, on the one hand, and the purificationist’s tendency to see exposure mechanisms in chemistry, ecology and society creating wholly discrete categories of hazard—and the administrative approach to regulation creates a powerful, one can even say insurmountable, obstacle to those who cannot or will not express their concerns and arguments in the prevailing rhetorical form.

12.9 Conclusion In concluding this chapter for the first edition of this book in 1997, I wrote that while the future of food biotechnology is bright, and while the ethical issues seem far less than overwhelming, some of us who are optimists and would-be boosters remain cautious. “Even if biotechnology fails to live up to its ethical and social promise, it will not bring about disaster. But it could live up to its promise as the first important science of a 21st century, a century where science recovers its moral compass and its position of leadership in social issues.” Writing a decade later for the second edition, I wrote that “whether or not this future will be realized still remained to be seen.” In writing what will certainly be the last revision, I note that some things do seem to be changing. Philosophers of science are taking an interest in the questions I was struggling with in the mid-1990s. Throughout Europe, the controversy over agrifood biotechnology spawned and uptick of interest in ethical study of technical innovation, and the European Union has developed programs to support work in this vein. A spurt of similar interest in “anticipatory governance” accompanied the

12.9 Conclusion

339

National Nanotechnology Initiative (NNI) in the United States, and The Journal of Responsible Innovation launched in 2014. While these are salutary developments, my concluding observation is that too much of this work continues to be focused on a naïve model of public engagement. Scholars in science studies have taken the deficit model to heart, and much of the work supported by responsible innovation grants in Europe or the NNI in the United States has used creative means to pique a non-scientists interest through games, video treatments, citizen science or science museums. Yet while some of these efforts do help people learn about science, and they may also make people more accepting of potentially scary technology, I would question their engagement with ethical issues. In the concluding chapter for the 2020 edition, I will offer a framework for addressing issues in a more systematic manner.

References Ames, B.N. 1983. Dietary carcinogens and anticarcinogens: Oxygen radicals and degenerative diseases. Science 221: 1256–1263. Baier, A. 1994. Moral Prejudices: Essays on Ethics. Cambridge, MA: Harvard University Press. Baier, A. 1986. Poisoning the wells. In Values at Risk, ed. D. MacLean, 49–74. Totowa, NJ: Rowman and Allanheld. Bauer, M.W., and G. Gaskell (eds.). 2002. Biotechnology: The Making of a Global Controversy. Cambridge, UK: Cambridge University Press. Borlaug, N. 2001. Ending world hunger. The promise of biotechnology and the threat of antiscience zealotry. Plant Physiology 124: 487–490. Borlaug, N. 2000. The Green Revolution Revisited and the Road Ahead. Special 30th Anniversary Lecture. Norwegian Nobel Institute, Oslo. https://www.biologyjunction.com/green%20revolut ion.pdf. Accessed 15 March 2020. Brossard, D., and B.V. Lewenstein. 2010. A critical appraisal of models of public understanding of science: Using practice to inform theory. In Communicating Science: New Agendas in Communication, ed. L.A. Kahlor and P. Stout, 11–39. New York: Routledge. Buchman, M. 1995. The impact of resistance to biotechnology in Switzerland: A sociological view of the recent referendum. In Resistance to New Technology: Nuclear Power, Information Technology, and Biotechnology, ed. M. Bauer, 207–224. Cambridge, UK: Cambridge University Press. Busch, L., W.B. Lacy, J. Burkhardt, and L.R. Lacy. 1991. Plants, Power and Profit: Social, Economic, and Ethical Consequences of the New Biotechnologies. Cambridge, MA and Oxford: Basil Blackwell. Chess, C. 1996. The public and agricultural biotechnology: Key questions. In Agricultural Biotechnology: Novel Products and New Partnerships, eds. R.W. Hardy and S.B. Segelken. Ithaca, NY: National Agricultural Biotechnology Council. Chess, C., and B.B. Johnson. 2007. Information is not enough. In Creating A Climate For Change: Communicating Climate Change And Facilitating Social Change, eds. S.C. Moser and L. Dilling, 223–233. New York: Cambridge University Press. Cook, G., E. Pieri, and P.T. Robbins. 2004. ‘The scientists think and the public feels’: Expert perceptions of the discourse of GM food. Discourse & Society. 15: 433–449. Cranor, C.F. 1993. Regulating Toxic Substances: A Philosophy of Science and the Law. New York: Oxford University Press.

340

12 Communication, Education and the Problem of Trust

van Dommelen, A. 1995. Quality of risk assessment: Artificial and fundamental controversies. In Contested Technology: Ethics Risk and Public Debate, ed. R. von Schomberg, 193–208. Tilburg NL: International Centre for Human and Public Affairs. Douglas, H. 2000. Inductive risk and values in science. Philosophy of Science 67: 559–579. Douglas, H. 2009. Science, Policy and the Value-Free Ideal. Pittsburg, PA: University of Pittsburg Press. Durant, J., M.W. Bauer, and G. Gaskell (eds.). 1998. Biotechnology in the Public Sphere. London: The Science Museum. Elliott, K.C. 2017. A Tapestry of Values: An Introduction to Values in Science. New York: Oxford University Press. Forsman, B., and S. Welin. 1995. The treatment of ethics in a Swedish Government Commission on Gene Technology. Studies in Research Ethics No 6. Centre for Research Ethics, The Royal Society of Arts and Sciences in Gothenburg, Sweden. Freeman, A.M., and P.R. Portney. 1989. Economics clarifies choices about managing risk. Resources 95 (Spring): 1–4. Frewer, L. 1998. Consumer perceptions and novel food acceptance. Outlook on Agriculture 27: 153–156. Frewer, L.J., and R. Shepherd. 1995. Ethical concerns and risk perceptions associated with different applications of genetic engineering: Interrelationships with the perceived need for regulation of the technology. Agriculture and Human Values 12 (1): 48–57. Frewer, L.J., R. Shepherd, and P. Sparks. 1994. Biotechnology and food production: Knowledge and perceived risk. British Food Journal 9: 26–33. Friedman, M., and L.J. Savage. 1948. The utility analysis of choices involving risk. Journal of Political Economy 56: 279–304. Gardner, J. 2015. The many faces of the reasonable person. Law Quarterly Review 131: 563–584. Gaskell, G., and M.W. Bauer (eds.). 2001. Biotechnology 1996–2000: The Years of Controversy. London: The Science Museum. Griffiths, A. 1996. Implementation of biotechnology research and development policy: implications for agricultural sustainability. Dissertation submitted in partial fulfillment of the requirements for the Doctor of Philosophy in Environmental Studies. Vancouver, CA: University of British Columbia. Gross, P.R., and N. Levitt. 1994. Higher Superstition: The Academic Left and Its Quarrels with Science. Baltimore: Johns Hopkins University Press. Hallman, W.K., and J. Metcalfe. 1994. Public Perceptions of Agricultural Biotechnology: A Survey of New Jersey Residents. New Brunswick, NJ: New Jersey Agricultural Experiment Station. Hamilton, C., S. Adolphs, and B. Nerlich. 2007. The meanings of ‘risk’: A view from corpus linguistics. Discourse & Society 18: 163–181. Hansen, J., L. Holm, L. Frewer, P. Robinson, and P. Sandøe. 2003. Beyond the knowledge deficit: Recent research into lay and expert attitudes to food risks. Appetite 41: 111–121. Hansson, S.O. 2018. Risk. In The Stanford Encyclopedia of Philosophy, ed. E.N. Zalta (Fall 2018 Edition). https://plato.stanford.edu/archives/fall2018/entries/risk/. Hoban, T.J., and P. Kendall. 1993. Consumer Attitudes about Food Biotechnology. Raleigh, NC: North Carolina Cooperative Extension Service. Hornig, S. 1993. Reading risk: Public response to print media accounts of technological risk. Public Understanding of Science 2: 95–109. Kahneman, D. 2011. Thinking, Fast and Slow. New York: Macmillan. Kitcher, P. 2001. Science, Truth and Democracy. New York: Oxford University Press. Kneese, A.V., S. Ben-David, and W.D. Schulze. 1983. The ethical foundations of benefit-cost analysis. In Energy and the Future, ed. D. MacLean and P.G. Brown, 59–74. Totowa, NJ: Rowman and Allenheld. Krebs, J. R. 2000. GM foods in the UK between 1996 and 1999: Comments on ‘Genetically modified crops: risks and promise’ by Gordon Conway. Conservation Ecology 4(1): 11. https://www.con secol.org/vol4/iss1/art11/

References

341

Kuzma, J., and J.C. Besley. 2008. Ethics of risk analysis and regulatory review: From bio-to nanotechnology. Nanoethics 2: 149–162. Ludwig, D. 1993. Environmental sustainability: Magic, science, and religion in natural resource management. Ecological Applications 3: 555–558. Plein, C.L. 1991. Popularizing biotechnology: The influence of issue definition. Science, Technology & Human Values 16: 474–490. Priest, S.H. 2000. US public opinion divided over biotechnology? Nature Biotechnology 18: 939– 942. Priest, S.H. 2001. A Grain of Truth: The Media, the Public, and Biotechnology. Lanham, MD: Rowman & Littlefield Publishers. Priest, S.H., and J. Talbot. 1994. Mass media and the ultimate technological fix: Newspaper coverage of biotechnology. Southwestern Mass Communication Journal 10 (1): 76–85. Rabino, I. 1991. The impact of activist pressures on recombinant DNA research. Science, Technology and Human Values 16: 70–87. Rabino, I. 1994. How European and U.S. genetic engineering scientists view the impact of public attention to their field: A comparison. Science, Technology and Human Values 19: 23–46. Randerson, J. 2008. Arpad Pusztai: Biological divide. The Guardian. Accessed 10/27/2020 at https:// www.theguardian.com/education/2008/jan/15/academicexperts.highereducationprofile. Rescher, N. 1983. Risk: A Philosophical Introduction to the Theory of Risk Evaluation and Management. Lanham, MD: University Press of America. Ruckleshaus, W.D. 1983. Science, risk, and public policy. Science 221 (1026): 1028. Schechtman, M. 2011. The narrative self. In The Oxford Handbook of the Self , ed. S. Gallagher. New York: Oxford University Press. Schroeder, C.H. 1986. Rights against risk. Columbia Law Review 86: 495–562. Shrader-Frechette, K.S. 1991. Risk and Rationality. Berkeley: University of California Press. Sinnott-Armstrong, W. 2005. It’s not my fault: Global warming and individual moral obligations, Perspectives on Climate Change: Science, Economics, Politics, Ethics. Advances in the Economics of Environmental Research 5: 293–315. Slovic, P. 1999. Trust, emotion, sex, politics, and science: Surveying the risk-assessment battlefield. . Risk Analysis 19: 689–701. Sparks, P., R. Shepherd, and L.J. Frewer. 1994. Gene technology, food production and public opinion: A UK study. Agriculture and Human Values 11 (1): 19–28. Sparks, P., R. Shepherd, and L.J. Frewer. 1995. Assessing and structuring attitudes towards the use of gene technology in food production: The role of perceived ethical obligation. Journal of Basic and Applied Social Psychology 16: 267–285. Stanovich, K. 2011. Rationality and the Reflective Mind. New York: Oxford University Press. Starr, C. 1969. Social benefit versus technological risk. Science 165: 1232–1238. Starr, C., R. Rudman, and C. Whipple. 1976. The philosophical basis of risk analysis. Annual Review of Energy 1: 629–662. Sturgis, P., and N. Allum. 2004. Science in society: Re-evaluating the deficit model of public attitudes. Public Understanding of Science 13: 55–74. Taylor, K. 1991. Many voices: Citizens and the issues of biotechnology and food safety. In Agricultural Biotechnology, Food Safety, and Nutritional Quality for the Consumer, ed. J.F. MacDonald, 96–102. Ithaca, NY: National Agricultural Biotechnology Council. Teitelman, R. 1989. Gene Dreams: Wall Street, Academia and the Rise of Biotechnology. New York: Basic Books. Thompson, P.B. 1985. Risking or being willing: Hamlet and the DC-10. The Journal of Value Inquiry 19: 301–310. Thompson, P.B. 1986. Uncertainty arguments in environmental issues. Environmental Ethics 8: 59–75. Thompson, P.B. 2018. Communicating science-based information about risk: How ethics can help. In Ethics and Practice in Science Communication, ed. S. Priest, J. Goodwin, and M.F. Dahlstrom, 34–54. Chicago: University of Chicago Press.

342

12 Communication, Education and the Problem of Trust

Thompson, P.B. 1990. Risk subjectivism and risk objectivism: When are risks real? Risk: Issues in Health and Safety 1(1):3–19. Trachtmann, L. M, and R. Perrucci. 2000. Science under Siege? Interest Groups and the Science Wars. Lanham, MD: Rowman and Littlefield. Wandersman, A.H., and W.K. Hallman. 1993. Are people acting irrationally? Understanding public concerns about environmental threats. American Psychologist 48 (6): 682–686. Winner, L. 1983. Techné and politeia: The technological constitution of a society. In Philosophy and Technology, ed. P.T. Durbin and F. Rapp, 97–111. Dordrecht: D Reidel. Zimon, J. 1992. Not knowing, needing to know, and wanting to know. In When Science Meets the Public, ed. B.V. Lewenstein, 13–20. Washington DC: American Association for the Advancement of Science.

Chapter 13

Gene Editing, Synthetic Biology and the Next Generation of Agrifood Biotechnology: Some Ethical Issues

Abstract The chapter provides synoptic overviews on key developments in gene technology since publication of the 2nd edition in 2007. Synthetic biology is discussed briefly, and more attention is given to CRISPrCas9, and gene editing. Both techniques can increase the speed at which a new product would move through the R&D process, and both have the potential to increase systemic linkages between gene technologies for food and agriculture, and gene technology for biomedical purposes. Beyond this, lessons learned from the experience with GMOs continue to be relevant. The framework of novel and normal risk will be a useful amendment to the technological ethics framework developed in earlier editions of this book. Three case studies are discussed: alternative proteins, horizontal environmental genetic alteration agents and gene drives for agricultural pest control. Only the last of these involves truly novel risks. Keywords CRISPrCas9 · Synthetic meat · Insect allies · Gene drives · Platform technology Gene editing refers to a suite of tools and practices in genetic modification that started to be applied in product development around 2010. The suite includes zinc fingers, TALONS and CRISPrCas9, each of which has a somewhat different mechanism of operation. Gene editing followed on the heels of whole genome assembly and synthetic biology. I will argue that in the agrifood sector, experience with GMOs provides a sound basis for approaching the ethics of these new methods. We should not presume that even newer methods in gene technology are impossible, nor should we assume that differences in the power or mechanism of these yet-to-be-imagined tools lack ethical significance. Nevertheless, statements on either matter would be speculative at the present juncture. As such, this chapter will concentrate on what unites them under the heading of gene editing, with some brief prefatory remarks on synthetic biology. Synthetic biology has thus far had limited applications in the food sector. Each the three gene editing tools provides the researcher to make a change in a plant or animal genome at a known location in the sequence of base pairs. This chapter will first offer a synoptic discussion of what this ability means in the context

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_13

343

344

13 Gene Editing, Synthetic Biology and the Next Generation …

of agricultural and food biotechnology prior to reconsidering the risks that might derive from the use of gene editing in food and fiber commodities.

13.1 Synthetic Biology: The Old New Thing Even as the 2nd edition of this book was going to press, the attention of scholars was being drawn to new technical forms. A global push to develop technical capacities operating at the nanoscale dominated discussions of technological ethics from 2005 to 2010. A nanometer is one billionth of a meter; technologies that deploy mechanisms at the nanoscale exploit chemical and physical properties that function at a range between 1 and 100 nanometers. At the nanoscale, the surface area of a particle becomes more significant than its molecular weight and it becomes able to move through its environment in ways that even microscopically small particles could not. While there was an early surge of interest in food nanotechnology, the enthusiasm subsided quickly. For one thing, a lot of chemistry happens at the nanoscale, and hopes for food technology turned out to be hypes: little more than a re-description of longstanding projects in food and agricultural chemistry. Encapsulation, for example, allows for greater control of both agricultural chemicals and food ingredients. The timing or dose of a chemical might be controlled by encapsulating the active ingredient in a coating responsive to sunlight, to water or to some other chemical applied at the right time. Encapsulated flavors might be released at a specific moment in cooking or even eating, (Sanguansri and Augstin 2006). Some (not all) techniques for encapsulation exploit properties that function at the nanoscale. Yet encapsulation is not a new thing in food technology: Think, “melts in your mouth, not in your hand.” Truly novel forms of food nanotechnology ran into a regulatory barrier. Toxicological models for understanding how contaminants or food ingredients move through the body are calibrated for the microscale, but a nanoparticle (e.g. a constituent ingredient of less than 100 nanometers in size) would behave differently. Its reactive properties would differ as well, and the variation in size among nanoparticles would frustrate many methods for food safety risk assessment. Lacking methods for a reliable estimate of exposure and with the nanoscale posing the potential for unknown hazards, regulatory agencies have exhibited ethically appropriate caution in approving nanoscale technologies in food and agriculture, where there is a strong chance that they will interact with other organisms, including human beings. While encapsulation technologies probably have little ethical significance, nanoparticles challenge key methods in toxicology, illustrating the difference between normal and novel risk, discussed at more length in Chap. 14, (see also Thompson and Hannah 2012). This classification does not imply that normal risks are being managed in an ethically justifiable way. Nevertheless, to the extent that technological ethics aims to highlight unique or unexpected consequences from technological innovations, the questions shift from thinking through possible impact of a tool or technique to

13.1 Synthetic Biology: The Old New Thing

345

assuring the integrity of institutions designed to address routine matters of public health and environmental quality, (Thompson 2010a). Yet even as enthusiasm for food and agricultural applications of nanotechnologies had started to wane, the next wave in gene technologies appeared. Synthetic biology is manipulation of nucleic acids creating a novel genetic sequence, not found in nature. This definition leaves a number of interpretive questions open, and it is questionable as to whether anyone has actually succeeded in creating a functional synthetic organism. However, it is possible to develop strings of base pairs that, when inserted into the nucleus of single-celled organism, function similar to a molecule of DNA produced through mitosis. If this artificially created genome includes appropriate modifications, the organism might produce a specific protein or biologic, just like conventional bioengineering. Amyris describes itself as a scientific research company that successfully developed a biological platform for producing artemisinin, an important anti-malarial compound. Amyris’s achievement suggests that other similar techniques might be developed for producing food ingredients or biofuels, (Kendig 2014). More generally, the BioBricks project for developing sequences that function as interchangeable parts for assembling biological systems exemplify this approach to gene technology. Along with J. Craig Venter’s artificial assembly of a known sequence for a microorganism, these projects provide an ostensive definition for synthetic biology, as the term is being applied in practical contexts, (Thompson 2012). Like agrifood nanotechnologies, we should ask whether synthetic biology poses novel risk in the food sector. It is important to recall that current regulatory policy for food safety requires extensive toxicological and clinical testing for compounds that do not already appear in foods Generally Recognized as Safe (GRAS) (discussed in Chap. 4). There are thus regulatory filters that reduce the immediacy of any ethical imperative to ponder questions about the novelty of synthetic biology in agriculture and food production. As in previous chapters, I have argued that more compelling questions arise in conjunction with entirely unregulated socio-economic impacts of these technologies. For example, smallholder wormwood farms are the traditional source for artemisinin. Prior to the Amyris innovation, poor farmers in in Madagascar, Pakistan and elsewhere were developing wormwood plantings in an attempt to enter a market dominated by Chinese farmers, (Faurant 2011). This is not a new concern. As Chaps. 9 and 10 document, similar concerns about the impact of gene technology on the economic survival of smallholder farming reside at the heart of debates over the first generation of GMOs. If techniques of synthetic biology allow well-capitalized pharmaceutical companies to displace the livelihoods of people already on the margin of survival, that is an ethical issue. However, it is not an unprecedented issue, (Thompson 2015). Yet even as debates over synthetic biology were superseding debates over food and agricultural nanotechnologies, they were themselves overshadowed by the appearance of more precise methods for gene insertion, or gene editing. Synthetic biology was no longer the new new thing. The balance of this chapter emphasizes the debates that are attending to the emergence of gene editing, but the larger point is that whatever the next new thing turns out to be, it would behoove scholars working in the

346

13 Gene Editing, Synthetic Biology and the Next Generation …

philosophy and policy of emerging technologies to carry the lessons learned from previous debates into their scholarship, (see Scott 2018). Indeed, in examining the case studies that close the chapter, the particular tools for genetic modification or cellular manipulation matter little. Consistent with a comment made at the very beginning of the first edition, one can make sense of the ethical issues in agrifood biotechnology without knowing much about how gene technologies actually work. While I have walked back that claim a bit in Chap. 1 of the current edition, it is still true that much of what is ethically significant about a cellular technology, such as is being used to develop alternatives to meat, can be derived from debates over imagined applications of gene transfer. The moral imagination anticipates the science, and while we should not simply assume that that the details never matter, neither should we assume that a new technique vitiates the lessons that should have been learned from the last new thing (or the one before that).

13.2 Gene Editing and the Next Generation of Biotechnology Like the next new thing, the metaphor of new generations signals significant technical change in technical capability. New software releases distinguish between minor alterations, fixes and upgrades and large-scale rewrites that make significant improvements. It is not at all obvious how one applies the metaphor in biotechnology, but however many generations have gone before, it seems right to say that gene editing will usher in the next one. As noted in Chap. 1, GMOs in use prior to 2010 were developed using techniques that insert transgenes at random locations in the modified organism’s genome. This has a number of disadvantages that can be summarized in terms of a low rate of success. Transgenes might not be incorporated at all, or they may be incorporated at points where they are either not expressed or disrupt some other gene process crucial to the organism’s survival. These techniques required significant screening efforts to identify the singular individuals with appropriate levels of the desired transformation for developing a GMO. This screening is followed by generations of backcrossing with non-transgenic individuals to develop plant varieties or animal breeds with the full range or traits that farmers want. Importantly, the potential for disruption of existing gene functions entails the possibility of unknown effects, which were a prime source of concern for food safety and animal health. In the age of gene editing, this potential is called a non-target effect. The ability to make changes at a known locus gave researchers more control over these problems. Zinc finger nucleases and transcription activator-like effector nucleases (TALENs) are enzymes that cause breaks or cuts in DNA at specific locations. Genes can then be inserted or disabled at those loci, and the biochemical properties of nucleic acids will reconnect or bind the broken bits, retaining the change. Publications documenting practical applications of zinc fingers and TALENs began to appear around 2010. Transformation using clustered regularly interspaced short palindromic

13.2 Gene Editing and the Next Generation of Biotechnology

347

repeats (CRISPR) promises greater flexibility and lower cost. With this tool, Cas9 enzymes are coupled with distinctive base-pair sequences to enable transformation at a much larger range of locations on the genome, (Zhang et al. 2018). Gene editing tools also open up a theoretical possibility for non-germline modification. For example, there has been much discussion on the potential for using a CRISPR Cas9 construct as a therapeutic strategy for treating genetic disorders. Individuals who have inherited gene constructs associated with a disorder such as cystic fibrosis, Huntington’s chorea or sickle cell anemia could, in theory, be treated by CRISPR Cas9 systems that would eliminate the disease by replacing the offending gene sequence with a “normal” (e.g. non-disease causing) sequence in the relevant cells of that individual’s body. Notably, this therapy might not cause changes in that individual’s egg or sperm, so the “cure” would not heritable. That is, the individual’s progeny would still be subject to inheriting the genetic disorder (see FDA 2017 for more background). The ability to make non-heritable changes also has significant advantages in the context of agriculture, as will be discussed below. One of the alleged advantages of gene editing is that it allows for changes in the genome that do not involve DNA, (see Zhang et al. 2018), while a less dramatic claim states that they need not contain novel proteins (Van Eenennaam 2018). These claims are material to some of the ongoing debates over regulatory oversight, but they can be easily misinterpreted in the context of public attitudes and ethical issues. Alison Van Eenennaam makes her claim about novel proteins in reference to a geneedited animal, but the FDA’s approach to regulating GMOs has always insisted on a higher regulatory standard for plants that contain novel proteins. Animal knockouts (common in medical biotechnology) might also lack novel proteins, so even here there is no clear reason why gene editing makes a difference. As noted in Chap. 5, Van Eenennaam’s hornless cows were subsequently found to contain an unexpected gene sequence, undercutting the rhetorical force of her argument, (Regalado 2019). The actual issue is that while most agricultural GMOs involve the addition of a gene construct, there are instances where a desired change can be brought about by either deleting genes (e.g. knockouts) or regulatory sequences. In previous generation biotechnology, this was done by introducing base pairs to intentionally disable activity, to “mess up” the gene and render it incapable of initiating the performance of specific function. Gene editing allows sequences to be removed, leaving no “foreign” DNA left in the organism after modification. From a regulatory standpoint, this poses both practical and ethical issues. From a practical standpoint, it is no longer clear how one could use existing genomic tools to determine whether an organism had been modified. In the past, the base pairs introduced to disable gene function served as a marker that distinguishes a GMO from a non-GMO, but with a sub-class of gene edits, there is no marker. The ethical issues intertwine with regulatory practice, producing something of a tangle. The absence of a reliable marker means that, for the time being, at least, regulations might not be enforceable if some bioengineer was inclined to cheat. At the same time, since no DNA (the phrase is often “foreign DNA”) is introduced into the genome, some argue that there is no basis for regulation in the first place. If the modified sweet pea is just like any other sweet pea save for the absence of a few base

348

13 Gene Editing, Synthetic Biology and the Next Generation …

pairs, there is no substantive basis for claiming that it differs from every other sweet pea. Escaping regulatory hurdles is important for innovators, as simply generating the data to assure regulators is both costly and time consuming. Similar arguments were made for “cisgenic” or “intragenic” GMOs, that is, plants or animals altered using first generation tools to incorporate genes that were drawn from organisms of the same species as the target organism, (Schouten et al. 2006; Telem et al. 2013). A brief revisitation of the ethical issues in that debate will thus prove helpful. One factor in favor of this argument is that public opinion research supports the claim that many people do, in fact, associate problems with the fact that transgenic GMOs have introduced gene constructs from other species, (Delwaide et al. 2015). The question is, what problems, and how do they figure in regulatory decision making? Bjorn Myskja argues that simply respecting the opinions of non-scientists provides an ethical rationale for favoring cisgenic over transgenic GMOs. He supports this by arguing that the risks and uncertainties of cisgenic GMOs are lower, and that this, in turn, can be linked to an ethic of showing respect for the complexities of nature, (Myskja 2006). Wendy Russell and Rob Sparrow respond to Myskja’s claims by insisting that the “foreignness” of a gene is of little relevance, whatever the public might think. They list a number of actual and proposed cisgenic organisms that pose significant ecological risks due to phenotypic interaction with the environment. They also briefly note (but do not stress) the potential for unnoticed but potentially harmful changes in genes, (Russell and Sparrow 2008). These are what today would be called “off target” effects. In the era of gene editing, off target effects have become the focus of contention. There are several distinct uses of the words ‘target’ in risk assessment for genetic engineering and GMOs. The potential for harm to “non-target organisms” was one of the environmental hazards identified for earlier GMOs, discussed in Chap. 7. Here, the organism being transformed was the target, and the concern was that a GMO might have an impact on other non-target organisms, either through gene flow or some other exposure mechanism. In the case of Bt crops, exposure to the toxin could harm insects that were not crop pests, such as Monarch butterflies. Impact on nontarget organisms is a concern that ranks high in the evaluation of chemical pesticides. In the present case, an “off target” effect could be broadly defined as any change in the base pair sequence that is other than the one a plant or animal scientist intends to effect. More narrowly it is a sequence alteration that occurs at some locus other than the one targeted by the gene editing tool. The target is the location on the genome where an alteration is being attempted, and changes anywhere else count as off target or non-target effects. In a literal sense, earlier methods of transformation lack any precise target at all. Yet they certainly had the potential to disrupt the functionality of other genes, depending on where they did land. This was as true for cisgenic as transgenic alterations. The potential for hazards arising from these alterations are one focus of Russell and Sparrow’s critique, and they are, in fact, a central reason why regulatory agencies like to see data on the biochemical constituents of a GMO before they render an opinion. Food safety regulation of GMOs has proceeded under the assumption that genes derived from foods (or any GRAS organism, see Chap. 4) would produce the same

13.2 Gene Editing and the Next Generation of Biotechnology

349

substance (generally a protein) they produced in the organism from which they were derived when they were moved to a new food organism. However, companies and regulatory agencies recognized that the randomness of insertions could, at least in theory, make other changes (that is, in addition to the insertion of the transgene), and these could create hazards. This general philosophy for food safety would not suggest that there are no significant differences between cisgenic and transgenic GMOs. In both cases, there are reasons to suppose that the gene product is not inherently risky, and in both cases the potential for random mutations introduces the potential for hazards. Note also that although I have discussed food safety hazards to illustrate the point, off-target effects could also introduce environmental hazards or deleterious effects on animal health. The regulatory approach for dealing with this potential drew upon plant breeders with other advanced breeding techniques including embryo transfer and mutation breeding. These, too, had the potential to create unwanted genetic changes, but dozens of products from these advanced breeding techniques were already on the market when the first generation of GMOs started to be introduced in the 1990s. There is thus inductive support for thinking that minor alterations of the genome would not cause hazards that would be especially difficult to detect. In other words, the kind of monitoring and measurement already performed in the process of plant and animal breeding was able to detect hazards long before they appeared in the food chain. For example, maize breeders early recognized that genetic factors could affect aflatoxin contamination, (Nagarajan and Bhat 1972) and later discovered that high lysine varieties developed through conventional breeding were associated with this risk. The result was a signficant change in the effort to produce high lysine varieties, irrrespective of any action by a regulatory agency, (Betrán et al. 2006). Nevertheless, regulatory review of GMOs has required supporting data showing that the biochemical composition of a GMO falls within the range established for nonmodified examples of the relevant plant or animal type, and this is one of the key things that leads to additional expense. Does gene editing make any difference? The reason to think that it might is that more precise targeting reduces the potential for unwanted changes in the genome. It does not, however, reduce off-target effects to zero, (Zhang et al. 2015). Thus, there is one important respect in which the risk ethics are back to the same place as the earlier debate over cisgenic GMOs. It may be easier to sell the public on the idea that a gene edit which does not introduce “new” DNA is more acceptable, but all the reasons that Russell and Sparrow cite for opposing deregulation of cisgenic GMOs would apply to gene edited plants and animals, as well. Most significantly, there can certainly be phenotypic changes that will interact with other organisms in an environment, and it these interactions cannot be presumed to be risk-free on an ex ante basis. Secondly, if the potential for off-target (that is unintended) changes in genes is lower for gene edited organisms than for earlier GMOs, this probably strengthens the inductive argument for thinking that any unwanted phenotypic traits will be picked up by the usual procedures of plant and animal breeding. But does it provide a basis for exempting these plants and animals from all regulatory review whatsoever?

350

13 Gene Editing, Synthetic Biology and the Next Generation …

Before answering this crucial question, it is important to note a potential weakness in Russell and Sparrow’s argument. They rely heavily on their recitation of risks from cisgenic GMOs, but the phenotypic traits that are the source of these risks might well be introduced through advanced breeding techniques (such as mutation breeding) or even traditional crosses. For example, they cite the potential for ecological impact from large or fast-growing fish, but while the celebrated Aqua Bounty™ salmon is a prominent GMO analyzed for its possible environmental impact, this impact follows from traits that are achievable through totally unregulated traditional breeding techniques. In fact, the same goes for off-target effects on a genome. Minor changes in gene sequence are associated with other breeding technologies, including cellculture and mutation breeding. Now it is certainly possible that Russell and Sparrow would respond to these observations by saying that all of these products should be regulated, based on their environmental or food-related risks. However, the fact is that they are not regulated: not in the United States, not in Europe and not in Australia, which is the reference point for Russell and Sparrow’s analysis. There is thus an additional question to answer: on what basis would we regulate gene edited crops or livestock when we do not regulate crops or livestock produced by other means that have a similar risk profile? Writing with Anne Ingeborg Myhr, Myska argues that that a fair regulatory system must indeed treat products with similar risk profiles similarly. In the same paper, Myska and Myhr argue that social impacts can also be included in a fair assessment process, (Myska and Myhr 2020). This is consistent with the position I have been arguing since 1997, but I would note that I draw a distinction between a riskbased assessment undertaken for ethical review and the information that should guide regulatory decision-making. Social impacts will surface again in later sections of the chapter, but first, more needs saying about non-target effects. In brief, regulatory agencies generally lack the legislative mandate to extend their oversight beyond its current reach. Products of biotechnology fall under food safety governance in Europe because special provisions have been made to include them, and they are covered in the United States because the biotechnology industry has found it in their interest to voluntarily submit products for Food and Drug Administration (FDA) review. One ethical reason for continuing to exclude products of mutation breeding or cell culture is simply that we do not have an instance where members of the consuming public were exposed to significant risks from foods derived using advanced breeding techniques. Why fix it if it ain’t broken, especially when fixing it would raise the costs of innovation, and eventually, food itself? I have already sketched reasons why at least some gene-edited crops can reasonably be included under the same umbrella of reasoning that cautions against regulating products of cell-culture, mutation breeding and, indeed, traditional cross-breeding. Max Haeussler argues that this is true even when we evaluate the significance of off-target effects within the context of cell metabolism, (Haeussler 2019). However, this does not exhaust the tangle of ethical issues. The analogy to cisgenic GMOs applies to only a small percentage of the gene edited plants and animals we can expect to see emerging over the next decade. Many applications will involve transgenes (e.g. genes derived from the gene pool of some other organism), or,

13.2 Gene Editing and the Next Generation of Biotechnology

351

colloquially, “foreign genes.” This takes us all the way back to the claim that “not involving DNA” is an advantage for gene-edited crops. This is true only in the sense that no “new,” “foreign,” or transgenic DNA remains in the modified organism after the edit has been completed, and even then only in those cases where gene editing is used to delete a construct. It does not imply that nucleic acids (e.g. DNA or RNA) are uninvolved. As already stated, nucleases facilitate modification by breaking DNA at specific locations, whether new DNA is inserted at those loci or not. RNA is used to “guide” CRISPRs to the appropriate location for transformation. Finally, many applications of gene editing will be undertaken in order to introduce additional genes into the modified organism. I regard claims to the effect that gene editing “does not use or involve DNA” as misleading. This issue has become important within the community of molecular biologists because at this writing, the United States and Europe appear to be taking opposite paths, with FDA and USDA prepared to exempt gene-edited plants that do not involve transgenes, and Europe prepared to treat them just as they would any GMO. It is also important because FDA does not have a basis for similarly exempting modified animals, owing to the fact that they currently interpret any modification of metabolic function as triggering a review under their provisions for animal drugs. This includes a much higher level of data submission and review, including clinical data on animal health, as well as the food safety review. All of these regulatory tangles are subject to change, of course. An ethics treatment of the issues must emphasize what should be the case for regulatory oversight, rather than what legislatures have, in fact, demanded, or what regulators have interpreted their current mandates to require. The considerations reviewed thus far would appear to support lightening the burden for gene edited crops and livestock species that do not introduce transgenes. There is one more “however”, however. Russell and Sparrow are right to note the ecological risks associated with several examples of cisgenic GMOs, and my riposte has simply been to note that much of what we do in agriculture poses similar risks. This is not really an ethically adequate answer, and readers should be aware that I am quite critical of agriculture’s environmental impacts in other contexts. It would be reasonable to argue against additional scrutiny for products of biotechnology if it were the case that they were only exacerbating what is already a bad situation. That has not been the record for GMOs, however. Bt crops, in particular, appear to have lowered risk profiles in crop production (Yaqoob et al. 2016), while even herbicide tolerant crops may have done so, though the record here is much less clear (Bonny 2016). In turning specifically to animals, however, this principle of not adding risk has more force. As discussed in Chap. 5, Bernard Rollin’s Principle of Welfare Conservation holds that we should not undertake genetic modifications that worsen the lot of farmed animals, (Rollin 1995). It seems entirely reasonable to insist upon that principle even for gene-edited animals that do not involve transgenes, though how this could or should be enshrined in regulation is well beyond the scope of the present inquiry.

352

13 Gene Editing, Synthetic Biology and the Next Generation …

13.3 The Products of Gene Editing in Food and Agriculture Many of the most crucial ethical questions depend on the specific application or product made with gene editing. The speed and cost-efficiency of gene editing means that it will figure increasingly in all the types of agricultural crops and livestock. Gene editing will be used for future GMOs including herbicide tolerant crops, crops that incorporate resistance to pests (such as Bt) or to plant and animal disease. It also includes plant and animal foods modified to improve their nutritional content (like Golden Rice) or to serve as “bioreactors” that produce pharmaceuticals or other commercially valuable substances in farmers’ fields. In addition, biotechnology has been a tool in the thus-far unsuccessful quest for plants that are more efficient sources of biofuel. Like artemisinin (discussed above), these are the types of product that have been covered at length in the previous chapters. To the extent that gene editing is simply a more efficient tool for developing similar products, the ethical issues are similar. Not all of these applications involve food, but some of the most extensively grown GMOs are in cotton, and (cottonseed oil aside) that does not involve food, either. What they share is the use of the traditional land-based production platform (e.g. farms and ranches). Previous chapters have discussed ethical issues that occur in virtue of gene technology’s potential to transform the use of that platform and the socio-cultural meaning of farming through, for example, restrictions on a farmer’s ability to save seed. Gene edited crops and livestock breeds can be expected to both lessen and increase these problems. The sense in which they will be exacerbated is direct and obvious. Gene editing means more farm commodities coming under the regime of genomics and intellectual property, increasing the ability of corporations to develop a package of farm inputs (e.g. seed, chemicals, managerial product support). Gene edits will extend the product-development strategy of targeted genetic modification to more specialized crops or livestock breeds, increasing the number of farmers who fall under this regime. These packages will outcompete any strategy farmers choosing from the array of commercially available inputs can develop on their own. Farmers may remain economically viable, but they will lose a great deal of their autonomy and economic agency in order to do so. Gene editing’s potential for lessening the extent or seriousness of these ethical issues depends upon whether it will disrupt the oligopolistic power of biotechnology companies. If development costs are reduced significantly, small start-up companies or publicly funded agricultural research institutes may again become able to develop seed varieties or other products that producers can use. At present, any innovative discovery must be licensed to one of the giants in order secure the capital needed to cover development, market and regulatory costs. Thus, there is yet one more element to the regulatory quandary described above: Breeders and molecular biologists in non-profits (e.g. universities, government labs and international research centers) are hoping that a lightened regulatory load will put them “back in the game” with respect to developing products that can be distributed to farmers on terms that are more congenial. Yet there are also products that would have been impracticable (if

13.3 The Products of Gene Editing in Food and Agriculture

353

not impossible) to develop using the tools of agrifood biotechnology’s earlier iterations. Some of these products could provoke more extensive change to the practices, infrastructure and cultural meaning of agriculture than any GMO currently on the market. The balance of this chapter will consider three such product classes. First, gene editing will converge with other technological innovations to create humanity’s first serious competitor to the agricultural, land-based production platform we associate with farming and ranching in rural areas. While many food system activists rail against “industrial agriculture,” there is less-noticed trend toward more conventionally industrial, factory-like production systems that challenge the dominance of farm production. Not all of these challenges need to involve biotechnology. For example, industrialized food factories use engineered systems to manage light, water and temperature, as well as delivery of nutrients, control of pests and removal of wastes. The signature technologies of such “vertical” production systems are not biological, but it is very unlikely that any of them will succeed without adjustment to the genetics of the plants and animals being grown in them. Lab-based meats or alternative animal proteins are one of the most extreme industrial competitors to farming, and one where gene editing will almost certainly join with other forms of advanced biotechnology to test pasture, range and other forms of food animal production. For the second class of applications, gene editing and genomic analysis will be used to develop novel ecosystems for land-based agricultural systems. Unlike applications in the first group, they do not constitute farm-less modes of agriculture, nor do they differ from farming practices that involve extensive manipulation of the environment in which plants and animals live. In a sense, activities such as plowing, weeding or pest control just are modification of a local ecosystem (e.g. the farm) to favor the reproduction of a particular species (e.g. the crop). Alternatively, breeders modify the target organism (e.g. the crop) so that it will be more able to thrive in the managed ecosystem that is the farm. Thus far, GMOs exemplify the latter approach, but farmers have occasionally been able to merge or bridge these alternatives by discovering and exploiting complementarities among cultivated and uncultivated species in the farm ecosystem, (see Picasso et al. 2011). Nevertheless, the idea of intentionally designing an entire cropping system has been a dream. Simultaneous gene editing of multiple species (crops, microorganisms and other species, such as insects) might make this dream into a reality. Finally, gene drives represent a third class of advanced biotechnologies with significant potential to transform existing farm ecosystems. A gene drive is a construct that will move an allele through a population of organisms faster than Mendelian mechanisms of reproduction. Gene drives are already attracting a significant amount of attention from philosophers in connection with their potential applications in conservation and public health. While this literature provides a useful background for evaluating the ethical issues associated with gene drives in agriculture, the focus in this chapter will be to discuss how gene drives might be deployed in the production of traditional agricultural commodities for food and fiber consumption. Given that purely agricultural use of gene drives is unlikely to be the first application of the revolutionary biotechnology, we can expect that many of the current uncertainties that surround proposals for their deployment will have significantly different ethical

354

13 Gene Editing, Synthetic Biology and the Next Generation …

significance by the time that they are used for agriculture. In fact, the key contribution of an agricultural ethics of gene drives may be to use agriculture’s experience with controlling pest and beneficial species as an analogy to frame the gene drive debate as it proceeds in conservation and public health.

13.4 Alternative Proteins The class of meat, milk and egg alternatives is already large, but is expected to grow exponentially between 2020 and 2030. Plant-based foods that simulate the taste, texture and appearance of meat have existed for millennia, if seitan and tofu are included. Highly processed versions began to appear on grocery shelves in the 1980s. In 2016 Patrick Brown, formerly a researcher at Stanford University, introduced the Impossible Burger, which adds heme to an otherwise traditional plant-based meat substitute. Heme is a natural chemical compound containing an iron molecule that aids oxygen transport in blood. It is especially abundant in animal muscle tissue, though it can also be found in nitrogen fixing plants. Brown’s company Impossible Foods claims that heme is what gives meat products their distinctive taste, (Impossible Foods, No date). Brown engineered a microorganism that produces quantities of heme sufficient for industrial production, (Jacobsen 2017). The Impossible Burger represents one end of the spectrum of products that use or will use biotechnology to develop alternatives to meat, milk and eggs. The leading scientist at the other end of continuum is Mark Post. Post debuted a burger made from actual muscle tissue but grown in cell culture in 2013. Post’s product, cultured meat, combines stem cells, tissue engineering and development of a medium that can support cellular growth to produce a product that resembles meat produced “on the hoof”, (Datar and Luining 2015). In contrast to the Impossible Burger, Post’s burger is made entirely from animal tissues that are claimed to be biologically identical to conventional meat products. The phrase “biologically identical” may be subjected to debate, but let us lay that aside for the purposes of this discussion. The California company Perfect Day is also producing products claimed to be biochemically identical to animal proteins by genetically engineering microorganisms to synthetically produce all the constituent components of milk or eggs, then combining these components with water to make a marketable product, (Perfect Day 2019). Although these products do not necessarily depend on the precision associated with gene editing, it is very likely that gene editing will play a significant role in reducing the cost and time needed for developing products or for manufacturing them at commercial scales. Alternative proteins raise ethical questions in moral ontology: Just what are they, anyway? Alternatively, from a more practical standpoint, what should we call them? Can one use the word “meat” to describe them, or is that misleading in some ethically problematic way? The organizations that represent the interests of traditional animal products have already asserted that using the words, ‘meat’, ‘milk’ or ‘eggs’ to describe these products (as well as products such as soy or almond milk) misleads

13.4 Alternative Proteins

355

the public, (Troitino 2017). However, much of the motivation for developing these products is their appeal to consumers who wish to avoid conventional animal products. Thus the marketing approach, including the product label, needs to find some way to both differentiate the alternative animal protein from the product it is intended to replace, at the same time that it conveys that the alternative will be an adequate substitute in the various cooking or dietary applications of the conventional product, (Fuentes and Fuentes 2017). This suggests that apparent ethical issues in the labeling of these products will turn out to be less problematic that a superficial overview might suggest. More intriguing philosophical questions arise in connection with the reception that alternative proteins are receiving among ethicists. As discussed in Chap. 5, the early debate over genetically engineered food animals turned around the question of whether significant change in a species’ nature or telos constitutes a harm. Bernard Rollin argued that modifications leading to reduced welfare in the form of pain, suffering or frustration of genetically based drives could not be justified except by extremely compelling offsetting considerations. While adequately compelling considerations might attend some forms of bio-medical research, they would not be found in the food system, where more humane alternatives are available and animal products are already in chronic oversupply, (Rollin 1996). However, Rollin argued that there are no ethical reasons against modifying an animal genome to produce an organism that lacks a given genetically based drive or capacity altogether. He admitted that such experiments would inflame public opinion, making such a strategy commercially unattractive, but he denied that it raised any concerns from the animal’s perspective, (Rollin 1998). Rollin’s prediction proved correct. A large academic literature of outrage over genetically engineered animals has accumulated over the last twenty years. Many authors take particular exception to modifications that would redress problems in animal production by creating organisms that are simply incapable of enduring the pain, suffering or dysfunction that pigs, chickens, cows and other livestock species endure in industrial production systems, (see Chap. 5). My own entrée into this literature articulates agreement with Rollin up to a point. I argued that much of the visceral opposition to the use of genetic transformation in livestock was a reflection of a faulty intuition. People were imagining that an individual animal was somehow being deprived of a given capability, but in Rollin’s proposal, the individuals in question would have never had this capability in the first place. I argued that we could imagine two scenarios for blunting of pain and suffering through diminishment (or as I put it, disenhancement) of the phenotype’s capability for experiencing welfare of any kind. In the “dumb down” strategy, one starts with the genome of a cow, a pig or a chicken and then tries to strip away the potentially troublesome capabilities. In the “build up” strategy, one starts with a non-sentient life form and then tries to add the characteristics that made the organism valuable as a food source, (Thompson 2008). In the fullness of time, it seems clear that “build up” is exactly what the alternative protein sector is trying to do. Subsequent entries on the ethics of animal biotechnology include a few authors who either accepted or duplicated my analysis and tried to build upon it, (Shriver

356

13 Gene Editing, Synthetic Biology and the Next Generation …

2009; Palmer 2011; Schulz-Bergin 2017), but most entries reject it by claiming that there is no quandary regarding applications of gene technology that mitigate animal welfare problems in the food sector. It is, on the contrary, unequivocally wrong. The literature opposing diminished livestock can be summarized in terms of three strands. Writing separately, Arrianna Ferrari and Adam Henschke both argue that the thought experiment ignores the concrete context of contemporary animal production. It suffers from the stink of the lamp. The wrongs currently being visited on animals in industrial production settings orient our moral judgment. The history of abuse, injustice and disregard for animal interests is what created the problem; reform proposals must acknowledge this history and respond to it in its particulars. If they do not, they perpetuate the injustice, function as a form of dissembling and distract attention from reforms that actually need to take place, (Ferrari 2012; Henschke 2012). Katie MacDonald makes a similar argument against a less ambitious genetically engineered animal, the Enviropig, modified to reduce phosphorous pollution in hog manure. The project isolates a single dimension of what is problematic about pig production, enabling a technological response that ignores the way that phosphorous pollution is embedded with the larger systemic integration of the pork industry, (MacDonald 2018). Ferrari, Henschke and MacDonald have put their fingers on an enduring problem in the philosophy of technology. It was one that I associate with consequentialist ethics. To wit: When critics complain about unwanted outcomes, they invite technologists invent things that avoid that outcome, but otherwise keep doing what we are doing. This attitude prevents people working in agricultural science from asking more comprehensive and more reflective questions about our food system, (Thompson 2017). I also agree with Ferrari and Henschke’s specific point: To view genetic diminishment as the unilateral solution to animal welfare issues in livestock production is morally problematic. I understand the umbrage of someone who thinks that we should not be distracted from the suffering animals in industrial production facilities, and I am not trying to divert attention from that problem. Beyond that, I am not sure I understand how Ferrari and Henschke think they have objected to my analysis. What is more, the subsequent development of the build-up approach in the guise of alternative proteins suggests that it is not just an academic exercise. The lamp stinks less as companies like Impossible Burger and Perfect move into production. The enthusiasm for using gene technologies (along with other industrial technologies) to create organisms that lack consciousness is gaining adherents. A second, heterogeneous group of arguments attempts to defend the intuition supporting the distinction between dumb-down and build-up, though in point of fact almost all of the effort is dedicated to reproducing the moral horror associated with dumb-down. Traci Warkentin’s “Dis/integrating Animals,” preceded my paper, but her argument exemplifies how the intuition that troubles disenhancement can be amplified by extending the thought experiment that motivates it into a full-blown dystopian vision of the future, (Warkentin 2006). Arianna Ferrari also conjures the notion of techno-dystopia in an essay that cites mine as an instance of a colonizing mindset that seeks total domination, (Ferrari 2015). Zipporah Weinberg characterizes the prospect of intentional diminishment of livestock as “an ontological collapse,”

13.4 Alternative Proteins

357

(Weinberg 2015). Like Warkentin, Anne Franciska Pusch draws upon Margaret Atwood’s dystopian science fiction where genetically engineered animals such as the disenhanced Chickie-Nobs figure in an apocalyptic scenario. Pusch does not claim that apocalypse is a realistic possibility. Her point is that this way of provoking the moral imagination can stimulate empathy and authentic moral growth, (Pusch 2015). If these arguments are successful, the emotional outrage provoked by dumb down (but not build up) might be sufficient to distinguish the two strategies from a moral perspective. Finally, other authors argue that the intuitively repulsive aspect of disenhancement (and of dumb down in particular) testifies to the predominance of virtue-based moral considerations over consequentialist or rights-based ethical theories in the moral imagination of non-specialists. The argument contests the dominance of Peter Singer and Tom Regan in the domain of animal ethics (see Chap. 5). The problem with gene technologies is that they do violence to the relationship between humans and other animals, undercutting the interactions that give rise to caring husbandry. This strand of analysis supports the overridingness of the virtues position through both philosophical argumentation and engagements with non-philosophers, (Coles et al. 2015; Murphy and Kabasenche 2018; Bos et al. 2018). Bernice Bovenkerk and Hanneke Nijland make the case by examining how genetic change is a problem for the breeding of companion animals. Exotic breeds objectify the animal, vitiating the possibility of understanding the human-animal relationship in an ethically justifiable manner, (Bovenkerk and Nijland 2017). These arguments differ from those in the second group in that they do attempt to provoke outrage, and to my reading, count equally against dumb down and build up forms of gene technology. I agree that relational considerations matter, and that animal geneticists should be mindful of them, (Thompson 2010b). However, I am less sure that these arguments overrule the emphasis on animal suffering in actual production environments. In the present context, the question is how do these arguments opposing diminishment of animals play out in the context of gene technologies for meat or milk alternatives? The question must be considered in the context of a philosophical literature that has expressed overwhelming support for this technology, (Hopkins and Dacey 2008; Van der Weele and Driessen 2013; Schaefer and Savulescu 2014). Yet when this literature is viewed in light of the three argument strands opposing animal diminishment, only a defense of untutored intuition offers unambiguous support for cultured meats and other gene-based animal proteins. Those who, like Henschke and Ferrari, stress the need to address concrete animal abuse in a material way might argue that alternative proteins do so by undercutting the market for industrially produced animal products. Indeed, this claim is made by some authors who advocate cultured meat on ethical grounds, (Pluhar 2010). Yet it is also clear that the technology still faces challenges in economic feasibility and market acceptance. It would thus appear that if dumb-down is a morally problematic distraction from the real problems of animal welfare in industrial production systems, as Henschke and Ferrari claim, so is cultured meat. However it is the third, relational/virtues strand that evokes the greatest tension. It would certainly appear that any of the products reviewed above would rather

358

13 Gene Editing, Synthetic Biology and the Next Generation …

dramatically transform humanity’s relationship to livestock species. What is more, advocating such approaches (along with veganism itself) as a means to eliminate animal production would also eliminate the animals themselves. Such observations have led some pro-animal theorists to be less than enthusiastic about cultured meat, (see McKenna 2018). Jacob Metcalf even argues that the exterminism implied by cultured meat turns the dystopian intuitions on their head, enrolling the supporters of alternative animal proteins in a form of world-denying technological future, (Metcalf 2013). Others have questioned whether it is so obvious that these cellular products are so obviously devoid of moral considerability, (Carruth 2013). If we should question the character of someone who would use technology to modify an animal’s phenotype in order to solve moral problems with industrial animal production, why wouldn’t we also question someone who proposes to build a new phenotype in order to do the same thing? Both strategies seem to instrumentalize the human/animal relationship to an extraordinary degree. In summation, I do not see that the quandary noted in my 2008 paper has been resolved. In fact, the tide of support for alternative proteins only exacerbates it. It is, as a matter of practicality, unlikely that extreme versions of dumb-down will be pursued, though gene edits of livestock species have already been used to alter the phenotype in less radical ways, (Carlson et al. 2016). The question is, are there convincing ethical reasons (as opposed to psychological or sociological observations) that justify the use of gene edits in support of alternative proteins, but oppose it in the case of hornless cows? In conclusion, it is unclear to me how (or whether) this philosophical conundrum should be taken on by the bioengineers themselves in any substantive manner. They should expect their work to engender a high degree of confusion and condemnation, and if the past is a guide to the future, that will continue to damage the economic prospects for animal biotechnology. This is not, however, a reason to condemn it on ethical grounds.

13.5 Novel Agricultural Ecosystems The terminology of “novel ecosystems” is now used by ecologists to describe geographically constrained compositions of organisms and abiotic elements that have emerged after human perturbation. For example, new relations among predator and prey species emerge following the human introduction of an invasive species or human induced change in abiotic elements after active mining or industrial activities cease, (Hobbs et al. 2013). Ecologists have taken an interest in gene editing because it is seen as a way to modify species and thus restructure ecosystems in response to the conservation challenges of climate change, (Palmer 2016). Although this usage of the expression “novel ecosystems,” specifies species interactions that do not involve continuous human supervision, there is a related sense in which farmers have always developed novel ecosystems, both unintentionally and by design. Indeed, the introduction of abiotic elements such as pesticides and synthetic fertilizers aims to facilitate agricultural yields by modifying the abiotic elements of a farm environment.

13.5 Novel Agricultural Ecosystems

359

In fact, farmers and agricultural scientists have long recognized that plants in the field respond to change in both their abiotic environment (e.g. water, solar energy) and their biotic environment (e.g. virus, fungi, insects and other damaging or diseasecausing organisms). Whether in plants or animals, the immune system mobilizes adaptation within the genetic configuration of cells within an individual organism. Human beings experience this response in the gradual recovery from a cold or flu virus. Plants do something similar, but too frequently, the response takes too long to become effective from the perspective of a farmer who is trying to bring a crop to market in a timely manner. In addition, the virus can damage the fruit or grain of a plant, detracting from its food value or appearance, (see Jones and Dangl 2006 for a scientifically rich discussion). As scientists came to appreciate the genetic basis for a plant’s adaptive response to other organisms, they began to speculate on the possibility of harnessing the capacity for their own purposes. In one sense, this dream led to first generation GMOs. One of the first GMOs was a papaya engineered to resist papaya ring spot virus. The GMO was modified to produce antibodies that would combat the virus well before damage to the fruit’s appearance occurred, (Gonzalves et al. 2010). As with all early generation GMOs, the modification in the resistant papaya was to the germline. A more ambitious version of the dream is to trigger a response while plants are growing in the field, a strategy that would mimic the natural processes of immune response. Alternatively, more ambitious still, to mimic the adaptive response of the immune system in controlling other elements of the plant metabolism. For example, under conditions of drought, a plant might respond by channeling available moisture away from the process of producing the edible parts of value to the farmer. If the genes controlling this response could be manipulated, a farmer might be able to reduce the loss of food value. But farmers would not want this genetic response to be active unless the crop is, in fact, under drought stress. This suggests that there would need to be some means of affecting genes while the crop is growing in the field, and, again mimicking nature, one way to do this would be to incorporate the instructions for such a genetic response into a construct such as a virus. The modified virus, in turn, would need to be delivered to the field. In nature, plant viruses are spread by vector organisms such as insects. Perhaps insects could be coaxed into cooperating with farmers? This dream of a systematic modification not simply of the plant, but of its farm ecosystem involves such an extensive set of interlocking alterations in the genetics of plants, viral DNA and the vectors themselves that it remains a largely speculative idea. However, in the era of gene editing, there are scientists who are interested in proof of concept research to determine whether it could work. Gene editing might enable such projects both through the way it affords speed and precision for genetic modification, and also because the CRISPrCas9 construct can itself be incorporated into a plant or vector genome in ways that facilitate “real-time” (i.e. within a single generation) genetic change. Perhaps the most prominent and best-funded examples of this work are being pursued under the Advanced Plant Systems (APS) initiative at the United States Department of Defense (DOD) Defense Advanced Research Projects Agency (DARPA). The APS includes a number of initiatives, include some that would use

360

13 Gene Editing, Synthetic Biology and the Next Generation …

gene editing to create plants that would function as biosensors for the presence of specific chemical or biological agents. The DARPA program that attempts to mimic elements of immunity to enable modification of a plant’s genome within a single growing season (e.g. within a single generation) is called “Insect Allies.” The public-facing website for this program states: ...Insect Allies performer teams are leveraging a natural and efficient two-step delivery system to transfer modified genes to plants: insect vectors and the plant viruses they transmit. The program’s three technical areas—viral manipulation, insect vector optimization, and selective gene therapy in mature plants—layer together to support the goal of rapidly modifying plant traits without the need for extensive infrastructure. (Bextine, n.d.)

The website also indicates that the program is envisioned as enabling a defensive capability for food production that would provide some measure of protection from aggression by enemies of the United States. The defensive orientation expressed in this description is consistent with U.S. participation in the Biological Weapons Convention. Signatories pledge never to acquire “biological agents or toxins that have no justification for peaceful purposes,” (U.S. Department of State n.d.). What are the ethical issues here? Most obviously, the attempt to intervene in the interaction of insects, viruses and plants poses conceptual and analytic challenges to biotechnological risk assessment. In response to these concerns, DARPA’s current version of Insect Allies has made a number of special provisions. First, it is a proof-of-concept program only. Researchers are investigating the potential for such integrated and coordinated projects of gene technology. As indicated in program manager Blake Bextine’s statement on the program, the phrase implies pursuit of a significant breakthrough in the capabilities of gene technology. As Catherine Kendig has noted, “proof-of-concept” research is often associated with a vaguely defined notion of innovative advance (Kendig 2016). However, in the case of Insect Allies, the emphasis may be on the word only: the project will not involve testing of the engineered systems. Bextine also notes that, in this case, proof-of-concept also implies an improved understanding of the normative, risk-based criteria for continuing the project into phases that further develop these systems, (Bextine 2018). Second, proofof-concept also means that the current program will not involve field-testing or environmental release of any modified virus, plant or insect. Although the current project will produce genetically modified plants, vectors and viral sequences, the research will be done under strict biological containment, reducing if not eliminating exposure to natural ecosystems. Finally, as Bextine’s statement suggests, each research group operating under the aegis of Insect Allies is contractually obligated to investigate and include mechanisms for recall or termination of the agents (vectors, viruses or plants) they are developing. In each of these respects, the DARPA projects reflect a strong ethical commitment to limiting environmental risk, at least at the current stage of development, (Bextine 2018). Thus, for Insect Allies, proof-of-concept exemplifies a range of the meanings that Kendig discusses in her article on proof-of-concept research in synthetic biology. (Kendig 2016). The project has accepted at least some of the ethical responsibilities implied by Hans Jonas’ prinzip verantwortung. Perhaps even more than alternative protein engineering, novel ecosystems portend such dramatic transformation of agricultural landscapes and food system organization

13.5 Novel Agricultural Ecosystems

361

that is difficult to imagine what the future might look like. One can envision a future in which the ability to produce in a given environment presupposes the active intervention of whatever agency—be it the U.S. DOD or some unknown future security organization. This would presumably offend farmers’ sense of autonomy and independence from government oversight or intervention, yet there are so many technological and political developments pressing in this direction that it is ineffectual as an objection to novel agricultural ecosystems, alone. More generally, anyone with a strong conservative bias against disruption or transformation of traditional food systems will regard systematic intervention in the genetic ecology of the food system with the same prejudice that militates against drones, robots, artificial intelligence, alternative proteins or, indeed, first generation GMOs. In this respect, however, we are just reminding ourselves that novel agricultural ecosystems are not immune from the full range of issues covered in the first 12 chapters of this book. The (perhaps) unique issue that has spawned reaction is to bring the threat of “dual use” into the domain of agricultural innovation. The European Commission defines dual use items as “... goods, software and technology that can be used for both civilian and military applications,” (EU 2018). In October of 2018, a group of European molecular biologists published an essay characterizing Insect Allies as a technology based on horizontal environmental genetic alteration agents (HEGAAs). They are, as noted, projects that bring about non-heritable (e.g. non-germline) genetic change through vectors dispersed through the environment. The article pointed out unresolved issues in the risk assessment and regulation of HEGGAs, but the primary thrust was to argue that the program would inevitably be perceived as a weapons program, and as such, a violation of the Biological Weapons Convention, (Reeves et al. 2018). Consistent with the author’s speculation, the publication of this article in the Policy Forum of Science sparked a flurry of stories in mainstream media (not to mention social media) stating that the U.S. DOD was attempting to develop weaponized insects. The ensuing reaction to this article can be charted along a continuum that begins with the relatively mild concern that HEGGAs or similarly systemic gene technologies breach the distinction between dual-use and traditional agricultural technologies. At the opposite end are those who see DARPA funding as proof of nefarious intent. While the authors of the original piece in Science specifically dissociate their concern with the funding source, they do emphasize how the involvement of DOD heightens the need for public engagement and a robust accompanying effort to highlight the ethical issues in such a project. Thus, the ethical issues associated with the Insect Allies initiative include: (1) the potential for military use; (2) whether this particular project puts the United States in violation of the Biological Weapons Convention; (3) whether or in what sense DARPA sponsorship creates additional burdens for public engagement and ethical review. I will close this discussion of novel agricultural ecosystems by reviewing each question in turn. As the above quotation from the European Commission’s Directorate General for Trade suggests, recent dual use discussions have arisen in connection with “goods, software and technology,” a classification that would not necessarily bring seeds, farming equipment and other agricultural inputs to mind, even if it does not logically

362

13 Gene Editing, Synthetic Biology and the Next Generation …

exclude them. In fact, agricultural scientists have long been aware of the potential for using agricultural pathogens for military or terroristic purposes. Anthrax is a disease of sheep and cattle. It was used as a weapon in World War I, and researchers in both the United States and the United Kingdom developed programs to increase its military effectives from World War II into the 1950s. The Biological Weapons Convention was signed in 1972 to limit, if not eliminate, the development of agricultural pathogens such as anthrax for aggressive purposes, (CDC 2016). In short, there is nothing new about the potential dual use of agricultural goods. However, the Convention on Biological Weapons points to widespread agreement on the unethical nature of deploying such weapons. The Convention extends its ethical disapproval (and prohibition) to development of dual use biotechnologies with the intent to use them in an aggressive fashion. Thus, for the case at hand, the question devolves to whether DARPA statements of peaceful, purely defensive intent can be believed. This brings us directly to the second question. Here we should remind ourselves of the limited sense in which an organization (such as the Government of the United States) can be said to have intent. One can point to documents that express goals explicitly (and several have been cited), but one can also point to conversations, meetings and even memoranda, many of which will not be public, where government officials commit the organizational, financial and coercive resources of a government to specific ends. The Convention on Biological Weapons is a statement of explicit intent, but can it be believed? Could there also be secret, undisclosed efforts to continue the development of biological weapons, even as the face presented to the public disavows them? Only a thoroughgoing Pollyanna would accept public statements of any government (perhaps especially the U.S. Government) entirely at face value. Indeed, even from a purely ethical standpoint, citizens must hope that their governments are attentive to the potentials for developing biological weapons, even as they pledge to neither develop nor deploy them. The duty to protect its citizens would require governments to conduct or support such research as would be needed to understand and potentially limit such potentials. However, there are at least two ways in which this could be done. One would be to conduct a wholly secret research program, such as the one undertaken on anthrax in World War II. We should not be so naïve as to think that no such programs exist. Yet DARPA is not such a program. DARPA’s projects are at least somewhat open to public scrutiny (though whether they are adequately so broaches the third question). They thus reflect the second option, which is to conduct research not only in view of citizens to whom the duty of defense is owed, but also open to the inspection of potential enemies. I argue that this second option is preferable on ethical grounds, though it brings along its own further complications. The most obvious point in favor of the public or semi-public option is that it opens the door to some degree of public oversight. The German group would not have been able to voice cautions about Insect Allies if it were a secret activity, nor would the slew of media outlets that took up the discussion have been able to confirm that such research was, indeed, ongoing simply by looking at the DARPA website. Second, public disclosure enables the potential for discussion and

13.5 Novel Agricultural Ecosystems

363

political debate over the direction and general capabilities of key technological developments. Again, the publicity attending Insect Allies as a DARPA project is what enables the discussion of HEGAAs or novel agricultural ecosystems as an important case study in the applications of gene editing tools. A more secretive program would at least render such discussion speculative. There is also the philosophically weak but practically significant fact that defense budgets tend to be more popular than budgets for general science or agriculture. In the United States, at least, one is far more likely to get projects with large coordinated teams funded through defense funding than as agricultural research. Finally, there is probably some deterrent effect (possibly offset by a publicity effect) that comes from the fact that potential enemies know that a government (in this case, the U.S. Government) is broadly aware of some technical potentials, and is undertaking the research and development needed to combat and neutralize their aggressive use. All of these reasons become stronger when the publicly known program is conducted in conjunction with a government’s announced commitment to purely peaceful or defensive uses, such as the Biological Weapons Convention. The ensuing complications of a DARPA-like defense program include important restrictions on the extent to which it can be public without compromising its effectiveness. Quite obviously, one would not want all of the details of a dual use capability available on the Internet, since they would as available to terrorists or hostile governments as they would be to citizens. While democratic values weigh in favor of publicity and public discussion, one would also hope that this discussion could be conducted in a manner that did not also disclose sensitive information. I have no immediate guidelines to suggest as to how one negotiates this tension, but I have signaled the importance of this complicating factor by characterizing a DARPA-like investigation into the capability for weaponizing a biological system as “semi-public.” The Science policy article suggests that a more thorough ethical analysis should accompany the project (e.g. the third question) but such an analysis would also be subject to these complicating factors. Not only would there be limits on how much of the technical detail could be described in an ethical analysis, even strictly ethical issues would need to be discussed at a high level of generality so that competent readers from hostile groups cannot discover crucial technical details by inference. In concluding, I support the German group’s support of research that would accompany the development of gene technologies in agriculture. While dual use presents one set of ethical concerns, potential for disruption of the existing food system provides equally compelling reasons to begin detailed ethical investigations into the social and ethical implications of agrifood biotechnology, including and especially innovations that will restructure agrifood ecosystems. Virtually every industrialized nation, with the probable exception of formerly socialist countries such as Russia or China, is ahead of the United States in enabling such research. As the authors of the Science commentary assert, this is not helping the U.S. science community gain the confidence of non-specialist observers either at home or abroad. Nevertheless, that observation returns us to theme of Chap. 12, and we must depressingly conclude that

364

13 Gene Editing, Synthetic Biology and the Next Generation …

the culture and funding agencies of the United States have yet to take these lessons on board.

13.6 Gene Drives A gene drive is a genetic construct that moves through a population of reproducing organisms more quickly (e.g. in fewer generations) than would be possible through Mendelian inheritance and natural selection. Mendelian inheritance is the mechanism of heredity where sexually reproducing parents pass genetic traits to their offspring. Whether a given characteristic of a parent’s phenotype (such as eye color or baldness) is exhibited in the progeny depends on more complex factors, including genetic dominance. In addition, a population will have more genetic constructs from individuals who have many descendants than from individuals who have few. Oversimplifying a bit, one can envision how traits will move through a population based on the interbreeding of sexual partners. Natural selection affects the trajectory of heredity when some individuals have a greater chance of passing down their genes to later generations due to their fitness, their ability to survive and thrive in their given ecological niche. But there are genetic characteristics that can alter the trajectory of descent in ways that deviate from the twin drivers of Mendelian inheritance. Gene drives exist in nature and do not depend on gene editing. The potential for engineering a gene drive opened the theoretical possibility of introducing a genetic change into a population of naturally breeding organisms. The most widely discussed applications of gene drives are for controlling insect vectors for diseases such as malaria, dengue and zika. In this context, however, the debate originates with proposals to modify mosquitoes so that they cannot transmit the disease. If the modified strain were to displace unmodified mosquitoes in a given environment, transmission to humans (as well as other vertebrates including livestock) would be controlled, (Collins and James 1996). As with all applications for genetic modification, gene editing would be a useful tool in creating gene drives. However, gene editing has deeper significance for the prospect of gene drives. Given that much effort had been expended in showing why constructs introduced into GMOs would not move through populations in an uncontrolled manner, the idea of a gene technology that was intended to spread without continuous human involvement was regarded with suspicion (Macer 2005). Gene drives intended for public health interventions (such as controlling mosquitoes that vector malaria or dengue) needed to be promulgated with safety provisions, (James 2005). The CRISPR construct has enabled the creation of gene drives (in theory) with significant power and with features such as self-limitation, reversibility and recall. In this context, self-limitation means that the movement of a construct through a breeding population will stop short of dominating the entire population. Reversibility means that the genomes within a population can be returned to configurations characteristic of genomes prior to a genetic intervention. This also implies recall, the ability to remove modified organisms from a population.

13.6 Gene Drives

365

As will become apparent, these features have made the prospect of using gene drives in a number of applications much more attractive. The public health application of genetically modified mosquitoes hit a roadblock around 2010, as it became apparent that field-testing of the approach involved both environmental and ethical risks, (James et al. 2011). As a public health technology, the key ethical risks were associated with risk to humans living in the area of the field trial, and the difficulty of obtaining informed consent, (Lavery et al. 2008). It should be noted that these issues would arise in connection with any modification of the insect vector and without regard to whether the dispersal strategy involves a gene drive. One might, for example, achieve penetration by the modified organism adequate for the purpose of field testing simply by repressing the unmodified population with a chemical pesticide and releasing mosquitoes modified to disrupt disease transition, but lacking the additional gene drive. The balance of the section will emphasize how gene drive constructs should be viewed in ethical perspective, but it is important to recall that specific applications for gene drives will have their own ethical implications irrespective of whether they are coupled with a gene drive. There is already a significant literature on the ethics of gene drives. An early paper on the ethics of gene editing tools characterizes gene drives as “potentially more dangerous and controversial...” than other applications, (Caplan et al. 2015). Other entries stress specific applications of gene drives, especially the public health strategies already mentioned (El Zahabi-Bekdash and Lavery 2010; Resnik 2018). Other non-agricultural applications include the eradication of pest or invasive species for conservation purposes (Pugh 2016), and modifying wild organisms to aid their adaptation to climate change, (Palmer 2016; Preston 2019; Sandler 2019). More generally, Zahra Meghani and Christopher Boëte have argued that any ethically justifiable test of a gene drive will require a localized decision procedure which puts exposed populations firmly in control of the situation. They also stress the distribution of benefits and costs, suggesting that decisions should be made in light of more traditional methods for vector control, (Meghani and Boëte 2018). Although considerable attention is being given to the use of gene drives in public health and conservation, rather little has been focused on agriculture. However, if gene drives can be used to depress the population of a disease-transmitting insect (e.g. a mosquito) there is no principled reason why they could not also be used to repress the population of a disease-transmitting or crop-damaging insect in a farmer’s field. In fact, there are a number of agricultural pests that appear to be good targets for deployment of a gene drive. These include the spotted-wing drosophila, which is an invasive species that lays eggs in soft fruits (e.g. strawberries, blueberries, etc), rendering them inedible and unsalable. Gene drives have been proposed for control of cabbage moths and white fly (a vector for a number of plant diseases), (Scott et al. 2018) Cochliomyia hominivorax (e.g screwworm) is a particularly apt example. The flesh-eating larval stage of this fly attacks cattle, sheep and other pastured livestock, causing considerable pain and distress to the infected animal and significant financial loss for the producer. Screwworm has been controlled in the United States by a massive (and expensive) program of releasing enough sterile males to initiate a crash in the population, followed by a rigorous inspection program designed to forestall

366

13 Gene Editing, Synthetic Biology and the Next Generation …

reintroduction, (Kouba 2004). Such efforts have proven impracticable throughout Latin America owing both their cost and to the pervasiveness of screwworm in tropical climes. A gene drive could function much like the sterile insect technique, while eliminating the need for large scale radioactive sterilization of males and release in infected areas, (Scott et al. 2018). Virginie Courtier-Orgogozo, Baptiste Morizot, and Christophe Boëte note the potential for using gene drives in crop protection, and identify three classes of what they call “uncertain risks.” First, the drive can move beyond its target population. For example, a gene drive intended to eliminate invasive fruit flies in an agricultural region could spread to regions where spotted-wing drosophilae are native. Second, a different gene construct could accidentally be incorporated into the gene drive, with highly unpredictable effects. Finally, given the novelty of engineered gene drives, one cannot exclude the possibility of unanticipated ecological impacts, (CourtierOrgogozo et al. 2017). Without meaning to minimize the significance of these risks, it is worth noting that they are actually quite similar to risks that have been studied in connection with first generation agrifood GMOs. This points the ethical analysis of agricultural gene drives in two conceptually opposing (though non-contradictory) directions. On the one hand, narrowing the focus to agricultural applications of gene drives, the decades of experience with GMOs and crop protection tools provides a knowledge basis for approaching the ethics of gene drive risk assessment. On the other hand, there is little question that gene drives incite the imagination and reignite the fear of technology run amok. Observing how little the thrust of risk-based critique has changed over the course of nearly forty years, we can add the entire catalog of ethical issues discussed throughout the preceding chapters of this book to Courtier-Orgogozo, Morizot, and Boëte’s concerns. Gene drives are highly unlikely to be constructs drawn from the GRAS list (see Chap. 3), so there will be food safety concerns that require significant testing, especially if residues or insect parts are likely to contaminate the food chain. Insects are animals and as such, modified insects will fall under aspects of the regulatory regime intended to address welfare, (Bruce et al. 2013). Finally, the entire gamut of socio-economic risks would appear to be of obvious significance, CourtierOrgogozo, Morizot, and Boëte’s failure to note them notwithstanding. The question is whether existing or future methods of risk assessment pay adequate attention to the full range of risk-based ethical considerations. Although Courtier-Orgogozo, Morizot, and Boëte characterize these as “uncertain risks,” it is highly doubtful that an agricultural gene drive will be among the first application of an engineered gene drive to be tested in the field or deployed in actual use. We will have learned a great deal (one way or the other) from public health or conservation applications long before that happens. Furthermore, existing forms of pest control provide analogs for imagining the risks of agricultural gene drives. This is a complex topic, and only a few examples can be discussed in the present context. Chemical pesticides are a signature technology for industrial agriculture, but it is widely acknowledged that ethical issues are raised by the health hazards they pose to agricultural laborers and their unwanted

13.6 Gene Drives

367

ecological consequences, (Pimentel 1995). Such issues form the backdrop for arguments favoring some alternative approach to controlling insect pests, including gene drives. However, there is a less noticed issue with pesticides. In less industrialized agricultural systems, the introduction of pesticides displaces and deskills agricultural labor, and shifts a portion of the farming’s productivity from labor to capital (Vandeman 1995). The fact that farmers nonetheless choose to use chemical pesticides may reduce the ethical force of this observation, but it is worth noting that there is no reason to presume that gene drives would not also lead to similar socioeconomic consequences. Biocontrol agents and the sterile insect technique (SIT) represent two alternatives to chemical control. Biocontrol deploys a mix of intercropping, the use of flooding or other physical systems and release of a pest’s natural enemies. Farming systems that draw upon a number of these strategies are, in fact, integrated into the traditional methods of less-industrialized producers, who still make up the numerical majority of the world’s farmers, (Naylor and Ehrlich 1997). However, attempts to introduce new forms of biocontrol have occasionally gone awry. The cane toad is the poster child of a biocontrol strategy gone wrong. Introduced into Australia in the 1930s to control grubs and beetles in the cane fields, it became an invasive species with such a spectacular array of harmful consequences that they are still being investigated. Along with less well-known introductions, the cane toad fiasco has led to a significant debate over the wisdom of introducing new species into an agricultural ecology in the attempt to control pests, (Thomas and Willis 1998). This debate can provide a starting point for situating the ecological risks that are associated with gene drives. The sterile insect technique (SIT) was used to control screwworm in the United States, as mentioned above. It works by releasing large numbers of sterile males in a region, who in turn mate with normal females, producing no offspring. The effectiveness of SIT depends heavily on the biology of the species in question, but can significantly (but temporarily, in most cases) lead to a crash in the population. In addition to biological limitations, SIT is costly and has not been a preferred method for perennial efforts at pest control, (Scott et al. 2018). Abstracting from these limitations, the self-limiting nature of SIT is viewed by entomologists as a mark in its favor. The males that are released eventually die, and they introduce no progeny into the ecosystem. In comparison to chemical control or biocontrol, SIT poses less risk to human health and to the environment. Thus, to the extent that a gene drive can mimic these features of SIT, the risks of a gene drive can begin to be put in perspective. However, entomologists working in the control of tropical diseases are aware that SIT is not free of social risk. The SIT strategy used to control screwworm in the United States is one of several strategies that entomologists have investigated. Cytoplasmic incompatibility was a hypothesis holding that female sterility could be produced by breeding with a genetically incompatible male. Like the screwworm application, it requires breeding and release of a large number of male mosquitos. An attempt to test this SIT hypothesis in India has entered the lore of the tropical disease community. Robert S. Desowitiz (1926–2008) was a specialist in tropical medicine who recounts the incident in his 1991 book, The Mosquito Capers.

368

13 Gene Editing, Synthetic Biology and the Next Generation …

On a morning in 1975, a van bearing the blue-and-white logo of the World Health Organization on the door—a snake caduceus through a global map—drives into the village center. The villagers, who have a fear and loathing of snakes, regard the serpent van suspiciously. They begin to be even more suspicious when a peculiar collection of men emerges from the van—a few undoubted Indians, some strange Orientals, and some very white white men. An angry murmur of astonishment passes through the gathered group of villagers when these men remove large mesh-covered cages from the vehicle, open the cages—and out flies a cloud of mosquitoes, (Desowitz 1991, p. 89).

As the episode developed, suspicion spread to the national press ant to the highest levels of the government. The villager’s justifiable concern was amplified by speculation that the episode was a secret test conducted by the United States’ military. Reports circulated in the press, and some foreign scientists were forced to leave India. A lengthy debate on the intent, procedure and justifiability of the test was carried out in the pages of Nature. The episode continues to be controversial and is cited in opposition to testing gene drives in India, (IANS 2017). The 1975 incident points away from any focused discussion of agriculture and what agricultural scientists have learned about the risks of crop protection technologies. It calls to mind the sense in which gene technologies of any kind spark the fear of technology run amok. There is little question that gene drives do this. They do so because unlike earlier GMOs, they are actually intended to spread throughout the natural environment after their release by scientists. Consistent with Meghani and Boëte’s analysis of public health applications, the ethical response is to engage these fears through well-considered forums where all citizens have an opportunity to participate. As argued in Chap. 12, such fears are not irrational, and they must be engaged in a respectful manner. This is not a problem that is unique to gene drives, though a gene drive is likely to both pique the interest of journalists and inflame public outrage.

13.7 Conclusion: Should We Worry? The first edition of this book (written almost a quarter century ago) was envisioned more as roadmap for locating ethical issues in agrifood biotechnology than as an analysis or prescriptive treatment of them. That orientation has continued throughout two revisions, even as the target audience has shifted from scientists actively engaged in the development or application of gene technologies to people in any discipline (including agricultural sciences) who have recognized the need and value of ethical reflection. For this audience, a roadmap should highlight areas where more detailed or extended deliberations need to occur, and should offer some suggestions for bridging gaps in knowledge that have been produced by high level of specialization that is characteristic of 21st century science. To that end, this chapter has outlined where gene and why tools for gene editing promise to create a new generation of food and agricultural gene technologies.

13.7 Conclusion: Should We Worry?

369

Gene editing will increase both the precision and speed with which genetic alterations can be made in the plants and animals that are produced agriculturally and utilized for food and fiber. It will not substantially change the ethical questions that are raised by the products of gene technology. As such, the entire framework of this book remains relevant to food and agricultural biotechnologies in which gene editing is deployed. Speed and precision may make a difference to certain regulatory structures, but the end product will determine which ethical issues are relevant, and which are not. To cite only the most obvious example, concerns about the welfare of genetically altered organisms matter when the organism is sentient, and plants are not, to the best of our knowledge, sentient. Plants may evoke feelings of attachment and they may be even be regarded as kin in some belief systems, but extending the type of concern discussed in Chap. 5 to modified plants misunderstands the nature of ethical responsibilities that humans have to other sentient (primarily vertebrate) animals. The greater precision of gene editing tools to modify livestock may pose challenges to the way that animal biotechnologies are regulated, but it does not obviate the ethical significance of what it feels like to be an animal so modified. Gene edited crops and animals will, thus, raise ethical issues with respect to human health. These will include not only the potential toxicity of gene-edited foods, but also the question of who actually gets to decide what facts are relevant to the food that any individual chooses to put into their mouth. Gene edited animals will raise questions about whether the technology has made the animal itself worse off, either in terms of the animal’s veterinary health, but also in terms of the animal’s experience of pain, frustration or inability to cope with its environment. Gene technologies will continue raise not only the narrow environmental issues of unintended impact on other species and ecosystem function, but also the broader, almost inchoate, questions about how agriculture figures in the human species’ relationship to the natural world. Gene editing (especially alternative proteins, novel ecosystems and gene drives) may provoke more awareness of these issues, but it will not alter their character. Finally, we have devoted two full chapters to the myriad social impacts ensuing from agricultural technology. They range from farmer autonomy and control over the food system to hunger, agricultural development and the ability of the public sector to deliver food system technologies. Products of gene editing will enter this space of concern and contestation, just like GMOs. If we ask whether we should worry about products of gene editing, including gene drives, the answer is, “Yes.” We should worry because all agricultural technologies raise ethical issues. We should worry because we did not create the institutions and capability needed to take on these issues at earlier stages in the development of technology. The issues do not come to the fore until long after private interests have spent large sums on R&D, creating a stake in their commercial success, irrespective of the ethical concerns. A more robust public discussion could alert the developers of technology to possible roadblocks and points of disagreement early enough that they could make more informed decisions about the specific design or features of a product, and in some cases, the wisdom of investing in a line of research at all. Such discussions themselves face challenges, (Thompson 2010a; Buckley et al.

370

13 Gene Editing, Synthetic Biology and the Next Generation …

2017), but there has been much progress both in the commitment to undertake such a conversation and in the methods for doing so, especially outside the United States. The question, “Should we worry?” might also be asked in a more apocalyptic tone. These new applications of gene technology challenge some boundaries humans have been reluctant to breach: modification of the human germline; development of organisms that are intended to spread throughout the environment and maintain themselves without ongoing assistance from humans (e.g. farmers). We are reminded of Mary Shelley’s Frankenstein, Kurt Vonnegut’s Cat’s Cradle and Margaret Atwood’s Oryx and Crake. What should we say about the prospects for an application of gene editing that runs amok, that undercuts the very survival of the human species? I do not mean to imply that these are meaningless or silly questions, but they are not the questions that pertain to agrifood applications of gene editing. Applications that breach these boundaries may indeed occur in domains where the risks appear to be offset by a compelling rationale for pressing forward. That might include relief of human suffering from genetic disease. It might include the conservation of endangered species and ecosystems. It might well include adaptation to or attempts to forestall climate change. However, (except perhaps for this last category) it is very unlikely to happen in agriculture before it has already happened elsewhere. Although there is a brave new generation of agricultural scientists who are brashly promoting their gene technologies as a means for averting some future famine, they will face the same burden of proof that has been applied to GMOs and, indeed, to other agricultural innovations for generations before them. If we apply those burdens unevenly and sometimes halfheartedly, that does not controvert the way in which attending to the past will continue to provide many sources of insight into the risks, benefits and other moral considerations of gene edited crops and animals.

References Betrán, F.J., S. Bhatnagar, T. Isakeit, G. Odvody, and K. Mayfield. 2006. Aflatoxin accumulation and associated traits in QPM maize inbreds and their testcrosses. Euphytica 152: 247–257. Bextine, B. n.d. Insect allies. DARPA. https://www.darpa.mil/program/insect-allies. Accessed 16 Mar 2020. Bextine, B. 2018. Statement from Dr. Blake Bextine, DARPA Program Manager for Insect Allies. https://www.darpa.mil/attachments/DARPA%20Response%20to%20Science% 20Opinion%20Piece%20-%20Oct%204%202018.pdf. Accessed 13 Jan 2020. Bonny, S. 2016. Genetically modified herbicide-tolerant crops, weeds, and herbicides: Overview and impact. Environmental Management 57: 31–48. Bos, J.M., B. Bovenkerk, P.H. Feindt, and Y.K. Van Dam. 2018. The quantified animal: Precision livestock farming and the ethical implications of objectification. Food Ethics 2: 77–92. Bovenkerk, B., and H.J. Nijland. 2017. The pedigree dog breeding debate in ethics and practice: Beyond welfare arguments. Journal of Agricultural and Environmental Ethics 30: 387–412. Bruce, A., D. Castle, C. Gibbs, J. Tait, and C.B.A. Whitelaw. 2013. Novel GM animal technologies and their governance. Transgenic Research 22: 681–695. Buckley, J.A., P.B. Thompson, and K.P. Whyte. 2017. Collingridge’s dilemma and the early ethical assessment of emerging technology: The case of nanotechnology enabled biosensers. Technology in Society 48: 54–63.

References

371

Caplan, A.L., B. Parent, M. Shen, and C. Plunkett. 2015. No time to waste—the ethical challenges created by CRISPR. EMBO Reports 16: 1421–1426. Carlson, D.F., C.A. Lancto, B. Zang, E.-S. Kim, M. Walton, D. Oldeschulte, C. Seabury, T.S. Sonstegard, and S.C. Fahrenkrug. 2016. Production of hornless dairy cattle from genome-edited cell lines. Nature Biotechnology 34: 479–481. Carruth, A. 2013. Culturing food: Bioart and in vitro meat.". Parallax 19: 88–100. CDC. (U.S. Centers for Disease Control and Prevention). 2016. A history of anthrax. https://www. cdc.gov/anthrax/resources/history/index.html. Accessed 13 Jan 2020. Coles, D., L.J. Frewer, and E. Goddard. 2015. Ethical issues and potential stakeholder priorities associated with the application of genomic technologies applied to animal production systems. Journal of Agricultural and Environmental Ethics 28: 231–253. Collins, F.H., and A.A. James. 1996. Genetic modification of mosquitoes. Science and Medicine 3: 52–61. Courtier-Orgogozo, V., B. Morizot, and C. Boëte. 2017. Agricultural pest control with CRISPRbased gene drive: Time for public debate.". EMBO Reports 18: 878–880. Datar, I., and D. Luining. 2015. Mark Post’s cultured beef, New Harvest. https://www.new-harvest. org/mark_post_cultured_beef. Accessed 4 Jan 2020. Delwaide, A.-C., L.L. Nalley, B.L. Dixon, D.M. Danforth, R.M. Nayga Jr., E.J. Van Loo, and W. Verbeke. 2015. Revisiting GMOs: Are there differences in European consumers’ acceptance and valuation for cisgenically vs transgenically bred rice? PloS One 10 (5): e0126060. https://doi. org/10.1371/journal.pone.0126060. Desowitz, R.S. 1991. The Malaria Capers: Tales of Parasites and People. New York, W.W: Norton. El Zahabi-Bekdash, L., and J. Lavery. 2010. Achieving precaution through effective community engagement in research with genetically modified mosquitoes. Asia Pacific Journal of Molecular Biology and Biotechnology 18: 247–250. EU. (European Union). 2018. Dual use controls. Update May 18, 2018. https://ec.europa.eu/trade/ import-and-export-rules/export-from-eu/dual-use-controls/. Accessed 13 Jan 2020. Faurant, C. 2011. From bark to weed: A history of artemisinin. Parasite 18: 215–218. Fuentes, C., and M. Fuentes. 2017. Making a market for alternatives: Marketing devices and the qualification of a vegan milk substitute. Journal of Marketing Management 33 (7–8): 529–555. https://doi.org/10.1080/0267257X.2017.1328456. FDA (U.S. Food and Drug Administration). 2017. What is gene therapy? How does it work?. https:// www.fda.gov/consumers/consumer-updates/what-gene-therapy-how-does-it-work. Accessed 2 Jan 2020. Ferrari, A. 2012. Animal disenhancement for animal welfare: The apparent philosophical conundrums and the real exploitation of animals. A response to Thompson and Palmer. NanoEthics 6: 65–76. Ferrari, A. 2015. Animal enhancement: Technovisionary paternalism and the colonisation of nature. In Inquiring into Animal Enhancement: Model or Countermodel of Human Enhancement?, ed. S. Bateman, J. Gayon, S. Allouche, J. Goffette, and M. Marzano, 13–33. London: Palgrave Pivot. Gonsalves, D., S. Tripathi, J.B. Carr, and J.Y. Suzuki. 2010. Papaya Ringspot virus. The Plant Health Instructor. https://doi.org/10.1094/PHII2010100401. https://sipco.ica.gov.co:8080/bitstr eam/123456789/17511/2/Gonsalves_etal_2010.pdf. Accessed 13 Jan 2020 Haeussler, M. 2019. CRISPR Off-Targets. Cell Biology and Toxicology. https://doi.org/10.1007/ s10565-019-09497-1. Henschke, A. 2012. Making sense of animal disenhancement. NanoEthics 6: 55–64. Hobbs, R.J., E.S. Higgs, and C. Hall (eds.). 2013. Novel Ecosystems: Intervening in the New Ecological World Order. Hoboken, NJ: John Wiley and Sons. Hopkins, P.D., and A. Dacey. 2008. Vegetarian meat: Could technology save animals and satisfy meat eaters? Journal of Agricultural and Environmental Ethics 21: 579-596. IANS (Indo-Asian News Service). 2017. India to be testing ground for questionable genetic control of mosquitos? Business Standard, Nov. 17. https://www.business-standard.com/article/

372

13 Gene Editing, Synthetic Biology and the Next Generation …

news-ians/india-to-be-testing-ground-for-questionable-genetic-control-of-mosquitoes-117111 700346_1.html. Accessed 16 Jan 2020. Impossible Foods. n.d. Heme + Science: Behind the Impossible. https://impossiblefoods.com/ heme/. Accessed 4 Jan 2020 Jacobsen, R. 2017. The biography of a plant-based burger, Pacific Standard, Original Sept. 6, 2016, updated July 28, 2017. https://psmag.com/news/the-biography-of-a-plant-based-burger#. 1dzfvzx4b. Accessed 4 Jan 4 2020. James, A.A. 2005. Gene drive systems in mosquitoes: Rules of the road. Trends in Parasitology 21: 64–67. James, S., C.P. Simmons, and A.A. James. 2011. Mosquito trials. Science 334: 771–772. Jones, J.D.G., and J.L. Dangl. 2006. The plant immune system. Nature 444: 323–329. Kendig, C.E. 2014. Synthetic biology and biofuels. In Encyclopedia of Food and Agricultural Ethics, eds. P.B. Thompson and D. Kaplan. New York: Springer. Kendig, C.E. 2016. What is proof of concept research and how does it generate epistemic and ethical categories for future scientific practice? Science and Engineering Ethics 22: 735–753. Kouba, V. 2004. History of the screwworm (Cochliomyia hominivorax) eradication in the Eastern Hemisphere. Historia Medicinae Veterinariae. 29 (2): 43–53. Lavery, J.V., L.C. Harrington, and T.W. Scott. 2008. Ethical, social, and cultural considerations for site selection for research with genetically modified mosquitoes. The American Journal of Tropical Medicine and Hygiene 79: 312–318. MacDonald, K.M. 2018. Absolute hogwash: Assemblage and the new breed of animal biotechnology. In Ethical Tensions from New Technology: The Case of Agricultural Biotechnology, ed. H.S. James, Jr, 126–136, Wallingford, Oxfordshire, UK: CABI International. Macer, D. 2005. Ethical, legal and social issues of genetically modifying insect vectors for public health.". Insect Biochemistry and Molecular Biology 35: 649–660. McKenna, E. 2018. Livestock: Food, Family and Friends. Athens: University of Georgia Press. Meghani, Z., and C. Boëte. 2018. Genetically engineered mosquitoes, Zika and other arboviruses, community engagement, costs, and patents: Ethical issues. PLoS Neglected Tropical Diseases 12(7). https://www.ncbi.nlm.nih.gov/pmc/articles/PMC6062015/. Accessed 17 March 2020 Metcalf, J. 2013. Meet shmeat: Food system ethics, biotechnology and re-worlding technoscience. Parallax 19: 74–87. Murphy, K.N., and W.P. Kabasenche. 2018. Animal disenhancement in moral context. NanoEthics 12: 225–236. Myskja, B.K. 2006. The moral difference between intragenic and transgenic modification of plants. Journal of Agricultural and Environmental Ethics 19: 225–238. Myskia, B.K., and A.I. Myhr. 2020. Non-safety assessments of genome-edited organisms: Should they be included in regulation? Science and Engineering Ethics. https://doi.org/10.1007/s11948020-00222-4. Nagarajan, V., and R.V. Bhat. 1972. Factors responsible for varietal differences in aflatoxin production in maize. Journal of Agricultural and Food Chemistry 20: 911–914. Naylor, R., and P.R. Ehrlich. 1997. Natural pest control services and agriculture. In Nature’s Services: Societal Dependence on Natural Ecosystems, ed. G.C. Daily, 151–174. Washington, DC: Island Press. Palmer, C. 2011. Animal disenhancement and the non-identity problem: A response to Thompson. NanoEthics 5: 43–48. Palmer, C. 2016. Saving species but losing wildness: Should we genetically adapt wild animal species to help them respond to climate change? Midwest Studies in Philosophy 40: 234–251. Perfect Day. 2019. How it works. Making dairy with the magic of flora. Perfect Day. https://www. perfectdayfoods.com/learn-more/. Accessed 4 Jan 2020. Picasso, V.D., E.C. Brummer, M. Liebman, P.M. Dixon, and B.J. Wilsey. 2011. Diverse perennial crop mixtures sustain higher productivity over time based on ecological complementarity. Renewable Agriculture and Food Systems 26: 317–327.

References

373

Pimentel, D. 1995. Amounts of pesticides reaching target pests: Environmental impacts and ethics. Journal of Agricultural and environmental Ethics 8: 17–29. Pluhar, E.B. 2010. Meat and morality: Alternatives to factory farming. Journal of Agricultural and Environmental Ethics 23: 455–468. Preston, C.J. 2019. The Synthetic Age: Outdesigning Evolution, Resurrecting Species, and Reengineering Our World. Cambridge, MA: MIT Press. Pugh, J. 2016. Driven to extinction? The ethics of eradicating mosquitoes with gene-drive technologies. Journal of Medical Ethics 42: 578–581. Pusch, A.F. 2015. Splices: When science catches up with science fiction. NanoEthics 9: 55–73. Reeves, R.G., S. Voeneky, D. Caetano-Anollés, F. Beck, and C. Boëte. 2018. Agricultural research, or a new bioweapon system? Science 362: 35–37. Regalado, A. 2019. Gene-edited cattle have a major screwup in their DNA. MIT Technology Review, Aug 29. https://www.technologyreview.com/2019/08/29/65364/recombinetics-gene-edi ted-hornless-cattle-major-dna-screwup/. Accessed 7 July 2020. Resnik, D.B. 2018. Ethics of community engagement in field trials of genetically modified mosquitoes. Developing World Bioethics 18: 135–143. Rollin, B. 1995. The Frankenstein Syndrome: Ethical and Social Issues in the Genetic Engineering of Animals. New York: Cambridge University Press. Rollin, B. 1996. Bad ethics, good ethics and the genetic engineering of animals in agriculture. Journal of Animal Science 74: 535–541. Rollin, B. 1998. On Telos and genetic engineering. In Animal Biotechnology and Ethics, ed. A. Holland and A. Johnson, 156–187. London: Chapman and Hall. Russell, A.W., and R. Sparrow. 2008. The case for regulating intragenic GMOs. Journal of Agricultural and Environmental Ethics 21: 153–181. Sandler, R. 2019. The ethics of genetic engineering and gene drives in conservation. Conservation Biology. https://doi.org/10.1111/cobi.13407. Sanguansri, P., and M.A. Augustin. 2006. Nanoscale materials development–a food industry perspective. Trends in Food Science & Technology 17: 547–556. Schaefer, G.O., and J. Savulescu. 2014. The ethics of producing in vitro meat. Journal of Applied Philosophy 31: 188–202. Schouten, H.J., F.A. Krens, and E. Jacobsen. 2006. Cisgenic plants are similar to traditionally bred plants: international regulations for genetically modified organisms should be altered to exempt cisgenesis. EMBO Reports 7: 750–3. https://doi.org/10.1038/sj.embor.7400769. Schultz-Bergin, M. 2017. The dignity of diminished animals: Species norms and engineering to improve welfare. Ethical Theory and Moral Practice 20: 843–856. Scott, D. 2018. Technological pragmatism: Navigating the ethical tensions created by agricultural biotechnology. In Ethical Tensions from New Technology: The Case of Agricultural Biotechnology, ed. H.S. James, Jr, 112–125. Wallingford, Oxfordshire, UK: CABI International. Scott, M.J., F. Gould, M. Lorenzen, N. Grubbs, O. Edwards, and D. O’Brochta. 2018. Agricultural production: assessment of the potential use of Cas9-mediated gene drive systems for agricultural pest control. Journal of Responsible Innovation 5 (sup1): S98–S120. Shriver, A. 2009. Knocking out pain in livestock: Can technology succeed where morality has stalled? Neuroethics 2: 115–124. Telem, R.S., S.H. Wani, N.B. Singh, R. Nandini, R. Sadhukhan, S. Bhattacharya, and N. Mandal. 2013. Cisgenics—A sustainable approach for crop improvement. Current Genomics 14: 468–476. https://doi.org/10.2174/13892029113146660013. Thomas, M.B., and A.J. Willis. 1998. Biocontrol—risky but necessary? Trends in Ecology & Evolution 13: 325–329. Thompson, P.B. 2008. The opposite of human enhancement: Nanotechnology and the blind chicken problem. NanoEthics 2: 305–316. Thompson, P.B. 2010. Agrifood nanotechnology: Is this anything? In Understanding Nanotechnology: Philosophy Policy and Publics, ed. U. Fiedeler, C. Coenen, S.R. Davies, and A. Ferrari, 157–170. Germany: Akademische Verlagsgesellshchaft AKA GmbH, Heidelberg.

374

13 Gene Editing, Synthetic Biology and the Next Generation …

Thompson, P.B. 2010. Why using genetics to address welfare may not be a good idea. Poultry Science 89: 814–821. Thompson, P.B. 2012. Synthetic biology needs a synthetic bioethics. Ethics, Policy and Environment 15: 1–20. Thompson, P.B., and W. Hannah. 2012. Novel and normal risk: Where does nanotechnology fit in? In Environmental Ethics: What Really Matters? What Really Works?, ed. D. Schmitz and E. Willett, 609–622. New York: Oxford University Press. Thompson, P.B. 2015. From synthetic bioethics to one bioethics: A reply to critics. Ethics, Policy & Environment 18: 215–224. Thompson, P.B. 2017. The Spirit of the Soil: Agriculture and Environmental Ethics. New York: Routledge. Troitino, C. 2017. The Dairy Pride Act’s beef with plant-based milk, Forbes April 6. https://www. forbes.com/sites/christinatroitino/2017/04/06/the-dairy-pride-acts-beef-with-plant-based-milk/# b19c94b454dc. Accessed 4 Jan 2020. U.S. Department of State. n.d. Biological Weapons Convention. https://www.state.gov/biologicalweapons-convention/. Accessed 13 Jan 2020. Van der Weele, C., and C. Driessen. 2013. Emerging profiles for cultured meat; ethics through and as design. Animals 3: 647–662. Vandeman, A.M. 1995. Management in a bottle: Pesticides and the deskilling of agriculture. Review of Radical Political Economics 27: 49–59. Van Eenennaam, A.L. 2018. The importance of a novel product risk-based trigger for gene-editing regulation in food animal species. The CRISPR Journal 1: 101–106. Warkentin, T. 2006. Dis/integrating animals: Ethical dimensions of the genetic engineering of animals for human consumption. AI and Society 20: 82–102. Weisberg, Z. 2015. Biotechnology as end game: Ontological and ethical collapse in the “biotech century.” NanoEthics 9: 39–54. Yaqoob, A., A.A. Shahid, T.R. Samiullah, A.Q. Rao, M.A.U. Khan, S. Tahir, S.A. Mirza, and T. Husnain. 2016. Risk assessment of Bt crops on the non-target plant-associated insects and soil organisms. Journal of the Science of Food and Agriculture 96: 2613–2619. Zhang, X.-H., L.Y. Tee, X.-G. Wang, Q.-S. Huang, and S.-H. Yang. 2015. Off-target effects in CRISPR/Cas9-mediated genome engineering, Molecular Therapy - Nucleic Acids 4: e264, ISSN 2162-2531. https://doi.org/10.1038/mtna.2015.37. Zhang, Y., K. Massel, I.D. Godwin, and C. Gao. 2018. Applications and potential of genome editing in crop improvement. Genome Biology 19: 210.

Chapter 14

Biotechnology, Controversy and the Philosophy of Technology

Abstract This concluding chapter situates the previous 13 chapters in the book within themes in the philosophy of technology. I begin by asking how philosophy could contribute to the controversy over biotechnology, through critique and or defense of this new suite of tools and techniques. The chapter continues by discussing how two key themes in the book—the focus on agriculture and the risk-based approach—could enrich the work of other philosophers working on Technoscience. Themes on the ontology of risk are given particular emphasis, and the arguments in the previous chapters are developed as an extension of Don Ihde’s postphenomenology. The chapter concludes by summarizing themes from The Spirit of the Soil (Thompson in The spirit of the soil: agriculture and environmental ethics. Routledge, New York, 2017) and indicating how the previous chapters in this book relate to a more general philosophical analysis of agricultural research. Keywords Risk · Risk assessment · Public acceptance of biotechnology · Socially relevant philosophy of science · Techne · Philosophy of agriculture · Don Ihde · Postphenomenology The process of revising a text written almost 25 years ago inspires reflection on the larger significance of the debate over food and agricultural applications of gene technology. Looking back on several decades of work, I confess occasional annoyance when asked to contribute new essays on agrifood biotechnology. During the time that I was writing the first draft of this book and supplementing it with perhaps a dozen more journal articles on the subject, I saw myself as doing work on technological risk that could be generalized to debates over autonomous technology and the cultural tendency to approach technological innovation in Manichaean terms. As Hans Achterhuis has argued, for most of the 20th century, philosophers viewed technology as either a godsend, or a menace from hell, but in either case as a quasimetaphysical force divorced from any specific tools or techniques. American philosophers, however, were taking “the empirical turn,” (Achterhuis 2001). My book was part of this movement, but even as the first edition of was appearing in 1997, I was already turning my attention more directly to the philosophy of sustainability. By this

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6_14

375

376

14 Biotechnology, Controversy and the Philosophy of Technology

time, I had begun to think of agriculture itself as a philosophically rich topic for both environmental studies and for the philosophy of technology. Michael Pollan’s The Omnivore’s Dilemma came out in 2006, stimulating a new cadre of philosophers to undertake research and teaching in the area of food ethics. As such, I have neglected the task of situating my work on biotechnology more firmly in the philosophical literature. This chapter lays out some of the threads that connect my work on risk with philosophical pragmatism, on the one hand, and the influential approach to philosophy of technology pioneered by doctoral advisor, Don Ihde, on the other. I discuss of how both agriculture and risk assessment could become a topic of wider application and philosophical interest, and conclude by coming back to the philosophy of agricultural research developed in the book that preceded this one, The Spirit of the Soil (Thompson 2017).

14.1 The Global Controversy: What Role for Philosophy? As discussed in the Introduction and Chap. 1, the first GMOs went into wide production in 1997 and 1998. Farmers in the United States adopted them rapidly, but introduction into Europe did not go well. By 2002, John Durant, Martin Bauer and George Gaskel were able to summarize extensive social science research funded by European science agencies to explain this “global controversy,” (Durant et al. 2002). Viewed in retrospect, the half decade between the first plantings of GMOs and Europe’s rejection of the technology is an anomaly demanding explanation. Indeed, a mythos now pervades this episode which includes the supposition that the introduction of gene technologies went smoothly in the United States, while Europeans (wisely or irrationally, depending on your point of view) resisted them. The mythos supposes that the wisdom of introducing gene technologies into food production was never discussed or debated in the United States, and that the U.S. biotechnology industry was caught by surprise when they encountered problems in moving their new seeds through the European regulatory system. Rachel Schurman and William Munro’s sociology of the controversy goes some distance toward correcting this picture. They describe a small network of activists working collaboratively in the United States and Europe prior to the release of the first transgenic crops in the late 1990s. They also discuss the recombinant bovine growth hormone (rBGH) controversy (see Chap. 3) and illustrate how it laid the foundations for later activism by helping critics see that mobilization of the public would depend on raising questions about the safety of gene technology. Though forgotten by the late 1990s, American newspapers covered controversy over rBGH widely throughout the 1980s and into the 1990s. Schurman and Munro describe the coalition of individuals and organizations as motivated by a blend of concerns. They included agriculture-specific concerns about the welfare of livestock and threats to the livelihood of smallholders, but some participants in the network saw any application of gene technology as a step down the slippery slope to human eugenics. This network

14.1 The Global Controversy: What Role for Philosophy?

377

of social advocates distrusted the growing power of corporations and the processes of globalization, (Schurman and Munro 2010). Combined with controversies over the potential use of gene technologies in human medicine, the bovine growth hormone debate had prompted a “national conversation” orchestrated by the Keystone Center, (see Thompson 2008 for details). This was followed by the formation of the National Agricultural Biotechnology Council (NABC), discussed in Chap. 12. A rich scholarly literature on significance of agricultural biotechnology predates the release of crops or the European debate. In fact, even the European debate predates the European debate. The first European academic conference on agricultural ethics was held at the University of Nottingham’s Easter School in 1993, where gene technology was a highly prominent topic of discussion. No one who heard Maurice Lex of the European commission speak at that meeting was surprised by Europe’s reaction. Lex stood before an audience of mostly scientists and told them that the science notwithstanding, political authorities in Europe would not able to ignore the massive effect of public opinion that was already building, (Lex 1995). Although Schurman and Munro repeatedly express the importance of ethics in the debate over gene technology, they do not link this theme to the work of philosophers or bioethicists. Aside from a single reference to Sheldon Krimsky, (p. 121), who is not identified as a professor of philosophy, the role that philosophy might play in mounting and ethical critique goes unmentioned. Yet in addition to myself, there were at least three other philosophers at the Nottingham meeting in 1995, Alan Holland, R.G. Frey and Roger Straughan. Straughan’s book with Michael Reiss covered biomedical as well as agricultural gene technologies, and preceded the first edition of this book by a year (Reiss and Straughan 1996). As Chaps. 1 through 12 of this book make abundantly clear, philosophers including Eric Millstone and Vandana Shiva had already created an extensive literature on the ethics of agricultural and food biotechnologies. In addition to those of us at the Nottingham meeting, the 1997 edition cited widely read works by Mark Sagoff, Jeff Burkhardt, Bernard Rollin and Stephen Stich. What did these philosophers think they were doing? On the one hand, the philosophers cited throughout the 1997 and 2007 editions of this book were examining a technical practice: the integration of rDNA tools for manipulating plant and animal genomes in the context of agricultural science. They identified how values and interests were shaping the development and conceptualization of that practice. They were on the alert for assumptions that, made differently, might reshape the practice and its outcomes, and they were particularly attentive to assumed values that needed defense or revision on ethical grounds. On the other hand, philosophers were interested in the problem of unintended consequences. The phenomenon of unwanted outcomes had become obvious, and was being theorized from disciplinary social science perspectives as well as from emerging cross and interdisciplinary fields including policy studies, science studies and bioethics. Philosophy contributed elements from ethics and political philosophy to this new domain of theory, but metaphysics and epistemology were equally relevant. By 2010, this type of philosophical research was being done by scholars who called it socially relevant philosophy of (and in) science. The “of” implies using concepts and theories

378

14 Biotechnology, Controversy and the Philosophy of Technology

from philosophy of science to analyze science’s social relevance. The “in” implies working cooperatively with scientists to steer the social implications of their work in more socially appropriate directions, (that is, avoiding unintended consequences) (Fehr and Plaisance 2010; Tuana 2010). In an earlier study, Schurman and Munro portrayed academic professionals as key influencers in stimulating debate. They cite economists work on the impact of rBST and sociological work on farm structure as both motivating social advocates and providing them with data and analysis to support their position, (Schurman and Munro 2003). I do not know whether the philosophical studies had similar influence, but the fact that recent studies overlook this work suggests that it did not, (see, for example Hicks and Millstein 2016; Ankeny and Bray 2018), The presumption that food and agricultural gene technologies were developed and introduced into the United States without controversy may lead scholars to assume that no philosophers were writing on the subject. However, recent studies also emphasize philosophy of science, rather than philosophy of technology. Given philosophy of science’s traditional focus on explanation, inference and theory building, the new work on values in science continues to prioritize these aspects of science. The approach is ill-suited to the study of a technical domain such as the agricultural sciences. As discussed at more length below, simply explaining how plants or livestock grow was never the rationale for bringing scientific investigative techniques to agriculture. Helping farmers has always been the dominant criterion of success. This means developing something they can use, and the most effective translations from the experiment station to the laboratory have tended to involve things, rather than ideas (see Suppe 1987; Zimdahl 2003 for further exploration of this idea).

14.2 Philosophy of Technology If tools and techniques, rather than theoretical knowledge, are the key outputs for agricultural research, one would expect that the biotechnology controversy to be a hot topic in the philosophy of technology, and it has attracted some attention. However, philosophers who study technology view agriculture as case study. It is an appropriate topic and one with potentially significant policy implications. However, practitioners of the sub-discipline do not see it as a subject with the potential for lasting impact or widespread engagement. Unlike information technology or new fields in engineering, agriculture has been around for a long time. Changes in agricultural production do not appear to have the dramatic transformative effects of technological innovations in other fields. Indeed, philosophers who study the food system have struggled to make readers cognizant of the way that food is saturated by tools and techniques, (Kaplan 2012). As a case study, agriculture is viewed as an optional area of application, rather than central to the philosophy of technology. I believe that this is a philosophical error, and will summarize a few of the reasons why I think so. Larry Hickman notes the difference between technology, on the one hand, and tools and techniques, on the other. A hammer is a tool; the craft of metalwork is a

14.2 Philosophy of Technology

379

technique. The collection of tools and techniques in use will vary according to time and place, but all human cultures (as well as some nonhumans) will have some. Technology is a reflective theorizing practice or logos on tools and techniques. Hickman argues that John Dewey’s functionalist epistemology resituates knowledge creation (e.g. science) as a technique, making the reflective theory of knowledge (e.g. philosophy) into technology in Hickman’s differentiated sense, (Hickman 1990, 2001). In contrast to Hickman’s proposal, common usage often treats technology and tools and techniques synonymously. Even among philosophers, Hickman’s careful definitions are abused, but the conceptual point he makes in distinguishing an assemblage of tools and techniques from reflective consideration of its logic, function and metaphysics is widely accepted. Carl Mitcham’s foundational work in the philosophy of technology deploys this distinction in identifying the historical origin of technological thinking (e.g. Hickman’s technology) as opposed to tools and technical practices, which preceded the advent of written culture. Mitcham notes that the Greek word τšχνη (techne) indicates craftsmanship or art. Greek philosophers saw ´ it as a contrast case for ™πιστημη (episteme) and thus engaged in philosophical debate over the difference between what contemporary analytic philosophers would describe as knowing that (episteme) and knowing how (techne). Mitcham holds that Greek thought saw techne or art as an inappropriate object of theory or science. As such, a reflective practice of techniques, tools and technical methods does not emerge until the modern period (e.g. the sixteenth, seventeenth and eighteenth century). On Mitcham’s approach, technology as a practice guided by logos or reasoning is roughly modeled by what we now call engineering, (Mitcham 1994). As a colleague, Mitcham is fully accepting of the idea that studies of agriculture belong in the philosophy of technology, but his theoretical model sharply limits the significance that a philosophy of agriculture could have. To focus for a moment on the Greeks, a focus on τšχνη (techne) misleads philosophers of technology, the word’s status as the etymological root of the English word technology notwithstanding. In fact, Greek philosophy does undertake reflective theorization of technical practice within the polis. It is just not focused on the musicians, sculptors and craft artisans that are the exemplars of techne. Xeonophon’s Oeconomicus is a particularly striking example. It is, on the face of it, a discourse on household management, though the dialog exhibits an excess of complexities. The layered dialogic structure puzzles the identification of the narrator: is it Ischomachus, the farmer from whose mouth much of the text radiates? Alternatively, the narrator is Socrates, who recounts an earlier conversation with Ischomachus to Critobulus (b. c. 369 BCE), but we cannot discount the role of Critobulus, nor the role of Xenophon himself. Critobulus (sometimes rendered as Crito) was a friend of Socrates and a wealthy farmer. Xenophon (d. 354 BCE) would have thus portrayed Socrates talking to someone who knew quite a bit about the subject of the dialog. Leo Strauss (1899–1973) was among the earliest interpreters to read Oeconomiucs as a work saturated by irony, satire and misdirection, (Bruell 1984). Recent scholarship emphasizes the section of the dialog in which Ischomachus discusses his instructions to his young wife. Michel Foucault (1926–1984) characterized this component of Xenophon’s dialog as a display of macho posturing and male domination (Foucault 1985), and the

380

14 Biotechnology, Controversy and the Philosophy of Technology

Oeconomicus has become standard reading in many courses on feminist thought. Relatedly, Ken Stikkers analyzes Foucault’s treatment of Xenophon as a paradigmatic example of power relations functioning through a spiritual medium (Stikkers 2009). While some of these interpretations treat Xenophon’s text as an exemplar of male dominance, those that stress the ironic misdirection in the text suggest that at least one of the possible narrators intended to be exposing dubious practices in the worldview of the typical Greek patriarch. I see this reading as problematic in virtue of the fact that at least as much of the text is taken up with Ischomachus describing some basic practices of agricultural production. These are not part of Ischommachus’ instructions to his wife. What is more, this component of the text introduces yet another ironic turn that I have not seen discussed in recent studies of Xenophon. While Xenophon’s Socrates initiates the conversation by pleading total ignorance of household management, Ischomachus’s description of the tasks needed to prepare and steward soils, crops and the harvest process is conducted in a manner that parallel’s Socrates own instruction of the slave boy in Plato’s Meno. Ischomachus shows that Socrates already knows quite a bit about farming, the fact that he has never done it, notwithstanding. Perhaps Socrates is telling his friend Crito about the time he (Socrates) rediscovered an implicit knowledge. Xenophon’s Oeconomicus is thus a text that not only articulates a reflective characterization of a technical practice, but also suggests that knowledge of this practice permeates Greek culture, even if this knowledge remains implicit for some of the most brilliant Greek minds. What is more, other philosophers consider the centrality and proper institutionalization of farming within the Greek way of life. We see this, for example, in Aristotle’s Politics, where he praises the hoi mesoi (e.g. household farmers) as the best part of the city. These passages suggest a backstory to the philosophy being written in the aftermath of the Peloponnesian War (ended 404 BCE). Philosophers are pondering how the technological infrastructure of Athens (emphasizing shipbuilding and trade) might have affected Athenian personality and the functioning of its democracy, especially in comparison to that of other city states where governance was more firmly tied to the farming class, (Mitchell 2015). I should note that in calling them farmers, I am not also implying that they spent much time with their hands in the soil. The point, simply, is that when we notice the centrality of agriculture an entire discourse on technology and technical practice opens up to us. If we include agriculture among the assemblage of tools and technique, the philosophy of technology extends much further back into our history than Mitcham seems to think. As agriculture becomes a pivot for reflection on technical practice, the biotechnology controversy should be seen as something more central to the philosophy of technology than it has been. Indeed, the entire complex of agricultural practices, tools and techniques should be viewed as a core topic for philosophers with an interest in the connection between technology, sustainability and the potential catastrophic impacts of climate change. However, there is also a sense in which I myself view the biotechnology component of my life’s work as something of a sideshow. My deeper interests lie in recovering the philosophy of agriculture as a topic of philosophical

14.2 Philosophy of Technology

381

reflection, and establishing its place in environmental ethics and policy. As discussed below, my book The Spirit of the Soil provides a more comprehensive framework for criticizing the mindset that accompanies much advocacy for gene technologies than I have developed in this book. What this book adds to that critique (viewed from my retrospective vantage point) is a bridge between my larger interests in agriculture as a dominant feature of humanity’s sociotechnical environment and the concerns with uncertainty and risk that were at the core of Hans Jonas’s important book, The Imperative of Responsibility (Jonas 1984). Jonas has arguably been more influential for bioethics than for philosophy of technology, though aspects of his text that surfaced as an exploration of “the precautionary principle,” provide an obvious demonstration of the role that risk assessment can play in any comprehensive philosophy of technology.

14.3 Risk Assessment and Its Enemies Although its roots extend to Aristotle and its quantitative foundations were laid in the early decades of the 17th century, risk analysis is a comparatively young discipline. The Society for Risk Analysis was established in 1980, the year in which I defended a philosophy dissertation focused on risk assessments for nuclear power plants. Chapters 1 and 2 lay out a general picture of risk assessment organized in terms of four tasks: hazard identification, exposure quantification, risk management and risk communication. These tasks are defined and developed within the community of researchers and policy makers around the presumption that risk can be understood as a function of hazard (e.g. the bad things that might happen) and exposure (the likelihood that they will happen, given some initializing event or conditions). An array of scientific methods can be deployed in operationalizing this approach for decision-making. Thus, risk management is the point in the process where an individual, a policymaker or some other choice process adopts a program of action based on a scientific characterization of risk. Collectively, these three components risk assessment facilitate a planned, rational approach to the regulation of technological innovations. When society as whole is the subject that will bear the brunt of a new technology’s negative impact, these three components of risk assessment extend the expected-value or expected-utility approach to the problem of anticipating and managing unanticipated consequences. The expected-value approach to risk was developed along with mathematical theories of chance and probability throughout the 17th and 18th century. It was implicit in the earliest pages of Jeremy Bentham’s characterization of utilitarian moral theory published in 1787. Little if any of this theory utilized the word risk as a pivot for its central constructs or methods of quantification. The conceptual achievement was understood as an approach to the measurement of chance and uncertainty, and as refinements in the procedure for inductive inference. The word probable underwent a transformation from meaning something like plausible or likely to indicating relative frequencies or a numerical value between zero and one, (see Hacking 1990, 2006).

382

14 Biotechnology, Controversy and the Philosophy of Technology

Throughout this period, risks were actions known to involve danger and the potential for loss. Equipping ships for an ocean voyage, for example, was understood to be risky. The institution of insurance emerged as convention for investors to limit their exposure to a financial catastrophe by sharing the loss when the expected thing occurred, and some fraction of ships sent out failed to return. Insurers applied the mathematics that had been developed to model games of chance and combined it with the data they already had on the ratio of losses to compute the price of insurance: the loss that anyone equipping a ship should expect. The institution of insurance became a model for other forms of financial decision-making, and was generalized in economic theory during the late 19th and early 20th century, (Bernstein 1998). This attenuated story of risk assessment suggests at least two interpretive possibilities. One is that the players in this narrative were on the trail of an abstract quantity, a mathematical ideal. Each turn in the plot reflects an advance in knowledge, as the community of scientific inquiry hones in on the essential logical form of expected utility and develops more and more sophisticated tools for its measurement. From this perspective, the elusive concept that came to be known as risk is something like a Platonic form, something real. Depending on one’s metaphysical commitments, it might be even more real than the coffee I had for breakfast in virtue of its generality. Peter Bernstein promotes this way of thinking in his history of risk, Against the Gods. On my view, the Bernouli’s (the Swiss family of theorists who quantified chance), men such as John Graunt (who extended the theory to state planning, hence the term statistics), the empiricist philosophers from Hume through Mill who applied the theory to inductive inference, and denizens of the subscription room at Lloyds of London were inventing a succession of instruments for improving the conduct of human affairs. On this reading, which I associate with Ian Hacking, the plurality of inventive steps suggests that while later innovators drew from their predecessors, characterizing this trajectory as converging on an essential form is unwarranted, (see Hacking 1990, 2006). I am quite interested in this metaphysical debate, but a different point is more relevant in the present context. Even if one is committed to the ideal reality or essentiality of expectation, probability and rational choice, it is a mistake to associate this illusive quantity with risk. If we want to understand risk, we should examine how people use the word in ordinary language. When we do that, we will see that the word vaguely conjoins some usage that is well captured by expected-value along with other forms that are entirely distinct from it. This view (my view) opposes philosophies holding that mastery of the logos (e.g. rationality) is an elite practice. Given the research trajectory of risk assessment within the last quarter of the 20th century, scientists quantifying hazard and exposure for nuclear power, chemicals and, yes, biotechnology, convinced themselves that they knew what risk was, and how to manage it. The corollary to their view was that people conversing on matters of daily life were not competent judges of risk. The view is alive among dual-systems psychologists who believe that System 1 is error prone (albeit fast) while System 2 provides the paradigmatic instantiation of logical thought (see my discussion in Chap. 12).

14.3 Risk Assessment and Its Enemies

383

Grammar is the key to the concept of risk. To understand risk, we should look to the way that people talk about it. The more quantitatively precise notion of hazard and exposure is a reasonable extension of the concept, but only in some conversational contexts. The toxicologists, epidemiologists and ecologists who quantify hazard and exposure relating to gene technology are interested in hazard and exposure, and appropriately so. Nevertheless, they should not arrogate to themselves the authority to opine about whether ordinary people deceive themselves when they talk about risk. My view thus supports the impetus for including risk communication as a fourth component of risk assessment. However, I do not view this as an attempt to convince ordinary people that they should themselves consider the risks of new technology by considering its expected value, much less that they must accept the scientific establishment’s estimate of expected value. Hence, my view also supports many of the claims made by social scientists and communication theorists who have studied the deficit model, or who have advocated more robust and engaged forms of science communication. The details are laid out in Chap. 12. Nevertheless, the matter does not really end there, (though that is, indeed, where the 1997 and 2007 editions of this book did end). The philosophy of risk just sketched does not imply that people never use the word risk to mean the expectation of a bad outcome, nor does it imply that the tools of quantification developed over the 400-year history charted by Ian Hacking are useless. It does not imply that the risks of biotechnology are “social constructions,” except, of course, in the spirit of William James’ (1842–1910) assertion that the trail of the human serpent runs over everything, (James 1904). Yes, attempts to quantify what we should expect from deploying gene technologies in our food system will reflect a host of judgements about what matters, what the alternatives are and how we imagine ourselves able to cope with the unknowns. Yet gene technologies can both help and harm us. Toxins, allergens and damage to ecosystems are not figments; the risks are real. The expected value approach can help us anticipate outcomes, and it can give us a framework in which to share our vision and knowledge in making a collective decision about how we should go forward. If the mainstream view is insensitive to some crucial contingencies in the risk assessment framework, much of the scholarship and activism that has resisted the mainstream view is insensitive to the way that discussing hazards, exposure and our approach to managing risk can improve the quality of our conversation. Since the second edition of this book, Sven Ove Hansson has undertaken a more thoroughgoing exploration of these ideas (Hansson 2007, 2013), with several contributions directed specifically toward agrifood biotechnology, (Hansson and Joelsson 2013; Hansson 2014). Although I am in substantial agreement with Hansson on most issues, we have a fundamental disagreement on the concept of risk reflected by our linguistic analysis. This disagreement may reflect deeper philosophical differences, though I cannot say for sure. Hansson’s article for the Stanford Encyclopedia of Philosophy defines the word risk as follows: In non-technical contexts, the word “risk” refers, often rather vaguely, to situations in which it is possible but not certain that some undesirable event will occur. In technical contexts, the word has several more specialized uses and meanings. Five of these are particularly important since they are widely used across disciplines:

384

14 Biotechnology, Controversy and the Philosophy of Technology

risk = an unwanted event which may or may not occur. An example of this usage is: “Lung cancer is one of the major risks that affect smokers.” risk = the cause of an unwanted event which may or may not occur. An example of this usage is: “Smoking is by far the most important health risk in industrialized countries.” (The unwanted event implicitly referred to here is a disease caused by smoking.) Both (1) and (2) are qualitative senses of risk. The word also has quantitative senses, of which the following is the oldest one: risk = the probability of an unwanted event which may or may not occur. This usage is exemplified by the following statement: “The risk that a smoker’s life is shortened by a smoking-related disease is about 50%.” risk = the statistical expectation value of an unwanted event which may or may not occur. The expectation value of a possible negative event is the product of its probability and some measure of its severity. It is common to use the number of killed persons as a measure of the severity of an accident. With this measure of severity, the “risk” (in sense 4) associated with a potential accident is equal to the statistically expected number of deaths. Other measures of severity give rise to other measures of risk. Although expectation values have been calculated since the 17th century, the use of the term “risk” in this sense is relatively new. It was introduced into risk analysis in the influential Reactor Safety Study, WASH-1400 (Rasmussenn 1975, Rechard 1999). Today it is the standard technical meaning of the term “risk” in many disciplines. It is regarded by some risk analysts as the only correct usage of the term. risk = the fact that a decision is made under conditions of known probabilities (“decision under risk” as opposed to “decision under uncertainty”).

Hansson does not note that his fifth meaning derives from Frank Knight (1885– 1972), the University of Chicago economist who wrote Risk, Uncertainty and Profit (Knight 1921). Much of the theory behind each of these meanings was developed by economists between 1920 and 1960, but my disagreement with Hansson is more fundamental. None of the definitions he offers, including the vague reference to situations where possible but not certain that some undesirable event will occur, can be analytically substituted for the word risk in any of the following English sentences. A: You risk being sued if you loan your car to her. B: They risk annihilation with their reckless charge against Russian forces at the Battle of Balaclava. C: The deer risk attack from the wolves by grazing in that open meadow.

Notably, the word risk is a verb in each of these sentences. Alternative formulations of the verb illustrate further incongruities with Hansson’s definitions. D: The President’s concern with reelection risks the health of the nation. E: I risked my neck jumping off that rock ledge. F: The company will risk its reputation if it proceeds in this fashion.

To be sure, excepting E, all of these sentences involve situations where it is possible (but not certain) that an undesirable event will occur, (though when Alfred Lord Tennyson (1809–1892) poeticized B, annihilation was certain, while the loyalty he saluted depended on the troops risking their lives anyway). In E, we know that the

14.3 Risk Assessment and Its Enemies

385

undesirable event did not occur, yet that seems to have no bearing on whether or not the person in question risked his or her neck. The point is that in each of these cases, risk is an act, while in each of Hansson’s definitional cases, risk is an event, qualified by uncertainty, probability or some form of indeterminateness. Acts are, of course, events, but when the word risk appears as a verb, it cannot be parsed smoothly unless the subject of the sentence has agential features. To see this, consider the grammatical oddness of the following sentences: G: The hurricane risks death and destruction all across New Orleans. H: The motorcycle risked becoming involved in an accident. I: The chemicals risk softening the shell of songbirds.

These are all sentences where the word cause would parse smoothly, as would phrases like might cause, could cause and could possibly cause. Readers attempting to parse a sentence like H may find themselves reading it as “The motorcyclist risked becoming involved in an accident.” We subconsciously do the work of substituting an agent for a mindless object in the subject position. Events can be caused by virtually anything, but only subjects ontologically capable of action can risk. No self-respecting copy editor would let G, H or I pass without comment. I take this observation to have philosophical significance. There are other oddities associated with conversational usage of the word risk that do not hang upon the word’s basic grammar. Consider a few more examples: J: You risk your health cooking those peas. K: She risked her career chances by getting a college degree. L: The company risked its future by selling a commercially successful product.

In comparison to G, H and I, these are readily understandable as grammatical sentences. However, in comparison to A through F, they are odd because they express a something contrary to our conversational expectations. They say something opposite to what ordinary language generally implicates, to use terminology introduce by Paul Grice (1913–1988). Companies that sell commercially successful products don’t normally put their future at risk by doing so. Women risk their career chances by dropping out of school, not by getting their degrees. And why would something as routine as cooking peas have an uncertain but undesirable consequence? Uttered in conversation, J, K and L call for further explanation: What do you mean? A through F implicate contexts in which the use of the word risk seems unexceptional; J, K and L implicate a need for further explanation and contextualization. Now, I do not deny that the word risk can be used in all of the ways that Hansson discusses, nor would I disagree with his claim that technical risk assessment almost always deploys usage in what I have elsewhere characterized as the event-predicting sense, (Thompson 1999). Normal, non-scientific conversations will also deploy the word risk to discuss possible eventualities. Yet proficient speakers of English are also fluent with the grammatical conventions that would make them recognize J, K. and L as statements inviting further elaboration, and G, H and I as statements that could only be uttered as jabberwocky. These conventions involve the pragmatics of the word

386

14 Biotechnology, Controversy and the Philosophy of Technology

risk. I have argued that they reflect an implied categorical distinction between acts that are risks (or risky) and acts that are unexceptional, everyday occurrences, elements of what Pierre Bourdieu calls habitus. Acts in the first category are generally recognized as dangerous (or GRAD, to paraphrase the FDA). A conversational utterance of J, K or L is, in effect, to state than a non-GRAD action should be seen as GRAD. The follow-ups are natural: Why? What should I do about it? The ethical significance of my disagreement with Sven Ove is covered in Chap. 12. I take it that the scientific community’s over commitment to event-predicting senses of the word risk has made them insensitive to the perfectly rational questions that people ask when they are told that a normal part of their habitus has been moved into GRAD territory. Of course, the scientists do not regard GMOs as particularly dangerous, so they proceed as if there is no need for further explanation. They assure people that, on the one hand, the chance of an unwanted outcome from eating GMOs is quite low, and, on the other hand, that they were mistaken about unexceptional nature of eating peas, corn or what have you, in the first place. There is, after all, no zero risk! My view is that the biotechnology debate has rent a tear in the norms of conversational implicature, and that philosophers can play a role that theorists of social movements cannot in repairing them. Hansson probably agrees with me about much of this, by the way, though he comes at it from a rather different philosophical direction, (see Modin and Hansson 2011). However, the considerations just outlined have considerable influence on why I consider myself a pragmatist. It has been more typical to interpret pragmatism somewhat differently. Dane Scott has characterized pragmatism as an attempt to moderate between extreme technological optimism, on the one hand, and extreme pessimism, on the other, (Scott 2018a). While I do not disagree with this attempt, I do not believe one has to be a philosophical pragmatist to endorse it. The conceptual framework of identifying hazards, using science to estimate our exposure to hazards and then mapping how different ethical traditions figure in our efforts to cope with an uncertain future is underutilized in current philosophical work on social and political implications of science and engineering. When the framework is extended to include explicit attempts to craft fair and respectful relationships with affected and interested parties—and that is what risk communication should involve—philosophical questions on the meaning and function of conversing about risks will broaden and thereby strengthen the framework even more. That will, I think, take philosophy of technology into pragmatism.

14.4 Environmental Pragmatism: A Prolegomena I despair at the prospect of formulating a concise statement of pragmatism. Pursuant to the previous section’s discussion of risk, the classical formulation of pragmatism developed by Charles S. Peirce (1839–1914) emphasizes chance and contingency in its metaphysics and epistemology. Pragmatism views such regularity and

14.4 Environmental Pragmatism: A Prolegomena

387

stability as people observe and rely upon as founded upon a universe of indeterminacy, randomness and constant change. As Dewey said, “the world is a scene of risk,” (Dewey 1925, 43). This Heraclitean orientation notwithstanding, the potential for pattern resides implicitly within the random. Repeating or self-replicating pattern is a subclass that, once realized as a random event, obtains, perdures and repeats until disrupted. This is the sense in which Peirce saw physical constants and the regularities observed in the biophysical sciences as evolutionary product of pure chance. Once up and running, the system of interaction endures, at least for a while. Equally significant, doctrines formulated to describe, explain or exploit the order of physicality are symbolic, not physical systems. Atomism, Newtonian and quantum physics are semiotic constructs. They presuppose the emergence and development of complex capabilities for signaling, representation and abstraction. Hence, no one should suppose that the physical sciences get down to the “root reality” (whatever that might be), as the meaning systems they deploy depend upon language and communicative habits that took many centuries to evolve. Peirce did not take this to imply that the sciences were worthless, however. Quite the contrary, his foundational article “The Fixation of Belief,” argues both for the utility of a stable belief system and the advantages of relying upon a community of inquirers dedicated to testing the adequacy of each other’s’ hypotheses, (Peirce 1940). While some pragmatists have shown reluctance to follow Peirce’s path into the metaphysics of pure chance, virtually all pragmatists accept that theorization presupposes a complex system of communicative practice. Several features distinguish pragmatism from the unexceptional observation that the practice of the physical scientist depends upon language. Pragmatism views communicative practices among humans as the product of a long history, much of which remains shrouded from view. The guess, or hypothesis, that orients pragmatism holds that systems of discursive practice emerge out organism-environment interactions. The pragmatists I associate with hold that these systems are not uniquely human. As discussed in Chap. 9, to be environed is not simply to exist at some locus in a space-time matrix. An environment is a surrounding world equipped with affordances matching the organism’s poise. Correlatively, to be poised is to express or embody a reality that obtains stability or pattern with respect to the milieu in which an entity is situated. In this respect, a pragmatist conception of the environment is consistent with the umwelt theorized by Jacob von Uexküll (1864–1944). von Uexküll’s A Foray into the Worlds of Animals and Humans begins with a discussion of the tick, an animal equipped with the restricted perceptual ability to sense the heat being emitted from a warm-blooded creature in its immediate vicinity. In von Uexküll’s account, the tick releases its hold from the branch or leaf on which it resides, perchance to land upon the heat source from which it indulges a blood meal. von Uexküll is not a pragmatist, but his discussion of the tick’s surrounding world accords nicely with the way that Dewey characterizes an organism’s orientation toward its environment. In von Uexküll’s treatment, the surrounding world of the tick is exhausted by the affordance of its limited, heatseeking, sensory capability. It is poised to respond to this (and only this) feature of its spatial location. Other elements that might coexist within the location are not part of the tick’s surrounding world.

388

14 Biotechnology, Controversy and the Philosophy of Technology

While von Uexküll stresses the fixed sensory capacity of the tick, arguing in quasi idealist fashion that the exteriority of the surrounding world is determined by the limited subjectivity of its perceptual apparatus, Dewey sees the organism and its environment as a unified field of mutual interdependences. This allows Dewey to see the activity of theorists (e.g. von Uexküll and himself) as itself meaningful within a context of organism-environment interaction. Scientists and philosophers are, of course, living in surrounding worlds drenched with symbolic significance, and they are poised for many more environmental interactions than a blood meal. Their perception is enriched not only with sensory affordances, but also with language, with history and with the rituals of habitus. Their worlds extend beyond their spatial location. The environment in which they act may orient them with greater responsiveness to dinosaurs or long dead stoics than to other members of their household. Yet like the tick, they are attuned toward some elements of their environment—elements that the psychologists of Dewey’s time would have called stimuli—and oblivious to potential features that would require the development of alternative affordances. The pragmatism developed by Peirce, Dewey and James is deeply environmental in its fundamental ontology. Peirce, James and Dewey are also thoroughgoing fallibilists, while this simply does not come up in von Uexküll’s theory of meaning. On the surface, falliblism just means that they (and all pragmatists) admit that they could be wrong about any element of their worldview. More deeply, it means that pragmatists abandon the quest for certainty. Falliblism is supported by the Heraclitean metaphysics, for change can upend every stability. Yet falliblism functions as a crucial element in pragmatist epistemology apart from Peirce’s views on chance. Programmatically, it puts pragmatists at odds with philosophers of the modern period who sought to wipe the slate clean and build an edifice of certain knowledge from scratch. In the first place, there is no wiping of the slate; a person’s cognitive inheritance is what they have to work with, and the same goes for the collective capacities reflected in interpersonal relationships. Any of these capacities can be challenged, but the ability to challenge depends upon critical competencies that are themselves part of history’s deposit. In the second place, these modernist programs were founded on sham doubts. Doubt is natural when some element in one’s fabric of habituated practice is shattered, when risk intrudes upon the calm flow of conversational implicature. The ability to place an arbitrary element of one’s way of life into genuine doubt is a hard won competence, and one to which the disciplines of philosophy are highly committed. Practically, falliblism means that pragmatists strive to consider any challenge to their worldview thorough reflective and communally discursive thinking. David Hume (1711–1776) had already provided excellent reasons to doubt the putative coherence of the Cartesian subject as experienced by human beings, so why would we extend it to ticks?

14.5 Technological Pragmatism: Ihde

389

14.5 Technological Pragmatism: Ihde Consistent with the preceding, I do not regard technology as a uniquely human achievement, but for conciseness, my discussion in this section is limited to the human condition. Tools and techniques are pervasive features of the environment in which human organisms have found themselves situated since well before the dawn of recorded history. They play a decisive role in shaping the affordances with which people engage their surrounding worlds. Don Ihde provides a schematic framework of characterizing this role in terms of embodiment relations, hermeneutic relations, alterity relations and background relations. Significantly, it is only in alterity relations, when a tool or technique confronts us as an “other” that technology stands out as a feature or thematic element of the environment, as something to which we might be poised. In other modes, tools and techniques recede into habitus, enabling and even encouraging some interactions, while discouraging or concealing the potential for other ways of acting. Ihde credits Heidegger’s discussion of the hammer from Sein und Zeit with inspiring this observation. In use, the hammer disappears from our field of intention, becoming absorbed into the act of hammering. Only when it breaks (or when we cease to hammer) does it stand out as an object before us, present at hand for inspection and theorization through concepts such as weight and form, (Ihde 2010). Embodiment relations and hermeneutic relations characterize the most prominent features of technology in use. The hammer again exemplifies the embodiment relation, as the tool becomes absorbed entirely in the act. The person who hammers is focused on the nail or stake to be driven, the object to be crushed. She is no more aware of the tool itself than she is of her own body, the movement of her arm and shoulder and the subtle shifting of weight. In fact, the tool is absorbed into the body image. Body image, in turn, is the sense that a person has of his or her own body in interacting within an environment. The body image situates the agent within a place, orienting the potentialities of action, the organism’s poise. Ihde had originally characterized these as noetic relations, connoting the incorporation of the tool into the act of perception or knowing, (Ihde 1979). This earlier usage suggests a deep connection to the extended mind hypothesis proposed by Clark and Chalmers (1998). Ihde’s switch to the term embodiment connotes a more visceral sense of incorporation, but even Otto, who relies on his notebook to do the work of his memory, deploys a tool in a manner that absorbs the artifact entirely within the orbit of an otherwise directed intention. Hermeneutic relations, (also called noematic relations) are the correlate. Here the tool is absorbed into the object of one’s attention, rather than the act itself. The object is “read” through the tool. The tool does unnoticed interpretive work in highlighting certain features of the object, while obscuring others, (Ihde 1979, 2009). A hammer communicates a hermeneutic relation to the extent that it highlights certain features of the object being hammered—its resistance to force, its ductility, in the case of nail, its ability to penetrate another object. In early work, Ihde had argued that hermeneutic relations invite the agent to associate amplified features

390

14 Biotechnology, Controversy and the Philosophy of Technology

with the essential reality of the object, and to ignore aspects that tool is less capable of communicating, (Ihde 1979). Background relations name the tools and techniques that operate at the horizon of a person’s experience. The central heating or air conditioning, the lighting, the machinery that permeates the infrastructure of contemporary life, (Ihde 2009). These receive less direct discussion from Ihde, but he is aware that they form a crucial element of the organism-environment interaction for contemporary human beings. What is more, when background tools and techniques fail, they are brought suddenly into the foreground. In the blink of an eye, they assume alterity relations, dominating the focal orientation of the human agent. It is then that tools and techniques stand forth as objects of a person’s attention. In Sein und Zeit, Heidegger described this as the origin of a peculiar ontological stance called presence at hand. Here, some feature of the world is highlighted, “lit up,” so to speak, occupying the center of person’s attention. What is more, it is the workings, functionality and mechanics of the broken tool that emerge as the object of concern. In Ihde’s reading of Heidegger, this becomes the ontological starting point for a worldview that privileges mechanical causes as the essential or “most real” features of being. The scientific attitude becomes a thoroughgoing projection of presence at hand that represses aspects of Dasien (the being characteristic of humans) while elevating favored aspects to become the standard for actual existence and the ultimate criteria for truth. As these tendencies become more and more pervasive in human experience, they come to dominate the human condition so thoroughly that they are taken to be definitive of it. Ihde views these elements of Heidegger’s thought as the residue of a romantic outlook that cannot be philosophically sustained, (Ihde 2010). Ihde’s engagement with the phenomenological tradition in philosophy will be obvious to readers having prior familiarity with the work of Heidegger, Edmund Husserl (1859–1938) and Maurice Merleau-Ponty (1908–1961). Yet Ihde makes a break with a strand of phenomenology and postmodernism that dominated the last quarter of the 20th century, the period in which he was most active. Authors working in this strand would substitute the words subject where I have used the words, person, human being, or agent, and would speak of subjectivity where I use phrases like experience, poise, the human condition or organism-environment interaction. This difference is diagnostic of the distinction between pragmatism, (which unlike some forms of analytic philosophy does take phenomenology quite seriously), and views that philosophize on the foundation of a Cartesian subject, a mental substance individuated in minds, a spirit detached from the body or some other speculative entity reminiscent of a soul that survives after bodily death. Phenomenological postmoderists associated more contingency with subjects than their philosophical forebearers, but often did so in a manner that assigns ontological privilege in much the same way as earlier generations of philosophical idealists. This is the same distinction that separates Dewey’s approach to the organism-environment interaction from that of von Uexküll, who, in my reading, writes as if the surrounding world of various creatures is entirely emergent from features that define the essential nature of their subjectivity. von Uexküll is silent on the question of where their subjectivity itself might have come from. Subjects are either grounding sources of the real, as they seem to be in

14.5 Technological Pragmatism: Ihde

391

some versions of German idealism, or expressions of a solipsistic will to power that subverts the question of truth to contests of identity politics. Ihde aligns explicitly with pragmatism, calling his approach postphenomenology, which he defines simply as phenomenology + pragmatism, (Ihde 2016). His argument draws heavily on the pluralism and anti-foundationalism he finds in Dewey, and on the commitments to the social democratic epistemology that he takes from Richard Rorty (1931–2007). Rorty establishes a pivot between the ironic, solipsistic standpoint of reflective subjectivity, always reflexively oscillating between self-deception and self-enlightenment, and solidarity, a commitment to collaborative furtherance of liberal tolerance and the mutual encouragement of opportunity. For Rorty, the latter rests on little more than the current era’s intellectual inheritance, which includes a substantial capability for collective critique, learning and reform, but as a pragmatist, that’s enough, (Rorty 1998). Besides, what more could one ask? Ihde’s pragmatist commitments emerge significantly in response to his generation’s proclivity toward an anti-science standpoint that he finds unwarranted, (Thompson 2020). This feature of postphenomenology will bring us back to the biotechnology debate.

14.6 Technological Pragmatism: Postphenomenology Expanded My graduate education in philosophy was undertaken under Ihde’s supervision, and I am proud to identify my work as an extension of his general approach. My early work on risk assessment emphasized nuclear power. I argued that risk assessment was itself a technique that should be subjected to the type of critical analysis that Ihde was developing with respect to hermeneutic relations. Independent of its scrutiny by the techniques of scientific risk assessment, risk is a phenomenon observable in discourse. It is in the way people talk about risk, in communicative practice, that the phenomenology of risk can be most successfully dissociated from the scientific risk assessments that, in the 1970s, were coming to dominate the ethics and policy debates over nuclear power. To be sure, people may have significant feelings, and non-humans, too, almost certainly experience the phenomenon of danger. But risk is not simply danger. As discussed above, I came to view it as intimately associated with the problem of action, as the phenomenon that arises when an agent chooses to embark on a dangerous path, or, alternatively, when the background habitus is disrupted and some feature of it (cooking peas?) becomes a focus of attention calling for further intentionality. It is only in further deliberation that the resources of expected-value analysis become meaningful, so identifying the phenomenon of risk with those features highlighted by the techniques of risk-assessment, as Hansson does, not only occludes elements that are integrated into communicative practice, it suggests that there is no difference between action and the mere potentiality of an event. This is a philosophical error, and it is likely to produce ethical mistakes, as well, (Thompson 1980).

392

14 Biotechnology, Controversy and the Philosophy of Technology

I maintained this orientation in my early papers on agricultural biotechnology, arguing that framing issues in terms of risk would introduce subtle forms of ethical bias, (Thompson 1987, 1988). However, I was also starting to appreciate some limitations of Ihde’s model. While Ihde was breaking ground by exposing the systemic ways that technologies influence perception—what a mainstream phenomenologist might call the ontology of the subject—it was simply inadequate as a general philosophy of technology. One problem was Ihde’s lack of engagement with normative questions. I spent much of the 1980s experimenting with linkages to mainstream thrusts in ethical theory. In the 1980s, I was treating risk assessment as a tool. The normative significance was that the tool’s hermeneutic relations obscure the active, intentional structure of risk, (Thompson 1985, 1986). Later, my pragmatism came to the fore, and I read Ihde as expositing the organism-environment nexus. As poised to act, virtue ethicists address a person’s disposition. Better virtue ethicists interpret dispositions contextually and relationally. A fully pragmatist virtue ethics would see dispositions as at best half the story. A person’s normative disposition is unified with the situation’s environmental affordances. Ethics is always a chicken and egg problem, where the agent attempts to query the situation (what the environment affords, combined with his or her own disposition), recognizing that the ideas and experiences one uses to conduct this inquiry may well be the source of fatal error, (see especially, Thompson 2015). We do what we can. I also concluded that Ihde’s quartet of technoscientific relations was going to be of limited helpfulness in analyzing the issues raised by gene technologies. There are postphenomenological questions that still await analysis with respect to technoscientifc models of the genome and its hermeneutical implications for the problematization of situations in both medicine and agriculture. Nevertheless, the phenomenological ontology implied by Ihde’s approach was not the right instrument for analyzing the issues being debated in the debate over bovine growth hormone. Some were issues of potential harm and precaution, topics that a decision-theoretic approach to risk questions was quite prepared to address. Others dealt with forms of relationality that intersect with the institutions for governance and policy. These were questions that had prompted Karl Marx’ work in the philosophy of technology as early as the composition of his 1844 manuscripts. In short, there was already a large body of philosophy on the role of technology in shaping social relationships when Ihde started his work in the 1970s. Ihde has not done much to incorporate this body of theory into postphenomenology, but there is no reason why one cannot do so, (Thompson 2006). My early work on agricultural biotechnology was taking place against the backdrop of the U.S. farm crisis of the 1980s, as discussed in Chap. 8. In any industry, technological innovations refine the efficiency of the production process, prompting a social philosophy debate between those who stress the benefits of competition and lower prices for consumers, and those who stress the exploitation, bankruptcy and unemployment that accompanies such transitions. In agriculture, however, this technology treadmill has provoked what some see as a pervasive social transformation, driving poor but self-reliant agrarians into wage labor and dependence upon the forces of capital for the provision of their livelihood, (Lewontin and Berlan 1986; see also Chap. 8). Schurman and Munro discuss how advocates for smallholders in less

14.6 Technological Pragmatism: Postphenomenology Expanded

393

industrialized agricultural systems saw agricultural research as prompting a treadmill inimical to the interests of their clients. Small farm advocates collaborated with activists motivated by animal protection, the fear of eugenics or religious objections to form the core network that would lay the foundations for a movement against food biotechnology in the first decade of the 21st century, (Schurman and Munro 2010, 51–81). Nothing in Ihde’s version of postphenomenology prepares the ground for this kind of concern, but the risk-based approach inspired by Jonas was capable of incorporating it, if only one takes a pluralistic view of what can count as a hazard. Gene technologies change social and governance technologies in other ways as well. Most basically, they allow an agronomic trait to be separated from the seed—the traditional vehicle for capturing value in agriculture—and treated as a commodity good in and of itself. When combined with patent protection, biotechnology companies had incentives to protect their claim on these novel goods, innovating a radical shift in the social relations between farmers and seed companies, as discussed in Chap. 10. Such a change would have had an obviously political and moral significance if it were done through a change in law or policy. As a technological innovation, it slipped by many social theorists, as if it were an act of nature. The attempt to create genetic use restriction technologies (GURTs), also discussed in Chap. 10, transforms the seed from something that can, in a specific sense, be used over and over again, with farmers replanting seed from the previous year’s crops, into a good that is consumed in use. Farmers must return to the seed company every year. This transformation had already taken place in some crops, like maize, exhibiting hybrid vigor. Hybrid varieties do not “breed true” and saved seed will be so much less productive that farmers in industrialized economies had become accustomed to buying seed every year, well before biotechnology appeared on the scene, (Fitzgerald 1990). Again, first generation postphenomenology did not incorporate these dimensions of sociotechnical relationality. If the first edition of Food Biotechnology in Ethical Perspective had been written with an audience of philosophers in mind, the risk-based approach might have been framed as a way to build upon the philosophical approach that Ihde was developing. As it happened, philosophers working in phenomenology during the 1990s could not have been less interested in agricultural technology. Along with closely allied feminist approaches, these scholars were working from a bias that Lisa Heldke has analyzed. Turning a feminist critique on its head, Heldke shows how elites (in this case, academic philosophers) fail to recognize the knowledge and identity of farmers, especially women farmers. The failure typifies elite’s disrespect of people from socially marginalized groups. This bias against farmers extends to anyone, including researchers, who take agriculture seriously. Food production is not a subject worthy of the elite intellectual’s attention. Writing in 2006, the bias was as evident among feminists as anywhere else, (Heldke 2006). Heldke and I both affiliate with pragmatism, perhaps because in our formative years, they were the only philosophers willing to view our work as a contribution to philosophy.

394

14 Biotechnology, Controversy and the Philosophy of Technology

14.7 Gene Technology as a Moral Apocalypse One last topic deserves consideration before closing. Heidegger’s Question Concerning Technology articulates a moral concern with technoscience. Although Ihde sees this period in Heidegger’s thought as a reversion to romanticism, Ihde’s own ontology helps us articulate Heidegger’s concern in moral terms. As language enabled organisms embedded within a technically dense environment, human beings risk losing their humanity as the hermeneutic effects of technoscientific practice incline steadily toward reductive instrumentality. I will unpack the substantive elements in this philosophically dense risk presently, but first I should stress that I do not see it as a version of the claims made by Francis Fukuyama in Our Posthuman Future: Consequences of the Biotechnology Revolution, (Fukuyama 2002), nor do I think that Heidegger’s fears track closely to Fukuyama’s. When I talk of human beings losing their humanity, I do not mean to engage the philosophical debate over genetically engineered enhancements of intelligence or perceptive ability. With Heidegger, I am talking about a possibility that could have been fully realized given the genetic and phenotypic endowment typical of human beings throughout recorded history. As Ihde says, techniques establish hermeneutic relations that amplify some elements in human experience, while dampening others. Presence-at-hand emerges in human experience when the tool breaks. This alterity relation invites engagement with the tool’s workings as a focal project of human activity. As this very project assumes the status of a habituated context for action, technology as a mode of seeing and revealing takes on systematic hermeneutic characteristics. In natural science, the surrounding world, the environment of the human being, becomes the object of knowledge production, but this object is increasingly identified in terms of those features amplified by the project of knowledge production itself: the stable, enduring, predictable and dead properties that can be known in advance of our encounter with things. As the project increases in power, a physical scientific notion of being as spatio-temporal location drives all the other ways in which being is said from the field. If this were not a moral problem in itself (and I think that it is), the fact that human beings, we, ourselves, come to experience ourselves as in the environment much as a marble is in a box allows the reductive gaze of this technoscientific project to be turned upon the human condition, upon Dasein. Each individual risks becoming lost within this project, of becoming unable to recover openness toward life, but because humans are social beings, the situated relationality of the human condition is placed at risk, as well. For pragmatists, falliblism is intended to guard against this possibility, but a thoroughgoing pragmatist will admit that even falliblism is fallible. The risk is real. While Jonas, Ihde and I are willing to face this risk by continuous monitoring of our technical practice, we cannot claim certainty that the moral collapse into pure instrumentality will be avoided. What is more, the moral collapse is itself clothed in further risks, the risks of genocide, of environmental devastation, of nuclear annihilation. The risks noted by Fukuyama could be added to this list. A world where the humanity of human beings has been erased is a world where anything is possible. One response—perhaps it is a precautionary response—advocates abandonment of

14.7 Gene Technology as a Moral Apocalypse

395

the technoscientific project in all of its dimensions. This abandonment might be releasement from spell of advancing technoscientific efficiency. It might take the form of retreat into a worldview of the past, a re-enchantment of nature or an embrace with cultural forms marginalized by the enlightenment project, the strategy that Ihde ties to romanticism. However, it might also take the form of active resistance to technoscience. One finds signs of this resistance everywhere among people who have been awakened to these existential risks and to their potential linkage to exterminist practices of violence and oppression, (see, for example, Trzak 2020, 76–79). Genetic engineering shoulders the sad legacy of the eugenic sterilization of disabled and racially stigmatized women in the construction of narratives of resistance to the project of technoscientific knowledge production. Racial, gendered, humanist and ableist genocides are real possibilities, gene technologies become fair targets for extreme acts of resistance. Schurman and Munro (2010) argue that these existential risks did motivate at least some of agrifood biotechnology’s early opponents. Yet if a pragmatist such as myself agrees that these extreme outcomes are possible, that is not to say that they are likely. Nor does it imply that we should mobilize all of our philosophical resources in a death-battle against the very idea of technology. Better approaches are available in the exposure quantification and risk management phases of the risk-based approach. Dewey was arguably one of the first critics of the way that Enlightenment philosophers deployed standards of rationality in oppressive ways, (Seigfried 2002), yet Dewey was not one to adopt an essentializing conception of instrumentality. It is ironic that the philosophical movement (feminism) that has done more to a upend philosophy’s tendency to elevate contingent elements of a phenomenon or concept into essential features should adopt an essentializing attitude toward instrumentality. For example, an article that discusses legal definitions for service animals from a feminist perspective, Kelly Oliver writes, “Obviously, defining service animals as equipment reduces them to disposable commodities that exist for our benefit,” (Oliver 2020, 115). Really? This is obvious ? Don’t service animals serve as both equipment and significant others? Indeed, don’t all life partners equip one another to meet their situatedness? Why would recognizing these functions reduce a dog or a human being to a commodity, much less a disposable one whose existence is exhausted by its benefit to another being? Most significantly, how can Oliver classify any of these (false) presumptive suppositions as obvious? Beyond the real possibility that this is nothing more than a rhetorical flourish for Oliver, my guess would be that those who insist upon resistance to technoscience as the only valid standpoint detect an unsettling stench amidst the plethora of gadgets and gimcrackery that keep citizens of industrial democracies distracted from more important matters. I can understand this, but I can’t endorse it. Indeed, there may even be a sense in which the crisis of reification has already occurred. Noelle McAfee combines elements of psychoanalytic theory with pragmatism to argue that a fear of crisis creates a psychosis within the collective consciousness of late capitalism. The repressed memory of past trauma triggers a reflexive response that creates not only fear, but also reinforces the social institutions and forms of relationality that reproduce this response. For McAffee, citizens of liberal democracies could acknowledge that they have projected an apocalyptic crisis that

396

14 Biotechnology, Controversy and the Philosophy of Technology

has, in fact, already happened into the future. It is their act of projecting the destabilizing and dehumanizing experience of crisis into the world that is yet to come that reproduces a fear response, in turn triggering the continuous reconstruction of defensively motivated repression. Recognizing this would be the first step in a process of reconciliation and reform, (McAfee 2019). McAfee is recommending a hopeful kind of pragmatism in which acknowledging the traumas that have already happened through technoscientific occlusion of the human condition would form the basis for a new project of reconciliation and community reconstruction. I would hope that a risk-based approach could be part of McAfee’s project.

14.8 Conclusion: Technological Pragmatism and World-Feeding Ideology My book The Spirit of the Soil argues that by the end of the 20th century, the ideology of productionism dominated the agricultural sciences. Briefly, productionism is a philosophy of agriculture placing singular value on increasing the output of the global capacity for food and fiber production. Productionism may have made sense when Abraham Lincoln campaigned on a promise to use the resources of government to “push the soil up to its full potential,” (Lincoln 1859). Perhaps 80% of the U.S. population were farmers then, just as in some African countries today. In addition (and just as in Africa today) the majority of them were living in grinding poverty. For Lincoln, increasing farm production was as much or more about helping farmers as it was producing enough for people to eat. Lincoln’s century saw the fundamental innovations in mechanizing farm work, and the inauguration of commercially successful plant breeding. Chemical technologies would be added to this mix after defense appropriations during World War I expanded the capacity for producing nitrogen fertilizers through the Haber-Bosch process. Research on pesticides was subsidized as a defense expenditure in World War II. By 1950 farmers who had access to this full suite of technologies found themselves able to produce more than the world needs. The problem was finding markets for that surplus. An exodus from farming ensued and by 1975, less than 5% the industrialized world’s population were farmers. Yet the early decades of the 20th century saw the emergence of a counterforce to productionism. Dr. Harvey Wiley (1844–1930) campaigned against adulteration of foods in the manufacturing process. His work led to the creation of the U.S. Food and Drug Administration (FDA) in 1906. Wiley’s FDA legitimated the practice of government regulation to protect the public from hazards introduced by technological innovations in the food system, (Stirling 2002). Yet FDA’s work in food safety had little impact in farm country and the drive to push the soil’s potential was not satiated. Population ecologists predicted global famines in the 1960s, and the agricultural sciences responded with the Green Revolution, a productionist strategy to defeat the Red Revolution by bringing farmers in less developed regions into the capitalist economy. The agricultural research establishment was mostly unfazed by

14.8 Conclusion: Technological Pragmatism and World-Feeding Ideology

397

the publication of Rachel Carson’s Silent Spring in 1962. The U.S. Federal Insecticide, Fungicide and Rodenticide Act (FIFRA) was passed initially in 1947, and underwent significant revisions that brought it under the control of the U.S. Environmental Protection Agency in 1972. It would be another decade or more before very many leaders in farm country or agricultural research organizations would question whether environmental considerations should constrain productionism. This summary characterization of productionism would need qualification before it could be said to describe the mindset of agricultural scientists in the second decade of the 21st century. Nevertheless, I have sat through countless presentations by agricultural scientists who begin with projections of world food needs looking 50 years into the future, and then arguing that their science—often a gene technology—is absolutely essential in order to save the day. The World Feeders Fallacy introduced in Chaps. 2 and 9 offers a more extended analysis of gene technology’s salvific potential among the people most vulnerable to hunger. The point to stress is that a rapid shift—and by rapid I mean within a single generation (or about 30 years)—to a full scale, industrial style agriculture with full deployment of mechanical, chemical and genetic inputs could be disastrous for many poor farmers, even if it did push aggregate food production up the quantities needed to feed everyone. This is not an uncontroversial claim. It was not the view expressed by Norman Borlaug nor Robert Paarlberg, to mention two individuals I have known and respected (see Borlaug 2001; Paarlberg 2009). Unpacking the philosophical and empirical dimensions of it would be a large task. Yet two Wall Street Journal reporters who followed Borlaug’s trail into Africa and support the promise of the Green Revolution document the institutional gap between places like India or Mexico, where Green Revolution varieties could be rapidly adopted, and remaining areas where the majority of the world’s poor are still farmers, (Thurow and Kilman 2009). David Rieff, another journalist, documents how globally adequate amounts of food production fail to alleviate the hunger of urban and rural poor alike, (Rieff 2015). Unlike Vandana Shiva, these journalists have no intrinsic complaint with gene technology, yet they provide a powerful argument for being skeptical about whether increasing the Earth’s capacity for production will actually help the poor. Dane Scott has subjected the technological optimism of scientists who promote their own research through a world-feeding ideology to a deep and powerful critique from the perspective of the philosophy of technology. He shows how campaigns both for and against GMOs have drawn upon comprehensive ideological visions of technology’s potential for good, as well as its proclivity for evil, (Scott 2018b). In many respects, his book should be read before this one, for the measured attempt to shore-up weaknesses in the risk assessment framework, and to understand its limits, demands a reader liberated from utopian and dystopian visions of technology, alike. Scott concludes that techno-optimism occludes the ability of many technically competent scientists when it comes to agrifood biotechnology, causing them to overlook ways in which the most likely implementation of their dream vision could (at the least), fail to help the people they identify as its beneficiaries, and could potentially do them harm. Ihde’s postphenomenology is the philosophical theory that explains this particular form of concept blindness.

398

14 Biotechnology, Controversy and the Philosophy of Technology

At the same time, Scott questions whether some of biotechnology’s sharpest critics know enough about farming or farmers to evaluate the promise of new seeds within the production systems currently in use. Scott’s book, better than mine, locates the right tone for criticizing the biotechnology establishment for their failure to be adequately inclusive in anticipating the unwanted consequences of their innovations. Having a rough prototype of the risk-based approach will allow a more productive conversation between those who know something about agriculture (including, of course, farmers themselves) and people who regard agriculture as something of a sideshow. For the apocalyptic prophets, agriculture is a worthy case study, but one that can be sacrificed in the more important battle over the soul of humanity. Those of us who follow Ihde and Merleau-Ponty take bodies seriously, and we evaluate the trade-off between eating and soul saving differently. Yet with this observation, the philosopher of technology’s potential contribution to genetic technoscience comes full circle, and the analysis offered in the successive editions of this book can be properly located.

References Achterhuis, H. 2001. Introduction: American philosophy of technology. In American Philosophy of Technology: The Empirical Turn, ed. H. Achterhuis, 1–10. Bloomington: Indiana University Press. Ankeny, R., and H.A. Bray. 2018. Genetically modified food. In The Oxford Handbook of Food Ethics, ed. A. Barnhill, M. Budolfson, and T. Doggett, 95–111. New York: Oxford University Press. Bernstein, P., and L. . 1998. Against the Odds: The Remarkable Story of Risk. New York: John Wiley and Sons. Borlaug, N. 2001. Ending world hunger. The promise of biotechnology and the threat of antiscience zealotry. Plant Physiology 124: 487–490. Bruell, C. 1984. Strauss on Xenophon’s Socrates. The Political Science Reviewer 14: 263–315. Clark, A., and D. Chalmers. 1998. The extended mind. Analysis 58: 7–19. Dewey, J. 1925[1981]. John Dewey The Later Works, 1925–1953, V. 1: Experience and Nature. Carbondale: Southern Illinois University Press. Durant, J., M. Bauer, and G. Gaskel (eds.). 2002. Biotechnology: The Making of a Global Controversy. New York: Oxford University Press. Fehr, C., and K.S. Plaisance. 2010. Socially relevant philosophy of science: An introduction. Synthese 177: 301–316. Fitzgerald, D. 1990. The Business of Breeding: Hybrid Corn in Illinois, 1890–1940. Ithaca, NY: Cornell University Press. Foucault, M. (1985) [1990] The Uses of Pleasure: The History of Sexuality Vol. 2. R. Hurley, (tr). New York: Vintage Books. Fukuyama, F. 2002. Our Posthuman Future: Consequences of the Biotechnology Revolution. New York: Farrar, Straus and Giroux. Hacking, I. 2006. The Emergence of Probability, 2nd ed. New York: Cambridge University Press. Hacking, I. 1990. The Taming of Chance. New York: Cambridge University Press. Hansson, S.O. 2007. Social decisions about risk and risk-taking. Social Choice and Welfare 29: 649–663. Hansson, S.O. 2013. The Ethics of Risk: Ethical Analysis in an Uncertain World. Dordrecht, NL: Springer. Hansson, S. O. 2014. Agricultural Biotechnology for health and the environment. In Biotechnology and Biodiversity. Sustainable Development and Biodiversity, ed. M.Ahuja and K. Ramawat, vol 4, 67–76, Dordrecht, NL: Springer.

References

399

Hansson, S.O., and K. Joelsson. 2013. Crop biotechnology for the environment? Journal of Agricultural and Environmental Ethics 26: 759–770. Heldke, L. 2006. Farming made her stupid. Hypatia 21: 151–165. Hickman, L.A. 1990. John Dewey’s Pragmatic Technology. Bloomington: Indiana University Press. Hickman, L.A. 2001. Philosophical Tools for a Technological Culture: Putting Pragmatism to Work. Bloomington: Indiana University Press. Hicks, D., and R.L. Millstein. 2016. GMOs: Non-health issues. In Encyclopedia of Food and Agricultural Ethics, eds. P.B. Thompson and D.M. Kaplan. New York: Springer. Ihde, D. 1979. Technics and Praxis: A Philosophy of Technology. Dordrecht, NL: D. Reidel. Ihde, D. 2009. Postphenomenology and Technoscience: The Peking University Lectures. Albany, NY: The SUNY Press. Ihde, D. 2010. Heidegger’s Technologies: Postphenomenological Perspectives. New York: Fordham University Press. Ihde, D. 2016. Husserl’s Missing Technologies. New York: Fordham University Press. James, W. 1904. The pragmatic method. The Journal of Philosophy, Psychology and Scientific Methods 1: 673–687. Jonas, H. 1984. The Imperative of Responsibility: The Search for Ethics in a Technological Age. Chicago: U Chicago Press. Kaplan, D. 2012. What’s wrong with techno food? In The Good Life in a Technological Age, ed. P. Brey, A. Briggle, and E. Spence, 241–250. New York: Routledge. Knight, F.H. 1921. Risk, Uncertainty and Profit. Boston: Houghton Mifflin. Lewontin, R.C., and J.-P. Berlan. 1986. Technology, research and the penetration of capital: The case of U.S. agriculture. Monthly Review 38 (July-August): 21+. Lex, M. 1995. Public acceptability of biotechnology, In Issues in Agricultural Bioethics, eds. T.B. Mepham, G.A. Tucker, and J. Wiseman, 387–401. Nottingham UK: University of Nottingham Press. Lincoln, A. 1859 [2018]. Address before the Wisconsin state agricultural society, Abraham Lincoln Online. https://www.abrahamlincolnonline.org/lincoln/speeches/fair.htm. Accessed 29 July 2019. McAfee, N. 2019. Fear of Breakdown: Politics and Psychoanalysis. New York: Columbia University Press. Mitcham, C. 1994. Thinking through Technology: The Path between Engineering and Philosophy. Chicago: University of Chicago Press. Mitchell, T.N. 2015. Democracy’s Beginning: The Athenian Story. New Haven, CT: Yale University Press. Modin, P.G., and S.O. Hansson. 2011. Moral and instrumental norms in food risk communication. Journal of Business Ethics 101: 313–324. Oliver, K. 2020. Service dogs: Between animal studies and disability studies. In Disability and Animality: Crip Perspectives in Critical Animal Studies, ed. S. Jenkins, K.S. Montford, and C. Taylor, 111–128. New York: Routledge. Paarlberg, R. 2009. Starved for Science: How Biotechnology Is Being Kept Out of Africa. Cambridge, MA: Harvard University Press. Peirce, C. S. 1940. The Philosophy of Peirce: Selected Writings. In J. Buchler, ed. Harcourt. New York: Brace and Company. Rasmussenn, N.C. 1975. Reactor safety study: Assessment of accident risks in US commercial nuclear power plants, WASH-1400 (NUREG-75/014). Washington, DC: US Nuclear Regulatory Commission. Rechard, R.P. (1999). Historical relationship between performance assessment for radioactive waste disposal and other types of risk assessment. Risk Analysis 19(5): 763–807. Reiss M.J., and R. Straughan. 1996. Improving Nature? The Science and Ethics of Genetic Engineering. Cambridge, UK: Cambridge U Press. Rieff, D. 2015. The Reproach of Hunger: Food, Justice and Money in the Twenty-First Century. New York: Simon and Schuster. Rorty, R. 1998. Achieving Our Country: Leftist Thought in Twentieth-Century America. Cambridge, MA: Harvard University Press.

400

14 Biotechnology, Controversy and the Philosophy of Technology

Scott, D. 2018a. Technological pragmatism: Navigating the ethical tensions created by agricultural biotechnology. In Ethical Tensions from New Technology: The Case of Agricultural Biotechnology, ed. H.S. James, Jr, 112–125. Wallingford, Oxfordshire, UK: CABI International. Scott, D. 2018b. Food, Genetic Engineering and Philosophy of Technology: Magic Bullets, Technological Fixes and Responsibilities to the Future. New York: Springer. Schurman, R., and W.A. Munro. 2003. Making biotech history: Social resistance to agricultural biotechnology and the future of the biotechnology industry. In R. Schurman and D.D. TakahashiKelso (eds) Engineering Trouble: Biotechnology and Its Discontents, 111–129. Berkeley: University of California Press. Schurman, R., and W.A. Munro. 2010. Fighting for the Future of Food: Activists Versus Agribusiness in the Struggle Over Biotechnology. Minneapolis: University of Minnesota Press. Seigfried, C.H. 2002. John Dewey’s pragmatist feminism. In Feminist Interpretations of John Dewey, ed. C.H. Seigfried, 47–77. University Park PA: Penn State University Press. Stikkers, K.W. 2009. Persons and power: Max Scheler and Michel Foucault on the spiritualization of power. The Pluralist 4: 51–59. Stirling, D.A. 2002. Harvey W. Wiley. Toxicological Sciences 67: 157–158. Suppe, F. 1987. The limited applicability of agricultural research. Agriculture and Human Values 4: 4–14. Thompson, P.B. 1980. The Concept of Risk, a dissertation submitted in partial fulfillment of the requirements for the Ph.D., Department of Philosophy. Stony Brook NY: State University of New York at Stony Brook. Thompson, P.B. 1985. Risking or being willing: Hamlet and the DC-10. the Journal of Value Inquiry 19: 301–310. Thompson, P.B. 1986. Uncertainty arguments in environmental issues. Environmental Ethics 8: 59–75. Thompson, P.B. 1987. Agricultural biotechnology and the rhetoric of risk: Some conceptual issues. The Environmental Professional 9: 316–326. Thompson, P.B. 1988. Agriculture, biotechnology, and the political evaluation of risk. Policy Studies Journal 17: 97–108. Thompson, P.B. 1999. The ethics of truth-telling and the problem of risk. Science and Engineering Ethics 5: 489–511. Thompson, P.B. 2006. Ihde and Technological Ethics. In Postphenomenology: A Critical Companion to Ihde, ed. E. Selinger, 109–116. Albany: State University of New York Press. Thompson, P.B. 2008. Nano and bio: How are they alike? How are they different? In What Can Nanotechnology Learn from Biotechnology: Social and Ethical Lessons from the Controversy over Agrifood Biotechnology and GMOs, ed. K. David and P.B. Thompson, 125–155. Burlington MA: Academic Press. Thompson, P.B. 2015. From Field to Fork: Food Ethics for Everyone. New York: Oxford University Press. Thompson, P.B. 2017. The Spirit of the Soil: Agriculture and Environmental Ethics, 2nd ed. New York: Routledge. Thompson, P.B. 2020. Ihde’s pragmatism. In Reimaging Philosophy and Technology Reinventing Ihde, ed. G. Miller and A. Shew, 43–62. New York: Springer. Trzak, A. 2020. Disability and the ahuman: A story about a dog, a duck, and the woman who cared for them. In Disability and Animality: Crip Perspectives in Critical Animal Studies, eds. S. Jenkins, K. Montford, and C. Taylor, 75–91, New York: Routledge. Thurow, R., and S. Kilman. 2009. Enough: Why the World’s Poorest Starve in an Age of Plenty. New York: PublicAffairs. Tuana, N. 2010. Leading with ethics, aiming for policy: New opportunities for philosophy of science. Synthese 177: 471–492. Zimdahl, R.L. 2003. The mission of land grant colleges of agriculture. American Journal of Alternative Agriculture 18: 103–115.

Index

A Agrarianism, 185, 215 Agricultural development, 227, 229, 244, 246, 255, 369 Agricultural innovation, 233, 361, 370 Agricultural research, 46, 60, 109, 187, 198– 200, 202, 224–229, 234, 238, 245, 274, 352, 363, 375, 376, 378, 393, 396, 397 Agriculture, 1, 2, 7, 13, 17–19, 26, 32, 34, 35, 40, 47, 48, 53, 66, 67, 69–72, 81, 82, 87, 98, 111, 121, 146, 147, 151, 161, 167–170, 172, 176, 182, 183, 186– 188, 196, 197, 199, 203, 205, 211, 213, 223, 224, 226, 228, 229, 231, 232, 234, 235, 237, 238, 241, 257, 282, 283, 289, 299, 305, 307, 324, 343–345, 347, 351–354, 363, 365, 366, 368–370, 375, 376, 378–381, 392, 393, 397, 398 Animal breeding, 1, 6, 8, 9, 14, 41, 42, 84, 85, 102, 334, 349 Animal ethics, 120, 357 Animal health and welfare, 109, 116, 172 Animal integrity, 117, 125 Animal natures, 125, 129, 132 Animal welfare, 2, 9, 58, 60, 63–65, 68–71, 74, 80, 90, 113–117, 120–123, 132, 141, 201, 236, 287, 307, 335, 336, 356, 357 Anthropocentrism, 148, 175, 177

B Bioethics, 27, 257, 280, 287, 289, 290, 296, 304, 377, 381

Biotechnology, 1, 2, 4–21, 25, 26, 28–38, 40–49, 53–61, 64, 67, 69–71, 74, 79–91, 97–103, 109–114, 116, 117, 119, 121–123, 125–128, 131, 132, 137–139, 142–146, 148–161, 168– 172, 174, 176, 177, 181–188, 193– 200, 203, 206, 207, 209, 210, 212– 217, 223, 227–236, 238–240, 242– 246, 251–261, 263, 266–280, 282, 283, 287–292, 294, 297–302, 304– 309, 313–317, 319, 321–325, 329, 330, 332–336, 338, 344, 346, 347, 350–355, 358, 362, 363, 368, 369, 376–378, 380, 382, 383, 386, 391– 393, 395, 397, 398 GMO, 6

C Charles Darwin, 3 Charles Mann, 8 CRISPrCas9, 308, 343, 359

D Deficit model, the, 160, 314–317, 323, 335, 339, 383 Discourse ethics, 175 Distributive justice, 47, 58, 71 Don Ihde, 375, 376, 389

E Eco-feminism, 185, 186 Environmental impact, 2, 17, 31, 32, 60, 62, 66, 67, 71, 73, 74, 82, 137–141, 150,

© Springer Nature Switzerland AG 2020 P. B. Thompson, Food and Agricultural Biotechnology in Ethical Perspective, The International Library of Environmental, Agricultural and Food Ethics 32, https://doi.org/10.1007/978-3-030-61214-6

401

402 151, 160, 161, 168, 169, 175, 179, 188, 194, 212, 228, 236, 259, 307, 333, 334, 336, 350, 351 Environmental risk, 1, 18, 31, 42, 67, 70, 89, 126, 137–143, 146, 150–152, 154, 155, 159, 161, 167, 168, 172, 174, 176, 187, 201, 204, 212, 258, 306, 360 Environmental values, 168, 171, 175, 177, 187, 207 Environmental virtues, 186 Ethics, 2, 17–19, 25–27, 31, 34, 38, 48, 49, 53, 54, 56, 57, 59, 61, 64, 66, 74, 75, 79–81, 83, 86, 87, 96, 100, 101, 109, 110, 113, 114, 122, 124, 126, 137, 139, 141, 142, 144, 147, 156, 160, 161, 167, 168, 170–172, 174– 182, 185, 186, 196, 200, 201, 206, 207, 215–217, 224, 227, 228, 234, 236, 239, 243, 252, 256, 257, 264, 270, 275, 281, 283, 288, 291, 292, 294–297, 304, 313, 314, 321, 322, 324, 325, 329, 330, 332, 333, 336, 343, 349, 351, 354–356, 365, 366, 376, 377, 381, 391, 392 Expected-value, 139, 140, 149–154, 156, 160, 161, 167, 168, 173–176, 187, 201, 216, 325, 328, 331, 333, 381, 382, 391

F Fairness, 19, 66, 85, 100, 151, 201, 212, 279 Feminism, 185, 201, 208, 209, 395 Food justice, 208, 217 Food safety, 19, 30–32, 35, 42, 53, 55, 60– 65, 69–71, 73–75, 79–86, 88, 90–93, 95–98, 102, 103, 113, 141, 150–152, 161, 174, 201, 298, 313, 325, 328, 329, 333–337, 344–346, 349–351, 366, 396 Food safety regulation, 81, 92, 348 Food security, 208, 237 Food sovereignty, 234, 236–238, 245 Francis Crick, 4

G Gary Comstock, 18 Gene drives, 111, 161, 290, 307, 343, 353, 354, 364–369 Gene editing, 11, 12, 14, 16, 83, 100, 111, 113, 116, 121, 169, 170, 274, 290,

Index 307, 308, 343–349, 351–354, 358– 360, 363–365, 368–370 Genetic engineering, 1, 2, 10–13, 38, 40, 42, 54, 57–61, 64, 84–87, 100, 109–113, 116, 117, 119, 121, 122, 124, 125, 129, 131, 139, 142–144, 148, 149, 154, 155, 159, 171, 184, 186, 187, 199, 200, 212, 227, 232, 270, 271, 278, 287, 290–292, 294–307, 309, 313, 321–323, 348, 395 Genetics, 1–13, 15, 18, 20, 65, 68, 109– 113, 115–119, 121, 122, 124, 126, 127, 129, 130, 138, 139, 144, 147, 171, 172, 184–186, 188, 242, 252– 255, 257–260, 262, 269–273, 277– 280, 282, 288, 289, 292, 294–302, 304–306, 308, 317, 329, 343, 345– 347, 349, 351, 352, 355–357, 359, 361, 364, 369, 370, 393, 397, 398 animal breeding, 7 mutation breeding, 8 plant breeding, 8 Gene transfer, 8–12, 16, 79, 81, 85, 99, 103, 111, 113, 121, 122, 157, 170, 233, 239, 244, 253, 273, 288, 289, 300, 346 Gilles-Éric Séralini, 16 Global food needs, 19, 223 Glyphosate, 16 Grammar of risk, the, 327, 328, 330, 334 Green revolution, 45, 46, 184, 224, 225, 227, 229, 230, 233, 238, 244, 246, 301, 396, 397 Gregor Mendel, 7 Gregory Pence, 18

H Hugh Lacey, 18

I Insect allies, 360–363 Intrinsic objections, 291 Intrinsic value, 147, 148, 168, 170–172, 176–179, 181, 291

J James Watson, 4

L Libertarianism, 27, 263, 282

Index Logical fallacies, 40 Luther Burbank, 7

M Mandatory and voluntary labeling, 101 Marxism, 201 Maurice Wilkins, 4 Michael Reiss, 18 M.S. Swaminathan, 8

N Natural law, 262, 264–273, 278

403 Risk, 10, 11, 15, 16, 18–21, 25, 26, 28–33, 35–44, 48, 53, 54, 56–58, 60–69, 72– 75, 79, 80, 82–84, 86–97, 99, 102, 103, 112, 115, 116, 119, 129, 131, 137, 138, 140–146, 148–161, 167, 172, 174–176, 183, 184, 187, 194, 195, 204, 209, 216, 223, 245, 251, 253, 256, 258, 259, 261, 278, 283, 287–291, 295, 298–300, 307, 313– 316, 319, 322–336, 338, 343–345, 348–351, 360, 365–368, 370, 375, 376, 381–388, 391–396, 398 Risk assessment, 20–22, 25, 26, 29, 30, 32, 42, 44, 61, 63, 65, 74, 84, 86–88, 90, 112, 121, 137, 139–143, 146, 148– 154, 156, 157, 160, 161, 170, 172, 174–176, 188, 195, 244, 245, 256, 295, 324, 331, 344, 348, 360, 361, 366, 376, 381–383, 385, 391, 392, 397 Risk communication, 30, 32, 60, 137, 156, 159, 314, 381, 383, 386 Roger Straughan, 18 Rosalind Franklin, 4 R. Paul Thompson, 19

P Patents, 54, 223, 244, 252, 254–261, 263, 266–269, 271, 273–277, 279, 280, 282, 283, 298, 302, 309, 393 Philosophy of agriculture, 379, 380, 396 Philosophy of biology, ix Philosophy of technology, 25, 34, 149, 157, 194, 278, 356, 375, 376, 378–381, 386, 392, 397 Plant breeding, 10, 85, 255, 396 Platform technology, 13 Playing God, 294, 300, 307 Political theory, 81, 87, 201, 264, 334 Postphenomenology, 375, 391–393, 397 Principle of responsibility, the, 25–28, 36, 48, 194 Property rights, 20, 204, 206, 251–253, 255, 256, 258, 260–264, 266–279, 281– 283, 288, 301, 302, 304, 306, 309, 313, 321, 335 Public acceptance of biotechnology, 48 Public attitudes, 244, 296, 314, 335, 347 Public understanding of science, 319

S Sanctity of life, 298 Social impact, 56, 61, 67, 71, 73, 97, 193, 195, 196, 210, 213, 216, 232, 255, 350, 369 Social justice, 39, 193, 195, 196, 201, 203, 207, 209, 216, 307 Socially relevant philosophy of science, 377 Structural injustice, 193, 201, 207, 208, 210, 211, 217 Synthetic biology, 7, 12–14, 127, 161, 172, 290, 307, 343–345, 360

R Raphaël Dubois, 14 Rights, 2, 33, 61, 63, 65, 69, 70, 94–101, 103, 109, 112, 115, 120–125, 127, 141, 145, 148, 157, 159, 173–175, 177, 180, 182, 183, 185, 193, 195, 196, 201–207, 209–215, 228, 230, 231, 239, 245, 252, 254–256, 258, 260, 262–265, 267–280, 282, 289– 291, 293, 295, 297, 305–308, 316, 317, 323, 331, 344, 346, 351, 357, 392, 398

T Techne, 379 Technological ethics, 19, 25, 26, 29, 37, 42, 54, 69, 88, 160, 261, 290, 292, 294– 296, 298–302, 304–307, 313, 343, 344 Telos, 64, 65, 119, 122–128, 130–132, 147, 355 Terminator, 251, 257–259, 261, 271, 277, 278, 283 Transgenic animals, 65, 126, 131, 270, 274, 302, 306

404 U Uncertainty, 29, 60, 71, 73, 74, 86, 89, 90, 96, 137, 149–153, 155–158, 196, 251, 274, 275, 324, 334, 348, 353, 381, 384, 385 Utilitarianism, 27, 96, 123, 130, 141, 161, 173, 174, 177, 193, 195, 201, 207, 215, 226, 241, 242, 273, 295, 307

Index Utilitarian optimization, 96

W Waclaw Szybalski, 12 World hunger, 38, 44, 45, 47