Are There Limits to Science? [1 ed.] 9781527500419, 9781443895811

This book is the result of the 2016 conference of the UK’s Science and Religion Forum which brings together leading scie

144 18 730KB

English Pages 201 Year 2017

Report DMCA / Copyright

DOWNLOAD FILE

Polecaj historie

Are There Limits to Science? [1 ed.]
 9781527500419, 9781443895811

Citation preview

Are There Limits to Science?

Are There Limits to Science? Edited by

Gillian Straine

Are There Limits to Science? Series: Conversations in Science and Religion Edited by Gillian Straine This book first published 2017 Cambridge Scholars Publishing Lady Stephenson Library, Newcastle upon Tyne, NE6 2PA, UK British Library Cataloguing in Publication Data A catalogue record for this book is available from the British Library Copyright © 2017 by Gillian Straine and contributors All rights for this book reserved. No part of this book may be reproduced, stored in a retrieval system, or transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise, without the prior permission of the copyright owner. ISBN (10): 1-4438-9581-4 ISBN (13): 978-1-4438-9581-1

Dedicated to the Memory of Dr. Kenneth Wilson (1937-2017) Former Chair of this Forum.

CONTENTS

The Science and Religion Forum ................................................................ x Acknowledgments ..................................................................................... xii Contributors .............................................................................................. xiii Introduction ................................................................................................. 1 Part One Chapter One ................................................................................................ 8 God, Naturalism, and the Limits of Science Fiona Ellis Chapter Two .............................................................................................. 21 From the Limits of Science to the Limits of Metaphysics: A Philosophical Evaluation of Fiona Ellis’ Naturalistic Theology Mikael Leidenhag Chapter Three ............................................................................................ 40 An Elephant in the Room: Why the Causal Joint is Still Worth Talking About Sarah Lane Ritchie Chapter Four .............................................................................................. 61 The Wild Experiment: Emotion, Reason, and the Limits of Science Donovan O. Schaefer Chapter Five .............................................................................................. 80 What Should Christian Theology (Not) Learn from Science? The Case of the Human Brain Neil Messer

viii

Contents

Part Two Chapter Six .............................................................................................. 102 Physics: The Endless Frontier Nathan Aviezer Chapter Seven.......................................................................................... 113 The Case of Thales’ Ox: An Historical Limit to Science (and its Relation to Religion) Michael Fuller Chapter Eight ........................................................................................... 121 John Spencer and the Limits of Natural Causation in Early Modern England Peter N. Jordan Chapter Nine............................................................................................ 131 Naturalism, the Limits of Science and the Case for Non-scientific Knowledge Emmanuel Nartey Chapter Ten ............................................................................................. 144 The Limits to Science: A Barthian Approach Philip Chapman Chapter Eleven ........................................................................................ 154 Ways to Develop Students’ Appreciation of the Power and Limitations of Science Berry Billingsley and Mehdi Nassaji Part Three Chapter Twelve ....................................................................................... 168 A Christian Response to Transhumanism David Ashford Chapter Thirteen ...................................................................................... 177 Among the Limits to Science: Some After-dinner Thoughts Neil Spurway

Are There Limits to Science?

ix

Chapter Fourteen ..................................................................................... 183 Closing Reflections Mark Harris

THE SCIENCE AND RELIGION FORUM

Growing out of informal discussions which began in 1972, around the key figure of The Revd. Dr. Arthur Peacocke, the Science and Religion Forum was formally inaugurated in 1975. Its stated purpose was “to enable and encourage further discussions of the issues which arise in the interaction between scientific understanding and religious thought”. These issues, together with the social and ethical decisions demanded by scientific and technological advances, have remained the subject of the Forum’s meetings since that date. In 2005 the Forum merged with the Christ and the Cosmos Initiative. This had been founded by The Revd. Bill Gowland, a past President of the Methodist Conference, with the intention of bringing the latest knowledge of scientific thinking within the orbit of the enquiring layperson. Thus enlarged, the Forum is open to all, of any personal faith or none, who are concerned to relate established scientific knowledge and methodology to religious faith and theological reflection. Implementing its broad objectives, it seeks: 1) to encourage scientists with limited knowledge of religion, and religious people with limited knowledge of science, to recognise and appreciate the contributions of both disciplines to human understanding of life in the world; 2) to provide an interface between academics active in sciencereligion work, and public communicators – notably teachers, clerics, and those training future members of these professions. At every point, the Forum strives to extend recognition that science and religion, properly understood, are not antagonists, but complementary in the quest for truth. The Forum holds a regular annual conference, plus occasional smaller ad hoc meetings, and publishes a twice-yearly journal, Reviews in Science and Religion. Since 2008 it has also published edited proceedings of its annual conferences, under the series title Conversations in Science and Religion.

Are There Limits to Science?

xi

At the date of publication, the Forum’s President is Prof John Hedley Brooke (Oxford) and its Chairman Revd Dr Michael Fuller (Edinburgh). www.srforum.org @SciRelForum

ACKNOWLEDGEMENTS

The conference on which this book is based was organised by The Revd. Dr. Mark Harris and Mrs. Hilary Martin. The SRF forum conferences are a unique blend of high academe and non-specialist engagement, and as such get into the heart of the field, asking important questions about the meaning and implications of our key topic. Thanks must go to both Mark and Hilary for their work in creating this distinctive event, and to the whole committee for sustaining our organisation, particularly The Revd. Dr. Michael Fuller who chairs with insight and vision. I would like to give sincere thanks to Daniel Rodger for his work as subeditor of this text, and thanks also to Victoria Carruthers at Cambridge Scholars Publishing for enabling this book. Gillian Straine

CONTRIBUTORS

Gillian Straine is the Director of the Guild of Health and St. Raphael, an ecumenical organisation promoting, resourcing and researching the healing ministry. Her degree and doctorate are both in Physics which she studied at Imperial College London. She is a priest in the Church of England and during preparation for ordination was awarded a MA in theology from the University of Oxford. Her books include Science and Religion: a path through polemic (SPCK 2014) and Cancer: a pilgrim companion (SPCK 2017). Fiona Ellis is Reader in Philosophy at Heythrop College, University of London, and Director for the Centre for Philosophy of Religion. Her most recent book is God, Value, and Nature (Oxford University Press 2014, 2016), and she is currently editing a collection of papers New Models of Religious Understanding (Oxford University Press 2017). This collection originated from a project which was a sub-grant of the Templeton funded Varieties of Understanding project, and Ellis was the director. She is currently co-directing a project on Religious Experience and Desire for which she has a research fellowship at the University of Notre Dame. Mikael Leidenhag earned his PhD in Philosophy of Religion at Uppsala University (Sweden) in 2016. His PhD thesis critically evaluated Religious Naturalism as a position in the dialogue between science and religion. His other research interests include Panentheism, Transhumanism, Panpsychism and the relevance of humility for the science-religion dialogue. He currently resides with his wife, Joanna Leidenhag, in Edinburgh. Sarah Lane Ritchie is a postdoctoral Research Fellow in Science & Theology at the University of St Andrews. Her PhD thesis is on divine action and human consciousness, while her postdoctoral work focuses on the various brain sciences and their relationship to theology and mindrelated issues of human flourishing. She holds an MSc in Science and Religion from the University of Edinburgh, an MDiv. from Princeton Theological Seminary, and an undergraduate degree in Philosophy and Religion from Spring Arbor University.

xiv

Contributors

Donovan Schaefer is an assistant professor in the Department of Religious Studies at the University of Pennsylvania. He earned his B.A. in the interdisciplinary Religion, Literature, and the Arts program at the University of British Columbia in Vancouver, Canada. His master’s and doctoral degrees are from Syracuse University in New York. After completing his doctorate, he held a Mellon Postdoctoral Fellowship at Haverford College, then spent three years as a departmental lecturer in science and religion at the University of Oxford. His first book, Religious Affects: Animality, Evolution, and Power (Duke 2015) challenges the notion that religion is inextricably linked to language and belief, proposing instead that it is primarily driven by affects. Neil Messer gained his PhD in molecular biology from Cambridge University before studying theology in Cambridge and King’s College London. He has held various academic posts in theology and ethics, and is currently Professor of Theology at the University of Winchester. His publications include Selfish Genes and Christian Ethics: Theological and Ethical Reflections on Evolutionary Biology (SCM, 2007), Respecting Life: Theology and Bioethics (SCM, 2011), Flourishing: Health, Disease and Bioethics in Theological Perspective (Eerdmans, 2013), and Theological Neuroethics (Bloomsbury T & T Clark, forthcoming). He is an ordained minister of the United Reformed Church, and a member of the Expert Group on Ethics of the Community of Protestant Churches in Europe. Nathan Aviezer is Professor of Physics and former Chairman of the Physics Department of Bar-Ilan University in Israel. Aviezer is the author of more than 140 scientific articles on condensed matter physics. In recognition of his important research contributions, Aviezer was elected as a Fellow of the American Physical Society. In addition to his scientific research, Aviezer has a long-standing involvement in the relationship between Torah and Science. He is the author of three books on this subject: In the Beginning (translated into nine languages), Fossils and Faith (translated into four languages), and Modern Science and Ancient Faith (recently published). Aviezer teaches a course at Bar-Ilan University on “Torah and Science,” which was awarded the prestigious Templeton Prize. In addition, Aviezer is active in the organization of an annual Torah and Science Conference which attracts hundreds of participants from all over Israel. Finally, Aviezer was recently awarded a grant from the Templeton Foundation to develop a teaching unit on Torah and science for Orthodox Jewish high schools.

Are There Limits to Science?

xv

Neil Spurway studied at Cambridge, but has worked in the University of Glasgow ever since, and is now Emeritus Professor of Exercise Physiology. He has chaired Glasgow’s Gifford Lectureships Committee, as well as the present Forum, been President of the Royal Philosophical Society of Glasgow and the Scottish Church Theology Society, and VicePresident of the European Society for the Study of Science And Theology, and also edited the latter’s journal, ESSSAT News. Alongside considerable scientific writing he initiated the present series and edited four of its previous volumes – most recently Forty Years of Science and Religion (2016). Berry Billingsley and Mehdi Nassaji work together on the LASAR (Learning about Science and Religion) project - now based at the Faculty of Education at Canterbury Christ Church University. Most recently LASAR has begun to investigate and discuss Epistemic Insight - where Epistemic Insight is 'knowledge about knowledge' and particularly knowledge about how disciplines interact and their strengths and limitations in real world and multidisciplinary arenas. For more information, please see www.lasarcentre.com and www.epistemicinsight.com. Emmanuel Nartey, BA (Legon), Ph.D. (Fordham), is Associate Professor of Philosophy at The City University of New York. His research interests include Modern Philosophy, Philosophy of Mind, and Philosophy of Religion. He has published papers on Integrating Science and Religion, Hylomorphism, Philosophical Perspectives during the Middle Ages, Beyond the Secular, Omniscience and Free Actions, Descartes and MindBody Problems. His book, Nature, Mind and Hylomorphism (SpringerVerlag), is forthcoming. Philip Chapman is a Methodist Minister. He attended lectures from Paul Feyerabend at Bristol in 1956-7. They sparked an interest in philosophy of science. In them, Karl Popper's doctrine of falsifiability as a criterion for scientific propositions featured prominently. In 1963, he met Karl Barth at a book launch near Paris. Towards the end of 40 years of pastoral work in several countries, he began the study of Barth at Leeds under Jacqui Stewart. In 2014, his work culminated in a PhD. featuring both Barth and Darwin, under the supervision of Neil Messer at Winchester. David Ashford is founder and managing director of Bristol Spaceplanes Limited, an innovative small company pioneering low-cost access to space. He graduated from Imperial College in aeronautical engineering and spent one year at Princeton doing post-graduate research on rocket

xvi

Contributors

motors. His first job, starting in 1961, was with the Hawker Siddeley Aviation spaceplane design team. He has since worked as an aerodynamicist, project engineer, and project manager on various aerospace projects, including the DC-8, DC-10, Concorde, the Skylark sounding rocket, and various naval missile and electronic warfare systems at Douglas Aircraft and at what is now BAE Systems. He co-authored with Prof. Patrick Collins the first serious book on space tourism Your Spaceflight Manual How You Could be a Tourist in Space Within Twenty Years (Headline, 1990), and wrote a follow-up book Spaceflight Revolution (Imperial College Press, 2002). His most recent book, Space Exploration: All That Matters, was published by Hodder in 2013. He has published about twenty papers on space transportation in the professional press. He is a Fellow of the Royal Aeronautical Society and is a Rolt Fellow at the Centre for the History of Technology at the University of Bath. Michael Fuller is the Forum’s chair, studied chemistry at Oxford and theology at Cambridge. He served as a priest in the dioceses of Oxford and Edinburgh, and for 14 years oversaw ministerial training for the Scottish Episcopal Church. He is a teaching Fellow at New College, Edinburgh, and an Honorary Canon of Edinburgh Cathedral. Alongside many papers he has written or edited seven books in the science and religion field, three in the present series. Peter Jordan is currently a research coordinator at the University of Oxford. After obtaining a degree in engineering from Queensland University of Technology and a PhD in physiology and biophysics from Cornell University, he was a postdoctoral research fellow at the National Institutes of Health. He subsequently obtained a master’s degree from the Divinity School at Duke University, followed by a PhD in religious studies from the University of Queensland, where he worked on natural philosophy and its relations to theology in early modern England. Mark Harris, the Forum’s Conference Secretary, is Senior Lecturer in Science and Religion at Edinburgh and runs the MSc and PhD programmes in that field. A former physicist, he is interested in the complex and subtle relationship between physics and religious belief, and is working on a long-standing project on the question of scientific and theological views of physical reality and the laws of nature. He is the author of The Nature of Creation: Examining the Bible and Science (Acumen 2013).

INTRODUCTION

Since its establishment over 40 years ago, the Science and Religion Forum has brought together people who grasp the importance of this subject. Whether academic or non-specialist, lay or ordained, theist or atheist, all are united in a common pursuit: to ask questions and listen to each other as we explore this vital area of thought. In 2015 the Science and Religion Forum marked its fortieth anniversary by considering the history of our field and surveying what might lie ahead. Critiquing its current position, several speakers pointed to an ambivalence about how the study of science and religion is considered both inside and outside the academy. There was some frustration that the work of the Forum in particular, and the field of science and religion in general, were not more widely recognised. Attempts were made to diagnose the reasons for this situation and where new work might promote positive change. But it was not all loss and despair; in a wide ranging event, a rallying cry was heard to carry on the work, and to explore new areas of engagement for the continued promotion and elevation of our interest area. The following year, at the Woodbrooke conference centre in Birmingham, the Science and Religion Forum made a response to this invitation, and began to survey the boundaries of our subject. A wise first move, the aim of the conference was to build up confidence about the journey ahead and to determine the way in which new grounds for exploration could be uncovered. Therefore limits were explored in this conference, and subsequently this book: how far can the conversation about science and religion go? Is it science or religion which determines the boundary? What has happened at these limit areas in the past? How does the evolution of science change the conversation? By necessity, this book deals in philosophy – what do we know, how do we know it, and how do we talk about it? But fascinatingly, these chapters are not a collection of dusty arguments over semantics. For these limit areas of knowledge are not the dwelling place of rarefied ideas, but rather they are bustling and noisy places of discussion and engagement. The limit areas, the boundaries between science and religion, are personal and urgent; it is where we ask important questions, like how do we know what we know? Can we make a distinction between the physical and the

2

Introduction

mental? And, the chief question for the whole field, what is the relationship between Creator and the creation? The discussion that emerges is captivating, and demands not only engagement of the mind, but choice about our own response – a book focused on the limits of science does not itself limit us or the field, but rather demands that we ask what this means for us? Where do we put the limits in our own self-understanding, and in our relationship with the whole of Creation, seen and unseen.

Part One The first section of this book contains chapters from the invited speakers to the conference. Fiona Ellis in her carefully argued and insightful paper takes us into a key limit area – the involvement of God in naturalism. Citing a distortion that leads to the regency of science in setting the tone of the conversation, her expansive naturalism offers an enlarged view of nature which opens a philosophically robust space for God, and for human experience to be included in epistemology. Cautioning against a “dazzlement with science”, encouragement is given to come to these limit areas with “dimensions of experience which elude science”, and to do so with confidence. Ending with a contemplative note, she suggests that a limit area is enhanced by seeing God in all things, a view which leads to better answers about how we should ascertain reality. In the following chapter, Mikael Leidenhag takes another look at theological naturalism and offers a critique of the naturalistic foundations of Ellis’ theology, suggesting a look in a different direction would not be so risky for metaphysics. Openmindedness and resistance to scientism are not enough, Leidenhag argues, in an expansive naturalism where the real problem is in defining physical and mental causality. He pinpoints the problem for the field in achieving a harmonisation without reductionism, and establishing an adequate theistic naturalism which includes value. His solution – a holistic, integrative dualism. Sarah Lane Ritchie makes an important contributing to this book in chapter three by taking us into the question of the casual joint, one of the defining limit areas of the science and religion field. She critiques the Divine Action Project as misframing the limit area which has looked for spaces in which God might act. A recent “theological turn” in the study of the causal joint has emerged that, on first reading, gives theology and metaphysics the chance to set the tone of the conversation. Lane explores whether this “turn” has potential for describing the causal joint, and finds that limit questions remain despite the potential for God involvement. But

Are There Limits to Science?

3

our eyes are drawn to a key theme: the one who sets the limits of the conversation, holds the power. Donovan Schaefer’s beautifully crafted chapter introduces into the book a new angle on our attempts to assess the limits of science: the impact of the emotions on science. In his fascinating piece, he critiques a self-perception of science as being able “to produce truths that are uncontaminated by the murk of emotion”, by looking at first at David Hume and William James before moving into contemporary affect theory to reveal science as instead guided by a “intransigent tissue of emotional priorities, micro-threads that lace cognition into a structure wave of information and inspiration”. He significantly opens up the horizon and leads us into new grounds and limit places. In the following and final chapter in part one, Neil Messer highlights another new and incredibly important limit area: the case of the human brain. Neuroscience, evolution and cognitive science are often seen as firm boundary markers in understanding religion and moral decision making. Typically, science holds the monopoly on the conversation which leads to an ethic precluding theistic accounts. Investigating this area, Messer observes that there might be confusion over scientific epistemology that, when sorted, can allow other matrices of relationship. He leads us through various interpretations of science and religion dialogue in a series of typologies to understand what it means to be human, balancing theology and evolution, cognitive and neuroscientific accounts. His struggle to honestly present a worked example, showing how science can act as a commentary to real life examples, invites the reader into the conversation, and thus takes the limit areas out of the ivory tower and into real life.

Part Two This section of the book contains some of the short paper sessions created following an open invitation to conference delegates. The range and depth of the articles broadens the discussion in useful, vital ways and, following the ethos of the Science and Religion Forum, gains valuable perspectives from all who are interested. Part two opens with an invitation from Nathan Aviezer to take a step back and consider the historical ground that lies underneath all the science and religion conversation. His article surveys what has changed over the years in science, beginning at the start of the 20th century, when there was calm and confidence about the power of science to explain all natural phenomena. His chapter reminds us usefully that science changes, and today’s limits may have a different interpretation in years to come. The

4

Introduction

unknowns remain, and the degree to which the unknowns become knowns is not predictable – this is both a note of caution and an invitation to participation. In chapter seven, we delve more deeply into the historical roots of the science, in a chapter written by the Chair of the Science and Religion Forum, Michael Fuller. Exploring the historical limit to science, an area much misunderstood in the popular imagination, he presents “The Case of Thales’ Ox”. Looking at ancient sources and expertly showing the caution required in historical investigation, he nonetheless presents at the limits of historical veracity an early man of science and religion. As the previous chapter showed, to explore present day limit areas, we must survey the ground on which we trend. Therefore, we remain with history in chapter eight, wherein Peter Jordan also debunks the conflict myth, seeing integration of naturalistic and theological knowledge this time in seventeenth century England. He presents the work of John Spencer who held together a commitment to natural philosophy and a faith in God, interestingly suggesting that the former can be used to construct good theology. While never suggesting that this was an easy task, Jordan’s chapter shows that historical study can lead to enlarged horizons, challenging the dominance of scientism, today and in the historical view. Chapter nine returns us to the contemporary naturalism scene. In this chapter, Emmanuel Nartey critiques the limits which are imposed on both science and religion through the types of questions they ask. In a discussion involving both the scientific method, and the types of naturalism possible, he postulated that questions of meaning and value cannot be ignored by science. This is an important limit area for the field, especially around theories of the mind. The theologian Karl Barth is the subject of chapter ten, which explores Barth’s theological idea that science and theology have nothing to say to one another. Philip Chapman contends that Barth’s writing suggests that it is not possible or permissible to use science and what science says about the world, to make any type of theological discovery. Science is limited by definition, and theology is determined by revelation alone. It is a description of the limit area for the science and religion dialogue that is worth taking note of, and which many find appealing. Science, Chapman concludes, should be modest in its limits, and religion should be certain in its origins though humble too in its self-perception. The final paper in Part 2 is by Berry Billingsley and Mehdi Nassaji and presents research they have conducted with teenagers around robotics. The education of the next generation should be of concern to anyone interested in science and religion. Young people are relentlessly subject to the myth

Are There Limits to Science?

5

of conflict between science and religion, which involves not only poor history but inadequate theology. In this chapter, the authors show how the study of robotics led to metaphysical questions about existence, causality, the scientific method and theory of mind. It is an important project, and one that exposes young people to the power and the limits of science, and what this means in their own lives.

Part Three The Science and Religion Forum is open to all, and values congeniality, conversation and openness. As such there is room for very different conversations to happen and be challenged in a friendly environment. Chapter twelve contains an interesting thought experiment in a key new area: transhumanism and human enhancement. This is a field generating a great deal of interest, with some confidently predicting the imminent arrival of machines with capabilities far exceeding those of humans. What religious responses might be made to such imaginary scenarios? Ashford's contribution offers one imaginative response. Neil Spurway, a former editor of this book and chair of the Forum, spoke after the conference dinner. His words are captured in chapter thirteen but are more than just post-meal entertainment. They contain a distillation of not only a life spent working at the limits of science and religion, but a considered reflection on that life by a man with a poet’s eye for truth. He reminds us that people who work at these frontiers of thought, show us fragility, truth and depth, knowledge gained the hard way that can transcend the physicality that they seek to understand. In many of the chapters of this book, the roles of human experience, emotion, brain studies, theories of mind and consciousness, appear at the limit areas of science and religion. And so fittingly, at the end of his reflections summing up the conference found in chapter fourteen, Mark Harris, announced the subject for next year’s conference: Neuroscience, mental health and religion. Indeed, mental health is increasingly seen as a key area in healthcare, a vital area of research, and of concern to those involved in ministry. Similarly, the study of the mind is one of the most controversial areas of science, and a contentious area of science and religion conversations. As a leader in the field and conference secretary to the Science and Religion Forum, Harris’ reflections on the conference highlights that at the limits of science and religion we find an invitation to engage: Our conference may have been about limits, but it was in no sense limited. I commend this book to its readers and highlight the invitation that it offers

6

Introduction

to engage at the limits of science and religion, a proposition which is ultimately an exploration of our humanity.

PART ONE

CHAPTER ONE GOD, NATURALISM, AND THE LIMITS OF SCIENCE FIONA ELLIS

1. Introduction The question of whether there are limits to science can be variously interpreted, and I shall begin by distinguishing the different things that could be at issue in this context, my eventual aim being to challenge an approach which suggests that science is unlimited in its scope in the sense that it is 'the measure of all things, of what is that it is, and of what is not that it is not' (Sellars 1963). The words come from Wilfrid Sellars, they sum up the position which has come to be known as scientific naturalism (or just plain “naturalism” or, more pejoratively, “scientistic naturalism”), and this form of naturalism has been described as the main programmatic orientation of contemporary Anglo-American philosophy.1 So a lot hangs on the question of how this position is to be understood and assessed. I shall argue that there are good reasons for resisting it, and that there is an alternative, more liberal, conception of naturalism which grants us the right to impose explanatory limits upon science whilst respecting the significance of its findings. This is expansive naturalism, it has been articulated and defended by philosophers like John McDowell, David Wiggins, and James Griffin, and it raises the question of what it really means to be a naturalist. I have argued elsewhere that expansive naturalism can be further expanded in the direction of God, and hence, that the naturalist can be a theist. The position is contentious, it challenges the received philosophical wisdom in various respects, and it has important implications for how we think about the science versus religion debate. In

1

This statement was prepared for a workshop on naturalism by Joseph Margolis and Mark Gottlieb, and it is cited by Bernstein 1995, p.58

God, Naturalism, and the Limits of Science

9

what follows I want to give a sense of my overall strategy, fend off some objections, and make explicit the implications for how we think about the limits of science and its bearing upon the science versus religion debate.

2. The limits of science: some clarifications When we talk of the limits of science one option is to take the relevant limits in an epistemological sense. Science is limited in this respect, for it is practised by scientists, and scientists, being human like the rest of us, are susceptible to ignorance and error. This is a perfectly general point about our epistemological limitations, and there is nothing as yet to suggest that science is better or worse off in this respect than any other mode of investigating reality. The idea that science per se is to be distinguished epistemologically from other modes of investigating reality is a familiar enough thought, one idea being that it carries some kind of epistemological advantage. There is an insight in this response-consider the explanatory and predictive power of modern science and contrast it with the pre-scientific superstitions and speculations which predominated before its rise. However, the point about ignorance and error must be accommodated, and we need to be clear in any case about what kind of science is at issue here. Presumably, the focus is Modern science, but science is not any one thing, and there is a question about its limits in this ontological sense. We can note also that even if there are good reasons for privileging science (or a particular kind of science) over some other modes of investigating reality, it does not follow that it is the only way of investigating reality, nor that it is superior to all other modes of investigation, nor that it is guaranteed to supply the understanding we seek. According to an extreme version of scientific naturalism, we should all be reductive materialists or reductive naturalists. This position can be variously interpreted, but D.M. Armstrong’s, “the natural world contains nothing but the entities recognised by physics” (Armstrong 1980, 156) suggests that the aforementioned measure of reality belongs exclusively to physics. The implication here is that physics determines what there is-as far as the natural world is concerned at least, although the typical naturalist will deny that there is anything beyond it. I shall return to this point below. For the moment, we can note that there is disagreement about the prevalence of reductive materialism2, and that its authority has been 2

Thomas Nagel claims that 'among the scientists and philosophers who do express views about the natural order as a whole, reductive materialism is widely assumed

10

Chapter One

justifiably contested. After all, the fact that physics deals with the ultimate constituents of nature does nothing to show that everything is reducible to those constituents, and it is difficult to see what could lend credence to such a vision. We can note that this objection to reductive naturalism is pitched at the level of philosophy rather than physics, or, more generally, science, this in itself exemplifying a clear enough sense in which science has its limits. It is limited in this context by philosophy, philosophy being that which poses a challenge to its more fundamentalist leanings. It is worth noting finally that these extravagant claims about the scope of physics sit rather awkwardly with a theme which is central to the naturalist's “philosophy”, namely, the desire to deliver us from the suspect entities of transcendent metaphysics and religion to a rather less mysterious earth. It has a “bias towards the unpuzzling” in this respect, as James Griffin has put it.3 Why is this a problem for physics? Well, if physicists are anything to go by there is nothing so mysterious as the constituents of reality. The reductive materialist takes the measure of nature to be physics. A more moderate naturalism defines this measure with reference to a broader conception of science (Why just physics? How on earth could that explain everything? And what reason could be given for insisting upon this restriction?), and an even more moderate position challenges the assumption that the offending restriction can be lifted only in terms which are themselves restricted by science (Why just science? How on earth could that explain everything? And what reason could be given for insisting upon this restriction?). It is in the context of giving expression to these latter complaints that we find John McDowell recommending that we “discourag(e) this dazzlement by science”which leads us to suppose that “genuine truth is restricted to what can be validated by their methods” (McDowell 2002, 295). McDowell is one of several recent philosophers who have sought to defend a more liberal or “expansive” form of naturalism. The position is opposed to scientism but it is not opposed to science, and the expansive naturalist gives due weight to its epistemological significance. What he denies, however, is that science is the only legitimate way of investigating reality, insisting to the contrary that there are dimensions of reality for which a purely scientific account is inadequate, and philosophical questions to be the only serious possibility', (Nagel 2012, 13). Dupré disagrees, but grants that the reductive spirit of this picture continues to animate philosophical thought (Dupré 2004 , 46-47). 3 See Marcus Du Sautoy's (2016) for a spelling out of science's epistemological limitations.

God, Naturalism, and the Limits of Science

11

to be raised about what we should be saying in this context. The moral dimension of reality is an important case in point for the typical expansive naturalist, but we might talk equally of love, beauty, meaning, and all of the other things for which a purely scientific account seems inadequate. The point is well summed up in a review of Marcus Du Sautoy's recent book What We Cannot Know: Explorations at the Edge of Knowledge (Du Sautoy is the Simonyi Professor for the Public Understanding of Science at Oxford). The reviewer is Jonathan Rée, and having made the point that Du Sautoy seeks to map some of the limits or “edges” of science, complains that an implicit scientism lurks in his approach. Hence: He fails to recognise that vast swaths of human knowledge are concerned with human meaning and interpretation rather than scientific fact and explanation: with beauty, for instance, or with history, poetry and memory, or love, ageing and mortality, or what words can and cannot express. He assures us in passing that judgements of beauty arise from dopamine rushes controlled by our genes, but if we disagree about whether something is beautiful we ought really to have a discussion about how it looks rather than trying to compare our dopamine levels. And if we want to investigate the significance and viability of religious belief, we might be well advised to start from the ambiguities of human experience rather than following Du Sautoy in appealing to the latest results of scientific research.4

The expansive naturalist agrees that we need to go beyond scientific fact and explanation if we are to come to a proper understanding of goodness, love, and beauty. However, he is uninterested in the significance and viability of religious belief, agreeing with the typical scientific naturalist that the natural world leaves no room for God. I have said already that I part company with him in this respect. I want now to give a sense of my position, taking my cue from Rée's suggestion that we “start from the ambiguities of human experience”. It should go without saying that the considerations I rehearse in this context are pitched at the level of philosophy (the limits of science again), although there will be a question of where this leaves the relation between philosophy and theology.

3. Expansive naturalism and God The expansive naturalist imposes limits upon science by denying that it has a monopoly on explanation and reality. However, he agrees with the scientific naturalist that we can explain what needs to be explained without 4

The Guardian, Saturday 2 July 2016.

12

Chapter One

introducing a second, supernatural, realm, taking the postulation of such a realm to introduce the metaphysical and epistemological difficulties a naturalistic standpoint is intended to avoid. As one philosopher sums it up, “[t]here is no room for any supernatural in naturalism...naturalism finds itself in thoroughgoing opposition to all forms of thought which assert the existence of a supernatural or transcendental Realm of Being and makes knowledge of that realm of fundamental importance to human living” (Krikorian 1944). Talk of a transcendental or supernatural realm of being calls to mind a theistic framework, and I have noted already that the typical naturalist stands opposed to theism. However, supernaturalism is said to get a grip in other contexts too, when, for example, we seek to comprehend the things for which a scientific account seems inadequatelove, beauty, goodness and so forth. James Griffin responds as follows, his focus being the more general question of value: “[v]alues do not need any world except the ordinary world around us-mainly the world of humans and animals and happenings in their lives. An other-worldly realm of values just produces unnecessary problems about what it could possibly be and how we could learn about it” (Griffin 1996, 43-44). He quickly adds that to defend his preferred position “one does not have to adopt a reductive form of naturalism”. The expansive naturalist seeks to vindicate the idea that this world-the world of humans and animals and happenings in their lives-is valueinvolving. He grants with the scientific naturalist that an other-worldly realm of values can make no sense of our evaluative practices (and more generally, all of those things which are “of fundamental importance to human living”), and agrees also that its postulation is unnecessary. (The objection here is that such a realm can shed no light upon the values we respond to when, for example, we see the generosity of an act, or respond to a loving encounter. Or to put it another way, an “other-worldly” realm of values floats free of our evaluative practices.) He insists, however, that this value-involving world-the only world there is-cannot be wholly comprehended in scientific terms, and that a proper understanding of our evaluative practices must take us beyond such a standpoint. So the limits of nature must be expanded beyond scientific parameters. It is worth noting finally that once this move is made, then the motive for supernaturalism is lost. After all, supernaturalism is motivated by the thought that the relevant items cannot be contained within the natural world, and they cannot be so contained if scientific naturalism is true. If this is so, then it looks as if the real culprit in this debate is the scientific naturalist, for if what I have said is correct, then his position invites supernaturalism rather than being its nemesis. At least this is so if

God, Naturalism, and the Limits of Science

13

supernaturalism involves postulating a world in addition to this world, and in such a way that the two worlds cannot be intelligibly related. The scientific naturalist's response to this is that we commit to supernaturalism the moment we exceed the limits of scientific naturalism, but this is because he interprets naturalism in a narrow, scientific, sense, taking “supernaturalism” to be the logical complement of “naturalism” in this narrow sense. So expansive naturalism is eo ipso a form of supernaturalism if we accept his terms of debate. I have suggested that we should not. The expansive naturalist operates with a conception of nature which is rich enough to accommodate value, but he has no interest in accommodating God, and we are to suppose that such a move would involve a retreat back towards the suspect supernaturalism his position is designed to avoid. The assumptions here are that there is no sense to be made of the idea that the natural world is God-involving as well as value-involving, and that God can be brought into the equation only with the postulation of a supernatural realm in addition to nature, where nature, we have been told, is the only world there is. According to my position these assumptions must be questioned. That is to say, I hold that there is sense to be made of the idea that the natural world is God-involving, and that we must reject the assumption that a theistic framework involves the postulation of a second, supernatural, realm. The idea that the natural world is God-involving will be familiar to anyone who takes God to stand in the most intimate connection with all things, i.e. it will be familiar to those who are situated within the JudeoChristian tradition, although this is not to deny that the nature of this intimate connection raises deep and perhaps irresolvable theological issues. It is, after all, God we are talking about, and God surely has title to be even more mysterious than the most intractable particle of physics. As for rejecting the idea that God inhabits some second, supernatural, realm, well theologians and philosophers have long emphasised that this is picture thinking at best, that it undermines God's infinitude (Hegel), and that it makes of Him a mere “being amongst beings” rather than being, as Rahner puts it “the most radical, the most original, and in a certain sense the most self-evident reality” (Rahner 1978, 63). The idea that God is the most self-evident reality has an air of paradox to it, and it would certainly be denied by the typical naturalist. However, it is really just another way of making the point that He stands in the most intimate connection with things, although this is not to deny that the position can be disputed, as it must be given that there can be no demonstration of God's existence. It is fundamental to this picture that there is a distinction between God and the things with which He is

14

Chapter One

intimately connected, this being the reason for denying that God is to be identified with anything within or beyond the world (which would make him a mere “member of the household of all reality” as Rahner puts it). So the idea that nature is God-involving does not mean that God is reducible to nature, and this anti-reductive point is one reason for thinking that talk of a supernatural Realm of Being is not entirely misplaced, provided that we do not take it literally. The typical expansive naturalist can make no sense of these thoughts. That is to say, he denies that the natural world could be God-involving, and offers a conception of value which makes no reference to God. To put it another way, he assumes that the introduction of God could only ever amount to suspect supernaturalism. But what if his objections are premised upon a faulty understanding of God, and what if they are on a level with the scientific naturalist's objection to expansive naturalism (remember that the scientific naturalist thinks that anything beyond scientific naturalism involves the postulation of a second, supernatural, realm, and hence, that expansive naturalism is a form of supernaturalism)? The expansive naturalist could try to block this parallel by objecting that the two cases are completely different – God introduces a suspect supernaturalism in the way that value does not. Yet this response simply begs the question against the possibility of an alternative framework – one which challenges the assumption that God must be viewed in these pejorative terms and hence, that nature must exclude Him. We might note that the secular expansive naturalist has every reason for taking this possibility seriously given his desire to raise anew the question of the limits of nature, and to challenge a prevalent way of drawing them.

4. Ambiguities of human experience This is all very well, but I have said little about what it means to describe nature in God-involving terms, other than to say what it is not and to hint that I shall be taking my cue from Rée's suggestion that we begin from the ambiguities of human experience rather than the latest scientific research. This was my approach in God, Value, and Nature (Ellis 2014), my strategy there being to further narrow the gap between secular and theistic expansive naturalism by focusing upon the case of moral experience and making a case for the claim that we encounter God in this context. My protagonist was Emmanuel Levinas-a significant figure in the dialectic given his insistence that being moral is the only way of relating authentically to God. As he puts it, “to know God is to know what is to be done” (Levinas 1990, 17), and “there can

God, Naturalism, and the Limits of Science

15

be no ‘knowledge’ of God separated from the relationship with men” (Levinas 1969, 78). Levinas is anxious to deflate our cognitive pretentions in this context, an obvious target being those who take a purely theoretical approach to God. God's status is so reduced on the assumption that knowledge in this context is to be found at the level of scientific research, and this assumption goes unquestioned by Du Sautoy, and, more famously, Richard Dawkins (“God's existence or non-existence is a scientific fact about the universe...the presence or absence of a creative super-intelligence is unequivocally a scientific question” (Dawkins 2006, 58)). His criticisms apply equally, however, to any approach which favours theory over praxis, a further target being the kind of philosophy of religion which insists that “understanding the world religiously is...an attempt to dissect and analyse and explain it in the manner of modern science” rather than engaging with the world in a manner which demands a moral and spiritual opening of the self5. On Levinas's position then, we engage with God at the level of morality-to know God is to know what is to be done. What does this have to do with the ambiguities of human experience? Well, given that God is irreducible to anything within or beyond the world, then there are no prospects for modelling our experience of God on our experience of any such item. This would be to return us to the very picture the naturalist is seeking to avoid. I have noted also that it is built into the tradition that God is intimately connected to all things, and that this connection helps to make sense of Rahner's rather gnomic remark that his reality is the most self-evident of all. Now there is a question of how and where this self-evidence is manifested, but we might bear in mind Augustine's claim that God is closer to me than myself, and remind ourselves of Levinas's insistence that I relate to God (and know Him) by virtue of standing in moral relations to others. We might also consider the following remarks of Simone Weil: The thought of God must not interpose itself between us and other creatures. It must not make the contact between us and them less direct. On the contrary, through it the contact must be made more direct. The real aim is not to see God in all things; it is that God through us should see the things we see. God has got to be on the side of the subject and not on that of the object during all those intervals of time when, 5

See John Cottingham's 'Transcending Science: Humane Models of Religious Understanding', forthcoming in Ellis, F. (ed.) 2017. New Models of Religious Understanding, Oxford: Oxford University Press.

16

Chapter One forsaking the contemplation of the light, we imitate the descending movement of God so as to turn ourselves towards the world (Weil 1956, 358).

The implication in all of this is that I experience God when I stand in moral relations to others. There is a question of where God is to be pinpointed in all of this, but this is surely the wrong question to ask if taken to mean that He is some kind of identifiable entity. Indeed, this point tells equally against Weil's dismissal of the idea that God could be “on the side of the object” (omnipresence again). Where does this leave the ambiguity of the relevant experience? Well, it is ambiguous for the simple reason that, in one clear enough sense, it does not involve God. Hence the perennial temptation towards atheism-which is no bad thing from Levinas's point of view given that atheism clears the sky of false gods, and puts us in a position to be morally responsible (i.e. puts us in a position to relate to God). Indeed, it is a radical implication of Levinas's line of thought that God can be truly present only by withdrawing (He is present in His absence). The relevance of this to the question of the limits of expansive naturalism should be clear. For if we accept this moral conception of what it means to relate to God, then there is scope for allowing that the secular expansive naturalist is already operating with a God-involving conception of nature, in spite of himself, as it were. To put it another way, there is a knife edge between secular and theistic expansive naturalism-the kind of knife-edge exploited by Levinas when he treats atheism as a route to God. The secular expansive naturalist's opposition to scientism goes hand in hand with the idea that approaching value in purely scientific terms cuts us loose from the very thing we are trying to understand, and that we can do justice to our subject-matter only by availing ourselves of the concepts which are operative at the level of ordinary moral engagement. We must be informed about morality “from the inside” in this respect, and we are to suppose that this kind of insider knowledge eludes the scientist. The social scientist's approach to morality is a target for David Wiggins, and he distinguishes its “clinical or scientific ideal of understanding” from his own more “participative” approach, even whilst granting that there are versions thereof which come close to eliminating the offending explanatory gap (Wiggins 1991). He talks in this context of the “humanization” of social science (“the humanization of human science”!), implying that we have left behind science in all but name. We can note that a structurally identical line of argument is rehearsed by the theologian Michael J. Buckley in the context of criticising approaches to God which proceed “from the outside” by abstracting from any experience of God, and which

God, Naturalism, and the Limits of Science

17

fail thereby to take account of the “personal engagement” which alone can guarantee relationship with God. Buckley's target here is the kind of natural theology which takes its lead from science and can make no sense of the idea that nature is God-involving (Buckley 1987, 348). There remains a question of what personal engagement with God really amounts to, and I have noted already that there is a knife edge between Levinas's position and one which seems to belong more properly to an atheistic framework. The knife-edge is to be expected given the nature of the case, and it is for this reason that moral experience can be said to wear an ambiguous face. It is reasonable to suppose that a similar ambiguity characterises other experiences which are pitched at the level of “human meaning and interpretation”, when, for example, we fall in love or judge that something is beautiful.

5. Science, again The theme which has emerged is that there are dimensions of experience which elude the perspective of science. The relevant experiences are value-involving, the value in question is not exclusively moral, and there is disagreement about whether and in what sense they can be said to involve God. Assuming that it makes sense to interpret the relevant experiences theistically, there is a question of whether it follows that science is incapable of revealing God. This follows if God is revealed only through the relevant value-involving experiences. However, there are two potential difficulties with this response. First, it seems difficult to reconcile with the idea that God is omnipresent to all things, for this suggests that He is capable of being revealed at every level of experience, including those of the scientist. Second, there is ample anecdotal evidence to suggest that science can and perhaps ought to be a gateway to God. So, for example, Einstein claims that “you will hardly find one among the profounder sort of scientific minds without a peculiar religious feeling of his own” (Einstein 1999), to which Werner Heisenberg adds that “the first gulp from the glass of natural sciences will turn you into an atheist, but at the bottom of the glass God is waiting for you”6. We are familiar, of course, with Dawkins' opposing view that the scientific fact about the universe is that God does not exist; or as Steven Weinberg puts it, “one of 6

This quotation is familiar, but there is a question concerning its veracity and source. It does not appear in Heisenberg's written works, but it is cited by Ulrich Hildebrand in 'Das Univerzum - Hinweis auf Gott?' in 'Ethos. Die Zeitschrift fur die Ganze Familie', Berneck, Schweiz: SchwengelerVerlag AG, No 10, Oktober 1988, 10. The quotation is not sourced.

18

Chapter One

the greatest achievements of science has been, if not to make it impossible for intelligent people to be religious, then at least to make it possible for them not to be religious. We should not retreat from this accomplishment” (Weinberg 1999). Du Sautoy challenges this viewpoint in one clear enough sense, suggesting that we equate God with “the abstract idea of things we do not know”. It is unclear, however, that he could be described as religious on this ground, for assuming that the knowledge in question is to be supplied by science, and that science is capable of delivering the goods, then this theoretical conception of God is just a further expression of scientism. Heisenberg's claim that the first gulp from the glass of natural science will turn you into an atheist helps to make sense of the attitude of Dawkins and Weinberg. This attitude comes about because modern natural science offers the prospects for comprehending the rules by which nature works. As such, it offers an alternative to the intervening gods of pre-scientific superstition, and God Himself seems to become an irrelevant and unnecessary extra, like the other-worldly values rejected by the expansive naturalist. (It is no accident that Du Sautoy is anxious to reject reference to a supernatural God.) We can note, however, that this line of argument exceeds the limits of science to become philosophy or metaphysics. It is bad metaphysics from Heisenberg's point of view for one who shares the aforementioned Heisenbergian sentiments, we are told that this atheistic attitude gives way eventually to theism: God is waiting at the bottom of the glass of natural sciences. We are to suppose that it is at this level that we encounter Einstein's “profounder sort of scientific mind”. The profounder sort of scientific mind will presumably reject scientism and suspect supernaturalism. That is to say, he will deny that science has a monopoly on explanation and reality, and that God is on a level with the intervening gods of pre-scientific superstition. We can note again that these conclusions arise from philosophical rather than scientific considerations. So the scientist is not bound to be an atheist, although this is not to say that he is bound eventually to be a theist, Heisenberg notwithstanding. Where does this leave the question of whether science can reveal God? It follows from what has been said that science per se remains neutral on the question of God's existence or non-existence, its focus being how things work and fit together within the realm of nature. However, the scientist can view these things from a broader, non-scientific, perspective, and does so, for example, when he is filled with a sense of wonder and compelled to broach the question of God. It is at this level of experience that he may feel that the world is charged with God's grandeur, and it is in this sense that he may come to feel God's intimate connection with all things. Where

God, Naturalism, and the Limits of Science

19

does this leave Levinas's idea that we relate to God only by standing in moral relations to others? The implication here is that Hopkins' mystical image is to be rejected, although one could respond perhaps that it is at the level of morality that God's grandeur is most properly discerned. Either way, I have suggested that Levinas and Weil are rather too quick to “forsake the contemplation of the light”. We can agree that God has got to be on the side of the subject, but it doesn't follow that we must abandon the aim of seeing God in all things, and it is integral to the naturalism I endorse that we do not.

References Armstrong, D.M. 1980. “Naturalism, Materialism, and First Philosophy”, in The Nature of Mind and Other Essays. St. Lucia:University of Queensland Press. Bernstein, R. 1995. “Whatever Happened to Naturalism?”, Proceedings and Addresses of the American Philosophical Association, 69, no.2. Buckley, M.J. 1987. At the Origins of Modern Atheism. Yale: Yale University Press. Dawkins, R. 2006. The God Delusion. London: Bantam Press. Dupre, J. 2004. “The Miracle of Monism”, in Naturalism in Question. Cambridge, MA: Harvard University Press. Du Sautoy, M. 2016. What We Cannot Know: Explorations at the Edge of Knowledge. London: Fourth Estate. Einstein, A. 1999. The World as I See It. Secaucus, NJ: Citadel Press. Ellis, F. 2014. God, Value, and Nature. Oxford: Oxford University Press. Griffin, J. 1996. Value Judgement: Improving our Ethical Beliefs. Oxford: Clarendon Press. Krikorian, Y.H. (ed.) 1944. “Epilogue: The Nature of Naturalism”, in Naturalism and the Human Spirit, New York: Columbia University Press. Levinas, E. 1969.Totality and Infinity, Lingis, A. (trans.), Pittsburgh: Duquesne University Press. —. 1990. “A Religion for Adults”, in Hand, S. (trans.), Difficult Freedom: Essays on Judaism. Baltimore: John Hopkins University Press. Rahner, K. 1978. Foundations of Christian Faith: An Introduction to the Idea of Christianity. Dych, W.V. (trans), London: Darton, Longman and Todd. McDowell, J. 2002. “Response to Charles Lamore”, in Smith, N.H. (ed.), Reading McDowell on Mind and World. London: Routledge.

20

Chapter One

Nagel, T. 2012. Mind and Cosmos: Why the Materialist Neo-Darwinian Conception of Nature is Almost Certainly False. New York: Oxford University Press. Sellars, W. 1963. “Empiricism and the Philosophy of Mind”, in Science, Perception and Reality. London: Routledge &Kegan Paul. Weil, S. 1956. The Notebooks of Simone Weil, vol 2, Wills, A. (trans.), London: Routledge and Kegan Paul. Weinberg, S. 1999. “A designer universe”, New York Review of Books, 46, 16, 46-48. Wiggins, D. 1991. “A Neglected Position?”, in Haldane, J. & Wright, C. (eds.), Reality, Representation, and Projection. Oxford: Oxford University Press.

CHAPTER TWO FROM THE LIMITS OF SCIENCE TO THE LIMITS OF METAPHYSICS: A PHILOSOPHICAL EVALUATION OF FIONA ELLIS’ NATURALISTIC THEOLOGY MIKAEL LEIDENHAG

Naturalism has traditionally been construed as the antithesis of both theology and a religious conception of reality. It has been argued that naturalism in viewing science as the only mediator of knowledge invites reductionism, and so excludes morality, values, free-will and any ultimate purpose to nature. Today, however, we see a multitude of ways that scholars are attempting to move naturalism beyond this reductionist story. Non-reductive versions of naturalism have been proposed which emphasize the limits of science, the emergent order of nature, and the layered character of physical reality. Because these new forms of naturalism reject a scientific monopoly on knowledge, they seem to create space for a theological understanding of the workings of nature and enable a more harmonious relationship between theology and science. This chapter will focus on Fiona Ellis’s theological proposal and project of fusing Christian theism with a non-reductive version of naturalism. Ellis turns to John McDowell’s expansive naturalism to not only demonstrate the possibility of moving beyond scientism, but also to show how the framework of naturalism is fully consistent with a theistic understanding of nature. I will critically evaluate this attempted synthesis and argue that McDowell’s naturalism, and way of doing philosophy in general, seems to be in tension with Ellis’s theological ambitions. Moreover, McDowell’s quietist approach to philosophy risks silencing metaphysics, which undermines theological attempts at explicating God’s relationship to nature. I, therefore, suggest that Ellis should look elsewhere, and perhaps even in a non-naturalist direction. As will be seen,

22

Chapter Two

my critique is not aimed at Ellis’ theological proposal per se, but the naturalistic foundation of her theology. This chapter will first describe Ellis’s theological naturalism, and her usage of McDowell’s naturalistic outlook. Thereafter I will explain what I call the standard naturalist explanatory strategy (SNES), and show how Ellis and McDowell seem to depart from this form of strategy. I will then outline some of the ambiguities and problems in Ellis’ theological naturalism. This chapter ends with a brief discussion regarding a dualist alternative.

Ellis’s Theological Naturalism: Beyond scientific naturalism and supernaturalism Fiona Ellis argues that contemporary understandings of naturalism carry an unnecessary commitment to scientism, according to which science is the only path to knowledge. Given that the question of the existence of God lies beyond the reach of science, God’s existence is either denied, or the question pertaining to God’s existence is deemed meaningless. Consequently, Ellis argues that if theism is going to be a real option we must find new ways of conceptualizing “naturalism” that takes us beyond “scientific naturalism”, with its strong commitment to scientism. Ellis’ project consists of two tasks: The first task is to evaluate different forms of non-scientific naturalism to see if any of these ontologies can account for values. The second task is to show that how we relate to values is both necessary and sufficient for relating to God (Ellis 2014, 4). In this way, Ellis seeks to show how “naturalism and theism can both be true” (Ellis 2014, 3). The aim, however, is not to argue that a naturalist has to move in the direction of theism, but, more modestly, that such a “move is worthy of serious consideration” (Ellis 2014, 7). Naturalism can be a tricky concept to grasp, and many have suggested that naturalism is no longer a distinctive “-ism”. Ellis is aware of the fluidity of “naturalism”, but suggests that it can be both negatively and positively defined. At the negative level, the naturalist rejects supernaturalism and the idea of beings or realities transcending the natural order (Ellis 2014, 10). Positively, the naturalistic position is characterized by a strong commitment to the scientific method. It is because of this latter commitment that theism, by theists and naturalists alike, has been considered incompatible with naturalism. Hence, “we are left with the claim that supernaturalism is to be rejected because it allows that there is more to reality than what the scientist can comprehend” (Ellis 2014, 11). Referring to Mario De Caro and David Macarthur, Ellis suggests that a

From the Limits of Science to the Limits of Metaphysics

23

naturalism that elevates science over all other disciplines – and that, therefore, seeks to reduce all other disciplines to the natural sciences – is problematic as such a view cannot be supported by science itself (Yes, it rests on philosophical assumptions that cannot be determined to be true or false by the scientific method. It’s interesting that logical positivism isn’t taken seriously anymore and yet a remnant of it still survives within the idea of scientism). Moreover, to suggest that we can explain all of nature and human behaviour in terms of purely physical stuff is to attribute “supernatural powers to the physical stuff”, and this goes contrary to the naturalistic position (Ellis 2014, 17). This is the true paradox of scientific naturalism. In order to fully understand the natural world we need to invite more disciplines to join this epistemological endeavour. Scientific naturalism, it seems, is too strict for its own good. We need a liberalized, relaxed, and expansive naturalism.

Expansive naturalisms The first version of expansive naturalism that Ellis considers is “expansive scientific naturalism”. This naturalism “grants that there is more to science than physics, and more to it than natural science” (Ellis 2014, 21). Thus, expansive naturalism is willing to include the human sciences within its methodology, and it rejects the epistemological monopoly of the natural sciences. However, despite being more open to an input from the humanities, scientific expansive naturalism cannot find a place for values. On this version of naturalism, which Ellis critiques, “human values are those which are valued by human beings” (Ellis 2014, 22). Thus, values are to be defined in terms of human interests. Ellis is sceptical towards such naturalism, suggesting that it is unable to account for the normativity, or “oughtness”, of values. It can, as Ellis points out, be in someone’s interest to be cruel (Ellis 2014, 34). Moreover, this kind of expansive naturalism, appealing to our interests, does not succeed in spelling out in detail the nature of values. There is an objective moral dimension which is missing from this account. Peter Railton, a proponent of expansive scientific naturalism, is aware of the problem of defining values in terms of human interests. He argues instead that values and moral norms should not be reduced to the opinions of individual persons, but should be judged from a social point of view. Railton seeks to maintain a realist view of moral discourse and practice, and avoid a relativist view of values, by introducing some independent standard of judgement. Railton takes values to be reducible to “complex social-psychological phenomena”, and the aim is to show “how the

24

Chapter Two

relevant phenomena “may bring into being in individuals a notion of obligatoriness that present itself to them as objective and independent of their personal inclinations” (Ellis 2014). In claiming so, Railton hopes to show how values, while being real, are not “spooky”, “intolerably odd”, or somehow ontologically different from other phenomena in nature. Ellis applauds Railton’s ambition to retain a realist view of values. Yet, Railton’s approach comes off as too weak. It has the negative effect, argues Ellis, of squeezing moral properties out of the picture by reducing such properties to social-psychological phenomena. Railton, therefore, is still on the reductionist side of the spectrum as he merely chooses to reduce values to something else (Ellis 2014, 48). Ellis moves on to consider John McDowell’s version of naturalism. Like other expansive naturalists, McDowell seeks to “defend the idea that moral properties are sui generis”, by demonstrating the objectivity of values and showing how values are “part of the fabric of the world” (Ellis 2014, 51, 52). For McDowell, values are an inherent part of an enchanted nature, which transcends the limits of science, but does not contradict the method or findings of science. Our practices, on this view, are “already irreducibly value-involving” (Ellis 2014, 66). In this way, McDowell’s naturalism is anti-scientistic (by denying that science has a monopoly on ontology), but not anti-scientific (as it still respects insights gained from science). However, by saying that values are irreducible to something physical we are not, according to McDowell, inviting supernaturalism or dualism. Our capacity to respond to the domain of values is a rational capacity that stems from our human nature, rather “than from any supernatural addition which gives us a mysterious separate involvement in an extra-natural world” (Ellis 2014, 63). Our capacity to engage with values, therefore, requires nothing extra-human, only our human nature. This capacity for recognizing values is fully natural, and thus consistent with naturalism. Indeed, this capacity refers to what McDowell calls second-nature. First-nature is simply our biology, our physical makeup, which falls within the realm of law and is studied by the natural sciences. Second-nature, whilst being connected to first-nature, goes beyond a purely scientific and law-focused description of reality and refers to our learned ability to respond to reasons and values. Ellis summarises this position well, “So the propensities we possess at birth or acquire through a process of biological maturation belong to our first nature, and those which are imparted by education, habituation, or training belong to our second nature” (Ellis 2014, 62). Nature is, in a sense, enchanted according to Ellis, who endorses McDowell’s expansive naturalism. McDowell’s naturalism takes us

From the Limits of Science to the Limits of Metaphysics

25

beyond a scientistic conception of nature, and towards a view of nature as infused with values. McDowell concludes that “there is more to explanation than scientific explanation, more to nature than what the scientist comprehends, and that once we concede these points we arrive at a partially re-enchanted conception of nature” (Ellis 2014, 80). Ellis bravely wonders if we should stop here and say that nature is enchanted with values, or if we could take this idea further and say that nature is divinely enchanted. An expansive naturalist, like McDowell, would be sceptical about “the possibility of extending his arguments in a theistic direction...” (Ellis 2014, 86). McDowell claims that such a broadening would lead to a worrisome ontological expansion. Ellis, on the contrary, maintains that naturalism has much to gain from a constructive engagement with theism and Christian theology. Thus, Ellis is not content with just pointing out the deficiencies of naturalism, but seeks to remedy these problems by fusing naturalism with theism. A theistic understanding of McDowell’s naturalism suggests that God has granted human beings the natural capacity of recognizing values in nature, and through this recognition we can relate to and be inwardly transformed by God (Ellis 2014, 91). Expansive naturalism as formulated by McDowell, argues Ellis, “grants us the right to talk about a divinely enchanted world...” (Ellis 2014, 147). Yet, there is the question of how far it is possible to go in a theistic direction and how much theism expansive naturalism can tolerate. The refusal by naturalists to fuse their ontology with theism is because God is imagined as a “separable something else which stands opposed to nature” (Ellis 2014, 178). This assumption, however, can be challenged according to Ellis who suggests that God is both immanent and transcendent. She writes: “The difference with God, however, is that He remains radically distinct from anything within the world even whilst retaining the most intimate connection with it” (Ellis 2014). God is not reducible to the world and values, and the values we encounter “have their source in God” (Ellis 2014, 193). Ellis further argues that a theistic position is compatible with science, and that science can neither demonstrate nor exclude the existence of God as God lies beyond the scope of scientific investigation (Ellis 2014, 151). Hence, if the expansive naturalist denies the existence of God on the basis of an alleged conflict between God’s existence and science, he/she is succumbing to the same philosophical mistake as the scientific naturalist who denies the existence of values. What difference does God make, then, for expansive naturalism? In one sense, God makes no difference at all, according to Ellis. But, in another sense, God changes everything. God

26

Chapter Two

“does so because He is the source of things” (Ellis 2014). God enriches nature, and provides the ground for values, and God’s activity is revealed at the level of values and morality (Ellis 2014, 189).

The Standard Naturalist Explanatory Strategy We have now seen how Ellis utilises McDowell’s framework in order to formulate a non-scientistic version of naturalism that can lend the conceptual space for bridging theological discourse with naturalism and so enable a fruitful dialogue between the two. McDowell, as I will explain later, seems to shy away from metaphysical issues, and he departs from what can be called the Standard Naturalist Explanatory Strategy (SNES). What does it mean to adopt SNES? To do philosophy in a naturalistic manner typically involves dealing with “placement issues”. As Huw Price puts it, “If all reality is ultimately natural reality, how are we to ‘place’ moral facts, mathematical facts, and so on? How are we to locate topics of these kinds within a naturalistic framework...?” (Price 2011, 187). This is the problem of understanding how a particular object, property, or fact can be a purely natural thing, and the kind of thing that is discoverable by science or at least consistent with a scientific take on reality. Another way of framing SNES is to say that naturalists seek to accommodate higher-level properties within the natural order. Naturalist projects in philosophy “are attempts to accommodate various kinds of discourse – e.g., moral discourse, mental discourse, mathematical discourse, semantic discourse – within a naturalistic worldview” (Horgan 1994, 303). A naturalistic moral realist would, for example, hold that moral facts can be characterized in terms of non-moral facts (purely physical facts). Indeed, these placement problems or accommodation projects give rise to a number of different strategies: the reductionist simply reduces higher-level facts to physical facts, the eliminativist eliminates higher-level facts from his/her ontology, the emergentist seeks to explain how higher-level facts are dependent on, yet irreducible to, physical facts. SNES, therefore, can lead to variety of strategies, including both reductionist and non-reductionist versions, for handling non-scientific discourses. Frank Jackson argues that this is not an optional task for a naturalist, but a serious naturalist will “seek a comprehensive account of some subject-matter – the mind, the semantic, or, most ambitiously, everything – in terms of a limited number of more or less basic notions” (Jackson 2000, 4). This, for Jackson, is to do serious metaphysics and such metaphysics always faces different placement problems.

From the Limits of Science to the Limits of Metaphysics

27

This is the standard way of conducting naturalistic metaphysics, but it is a method and philosophical approach that a few notable philosophers take issues with – John McDowell being one of them. McDowell takes this approach to be a case of “bad”, “shallow”, and “primitive metaphysics”, which has caused a great deal of anxiety and confusion in contemporary philosophy. We need to change our methodological approach if there is to be any real progress in philosophical debates. Drawing on Wittgenstein, McDowell wishes for the philosopher to adopt a therapeutic approach to metaphysical issues; philosophy is not about solving problems (at least not in the traditional sense of the word), but clarifying the problematic assumptions underlying our intellectual traditions. This Wittgensteinian therapy does engage philosophical problems, “even if not quite the problems that traditional metaphysics sets out to address” (Price 2015, 310). As McDowell writes, “Wittgenstein’s quietism is not a refusal to engage in substantive philosophy in the face of what everyone has to accept as genuine problems. It is an activity of diagnosing, so as to explain away, some appearances when we are confronted with genuine problems. The supposed problems disappear, leaving no need for theory constructive to make things ‘less mysterious’” (McDowell 2009, 371). In this way, McDowell refuses to take part in the traditional “metaphysical matchinggame”, and to adopt SNES. The naturalistic strategy, for McDowell, is to show how particular phenomena, such as values and intentionality, need no metaphysical explanation. Instead, they already belong to the natural order and are constitutive of human nature, and this realization can finally provide some well-needed peace to philosophy. Therapeutic philosophy is designed to relieve us of the travails of positive, or substantive, philosophy. As McDowell says, “Quietism does indeed urge us to engage in certain supposed tasks, but precisely because it requires us to work at showing that they are not necessary” (McDowell 2009). SNES asks us to engage in a form of positive philosophy, trying to square particular higherlevel phenomena with that of scientific discoveries and vocabularies, or explaining how such phenomena can be derived from the major theories of science. This task, however, is neither fruitful nor required, according to McDowell. I will argue, however, that this quietism is problematic both for the internal coherency of McDowell’s philosophy as well as for Ellis’ theological project.

28

Chapter Two

SNES and the Problem of Competing Ontologies McDowell proposes a non-reductionist version of naturalism. It is intriguing because it not only seeks to break the spell of reductionism, but it also seeks to correct a broader and more general faulty metaphysical thinking within philosophy. It refuses to play the game that, for McDowell, inevitably ends up in either scientistic or overly dualistic conclusions. The goal is to have a relaxed naturalism that recognizes the limitations of science while at the same time acknowledging the particularity of human creatures as embedded within the natural order. This is why Ellis turns to McDowell, because to achieve the proper vindication of theological notions requires no more than open-mindedness and a resistance to scientism – a resistance offered by McDowell’s metaphysically modest philosophy (Ellis 2014, 4). What is interesting about McDowell’s naturalism is that it does not simply emphasise the limits of the natural sciences, but it stresses the conceptual limitations of the metaphysical enterprise as a whole. Hence, McDowell’s and Ellis’ seeming rejection of SNES is to be seen as both a denunciation of a scientistic conception of science and with it the kind of metaphysics (i.e. substantive or positive philosophy) that enables scientism in the first place. But, is a McDowelltype philosophical practice successful? It should be noted that both McDowell and Ellis employ notions that are in need of metaphysical clarification. The point of McDowell’s naturalism is to show how “knowing and thought” can be considered natural phenomena, and how our knowledge of values is the product of our second nature. The reason for the often naturalistic rejection of mental phenomena is because of the contrast between nature, as described by science, and “the space of reasons” (McDowell 2004, 92). By “space of reasons”, McDowell refers to the possibility of justifying what one says, which is an essential feature of an adequate theory of rationality. Naturalism has, in emphasising the “spaces of causes” over “spaces of reasons”, turned a core feature of human rationality into a metaphysical anomaly. McDowell suggests that “To avoid conceiving thinking and knowing as supernatural, we should stress that thinking and knowing are aspects of our lives”, and “they are part of our ways of being animals” (McDowell 2004, 93). Yet, he claims that mental phenomena are not strictly physical. McDowell insists that such phenomena are sui generis with regard to the rest of nature, and they go beyond natural law and our current conception of the natural (McDowell 2004, 97). The space of reasons is sui generis and cannot be captured within the realm of law. McDowell, despite seeking to preserve the realness and causal efficacy of the mental, seems to suggest that the

From the Limits of Science to the Limits of Metaphysics

29

causality of the mental is ontologically different from causality at the physical level. Mind, on McDowell’s view, occupies its own level of reality that cannot be considered purely physical. I am not claiming that McDowell succumbs to dualism, but he leaves us with a view of the mental that many naturalists would find problematic. Naturalists generally seek to harmonize the level of the mental with that of the physical (as exemplified in SNES), but this attempt is resisted by McDowell who thinks that such harmonization opens up the door for reductionism. In this way, McDowell does not reconcile the mental and the physical, but leaves sui generis mental phenomena ontologically unexplained. Similarly Ellis in her discussion of the issue of queerness seems to leave values unexplained. As I have described, a key aspect of Ellis’ theological project is the idea that how we relate to values is similar or analogous to how we relate to God. Thus, Ellis’ theology is forced to deal with the question of how values fit into a naturalistic conception of nature. There is in her view of values a form of causal language that is in need for further clarification. Values do not exist in a separate supernatural realm, according to Ellis. Nevertheless, they influence us in some particular way. Values do not “interfere with the normal, causal run of things”, but they have some influence upon us. As Ellis writes, “we need to be able to make sense of the idea that values can have a transformative effect upon us, that they can lead us to reappraise what is and ought to be of importance to us...” (Ellis 2014, 67). The moral context involves, at least to an extent, causal language. Our capacity for knowing values, as Ellis states, is continuous with ordinary ways of knowing. Yet this form of knowing is not reducible to “non-moral ways of knowing” (Ellis 2014, 68). However, it is not obvious how this balance, between first-nature and second-nature, is possible and achievable within a naturalistic framework. It seems as if Ellis’s and McDowell’s proposals amount to a form of mysterianism; the emergence of second-nature from first-nature, and the place of values in nature, is simply a mystery. In the end, our ability to know values is simply asserted or taken for granted on this account, but no explanation is given for how this is possible within a naturalistic framework. A McDowell-type philosopher could, however, retort by saying that my critique (or, more precisely, call for explanation) is based on faulty metaphysics. That is, I am assuming in my critique the kind of SNES that McDowell/Ellis seems to reject. However, I suggest that SNES, or some similar strategy, is a necessary component for any adequate naturalism. A naturalism that leaves particular phenomena unexplained encounters what I call the Problem of Competing Ontologies (PCO). This problem describes the metaphysical and epistemological question of why we should

30

Chapter Two

prefer some ontology O1 over ontologies O2, O3, O4, etc. McDowell argues that mental phenomena are irreducible and sui generis. But why should we take this to support naturalism as opposed to some of its ontological rivals? The problem is that the irreducible and sui generis character of the mental remains fully compatible with other non-naturalist frameworks, such as substance dualism, supernaturalism, emergent dualism, hyperdualism, and panpsychism. The problem is that the claim of irreducibility for the mental does not necessarily give us a naturalistic understanding of mental phenomena. Hence, it is not enough for a naturalist to point at certain properties and show in what ways they can escape the reductionist story and how it is possible to resist scientism; the naturalist also has to explain why property X is a better fit for naturalism than other competing ontological frameworks, otherwise this problem remains. It seems, therefore, that SNES is not optional for the committed naturalist. There are ontologies that would, similar to McDowell’s and Ellis’ proposals, include mental properties in their list of things existing; hence, to engage in SNES seems to be necessary to avoid PCO. It further seems as if PCO becomes relevant in Ellis’ account of values. Ellis, through McDowell, attempts to explain values in an Aristotelian fashion, that our interaction with the domain of values requires nothing more than “ethical upbringing” (Ellis 2014, 63). Hence, our second nature allows us the capacity for moral reasoning and reflection. But, this way of explaining the reality of values seems insufficient. It seems as if Ellis’ account confuses ontology with epistemology. That is, in outlining how humans can epistemologically know certain values to be true, this account does not explain ontologically how it is that sui generis values can be a part of the natural order, or how they fit a naturalistic description of nature. This does not explain how to make sense of values in a purely natural world, or how it is that these values which we are brought up to believe in are not simply figments of our imagination, or constructs of our society. Far from exclusively supporting naturalism, the idea that we have the epistemological capacity for knowing values seems compatible with non-naturalist frameworks, and so PCO shows up again. Once again, it seems as if McDowell/Ellis are forced back into SNES. They are obliged to explicate the necessity of a naturalistic interpretation if naturalism is to be more than a mere option in the metaphysical debate about the realness of values. Yet, on McDowell’s philosophical quietism, to engage in SNES is not possible as the main goal, or the only coherent goal, of the intellectual discipline of philosophy is to uncover problematic assumptionsto provide therapy-and not to present a true account of the world, which is the concern for substantive philosophy. Indeed, it seems as if McDowell’s

From the Limits of Science to the Limits of Metaphysics

31

Wittgensteinian leanings prohibit this project from achieving one important thing: to show why it is that a naturalistic conception of values and mind is correct and superior to non-naturalist or anti-naturalist rivals.

McDowell’s quietism and God’s relation to nature We shall now turn to Ellis’ theistic proposal. What does she mean by “God” and how does God relate to the natural order? She makes it clear that her proposed theism goes beyond onto-theology. God is not just one more item within the “larger household of reality” (Ellis 2014, 115). God, in virtue of being the creator and sustainer of everything, is not a part of the natural world, and God cannot be identified with anything in nature (Ellis 2014, 95). Ellis stresses the otherness of God, saying that “He is neither within the world nor beyond it” (Ellis 2014, 151). She further argues that “we must resist thinking of God as a being who stands over and above the world in a realm which cannot be intelligibly related to it” (Ellis 2014, 197). Moreover, God “remains radically distinct from anything within the world even whilst remaining the most intimate connection with it” (Ellis 2014, 5). How is God’s activity imagined within Ellis’ theological proposal? Ellis claims that God does not interfere in the natural cause of things as God is not a competing cause among other causes. Yet, she seeks to defend a robust form of theism and avoid deism. Thus, God must be allowed to make some difference in reality and “nature is irreducibly open to His communicative action” (Ellis 2014, 193), which is manifested in the inwardly transformation of human creatures, a form of divine action enabled by a broader conception of nature. It seems as if Ellis seeks to combine three components in her theistic proposal: a) an adequate balance between transcendence and immanence, b) a non-interventionist view of God’s action, c) a moral-focused account of divine action that is connected to the ethical transformation of human creatures. I find it rather surprising that, given her theological concerns and her desire to fuse theism and naturalism, there is such a lack of engagement with the position commonly referred to as “panentheism”, as articulated by Philip (1997; 2004) and Arthur Peacocke (1993; 2004). Proponents of panentheism share many of Ellis’ concerns and ideas: the problem of divine interventionism, philosophical problems of dualism, and that science, to some extent, mediates a naturalistic conception of nature. In a similar way to Ellis, panentheism seeks to maintain a balance between, on the one hand, God’s transcendence in relation of to the physical, and, on the other hand, God’s presence, activity and immanence within the natural order. A more clear discussion in relation to panentheism would

32

Chapter Two

have helped to clarify Ellis’ view even more on naturalism and theism. This is not a critique of Ellis; rather it is to ask for a dialogue that can help to shed light on her philosophical and theological commitments and to place her proposal on the map of available naturalistic construals of theism. We can see that Ellis employs a form of divine action through her theistic account of values, the idea that “values have their source in God” (Ellis 2014, 193). God’s active role in the moral realm also becomes evident in the claim that “God can inwardly transform us, that our natural being is perfected in this manner” (Ellis 2014, 115). Ellis is aware that the appeal to divine action is rejected by naturalists. A naturalist would presumably say that it is unnecessary to invoke God-talk “because we can explain things without a theistic detour” (Ellis 2014, 150). Ellis argues that this “either/or framework” – the idea that we must either believe in divine action or natural causation – must be abandoned. The idea of God’s omnipresence, for Ellis, means that God does and does not make a difference in reality. Because of God’s omnipresence, God cannot be considered a cause among other causes. This, as was explained earlier, means that God makes no difference at all. Yet, in another and more significant way, God makes all the difference. How do we make sense of this position? I am very supportive of Ellis’ ambition to harmonize divine action and natural causation, but she seems unable to explain how this turn from either/or to both should be understood. Indeed, Ellis seems aware of the weakness of this explanation. She writes concerning the problems of explicating God’s active role in nature that “our inquiries must terminate inevitably and appropriately at the level of mystery...” (Ellis 2014, 151). I think that this answer might come across as unsatisfying for the concerned naturalist. A naturalist, including an expansive naturalist like McDowell, seeks to develop a unified view of nature, whereby everything is connected to the natural domain (conceptualized in a non-scientistic way). Ellis’ theistic explanation of values would be considered problematic according to a naturalistic framework as it now locates (the source of) values in an extra-natural realm. The reason for the existence of values is because God instantiated those values, and so we have some properties that transcend the physical domain (they are sui generis with regard to the physical) which cannot be explained solely in terms of the potentialities of nature. This explanatory split should worry a naturalist. What Ellis needs to do in order to release some of this “naturalistic anxiety” is not to appeal to mystery, but to metaphysically explicate God’s active role in relation to nature, to show how divine action does not introduce a problematic dualism. In some way, this would take Ellis’

From the Limits of Science to the Limits of Metaphysics

33

proposal back to SNES. As previously suggested, SNES, or something very close to it, is in fact a necessary part of naturalism, even an expansive interpretation of the commitments of a naturalistic worldview that we find in McDowell’s philosophy. This need to further illuminate the relationship between naturalism and theism is, however, resisted by McDowell’s approach. As he suggested, the main task for the philosopher is, through a therapeutic engagement, pinpoint problematic assumptions. To abandon positive/substantive philosophy and adopt a more therapeutic approach is McDowell’s advice. But, as I argued earlier, a complete abandonment of positive philosophy, trying to show how particular phenomena fit a naturalistic worldview, leads to the problem of competing ontologies; that is, naturalism merely becomes one possible alternative among the available of ontologies. To focus solely on a therapeutic approach leads to a form of philosophical quietism, which is devastating if one seeks to fuse theism and naturalism and to illustrate how it is possible (and even desirable) to allow God into a naturalistic conception of reality. Hence, we need to explain how McDowell’s naturalism can be joined with theism. Yet, such an explication is not allowed on McDowell’s own approach as it would take naturalism towards substantive metaphysics and the very positive philosophy that creates so much confusion in philosophical discourses. I suggest that McDowell’s quietism, whilst offering a sound critique of scientism, is not the way forward for theology. It not only stresses the limits of science, but metaphysics as such, although some metaphysical reflection is needed to demonstrate a relationship of harmony and mutual understanding between Christian theism and naturalism. In the next section I will consider another framework that can achieve some of Ellis’ theological ambitions by providing a more sound philosophical foundation. I will consider the integrative dualism of Charles Taliaferro which, in a similar way to expansive naturalism, takes seriously the embodied nature of human beings.

From expansive naturalism towards holistic dualism Expansive naturalism, while it at first glance appears to be a promising escape route from scientism, invites a problematic quietism. As was also argued, McDowell’s naturalism seems difficult to justify, that it remains unclear why we should prefer naturalism over some non-naturalistic alternatives. Expansive naturalism is a contender on the metaphysical market, but there is a striking lack of positive reasons for philosophy, let alone Christian theology, to embrace this proposed naturalism. In this

34

Chapter Two

section I will put forward a few reasons for why Christian theology should not fear, but wholeheartedly embrace, dualism. Moreover, I will seek to show how McDowell’s naturalism is fully consistent with a dualistic ontology. Despite opting for a naturalistic understanding of human nature, and our ability to know values, Ellis comments Taliaferro’s integrative dualism in fairly positive terms. Taliaferro’s dualism is not Cartesian, and it denies the strict separation between mind and body. While remaining distinct, the mind and the body are ontologically connected and constitute a holistic unity. Integrative dualism is summarised by Taliaferro in this way: “According to the version of dualism I accept, persons are themselves nonphysical individuals. As embodied beings, however, persons are not ghosts or mere accessories to their bodies. Persons are integrally related to their bodies so that the person and his or her body function as a singular unit mentally and physically” (Taliaferro 1994, 114-115). Hence, this dualism rejects the materialist reduction of personhood to the physical, mere dual-aspect theory, and a radical Cartesian division between mental life and physical constituents. The “person and the body differ metaphysically in the sense that they are separable individuals, and yet they function as a singular reality as an embodied person” (Taliaferro 1994, 116). Contrary to McDowell, and other naturalists, Taliaferro stresses the compatibility between substance dualism and human embodiment. So, if it is the case that Taliaferro’s dualism allows for a holistic unity of the person without introducing a problematic gap between mind and matter, then why is there a need for theology to move in the direction of naturalism? A naturalist could, of course, argue that dualism is too problematic and that it leads to a host of philosophical problems. However, as Ellis concedes, the “usual arguments – the problems of interaction and individuation – tend to depend on considerations which are persuasive only to one who is already convinced that substance dualism is a nonstarter” (Ellis 2010, 349). Moreover, as she also writes “it is not obvious that McDowell’s conception of human nature is without difficulty” (Ellis 2010, 349). It seems, therefore, as if Ellis leaves open the possibility of construing human personhood in dualist terms, albeit a dualism that respects the embodied character of human nature. McDowell not merely fails in justifying his preferred naturalism, but his dismissal of dualism is question-begging. McDowell suggests that second-nature is no less natural than our first-nature, and that there is “no need to offer to make a connection between them beyond their both being natural” (McDowell 2009, 221). This is a “relaxed naturalism” because it does not venture into the domain of explanations. Rather, as McDowell

From the Limits of Science to the Limits of Metaphysics

35

claims, it merely follows our philosophical intuitions. Nevertheless, as Howard Robinson argues, “It is not intuitively obvious that the human capacity for abstract thought, and for aesthetic and spiritual inspiration, are ‘phenomena that are biological’, especially when it is being specifically denied that they can be accommodated within the natural science of biology” (Robinson 2014, 165). It is not obvious why, given that McDowell denies that second-nature can be scientifically explained, phenomenal properties and our higher-order mental capacities cannot be construed along the lines of dualism. Dualism is simply ruled out from the get-go on McDowell’s naturalism. Yet expansive naturalism, as construed by McDowell, seems entirely consistent with dualism. He emphasises the sui generis character of the space of reasons, that it cannot be accounted for, or explained through the realm of law. It escapes scientific reduction. Whatever biology is able to describe, whatever phenomena fall within the confines of its methodology, our capacity for reasoning and to know values remains irreducible and beyond a scientific epistemology. The limits of science, in this way, help to save the mental domain from the acid of reductionism. But as I explained above, McDowell does little to explicate the ontology of the mental or how it can be construed within a naturalistic framework. Indeed, he merely seems to assume the validity and reality of a first-person perspective-that we can reflect, reason, and mentally “hook up” with reality – in his account of second-nature. He rejects what is often called the identity theory of mind, that mental processes are identical to brain processes. Given that McDowell affirms the reality of the mental he clearly rejects eliminativism, that the mental should be eliminated given that it cannot be accounted for in purely scientific terms. He also claims that, “Second nature could not float free of potentialities that belong to a normal human organism” (McDowell 1994, 84). Thus, given that mind is connected to and irreducible to the biological level, perhaps McDowell subscribes to some type of property dualism. On this reading, McDowell believes that even though the body is purely natural, it possesses both physical and mental properties. This position further stresses the dependency of mental properties upon the biological, in that mental happenings cannot occur without a physical occurrence (a particular brain state). Perhaps my intention to classify McDowell’s naturalism is to put too much metaphysics into philosophy, which goes against the spirit of quietism. However, some clarification is needed if McDowell wishes to show why naturalism is superior to dualism, or why his expansive naturalism is not a dualism in disguise. Indeed, McDowell’s case for mind seems to speak more for dualism than naturalism. His rejection of

36

Chapter Two

scientism and reductionism, while maintaining that the human mind is beyond the sciences and naturalistic explication, fits dualism quite well; in particular Taliaferro’s more holistic understanding of dualism that upholds human embodiment. Moreover, Taliaferro, contrary to McDowell, does not assume a Wittgensteinian quietism and so does not as easily encounter the problem of competing ontologies. Taliaferro stays within a more traditional metaphysical setting, but can avoid the problems of scientism and the various scientific attempts at reducing second-nature to the physical. McDowell provides a way to dialogue with theology through his rejection of scientism, according to Ellis. Yet, Ellis is aware of the fact that McDowell’s naturalism is not without its share of problems. Given Ellis’ openness to integrative dualism, it remains unclear why she seeks to address the problems in theology through specifically naturalism. Ellis adopts an expressive model of God, whereby the world is seen as “expressive of God as a mind” (Ellis 2010, 342), and the challenge is to convince the naturalist to take this assertion seriously. I suggest that a dualistic conception of mind that dares to talk about mental properties as nonphysical and sui generis with regard to the realm of law clearly demonstrates the relevance of a theistic understanding of reality. One way to show the positive relevance of the reality of mind for theism can be found in the writings of J.P. Moreland. His argument, simply called “the argument from consciousness” (AC), portrays the deductive route from consciousness to God through a number of metaphysical alternatives and steps. “Mental events are genuine non-physical mental entities that exist” and there must be some explanation for the correlation between mental occurrences and physical events. Moreland suggests that we have two explanatory alternatives before us: personal explanations and natural explanations. Thereafter, Moreland considers several possible natural explanations for consciousness, and suggests that they all fail in explaining the nature of mental states (Moreland 2008, 53-174). Hence, mental properties cannot be the kind of properties studied and examined by science, which means that a reductionist reframing of consciousness fails. This conclusion seems to be shared by both McDowell and Ellis given their resistance to scientism and reductionism. Moreland, however, does not espouse the kind of quietism that is associated with expansive naturalism, but instead takes metaphysics in a non-reductionist and theistic direction. Moreland is, in this way, doing what McDowell calls positive philosophy while seeking to save the mental from scientistic reductionism. Having established the failure of scientific explanations for the phenomenon of minded creatures, Moreland concludes that the alternative metaphysical conclusion must be correct: consciousness requires a

From the Limits of Science to the Limits of Metaphysics

37

personal explanation. Moreland writes, “By way of application, the adequacy of a personal explanation does not consist in offering a mechanism, but rather, in correctly citing the relevant person, his intentions, the basic power exercised, and in some cases, offering a description of the relevant action plan” (Moreland 2008, 50). Such an “action plan” would include a model of God’s intentions and God’s intentions for creating this world, a world particularly suitable for complex creatures. To go from a personal explanation to a theistic understanding is, according to Moreland, rather uncontroversial, even though a Humean objection could be raised at this point. This objection asks us how one should understand the type, size and number of deities involved in the becoming of consciousness? Despite this sceptical objection it is uncontroversial to claim that a personal explanation, all things being equal, better supports a theistic metaphysics than naturalism, be it reductive or nonreductive. A proper theory of consciousness that affirms the sui generis character of the mental and cognitive abilities of complex creatures thus invites theism into the picture and gives positive reasons for, as Ellis puts it, viewing nature as an “expression of God”. The purpose of this paper has been to challenge McDowell’s naturalism, and Ellis’ theological appropriation of it. One key conclusion of this discussion is that the problems that reductionist and scientistic frameworks pose for consciousness, and second-nature, are not best met by silencing metaphysics or completely abandoning substantive or positive philosophy. It is not an effective strategy and it leads to even more philosophical problems. What is needed is a sound metaphysical framework that can avoid a creeping reductionism while remaining honest about the epistemological limitations of the natural sciences.

Conclusions I have critically evaluated Fiona Ellis’ attempt at framing God’s relationship to the natural order, and God’s active involvement in the domain of values, through John McDowell’s naturalistic philosophy. This chapter sought to show that this is not a successful combination. While Ellis’ theological project in itself is admirable and needed, it seems as if expansive naturalism cannot lend the appropriate resources for furthering this project. Indeed, McDowell’s philosophy is committed to a Wittgenstein-inspired quietism that is not just philosophically problematic, but prohibits a theologian from metaphysically explicating and deepening our understanding of God’s involvement in the natural domain and in individual lives.

38

Chapter Two

I, therefore, suggested that Ellis should look elsewhere, including in dualist directions. She seems open to, in particular, Taliaferro’s integrative dualism, claiming that it has the potential to provide an ontological foundation for second-nature and the reality of the mental. It was seen that Taliaferro employs a metaphysical approach that is different to that of McDowell. Rather than shying away from constructive metaphysics, Taliaferro seeks to correct faulty metaphysics and establish a path between reductionism and naive dualism. Taliaferro’s dualism would be of great benefit for furthering Ellis’s theological vision.

References Clayton, P. 1997. God and Contemporary Science. Edinburgh: Edinburgh University Press. —. 2004. Mind and Emergence: From Quantum to Consciousness. New York: Oxford University Press. Ellis, F. 2010. “God and other minds”, Religious Studies, 46, 331-351. —. 2014. God, Value, and Nature. Oxford: Oxford University Press. Horgan, T. 1994. “Naturalism and Intentionality”, Philosophical Studies 76, 301-326. Jackson. F. 2000. From Metaphysics to Ethics: A Defence of Conceptual Analysis. New York: Oxford University Press. McDowell, J. 1994. Mind and World. Cambridge: Harvard University Press. —. 2004. “Naturalism in the Philosophy of Mind”, in De Caro, M., and Macarthur, D. (eds.) Naturalism in Question. London and Cambridge: Harvard University Press. —. 2009. “Wittgensteinian ‘Quietism’”, Common Knowledge 15, 3, 365372. Moreland, J.P. 2008. Consciousness and the Existence of God: A Theistic Argument. New York: Routledge. Peacocke, A. 1993. Theology for a Scientific Age: Being and Becoming – Natural, Divine, and Human. Minneapolis, MN: Fortress Press. —. 2004 “Articulating God’s Presence in and to the World Unveiled by the Sciences”, in Clayton, P., and Peacocke A. (eds.) In Whom We Live and Move and Have Our Being: Panentheistic Reflections on God’s Presence in a Scientific World. Michigan and Cambridge: William B. Eerdmans Publishing Company. Price, H. 2011. Naturalism Without Mirrors. New York: Oxford University Press, Inc.

From the Limits of Science to the Limits of Metaphysics

39

—. 2015. “Idling and Sliding Toward Philosophical Peace”, in Gross, S., Tebben, N., and Williams, M. (eds.) Meaning Without Representation: Essays on Truth, Expression, Normativity, and Naturalism. Oxford: Oxford University Press. Robinson, H. 2014. “Naturalism and the Unavoidability of the Cartesian Perspective”, in Lavazza, A., and Robinson, H. (eds.) Contemporary Dualism: A Defense. New York and London: Routledge. Taliaferro, C. 1994. Consciousness and the Mind of God. Cambridge: Cambridge University Press.

CHAPTER THREE AN ELEPHANT IN THE ROOM: WHY THE CAUSAL JOINT IS STILL WORTH TALKING ABOUT SARAH LANE RITCHIE

Introduction Divine action remains one of the key issues shaping the field of science and religion. As the explanatory power of the natural sciences renders so many phenomena explicable, the theological affirmation that God acts in the world has seemed, to many, increasingly problematic. Are there limits to what science can tell us about how God does or does not act in the world? How, exactly, can immaterial spiritual realities be said to affect natural processes, especially when these processes are apparently intelligible in physical terms alone? Perhaps infamously, this question has come to be discussed in terms of the “causal joint”: the theoretical nexus at which an immaterial God meets physical realities to bring about specific events in the natural world. In recent decades, much of the science and religion conversation has centered on the search for a scientifically plausible causal joint in underdetermined natural processes (quantum mechanics, chaos theory, emergence, complex systems theory, etc.). The focus of this discussion has been the so-called Divine Action Project (DAP), an impressively rigorous joint endeavor between the Center for Theology and the Natural Sciences (CTNS) and the Vatican Observatory. Spanning a period of fifteen years, DAP contributors produced a significant body of literature on the theme, “Scientific Perspectives on Divine Action”. While few conclusions were reached by the widely diverse scholars involved, it is clear that the general metaphysical assumptions of many in the DAP have significantly influenced the science and religion field more broadly. In particular, scholars have focused their efforts on achieving what Philip Clayton has

An Elephant in the Room

41

called “maximum traction” between theology and the sciences-deliberately bringing the disciplines into contact with each other, and (often) subjecting theology to scientific critique (Clayton 2008, 54f). In recent years, however, many have begun to question the value of such causal joint projects as those developed by DAP participants. These scholars challenge the metaphysical presuppositions inherent in the very questions framing the regnant divine action conversation, such as the paradigmatic “intervention vs. nonintervention” framework. These assumptions involve the fundamental relationship between God and nature, and the question of how to understand the material world: what does it mean to be natural? The argument goes that instead of using science to locate open spaces wherein God can act in an otherwise autonomous natural world, one should instead allow theology to (at least partly) shape the basic categories, distinctions, and fundamental assumptions involved. In other words, science is necessarily constrained: it cannot identify or articulate the fundamental relationship between God and the natural world. In these views, the search for a causal joint is wrong-headed, as it fails to understand the dynamics of the God-world relationship. This “theological turn” in the divine action debate has given rise to proposals involving pneumatology, eschatology, and a variety of strong theistic naturalisms (for example). Though such approaches differ greatly from each other, one commonality they share is the primary emphasis placed on theology and metaphysics in contextualising and limiting the scope of science – rather than the other way around. The upshot of this theological turn is that because such approaches are independent of scientific findings, they are immune to falsification and the advance of scientific knowledge. Of course, this is also their greatest challenge: these models are essentially divorced from the most successful knowledge-seeking enterprise known to humanity. While theological approaches to divine action do indeed seem to be an important step in the right direction, I suggest that they have not yet dealt effectively with the causal joint. That is, these models often deny the need for, or importance of, a causal joint; but this denial does not necessarily answer the question, “How can a transcendent, immaterial being interact with physical processes?”. My goal here is twofold; first, I survey the key issues in causal joint theories, and articulate the metaphysical commitments inherent (but rarely expressed) in the questions framing the debate. Second, I begin to explore the theological turn in science and religion, questioning whether various approaches have effectively dealt with the causal joint problem. I conclude that while the theological turn in divine action theories is a sign of real progress, more work is required in addressing the question of how

42

Chapter Three

transcendent, uncreated realities could affect the physical world – at least without radically reconceptualising the ontology of the natural world.

Part 1 If anything can be said of Christian theism, it is that God somehow acts in the world. However, contemporary science has increasingly revealed the natural world to be intelligible, seemingly explicable without recourse to divine agency and perhaps closed to supernatural causation. Hence the causal joint question: if a transcendent God interacts with physical processes, how exactly does that happen without undermining science? Though the literature on divine action is vast, my goal in this section is to highlight the general contours of this conversation in recent decades. Specifically, I suggest that approaches to the causal joint (and thus divine action in general) are largely determined by one’s perspective on three distinct, but related, debates: intervention/non-intervention, the laws of nature, and compatibilism/incompatibilism. By surveying these debates, it becomes clear that the divine action dialogue has been dominated by attempts to use science in the development of divine action theology. My intention here is not to analyse specific causal joint proposals, but rather to highlight the key questions shaping the various proposals, suggesting that these questions themselves are often question-begging or misleading. This section is thus meant to offer a helpful (albeit regrettably over-generalised) overview of what I will call the “standard” contemporary causal joint scene, the metaphysical and theological bases of which are challenged in Part 2.

Intervention vs. Nonintervention The question of intervention is the bedrock of divine action debates, and one’s stance on this issue largely determines one’s stance on the causal joint. Does God intervene in the laws of nature to bring about specific events within the natural world? This question is pressing because of the simple fact that science has been enormously successful at providing physical explanations for observed phenomena. In fact, the technical term “causal closure of the physical” has become somewhat of a bedrock assumption in the philosophy of science: all physical phenomena are at least treated as having a physical cause.1 Indeed, science has not only 1 David Papineau writes clearly about this: ‘all physical effects are fully caused by purely physical prior histories’ (2002, 17).

An Elephant in the Room

43

proved to be remarkably effective, but this effectiveness seems to be dependent on the intelligibility of physical regularities. It is worth questioning, then, whether it is appropriate to speak of God intervening in these regular, intelligible laws of nature, because God is non-physical. Divine intervention would, it is argued, undermine the basis of scientific inquiry; Philip Clayton thus urges “the presumption of naturalism” when examining any phenomenon (Clayton 1997, 171). It is thus unsurprising that the vast majority of science and religion scholars dismiss intervention as a viable option, given the body of scientific knowledge to the contrary. Beyond scientific concerns regarding interventionist divine action, there are significant theological barriers as well. Many theologians are extremely reluctant to affirm a model of divine action in which God seems to be in competition with the very laws of nature that God created. Taede Smedes explains that “to argue that God works against the laws of nature, or suspends them temporarily, would make the concept of God inconsistent. If in general providence God acted continuously through the nexus of secondary causality in accord with the laws of nature, and if at the same time God worked against these laws by putting them temporarily out of order, the action would be internally inconsistent: God’s special action would work against God’s general action” (Smedes 2008, 243). This theological conviction (that God would not work against the laws that God created) is enormously common in divine action conversations, and most notably so in the DAP. Note, however, that Smedes makes a clear theological distinction between different modes of divine action: general and special. He (and, it must be said, most scholars in the field) pairs the theological category of general divine action with the scientific category of “laws of nature”. The theological category of special divine action, then, is positioned as a competitor to the theological category of general divine action and the laws of nature. Here, it is helpful to briefly discuss the supposed distinctions between modes of divine action, particularly as they are so often assumed, but rarely challenged, and have everything to do with the causal joint. In his astute and immensely helpful analysis of the DAP, Wesley Wildman notes that participants in the DAP distinguish between general divine action (GDA), special divine action (SDA), and miracles. As delineated by Wildman, GSA “is the creation and sustaining of all reality in so far as this does not necessarily presume any specific providential divine intentions or purposes” (Wildman 2004, 37). This basically deistic category of divine action is intended to describe divine agency in upholding the regular laws of nature, and is not thought to require a causal joint. One implication of this is that science could never prove or disprove

44

Chapter Three

God’s general action; whatever science affirms to be true about the laws of nature is said to be sustained by God. That is, GDA is not thought to require specific, ongoing, and variable divine action; it instead describes a law-governed and autonomous world. SDA, on the other hand, refers to “specific providential acts, envisaged, intended, and somehow brought about in this world by God” (Wildman 2004, 37). It is important to clarify that SDA is generally used to indicate God’s noninterventionist action in and through natural processes. This is the domain of proponents of noninterventionist SDA, and of most scholars influenced by the DAP. Note that SDA is usually distinguished from miracles – at least insofar as miracles are defined in the Humean sense as violations of the laws of nature. Rather, SDA scholars seem to be motivated by a dual commitment to scientific knowledge and theological realities; thus, causal joint theories are often attempts to marry divine agency to natural processes. SDA theorists, then, have often been quick to develop and embrace causal joint theories that promise the best of both worlds: attention to scientific details and physical processes on one hand, and personal, specific divine action on the other. The relationships involved here are worth emphasising: SDA language is largely used by theorists committed to a noninterventionist framework for divine action, and is generally associated with attempts to locate specific causal joints in the natural world. The implication in such theories is that science has the power to render intelligible the specific mechanisms involved in divine action, at least insofar as SDA is assumed to occur through intelligible natural processes. While these distinctions between GDA, SDA, and miracles can be challenged as possibly arbitrary categories (Gregersen 2008), the main point here is that these distinctions are intimately bound up with a commitment to noninterventionism. Those embracing noninterventionist SDA, therefore, look for coherent ways for God to work in and through the laws of nature; there is thus a direct relationship between noninterventionism, the causal joint, and the limits of science (or lack thereof). The search for a noninterventionist causal joint belies a prioritisation of scientific knowledge, and essentially renders the natural sciences as “objective arbiters of ‘the way things really are’” (Smith 2008, 885). This includes whether, and how, God might act in the world. That is, there is a sort of logical progression from the prioritisation of scientific knowledge, to a commitment to noninterventionist divine action, to the necessity of a scientifically-identifiable causal joint in which this divine action might occur. If one begins with the assumption that God either cannot or does not intervene in natural processes, then it is only a short step to concluding that God somehow works through natural

An Elephant in the Room

45

processes themselves. But this “somehow” is problematic; how could God work with through natural processes, when these processes are increasingly intelligible by the natural sciences? The most popular option in recent decades has been to identify open spaces within the natural world that are seemingly underdetermined. To put this a different way: if the causal closure principle is correct and every physical event has a physical cause, then there seems to be no way for God to act without actually intervening in natural processes. If, however, there are areas of the world that are ontologically underdetermined by prior physical events, then God could fill this causal vacuum and bring about specific physical effects – or so the story goes. Thus, those involved with (or influenced by) the DAP have put forward proposals involving quantum mechanics, chaos theory, emergence, and complexity, arguing that ontological indeterminism essentially “makes room” for God to be a causal agent. Though such proposals differ greatly in specifics, they share a common conviction that science itself is able to locate (or at least point to) the specific locus of noninterventionist SDA. In this respect, science is given enormous theological power, insofar as it has been used to identify specific causal joints wherein God might act. This is a double-edge sword, however, for this commitment to “traction” between science and theology also gives science the power to say where the causal joint is not. Thus, for example, Polkinghorne’s use of chaos theory has been overwhelmingly critiqued for its misconstrual of deterministic processes, quantum mechanics proposals have been similarly challenged for not being scientifically plausible or actually noninterventionist, and emergent divine action theorists have struggled to defend top-down divine causation for the same reasons. There is, perhaps, a slight tinge of irony in all of this: proponents of noninterventionist SDA defend a very traditional theistic view of God and God’s action, while at the same time giving science almost unlimited power to determine the parameters of this action. A commitment to noninterventionist SDA seems to require specific underdetermined causal joints in the natural world, but this requirement renders such theories vulnerable to advances in scientific explanation. As Charles Coulson infamously wrote, “There is no 'God of the gaps' to take over at those strategic places where science fails; and the reason is that gaps of this sort have the unpreventable habit of shrinking” (Coulson 1955, 20). Indeed, betting against the explanatory success of scientific analysis seems increasingly inadvisable. Of course, it would be inaccurate to say that noninterventionist divine action is the only option. For one thing, it is a bit misleading to label any

46

Chapter Three

divine action as noninterventionist, at least to the extent that divine action indicates a state of affairs that would have been otherwise had God not acted in some way. This is an important point; the choice is not between intervention and nonintervention as a binary choice, but a question of degree, mechanism, and whether or not divine action can actually be discerned as such. If God is affecting the world in all divine action, then we need to be clearer about the terminology used to describe this. More importantly, it is worth noting that it is at least possible to develop a coherent interventionist approach. Alvin Plantinga, for example, challenges both the scientific and theological objections to interventionism described above, rightly identifying the causal closure principle as the underlying assumption in most scientifically-based objections to intervention. Importantly, he argues that the causal closure principle is indeed an assumption – not a scientific conclusion. Science itself is unable to prove the causal closure principle. Plantinga is also aware of the theological arguments against intervention, namely regarding divine consistency and integrity. To this theological challenge, Plantinga responds that “there would be arbitrariness and inconsistency only if there were no special reason for taking action contrary to the usual regularities” (Plantinga 2008, 388). That is, the most God-like or internally consistent divine actions might require that God act differently in different situations. This is a relational and personal understanding of consistency. One might well question whether Plantinga is taking scientific methodology seriously enough, but his commonsense challenge to contemporary assumptions may be a helpful corrective. The point here is that the interventionist question can be framed in multiple ways, and largely determines the specific ways in which one approaches the causal joint. The question remains, however, whether this framework might be potentially questionbegging and perhaps unhelpful.

The Laws of Nature The second central debate in divine action theories involves the laws of nature. To say that God does or does not intervene in the laws of nature would seem to presume a very specific understanding of these “laws” as legally binding and static. The classic (over-simplified) debate is between the laws as either prescriptive or descriptive. Prescriptive laws would possess an independent, idealised ontological status; they would “ontologically determine which possibilities are open to the world and which are not” (Saunders 2002, 66). Descriptive laws, on the other hand, would “merely” describe contingent regularities that need not be binding.

An Elephant in the Room

47

A great many divine action theorists subscribe to a descriptive view of the laws of nature; this makes sense, as it would seem prima facie to allow for unfettered divine action. Herein lies a fascinating paradox, however, for the entire interventionist/non-interventionist framework in which these theorists operate already presumes that the laws of nature are prescriptive. If the laws were merely descriptions of observed regularities, there would be no concerns about interventions, for there would be no ontological laws in which God could intervene. Clearly, then, we need to examine this “prescriptive versus descriptive” dichotomy a bit more closely. Gregersen is correct in noting that the presupposed prescriptive “concept of the laws of nature serves as a foil for shaping the idea of noninterventionist SDA” (Gregersen, 2008 191); many causal joint theories are developed under this framework. Indeed, this is where the indeterminism debate becomes vitally important for noninterventionist SDA. If the laws of nature are ontological, then it is supposed that the only possibility for non-interventionist SDA is in areas that are ontologically underdetermined. Ontological gaps may be “fair game” for divine action, as (apparently) no ontological laws are broken. In other words, theorists begin with an assumption that SDA must be non-interventionist, and this assumption necessarily presumes an ontological view of the laws of nature. In order for God to act, then, there must be features of the causal web that are ontologically undetermined by a law of nature. Thus, a necessitarian view of the laws of nature would seem to require a causal joint in underdetermined natural processes if noninterventionist SDA is to be possible. And indeed, causal joint theories developed along these lines generally use science to suggest where these spheres of indeterminacy might be – again giving science the authority to determine the possibilities and limits of divine action. It is true that the prescriptive “necessitarian” perspective would seem to have merit, insofar as it corresponds to the observable success of the scientific method and the seeming discovery of pre-existent physical relationships. It should be noted, though, that noninterventionist SDA may be possible even within this prescriptive view of the laws of nature. For example, necessitarian laws could exist, but only in localised settings; this is Nancy Cartwright’s “dappled world” view (Cartwright 1999). In her model, laws are acknowledged, but also recognised to be limited in scope and application – a so-called patchwork of laws. In this model, it is conceivable that God could act via some sort of lawful action in certain contexts, even if that lawful action remained inaccessible to science. Necessitarian options notwithstanding, however, SDA proponents have largely opted for theories utilising scientifically-identifiable spheres of

48

Chapter Three

indeterminism. The problems with this have been well described by Nicholas Saunders and many others; challenges often boil down to the conclusion that a causal joint in such areas would still entail intervention, or would be scientifically implausible. Not everyone subscribes to the necessitarian view of the laws of nature, however. As mentioned, one of the most striking (and puzzling!) features of the divine action conversation is just how many of those involved claim to hold a descriptive view of physical laws. Though the entire non-interventionist SDA project presupposes a prescriptive view of the laws of nature, many divine action theorists explicitly endorse a descriptive view of natural laws. This is interesting, as a truly descriptive view would render unnecessary the search for a causal joint in underdetermined processes – there would be no ontological laws in which God could intervene in the first place. It is clear that the descriptive view is thought to “incorporate enough flexibility and openness to accommodate the intentional actions of God” (Saunders 2002, 49), but what does that mean? In the philosophy of science, the descriptive view is generally taken to indicate a regularity account wherein emphasis is places on singular events, rather than on laws governing those events. That is, laws describe what actually happens on a regular basis, but not what can or cannot happen. Again, note that there is no interventionist problem under the regularist view. If there simply are no universal laws for God to violate, causal joints could basically exist wherever and whenever God decided to act (though exactly how the immaterial could affect physical regularities would remain opaque). But this does not seem to be what SDA theorists mean by the term “descriptive”. Indeed, it seems that there is a disjunction between the way theorists think about laws of nature when approaching them on a metaphysical level, and the way they develop actual SDA theories. That is, virtually everyone in the DAP affirmed the need for non-interventionist divine action (thus presupposing prescriptive laws of nature), but often use descriptive approaches in their respective theories. The lack of internal consistency in SDA theories highlights the way that certain assumptions regarding noninterventionism and the laws of nature determine causal joint theories, even when not consciously incorporated into the details of those theories. The debate (and confusion) regarding prescriptive and descriptive laws is perhaps made clearer by a second distinction: that between the laws of nature and the laws of science. This is an immensely important distinction between the laws of science as we currently understand and formulate

An Elephant in the Room

49

them, and the laws of nature as they function in reality. It is worth noting that many in the science and religion field use the term “descriptive” when speaking of the laws of nature, but seem to actually be speaking of the known laws of science. That is, many theologians will say that the laws of nature are useful descriptions of reality, while not being clear that they are actually speaking of the relatively more constrained laws of science. Nevertheless, some SDA theorists work constructively with this distinction, developing an instrumentalist, approximationist view of natural laws: “‘Our laws of nature’ are much more limited and uncertain than the full range of ‘the laws’ in themselves…there are also such processes and relationships which are in principle beyond the competencies of the natural sciences to investigate and to model” (Stoeger 2008, 237). This quote from William Stoeger is illuminative, demonstrating not only the view that scientific laws are contextualised approximations, but also the limits that this view places on science. Indeed, an approximationist view forbids science from claiming any ontological territory concerning the laws of nature or the causal joint. If, as William Alston argued, “none of our laws take account of all possible influences” (Alston 1993, 190), then science basically has no power to say where divine action does or does not occur. In an instrumentalist or approximationist model, God could act through “higher laws” of nature. The causal joint would simply be relegated to these higher laws that undergird or somehow interact with our known scientific laws. Of course, an approximationist or instrumentalist theory of divine action essentially divorces theology from scientific knowledge; science is rendered incapable of saying anything about possible higher laws or any hidden causal joints. It is worth pointing out that an approximationist approach to natural laws is compatible with the prescriptive approach to the laws of nature, as well as with the descriptive. As Stoeger explains, “God may act in a purely ‘natural’ way…but in a way which we see as supernatural intervention simply because we have not yet come to comprehend fully the relationships and regularities (the ‘higher laws’) which obtain” (Stoeger 2009, 124). This, of course, invites “God of the gaps” charges and also fails to explain how an immaterial, transcendent God could interact with physical processes. The causal joint problem is not removed, but simply pushed back beyond the reach of science. To summarise, then: one’s view on the laws of nature largely shape one’s approach to the causal joint. A prescriptive view pushes one to non-interventionist theories utilising ontological indeterminacy, while truly descriptive views could potentially allow innumerable causal joints (though without specifying mechanisms or the sort of God-nature relationship this would require). Approximationist .

50

Chapter Three

views, on the other hand, locate the causal joint in “higher laws” that are inaccessible to science.

Compatibilism/Incompatibilism The third key debate surrounding divine action and the causal joint involves compatibilism, or the degree to which divine action is compatible with a closed causal history. In other words, is it coherent to speak of God acting in the natural world, if a complete scientific account of that action is available? The contemporary debate often assumes incompatibilism, wherein natural causes and divine causes cannot be seen as affecting the same event. It seems intuitive to say that an event is either caused by natural processes, or it is caused by divine action through ontologically underdetermined processes – but not both. Thus, “in order for God to act, something in the natural order has to give way – hence the active search for irreducible ontological gaps in the causal nexus” (Smedes 2008, 256). If one can identify “gappy” areas in the natural world that are ontologically underdetermined by other physical causal antecedents, then it may be possible to include God as a causal agent in these realms. Unsurprisingly, then, incompatibilism has been a driving force behind the majority of causal joint proposals in recent decades. Incompatibilism, moreover, “enables one to rule out types of special divine action on the basis of our scientific understanding of the laws of nature – God may only act in a special, direct way in the ‘causal gaps’ opened by indeterminacies” (Stoeger 2008, 240). That is, incompatibilists privilege scientific knowledge in determining where divine action might or might not occur. This is consonant with the primacy of scientific knowledge highlighted in the above debates concerning indeterminism and the laws of nature. This is not to say that compatibilists are absent from the divine action conversation – far from it. Compatibilists affirm that divine action is “continuous with natural processes, present throughout the whole cosmos and entirely compatible with our descriptions of mathematical behavior” (Ward 2008, 260). The classic example of compatibilist divine action comes from Thomism, which addresses the causal joint by affirming Aquinas’ distinction between God as a primary cause, and created, secondary causes: “God is the cause of all actions, inasmuch as every agent is the instrument of the divine power acting” (Aquinas 1947, De Pot. Q. 3, Art. 7). The Thomist doctrine of double agency affirms that God and natural agents work on different levels: God, as primary cause, creates, sustains, and acts through all secondary causes, but these secondary causes also have their own causal agency. So, “the same effect is ascribed to a

An Elephant in the Room

51

natural cause and to God” (Aquinas 1976, SCG, 3.70) and fully so: it is not as though God does some of the work, and the creature does the rest. Rather, both the natural agent and God are fully causal – “the whole effect proceeds from each, yet in different ways” (Silva 2014, 285). Of course, all this raises the obvious question of how, exactly, God could act not through a causal joint, but through the material world of secondary causes that is already “complete on its own level” (Johnson 1996, 8). Thomists are univocal in their response to this: we cannot know how God acts through secondary causes. For Thomists, any search for a “causal joint” is wrongheaded; everything is a causal joint (which is actually quite similar to saying that there is no causal joint). Thomists ultimately appeal to divine mystery when it comes to the actual “how” of double agency and the “what” of the causal joint. But even if all divine action theories inevitably reach a point of mystery beyond which they cannot go, the Thomistic version of this mystery seems particularly susceptible to challenge. It is really difficult to affirm that a single event occurring in the physical, observable world has both a fully natural (secondary) cause and a fully divine (primary) cause. As John Polkinghorne sums it up, double agency seems like “an unintelligible kind of theological doublespeak” (Polkinghorne 1994, 81-82). Thomism highlights the seeming paradox all compatibilists face when it comes to dealing adequately with the causal joint. Moreover, compatibilist theories in general are essentially immune to scientific critique; because they are not dependent on “gaps” in the scientific account of the world, “the science and the theology pass by each other without much traction” (Wildman 2006, 148). Thus, there are at least two potentially problematic aspects of Thomistic compatibilism: it is immune from scientific dialogue or critique, and to many it simply seems illogical to say that a single effect is fully caused by two distinct causal agents. The upshot of all this is that the contemporary divine action dialogue has often preferred incompatibilist approaches. It should be noted that there are remarkably nuanced neo-Thomistic accounts of divine action. Denis Edwards, for example, emphasises the participatory, non-coercive basic relationship between God and nature. He argues that “God does not override the process, nor bypass the laws of nature. God accepts and works creatively with the limits of creaturely processes, lovingly respecting the integrity of creatures” (Edwards 2010, 66). Note that this sounds remarkably different from classic accounts of primary and secondary causality. Indeed, I have not focused on these more developed Thomistic accounts precisely because they are essentially participatory accounts of divine action, and as such they drift into a God-

52

Chapter Three

world model that seems (in my view) “other” than strictly Thomistic. Moreover, I fear these contemporary Thomistic accounts obscure the core paradox of double agency. In sum then, most SDA theorists have opted for an incompatibilist understanding of divine action. This is congruous with the simultaneous commitment to non-interventionism and a high view of the laws of nature. In this section, I have highlighted the contours of three critical debates in the divine action conversation: interventionism/noninterventionism, the laws of nature, and compatibilism/incompatibilism. By examining the way each of these debates render different approaches to the causal joint, we can see how the assumptions inherent in the dichotomies themselves have framed and shaped the perceived possibilities for divine action in recent decades. SDA theorists have overwhelmingly embraced an incompatibilist, noninterventionism framework, which presupposes an ontological view of the laws of nature. Critically important within this framework, it is science that finally determines where and how divine action could possibly occur (or be identified) in the natural world. While this is done in the name of “traction” between science and theology, it is often difficult to see how this standard approach to the causal joint allows any real credibility or authority for theology. It is in reaction to this that a theological turn has occurred in science and religion, and to which we now turn.

Part 2 The incompatibilist, noninterventionist framework for divine action was widely affirmed in the DAP and has continued to be assumed by many theorists influenced by the DAP. These approaches essentially make science the final arbiter of where God might possibly act in a lawgoverned world, with great emphasis placed on scientifically identifiable points of indeterminism. Many, however, are now challenging this approach and its implicit assumptions about the God-nature relationship. These theorists are instead turning to theology for alternative ways to think about the metaphysical doctrine of God in general, and the causal nexus between transcendent and created realities in particular. By reconsidering what it means for physical realities to be properly “natural”, critics of standard SDA models challenge the legitimacy of the metaphysical assumptions framing the debates considered above. The interventionist/ noninterventionist debate, for example, is discarded as implicitly assuming a deistic view of the God-nature relationship, in which nature is understood as basically autonomous and self-sufficient apart from God’s immanent presence and activity. As Aubrey Moore wrote, “a theory of

An Elephant in the Room

53

occasional intervention implies as its correlative a theory of ordinary absence” (Moore 1889, 184). Similarly, the compatibilism/incompatibilism debate is critiqued as being overly binary and un-nuanced, and various compatibilist accounts have been developed that (supposedly) go beyond neo-Thomistic double agency. Further, the debate over the laws of nature is critiqued as being misleading in its generally oversimplified presentation, and a view of natural laws is developed that includes the influence and participation of spiritual realities with physical processes. As Gregersen explains, “the ontology of divine action determines the understanding of the laws of nature, rather than the other way around” (Gregersen 2008, 194). In short – critiques of standard causal joint models reject the paradigms under which many theories have been developed. Instead of using science to point toward causal joint possibilities, they instead offer models that privilege theological categories reframing the God-nature model. Science, it is argued, is limited in what it can tell us about metaphysics and the basic ontology of nature. Science cannot point to a single indeterministic “causal joint” as being uniquely open to divine action, because all of nature participates in and with God at the most basic level. There are quite a few variations on this theological turn in divine action, and I can here merely highlight a select few. While each deserves extensive analysis in its own right, a brief snapshot of differing approaches will at least demonstrate how specific theological frameworks deal with the causal joint and the doctrine of God. First is the theistic naturalism developed by Christopher C. Knight. Knight has used his Eastern Orthodox perspective to develop a robust panentheistic2 naturalism that of necessity includes God’s active presence. That is, the panentheistic doctrine of God inherent in Knight’s Orthodox theology allows him to argue that the reality and influence of God is simply a brute fact about reality-without which knowledge of reality is incomplete. Knight rejects the natural-supernatural binary as a false dichotomy that renders the intervention/nonintervention framework inadequate. That is, there is no such thing as a self-sufficient pure nature that normally exists apart from God’s special action. This is where the nearly universal distinction 2

There are, of course, many others (even in the DAP) who embrace panentheism. The distinction between Knight and DAP-type panentheists is that people like Clayton and Peacocke still use science to highlight gaps in scientific knowledge (emergence and complexity theory, for example). In this sense, many panentheists unnecessarily utilise the standard noninterventionist framework – unnecessarily because panentheism theoretically erases then erases the need for ‘gappy’ areas in the causal web.

54

Chapter Three

between general and special divine action becomes important; the distinction presumes a model in which the default position for nature is one of autonomous natural processes. Knight’s panentheistic naturalism, on the other hand, would allow us to “transcend the need for any distinction between what nature can do ‘on its own’ and what can only be done through some special mode of action” (Knight 2011, 47). Because nature exists in God, creaturely participation with God and the experience of divine action actually serve to make nature more natural – not less (Knight 2007, 94). Divine action is thus not to be understood as an intervention of natural laws, but as a reflection of true nature and “higher” laws. Divine actions are “manifestations of ‘laws of nature’ that reflect, more fully than those laws of nature that are scientifically explorable, God’s presence in all things” (Knight 2007, 94). And what of the causal joint? For this panentheistic perspective, everything would be a causal joint, as God’s immanent presence and activity would inhere in every created entity (Gregersen 2008, 195). Ironically, it seems that the more naturalistic one is willing to go, the less of a problem divine action becomes; divine action in creation can be seen “as a transformation from the subnatural to the natural” (Knight 2007, 95). If everything is natural, including “higher laws” that are inaccessible to scientific methodologies, then divine action becomes, quite literally, the most natural thing in the world. Of course, there are serious questions that can be raised against Knight’s panentheistic naturalism. Significantly, this view drastically limits what science can say about divine action. If the causal joint is simply relegated to higher laws, this view invites “God of the gaps” charges: how exactly do these higher laws connect an ontologically immaterial being to physical realities? Presumably these higher laws are immune to scientific analysis, but if so, the causal joint question is not answered; it is merely pushed further back, forever out of reach of scientific inquiry. This epistemological boundary may not be a problem, and may in fact be inevitable. However, it is important to be clear that the preferential treatment given to theology here directly undermines any efforts to seek traction between theology and science. Beyond this, there is also a serious theological problem with this view, regarding divine transcendence. Namely, if there is no scientifically identifiable causal joint and everything exists in God, how can divine transcendence be maintained? This is one of the key issues with all panentheistic frameworks, in fact: they can be critiqued as eventually collapsing into pantheism. Indeed, this challenge highlights the core problem of the causal joint: if God is transcendent and non-physical, then at some point and in some very specific way, uncreated

An Elephant in the Room

55

divine will must relate to created, material processes. If all divine action occurs through “higher” laws that somehow exist in relationship with the known laws of science, what is the ontology of these higher laws? Are created and spiritual realities truly distinct, or do they somehow bleed or blend into each at some point in a gradual fashion? If there is no definitive causal joint because everything exists in God, it is extremely difficult to maintain the ontological distinction between God and nature. These challenges to Knight’s divine action model are directly related to its panentheistic framework. One might wonder, then, whether other theological emphases might be able to perform the metaphysical functions of panentheism, without threatening divine transcendence. One alternative gaining quite a bit of momentum comes from the field of pneumatology. Pneumatologists such as Amos Yong and James K.A. Smith have developed divine action theories akin to Knight’s theistic naturalism insofar as they challenge common assumptions of what is properly natural. They emphasise the Spirit’s presence and activity in all of nature as normative, rather than aberrant, rejecting the deistic dichotomy between God and the supposedly autonomous world of self-sufficient natural causes. Smith in particular develops what he calls an “enchanted naturalism”, stemming from a commitment to participatory ontology: “matter as created exceeds itself and is only insofar as it participates in or is suspended from the transcendent Creator” (Smth 2008, 889). Yong adds a “pneumatological assist” to this participatory ontology, affirming the Spirit as “the agent of ‘suspension,’ the Triune person in whom the material world is suspended” (Smith 2008, 890). Note that instead of relying on panentheism, these pneumatologists emphasise the person and work of the Holy Spirit as a legitimate and necessary participant in all physical processes; material creation is “charged” with the presence of the Spirit (Smith 2010, 12). Pneumatological divine action theories thus argue that the distinct role of the Holy Spirit in creation might preserve both the immanence and the transcendence of God, performing the same metaphysical function as panentheism without the attendant problems. Again, we see here a methodological reversal of the DAP, insofar as this approach privileges theology over scientific knowledge. How, then, does a pneumatological perspective handle the causal joint in relation to our three debates? Smith, for example, affirms “the elasticity of nature as always already inhabited by the Spirit” (Smith 2008, 881). But what exactly does that mean? As already mentioned, pneumatologists often reject the interventionist paradigm. In this way they are similar to panentheists, arguing that “all that is ‘participates’ in the divine through the animating power of the Spirit” (Smith 2007, 254). The normative

56

Chapter Three

position for creation is to be fundamentally involved with the Spirit who indwells all creation but remains fully divine and transcendent. Thus, talk of intervention or nonintervention becomes nonsensical, as these terms assume an insufficient God-world model. But how, exactly, does the Spirit interact with the laws of nature? Yong, for one, seems to embrace an approximationist or instrumentalist view of the laws of nature. In his view, “the laws of nature are amenable to the basic actions of God and sufficiently flexible” (Yong 2011, 131). Yong sees the known laws of science as corresponding to real tendencies, but also affirms “the Holy Spirit as working in and through nature and its laws” (Yong 2011, 125). Smith further suggests that we see special divine actions as “sped-up modes of the Spirit’s more regular presences” (Smith 2008, 892). Divine action, then, is not an either/or binary. Rather, a pneumatological approach sees the laws of nature as being flexible and inherently “involved” with the Spirit’s immanent presence and activity at all times, with varying effects dependent on creaturely response to God. We can see that pneumatological divine action theories reject the noninterventionist paradigm and advocate a flexible, open view on the laws of nature. This renders a perspective on the causal joint that is basically congruent with that of Knight’s Orthodox panentheism: there is no single causal joint as all nature is, somehow, inherently involved with the immanent presence and activity of God. But this “somehow” is the causal joint problem we must always come back to: if God is uncreated, transcendent, and ontologically distinct from physical processes, then in some way divine agency must interact with basic matter. We can look at this problem a bit closer through our third debate, involving compatibilism/incompatibilism. As discussed, DAP theorists were keen to advocate incompatibilist divine action, but critics have challenged the presumption of deism involved with incompatibilism and have developed theologically-based compatibilist positions. Indeed, both theistic naturalisms and pneumatological approaches could affirm a compatibilist position, insofar as they assert that the basic ontological status of matter involves participation with God. But what is this basic relationship between God and nature? The theological turn that has produced these “newer” compatibilist approaches inevitably raises the question of whether these various reimagined naturalisms and pneumatologies escape the challenges of Thomistic double agency – how can both God and natural creatures be efficacious in bringing about specific events? One way to deal with this is to make a distinction between Thomistic causation on one hand, and participatory ontology on the other. For example, in Smith’s enchanted naturalism, the

An Elephant in the Room

57

causal relationship between God and natural processes is one of interconnectedness, of free creaturely response to the Spirit’s persuasive “lure.” But then, how does God lure, influence, or relate to natural processes? Are we to imagine some common ontological ground at which God and created matter interact? If so, then how does a participatory ontology avoid slipping into pantheism? To put the point differently: if a compatibilist, naturalist theology of divine action rejects the causal joint and affirms basic interaction between God and nature, it is difficult to maintain the classical infinite distinction between transcendent divinity and created matter. Almost universally, theologically-based approaches to divine action appeal to mystery when it comes to the causal joint. Theistic naturalists, contemporary Thomists, and pneumatologists will often affirm that “true nature” is always in dynamic, active relationship with God – and that the mechanics of this are inherently unknowable by scientific methodologies. And yes, it seems inevitable that all divine action theories will eventually run up against a wall of mystery, behind which lies divine realities that are essentially unknowable to finite human beings. This may be a necessary frustration for science and religion but theologically acceptable and even desirable. And yet, even if we must finally include an element of mystery in our science and religion models, it is necessary to at least be aware of the serious challenges posed by the causal joint problem. Put simply, affirming some sort of higher spiritual laws or a mysterious paradox of double agency does not actually solve the problem of how a transcendent God could act in the physical world. This problem is actually made more acute when one recognises the danger even in participatory models – namely, the danger of a transcendent God being inherently involved with created matter to such an extent that it becomes unclear how these models avoid pantheism. The challenge, it seems, is to more fully explore what we mean when we talk about nature, matter, and physical processes. If all of nature is inherently involved with God (as the above models assert), then the ontology of matter itself needs to be further explicated. Smedes poses a similar challenge, noting that “Simple questions such as What is science? and What is religion? are hardly ever asked anymore” (Smedes 2008, 255). To conclude, then, it seems that there has been a shift away from standard causal joint proposals. These standard proposals, as exemplified by the DAP, often affirmed noninterventionist, incompatibilist divine action in underdetermined areas of the natural world; they also seem to presume an ontological status for the laws of nature. Critics of these approaches challenge the very frameworks in which these debates have

58

Chapter Three

transpired, denying the deistic God-world model implicitly assumed by them. These “newer” models deny that the causal joint problem is actually problematic, insisting instead that there is no single causal joint precisely because all the natural world participates constantly in divine activity. This theological turn seems to be gaining momentum and is, I suggest, a welcome change. But while I do find some version of theistic naturalism to be a move in the right direction, we should not imagine that this theological turn actually solves the causal joint problem, or that it answers the looming question of what, exactly, it means to be natural. Is this a scientific question or a theological, metaphysical one? The theological turn is a welcome development in science and religion, but we would do well to be wary of the limits it seems to place on science.

References Alston, W. 1993. “Divine Action, Human Freedom, and the Laws of Nature”, in Russell, R.J., Murphy, N., and Isham, C. (eds.) Quantum Cosmology and the Laws of Nature, Vatican City: Vatican Observatory. Aquinas, T. 1947. De Potentia Dei. Translated by Fathers of the English Dominican Province. Westminster, Maryland: The Newman Press. —. 1976. The Summa Contra Gentiles Bk. 3: Providence. United States: University of Notre Dame Press. Cartwright, N. 1998. God and contemporary science (Edinburgh studies in constructive theology). Grand Rapids, MI: Eerdmans. —. 1999. The dappled world: A study of the boundaries of science. Cambridge, UK: Cambridge University Press. —. 2008. Adventures in the spirit: New forays in philosophical theology. Simpson, Z. (ed). Minneapolis, MN: Fortress Press. Coulson, C.A. 1955. Science and Christian Belief. Oxford: Oxford University Press. Edwards, D. 2010. How God Acts: Creation, redemption, and special divine action. Minneapolis, MN: Fortress Press. Gregersen, N.H. 2008. “Special Divine Action and the Quilt of Laws: Why the Distinction between General and Special Divine Action Cannot be Maintained”, in Russell, R.J., Murphy, N., and Stoeger, W.R. (eds.), Scientific Perspectives on Divine Action: Twenty Years of Challenge and Progress. Vatican City: Vatican Observatory Foundation. Johnson, E.A. 1996. “Does God play dice? Divine providence and chance”, Theological Studies, 57, 1, 3–18.

An Elephant in the Room

59

Knight, C. 2007. The God of nature : Incarnation and contemporary science (Theology and the sciences). Minneapolis: Fortress Press. —. 2011. “Divine Action and the Laws of Nature: An Orthodox Perspective on Miracles”, in Buxhoeveden, D. and Woloschak, G.E. (eds.) Science and the Eastern Orthodox Church, Burlington, VT: Ashgate Publishing. Moore, A. 1897. Science and the faith : Essays on apologetic subjects, with an introduction. London: K. Paul, Trench, Trübner. Papineau, D. 2004. Thinking about consciousness. Oxford: Oxford University Press, USA. Plantinga, A. 2008. “What is ‘intervention’?”, Theology and Science, 6, 4, 369-401. Polkinghorne, J. 1994. Science and Christian belief: Theological reflections of a bottom-up thinker. London: SPCK Publishing. Saunders, N. 2002. Divine action and modern science. Cambridge, United Kingdom: Cambridge University Press. Silva, I. 2014. “Revisiting Aquinas on providence and rising to the challenge of divine action in nature”, The Journal of Religion, 94,3, 277–291. Smedes, T.A. 2008. “Beyond Barbour or Back to Basics? The Future of Science-and-Religion and the Quest for Unity”, Zygon, 43,1, 235–258. Smith, J. 2007. “The spirit, religions, and the world as sacrament: A response to Amos Yong’s Pneumatological assist”, Journal of Pentecostal Theology, 15, 2, 251–261. —. 2008. “Is the Universe Open for Surprise? Pentecostal Ontology and the Spirit of Naturalism”, Zygon, 43, 4, 879–896. —. 2010. Thinking in tongues: Pentecostal contributions to Christian philosophy. Grand Rapids, MI: William B. Eerdmans Pub. Company. Stoeger, W.R. 2008. “Conceiving Divine Action in a Dynamic Universe”, in Russell, R.J., Murphy, N., and Stoeger, W.R. (eds.), Scientific Perspectives on Divine Action: Twenty Years of Challenge and Progress. Vatican City State: Vatican Observatory Publications. —. 2009. “Describing God’s Action in the World in Light of Scientific Knowledge of Reality”, in Shults, F.L., Murphy, N., and Russell, R.J. (eds.), Philosophy, Science and Divine Action (Philosophical Studies in Science and Religion). Leiden: Brill. Ward, K. 2008. The big questions in science and religion. 2nd edn. Boca Raton, FL, United States: Templeton Foundation Press, U.S. Wildman, W. 2004. “The divine action project, 1988–2003”, Theology and Science, 2, 1, 31–75.

60

Chapter Three

—. 2006. “Robert John Russell’s Theology of God’s Action”, in Russell, R.J., Peters, T., and Hallanger, N. (eds.), God’s Action Nature’s World: Essays in Honor of Robert John Russell. Aldershot: Ashgate Publishing. Yong, A. 2011. The spirit of creation: Modern science and divine action in the Pentecostal-charismatic imagination. Grand Rapids, MI: William B Eerdmans Publishing Co.

CHAPTER FOUR THE WILD EXPERIMENT: EMOTION, REASON, AND THE LIMITS OF SCIENCE DONOVAN O. SCHAEFER

In a letter of early spring 1863, Charles Darwin, in a rare autobiographical mood, announced to his friend and ally Joseph Dalton Hooker, “I am like a gambler, & love a wild experiment.” His son, Francis Darwin, who 30 years later would edit the volume that would become Darwin’s Life and Letters, corroborated this impression, writing that “The love of experiment was very strong in him, and I can remember the way he would say, ‘I shan’t be easy till I have tried it,’ as if an outside force were driving him” (Browne 2002, 169). Janet Browne, Darwin’s pre-eminent contemporary biographer, describes how Darwin found his research to have a nourishing, healing force. Darwin, in her writing, “wanted to put his wits once again against the native ingenuity of animals and plants, to be able to follow a lucky hunch, and use all his guile and skill to arrive at a satisfactory conclusion… he loved the spirit of the game itself” (Browne 2002, 167). In his 2006 book The God Delusion, Richard Dawkins raggedly admonishes his readers: “I don't want to decry human feelings. But let's be clear, in any particular conversation, what we are talking about: feelings, or truth. Both may be important, but they are not the same thing” (Dawkins 2008, 395). For Dawkins, the foremost modern proponent of the Castor and Pollux theory of science and secularism, the horizons of science are unlimited. As we furnish the warehouse of truth, we steadily declutter it of the emotional relics of religion. It is precisely science’s ability to produce truths that are uncontaminated by the murk of emotion, to jump over our bodies and capture the world in its geometric exactitude, that renders science limitless. But, elsewhere, Dawkins acknowledges that the banners of emotion do attach themselves to science. Dawkins describes an episode in which an

62

Chapter Four

Oxford zoology professor confronted a visiting lecturer who challenged his skepticism on the existence of the Golgi apparatus: “the old man strode to the front of the hall, shook the American by the hand and said—with passion—‘My dear fellow, I wish to thank you. I have been wrong these fifteen years.’” Dawkins continues: “We clapped our hands red…. The memory of the incident still brings a lump to my throat” (Dawkins 2008, 321). In his earlier book Unweaving the Rainbow, an iteration of the welltrodden genre of odes to “wonder” in science, Dawkins recounts Sir Arthur Eddington's praise of Einstein, after demonstrating for himself the validity of the theory of general relativity. Dawkins: “I read those words with a catch in the throat” (Dawkins 1998, 42). But Dawkins’s argument in Unweaving the Rainbow, in linking science with wonder, shines the spotlight on the extravagance of discoveries. The wonder of distant galaxies. The wonder of our cells. The wonder of the fabulous complexity of organisms (Dawkins 1998, 8). “If history had worked out differently,” he asks, “and Michelangelo had been commissioned to paint a ceiling for a giant Museum of Science, might not he have produced something at least as inspirational as the Sistine Chapel? How sad that we shall never hear Beethoven's Mesozoic Symphony, or Mozart's opera The Expanding Universe” (Dawkins 2008, 111). The emotionality of science is, for Dawkins, a fringe benefit, a side-effect, the prize at the end of the race. Science is not itself emotional. This would be to contradict the wall of separation between emotion and truth, to locate the scientific method too close to the ambiguities of bodies. Dawkins deletes emotion from science. Science’s limitless purview remains intact. In this chapter, I want to consider an alternative to Dawkins’ perspective, one that I think more faithfully captures Darwin’s gambler’s enthusiasm for the wild experiment. Rather than showcasing an external limit to science—a border beyond which its questions lose all meaning, though I’m sure those exist—I want to suggest that it has an underattended internal limit. That limit is that science never manages to supersede the guiding emotionality of our bodies, and therefore our minds. The philosopher Lars Svendsen writes that “[p]hilosophy does not begin in a state of cognitive or affective neutrality, but rather in a state of disturbance or bewilderment” (Svendsen 2012, 424). I want to make a consonant argument with respect to science. Science is, I propose, always guided by a subtle, but intransigent tissue of emotional priorities, microthreads that lace cognition into a structured weave of information and inspiration. There are a number of ways that one might pursue the project of shading in emotionality as an inner limit to science. Even though I think

The Wild Experiment: Emotion, Reason, and the Limits of Science

63

it’s something that everyone, in a certain sense, knows, the process of creating a theoretical framework that can bring it to the surface has proceeded only partially and from a number of different directions. My efforts here will be focused on linking a particular philosophical genealogy of this idea—David Hume and the American Pragmatist William James— to contemporary work in affect theory. This complex will draw out what affect theorist Lauren Berlant has called the “sensualized epistemology” of knowledge production. It will show that science is not a bloodless undertaking, but a game, a field of play, a wild experiment that we are pulled back into again and again.

Hume: Hunting, Gambling, Philosophizing I want to start this story with Hume; that decision is partly arbitrary, but it also reflects the status of Hume as one of the first great modern philosophers of science, an empiricist, and a thinker uniquely concerned with the role of the emotions. Hume himself saw philosophical knowledgeproduction as affectively laden, calling philosophy “the ruling Passion of my Life and the Great Source of my Enjoyments” (Ayer 2000, 2). René Rosfort and Giovanni Stanghellini draw this out with a comparison to Spinoza. They note that “if one reads Spinoza’s geometrically and stringently structured Ethics (1677) together with Hume’s eloquent Treatise (1739-40), one cannot but notice the pronounced difference in atmosphere. Spinoza’s radical rationalism, with its logical denunciations of the passions, saturates the structure and expression of almost every sentence, and Hume’s no less radical empiricist skepticism towards our rational capacities allows his passions, sentiments, and taste to animate the text” (Rosfort and Stanghellini 2012, 411). I want to suggest that Hume’s particular version of empiricism leads not just to a close focus on the relationship between experiencing bodies and knowledge, but feeling bodies and knowledge. Science, Hume shows, is always attached to the vibrant grid of our affectively laced thoughts. Its internal limit is its nonseparability from our emotions. In his Enquiry Concerning Human Understanding, Hume lays out a model of scientific knowledge production as fundamentally detached from objective certainty. This is the famous critique of causality that has left so many undergraduate philosophy students agape. How do we know what we know, Hume asks? How do we know that the leaves will fall, that the moon will change phases, that water will quench our thirst? We don’t. We know what has happened well enough. The moon has changed phases many dozens of times in our lifetimes, and to the best of our knowledge

64

Chapter Four

for many lifetimes before ours. But Hume shows us that we can’t know that it will change phase again. Philosophers tend to make examples out of things around them— tables, chairs, and glasses of liquor on the desk. You can tell Hume is the worst person to invite to a party because his favourite example involves billiard balls. How many otherwise enjoyable evenings in the games room were ruined by Hume droning on to the players that they couldn’t know that the balls would move when hit? “Adam,” he concludes triumphantly, “though his rational faculties be supposed, at the very first, entirely perfect, could not have inferred from the fluidity, and transparency of water, that it would suffocate him, or from the light and warmth of fire, that it would consume him” (Hume 2007, 19). On this inductive principle, of course, rests the entirety of the empirical sciences. But there is no foundation, at least no certain foundation, and certainly no machine-like deductions to be made from formal reason. Hume introduces this as part of his point that all of our knowledge has its source in the things of the world. The battle is against causality, the war is against rationalism. Hume’s point is that if cause and effect can’t even be conclusively demonstrated by reason, what little could it offer by way of raw material for science? “What is the foundation of all our reasonings and conclusions concerning that relation [of cause and effect]?” Hume asks. “It may be replied in one word, EXPERIENCE” (Hume 2007, 23). We know only through the impressions left by the world in our bodies, which, for Hume, are of fundamentally the same substance as our minds. “Thus,” Hume closes, “the observation of human blindness and weakness is the result of all philosophy, and meets us, at every turn, in spite of our endeavours to elude or avoid it” (Hume 2007, 22). Empiricism is primarily concerned with spotlighting the internal limit of human knowledge production, its inability to transcend our finite bodies. But Hume is also confident that we do generate causal links, even in the absence of rational certainty (Hume 2007, 151). Immanuel Kant’s solution to this very problem a generation later is, I think, basically right: he suggests that the mind casts its own structuring grid over the objects of sense experience automatically arranging them in causal relations for us. But Hume’s own solution is not necessarily in contradiction with Kant’s, and its phrasing is more useful to us. For Hume, we feel causality. “If I see a billiard-ball moving towards another,” Hume states, blithely ruining yet another party, “I can easily conceive it to stop upon contact. This conception implies no contradiction; but still it feels very differently from that conception, by which I represent to myself the impulse, and the communication of motion from one ball to another” (Hume 2007, 35,

The Wild Experiment: Emotion, Reason, and the Limits of Science

65

emphasis mine). We feel it. We don’t know that a causal relationship is in effect through a line of rationally deduced propositions. We feel it. Hume goes even further than this. He contends that all knowledge is gauged by this emotion. He insists that “the difference between fiction and belief lies in some sentiment or feeling, which is annexed to the latter, not to the former, and which depends not on the will, nor can be commanded at pleasure. It must be excited by nature, like all other sentiments; and must arise from the particular situation, in which the mind is placed at any particular juncture” (Hume 2007, 35). Hume jettisons the cognitivist foundations of knowledge altogether, appealing to our shared intuitions about our own ways of knowing to insist that all belief is emotional. “BELIEF is the true and proper name of this feeling,” he affirms, “and no one is ever at a loss to know the meaning of that term; because every man is every moment conscious of the sentiment represented by it” (Hume 2007, 35). It is a radical proposal, an uncompromising refutation of rationalism. Belief is a feeling, a sentiment and it is ineradicably grounded in our bodies. This is both the limit and the horizon of the possibility of science. I want to pivot backwards in Hume’s canon, now, and retreat to his earlier Treatise of Human Nature. This is a somewhat controversial move among Hume scholars, of which I am not one, in that Hume repudiated this essay. But it’s an open question to me whether he really reversed course on the views contained therein, and regardless, I want to consider them as part of the same experimental ferment out of which the Enquiry grew. And just as Hume was forced to retreat from some of his intuitions—for instance, about the weakness of the argument from design (Dennett 1995, 32)—because the empirical science was still almost a century away from catching him. I think it is useful to revisit Hume with the openness to the possibility that some of his ideas may only now be accessible to us. Hume’s Treatise contains some of his most strident statements affirming the subordination of reason to passion. His mechanism for accomplishing this is to demonstrate that even though reason can illuminate the relationships between pieces of information, reason is feckless in the actual process of decision-making. Reason can’t discern a value. “Tis not contrary to reason to prefer the destruction of the whole world to the scratching of my finger,” Hume insists, before blustering that “[t]is not contrary to reason for me to chuse my total ruin, to prevent the least uneasiness of an Indian or person wholly unknown to me” (Hume 1960, 416). Reason simply can’t adjudicate between these preferences; it can only supply the information that will be animated by the force field of

66

Chapter Four

desire. Reason is inert without emotional energy animating it. This is where Hume deploys one of his most trenchant and controversial lines: “Reason is, and ought only to be the slave of the passions, and can never pretend to any other office than to serve and obey them” (Hume 1960, 415). In every arena, including knowledge production, reason is only the topsoil sprinkled over the tectonic plates of the passions beneath. What we think is reason is, for Hume—surprise—actually a passion. Specifically, it is a configuration of what he calls “calm passions.” “What we call strength of mind,” he proposes, “implies the prevalence of the calm passions above the violent” (Hume 1960, 418). The calm passions— passions that fixate on distant rewards or consequences rather than immediate rewards and pleasures—are arrayed in such a way as to discipline and outmaneuver the violent passions. “By reason we mean affections…; but such as operate more calmly, and cause no disorder in the temper: Which tranquility leads us into a mistake concerning them, and causes us to regard them as conclusions only of our intellectual faculties.” (Hume 1960, 437). The status of scientific knowledge production in this scheme is somewhat ambiguous—in part because Hume’s understanding of science and philosophy don’t precisely conform to our configuration of those categories. But in Part III of the Treatise, in a chapter entitled, “Of Curiosity, or the Love of Truth,” Hume points out that “many philosophers have consum’d their time, have destroy’d their health, and neglected their fortune, in the search of such truths, as they esteem’d important and useful to the world” (Hume 1960, 450). Scientific knowledge production is not an instance of the supremacy of reason over passion. The love of truth is, to its core, a peculiarly calibrated tissue of passions. And yet, whereas Hume emphasizes the calm passions in his account of the machinations of knowledge production, there is another element— drawn from the violent passions—that also seems to be part of the affective alchemy of science. This is surprise itself. Hume sees surprise as an emotion that not only has its own profile, but has an amplificatory effect on other emotions. “[T]ho’ surprise be agreeable in itself,” he writes, “yet as it puts the spirits in agitation, it not only augments our agreeable affections, but also our painful, according to the foregoing principle, that every emotion, which precedes or attends a passion, is easily converted into it” (Hume 1960, 423). Surprise adds an extra dimension of intensity to the landscape of knowledge production. This is why philosophical knowledge production is associated, for Hume, with two other pastimes: hunting and gaming. Hume muses that “there cannot be two passions more nearly resembling each other, than those of hunting and philosophy” (Hume 1960, 451). Only a moment later,

The Wild Experiment: Emotion, Reason, and the Limits of Science

67

though, he adds, “we may consider the passion of gaming, which affords a pleasure from the same principles as hunting and philosophy” (Hume 1960, 452). My contention would be that the synchrony aligning these three activities has everything to do with their convergent affective structures. Each is about the pleasurable navigation of a field of information that is occasionally punctuated by sharp spikes of surprise. Psychologists call this a “variable ratio response pattern,” a highly addictive schedule of rewards that essentially offers randomly spaced payouts, like a slot machine. It is, in short, the feeling of a wild experiment. Hume’s model of knowledge production is substantially different from Dawkins’. For Dawkins, the treasures of science are wondrous to behold. But the artifacts of truth have no internal emotional structure. Hume, by contrast, sees knowledge production itself as emotionally saturated. For Hume, truth and reason are in alignment with feeling, not external to the feeling apparatus. At the same time, there are parts of Hume’s framework that leave the Darwinian picture behind. For instance, Hume argues in a few instances that part of the pleasure of science must be its functionality. Writing of hunting (which is like philosophy, which is like gambling), Hume notes that “these actions must be attended with an idea of utility, in order to their having any effect upon us” (Hume 1960, 451). No-one would hunt crows or magpies because they cannot be eaten. And gambling is only pleasure because of the prospect of financial gain (Hume 1960, 452). Even philosophers, Hume writes, “[w]ere they convinc’d, that their discoveries were of no consequence, they wou’d entirely lose all relish for their studies” (Hume 1960, 450). I’m not sure that this principle of utility is sufficient for explaining the emotional urgency of science. It strikes me as trying to make science more reasonable than it actually is. Hume, in true Enlightenment fashion, can’t understand why natural philosophers would sacrifice their health or their wealth if it weren’t for some noble contrivance. But Darwin certainly did both, putting his health on the line and likely making very little, if any, profit from his scientific books (Browne 2002, 133). I think Hume underestimates the power of his own ideas. If the passions are sovereign over reason, it is not necessary to reroute scientific knowledge production through a calculative framework. Just as a gambler can become addicted to the dice, the roulette wheel, the first glance at a fresh hand of cards, so scientists can become addicted to their craft.

68

Chapter Four

William James: The Sentiment of Rationality This orientation to the emotionality of science continues in the postDarwinian period. Darwin’s late work The Expression of the Emotions in Man and Animals is designed to locate human emotions in continuity with animal emotions. He extends this even to thought itself. Paraphrasing Duchenne, he writes: “the senses, the imagination, and thought itself— elevated and abstract as we suppose it to be—cannot operate without arousing corresponding feeling” (Darwin 2009, 336). But Darwin develops this very little. It is the generation of philosophers and psychologists after Darwin, such as the American Pragmatists, who flesh this out. Although they are called “Pragmatists,” James, at least, was perfectly happy placing himself squarely in the lineage of Hume. In his 1907 essay collection The Will to Believe he explicitly referred to his method as “radical empiricism.” (James 1907, 7). Like Hume, he emphasized the priority of experience, rather than reflection, as the mode of knowledge production. And like Hume, James’ epistemology makes emotion central to the process of organizing information systematically in the generation of science. But James goes further than Hume in aligning the emotionality of knowledge production, of philosophy, and of religion in a single column. One of the best ways into understanding this skein of knowledge production and religion, wrapped together by emotion, is to start with his treatment of Pascal. In The Will to Believe, James lays out Pascal’s famous wager, from Pensées. James characterizes it as a coin flip, in which you must weigh up the costs and benefits of staking all on heads or all on tails. Once again, we are returned to gambling. For Pascal, this wager can only lead us to conclude that we must bet on salvation; the prospect of gaining an infinite reward justifies the low initial stake of faith, and so let there be masses and holy water for all. But whereas Hume and Darwin introduce gambling as a scene of delightful adventure, James finds nothing in it to celebrate. Instead, he sees it as a joyless settling of accounts. “We feel,” he writes, “that a faith in masses and holy water adopted willfully after such a mechanical calculation, would lack the inner soul of faith’s reality.” He continues, suddenly overtaken by a vindictive disposition: “and if we were ourselves in the place of the Deity, we should probably take particular pleasure in cutting off believers of this pattern from their infinite reward” (James 1907, 5). But this is the Puritan’s spiritual objection to Pascal’s wager as an arid, mathematical faith. For now, we are more interested in James’ objection to

The Wild Experiment: Emotion, Reason, and the Limits of Science

69

it on scientific grounds. James does not just think that Pascal’s faith is airless: he also emphasizes that to deduce masses and holy water from an arithmetic operation is absurd. “It is evident,” he insists, “that unless there be some pre-existing tendency to believe in masses and holy water, the option offered to the will by Pascal is not a living option” (James 1907, 5). He continues: Pascal’s “own personal belief in masses and holy water had far other springs” (James 1907, 5). The logical channel that Pascal opens for us is decidedly underwhelming for instilling the conviction of faith in masses and holy water—which is why, James notes, no Muslim and no Protestant was ever persuaded by it. Instead, James will emphasize that religion emerges not out of an arithmetic operation, but out of a deeper, living field of affective tissues. James is not the first thinker to affiliate religion and emotion, but he is the first to follow the Darwinian avenue to a comprehensive, versatile framework that grounded religion in our embodied, emotional life. Elsewhere in The Will to Believe, James will write that religion is the result of a “craving of the heart to believe that behind nature there is a spirit whose expression nature is” (James 1907, 40). Religion is not arrived at conceptually, but sourced in a felt urgency. “What philosophers call ‘natural theology,’” he proposes, “has been one way of appeasing this craving; that poetry of nature in which our English literature is so rich has been another way” (James 1907, 40). Theology, he elaborates in The Varieties of Religious Experience, is best understood not as conceptual machinery that leads to feeling, but as precisely the opposite: theology is an attempt to dress religious feeling in conceptual vestments: “in a world in which no religious feeling had ever existed, I doubt whether any philosophic theology could ever have been framed” (James 1961, 338). Pascal’s wager is misguided precisely because it tries to flatten religion to computation. The sources of our passion for religion can’t be corralled into a cognitive box. What we must bear in mind, though, is that for James it is not just religious information that is programmed by affective code emanating from the body. James will insist that “[w]hen [an organ, such as the liver] alters in one way the blood that percolates it, we get the methodist, when in another way, we get the atheist form of mind. So of all our raptures and our drynesses, our longings and pantings. They are equally organically founded, be they religious or of non-religious content” (James 1961, 30).

70

Chapter Four

Although reason has been cast in the starring role in the western epistemological drama, James affirms that the supporting cast of emotional dramatis personae are the real drivers of the plot. This leads to one of James’s key concepts—the sentiment of rationality. As James writes in his essay of that name, “[t]he transition from a state of puzzle and perplexity to rational comprehension is full of lively relief and pleasure” (James 1907, 63). As with Hume, knowledge production for James is achieved by emotional, rather than cognitive illumination. The scientific method, for James, is achieved through a balance between “eager interest” and “equally keen nervousness lest [we] become deceived” (James 1907, 21). Incidentally, this term and its definition bear striking similarities to what the neuroscientist Antonio Damasio has called “the passion for reason” (Damasio 1994, 245). As Rosfort and Stanghellini write, “[f]eelings motivate, orient, and sustain our rational thinking about the world, other people, and ourselves, and it is not possible to separate thinking and feeling in an actual human life” (Rosfort & Stanghellini 2012, 404). Again, this differs from the correlation of the fruits of science with wondrous gifts promised by Dawkins. Rosfort and Stanghellini affirm that “the intellectual emotions” are woven into the very act of science itself, conducting “an excitation of the will that allows us to find intellectual pleasure and even joy in pursuing a better understanding of that which may at first produce pain, surprise, boredom, sadness, or sorrow” (Rosfort and Stanghellini 2012, 406). And as with Hume, philosophy itself, in James’ framing, becomes an eminently affective operation. Richard Shusterman notes that James identified pragmatic philosophy as trafficking in a particular “strenuous mood”—”a mood of energetic striving and willingness to endure difficulties and risks in the melioristic effort to improve experience” (Shusterman 2012, 433). Pragmatism is a philosophy that has soberly reconciled itself to what I am calling the internal limit of knowledge production: the inability of our finite, physical souls to attain objective certainty. The rambunctious “strenuousness” of pragmatic philosophy is itself an effect of precisely this humbling and liberating recognition: “the Jamesian strenuous mood”. Shusterman argues that this humility “requires the pragmatist to live without the comfort of an Absolute that ultimately guarantees ultimate truth and harmonizes all conflict and difference” (Shusterman 2012, 449). As M. Gail Hamner writes, this is James’ notion of tough-mindedness, the quality of “individuals of conscience, persons who were not afraid to assert their differences from the status quo and yet who also did not hesitate to face their own limitations and failures” (Hamner 2003, 151). The spirit of pragmatist philosophy is one of

The Wild Experiment: Emotion, Reason, and the Limits of Science

71

modesty, then, but also one of exhilaration. At the intersection of these horizons emerges the affective profile of philosophical enquiry. At the same time, James, in his recapitulation of a Protestant, Pietist, even Puritan intellectual idiom that locates the resolute spiritual interior of the individual of faith as the paramount moral resource, ultimately returns us to an understanding of knowledge-production that is too much anchored to individual sovereignty. “James regards purpose through the lens of the individual,” Hamner observes, “as the means of negotiating moral interactions and personal interests” (Hamner 2003, 127). This becomes the prism through which James objects to certain varieties of atheism in parlance in his time. He repudiates the caution of atheists who insist on comprehensive proof before belief may be granted, pointing out that to withhold assent is just as much an assertion as to grant it. James’s defense of faith in Will to Believe ultimately comes down to a startlingly individualistic injunction that “No one of us ought to issue vetoes to the other, nor should we bandy words of abuse. We ought, on the contrary, delicately and profoundly to respect one another’s mental freedom: then only shall we bring about the intellectual republic; then only shall we have that spirit of inner tolerance without which all our outer tolerance is soulless, and which is empiricism’s glory; then only shall we live and let live, in speculative as well as in practical things” (James 1907, 30).

This liberal utopianism is an intriguing twist on the defense of faith, but ultimately it is inadequate as a model for how knowledge—whether religious or scientific—is fabricated. For James, the social dimension of the emotions is recklessly downplayed, ultimately leaving us in asocial cocoons of our own seemingly autogenerated affective dispositions. Pascal tries to reduce faith to a coin flip. But James shows that it is, of necessity, much more than this: there is an entire, glittering casino of games of chance calling our names. We’re pulled in multiple directions into the multiple ways of feeling and knowing the world. In this vortex of forces, we find ourselves knowing in ways both religious and nonreligious, both scientific and philosophical. What James doesn’t see is that this picture even outstrips the framework of Protestant individualism that he uses to enclose his own account of faith and knowledge production. This is why we need to consider one last perspective on how emotion sets up an internal limit of science.

72

Chapter Four

Affect Theory: Sensualized Epistemology The contemporary academic and cultural landscape has been increasingly given over to interest in the emotions—so much so that one of the waves in the humanities academy right now is known as the “affective turn.” Rosfort and Stanghellini divide scholarship on the emotions into two groups. The dominant group are the “cognitive theories,” represented by figures such as Sartre, Anthony Kenny, Robert Gordon, Martha Nussbaum, and Robert Solomon, who “consider cognitive capacities and rational dispositions to be at the heart of emotional life” (Rosfort and Stanghellini 2012, 396). Arrayed against this group are the “feeling theories,” typified by figures such as James, Antonio Damasio, Jenefer Robinson, and Jesse Prinz, who “argue that physiological reverberations are what makes an emotion emotional” (Rosfort and Stanghellini 2012, 396). I want to follow here a track that is more associated with the subordinate group, the contemporary critical field known as affect theory. Like the figures mentioned, affect theorists tend to make emotion, not cognition, primary in the productions of embodied life. In this way, they offer another corroborating voice in the story of the link between science and feeling. Moving from a specific figure to an entire subfield leaves one with a wealth of conversations to tap, so I will focus here on just a few axes within affect theory. One such axis is grounded in the work of the philosopher Martin Heidegger. Heidegger starts from a model of the self that is similar in many respects to that proposed by James and the pragmatists: it is a self that is fundamentally embedded in the world rather than detached from it, a self that is by doing and experiencing rather than by detaching and observing. Heidegger calls this “Being-in-the-world.” Equally important for later affect theorists, Heidegger also underscores that part of being-in-the-world is being “in a mood.” As embodied beings embedded in the world, we are always attuned, never in neutral: “it is not possible not to be in a mood” (Ahmed 2014, 14). Only the ice cube self that is walled off from the world can inure itself to the flows and impositions of the world on its emotional state. What matters for our purposes is that this moodiness becomes both the precondition and the structuring envelope of knowledge-production. Mood, as Rita Felski and Susan Fraiman point out, “is not optional, but a prerequisite for any kind of intellectual engagement…. Whether our attitude is ironic or irenic, generous or guarded, strenuous or languourous [sic] will help determine how we situate ourselves in relation to an object of study and what we find most salient.” As with Hume and James, they underscore that “critical detachment is not an absence of mood, but one

The Wild Experiment: Emotion, Reason, and the Limits of Science

73

manifestation of it” (Felski and Fraiman, 2012, vi). The scientific method is a specially trained configuration of dispositions that enables the emergence and organization of particular objects of knowledge within the information field of the world. Knowing, for Heidegger, is itself a process of “Being-in-the-world,” which necessitates concern, leaving us “fascinated by the world” (Heidegger 1962, 88). And so Felski and Fraiman write, “Mood impinges on method” (Felski and Fraiman 2012, vi). An even more precise perspective on the science-emotion relationship is offered by another background figure in affect theory, the 20th-century Princeton psychologist Silvan Tomkins, especially in the reading of his most astute contemporary interpreter, the queer theorist Eve Kosofsky Sedgwick. Tomkins is an exponent of a version of what psychologists would now call the Basic Emotions hypothesis: he believes that experience is patterned by a set of emotional templates that are embedded in human neurobiology. In Tomkins’ portfolio, there are 9: Surprise-Startle, Distress-Anguish, Anger-Rage, Enjoyment-Joy, Interest-Excitement, FearTerror, Shame-Humiliation, Dissmell, and Disgust. As you might guess, Interest-Excitement is the faculty that is of most concern to us. It is the Interest-Excitement pole, for instance, that Tomkins suggests is the “primary affective investment” of Darwin (Tomkins 1995, 75). For Tomkins, interest is a sort of affective master-switch. It is the animating force that moves us through the world, bearing marked similarity to what the neuroscientist Jaak Panksepp has called the mammalian brain’s SEEKING system, also known as the “foraging/ exploration/investigation/curiosity/interest/expectancy” system that “leads organisms to eagerly pursue the fruits of their environment—from nuts to knowledge, so to speak” (Panksepp 1998, 145). As Tomkins writes: “The interrelationships between the affect of interest and the functions of thought and memory are so extensive that absence of the affective support of interest would jeopardize intellectual development no less than destruction of brain tissue. To think, as to engage in any other human activity, one must care, one must be excited, must be continually rewarded” (Tomkins 1995, 76).

Not only our wild experiments, but our everyday modes of exploration and interaction with the world are guided by this system. The notion of a separation between truth and feeling that leaves science unfettered to soar over the world is absent here. “Reason without affect,” Tomkins says, “would be impotent, affect without reason would be blind” (Tomkins 1995, 37).

74

Chapter Four

Sedgwick takes this notion of the intrinsic affective lacing of knowledge production still further. Her concern, especially in her famous piece Paranoid vs. Reparative Reading, or, You’re So Paranoid You Probably Think This Essay is about You, is to show that strategies of knowledge-production are not only affective, but differently affective. Bias, in this view, is inescapable, but different biases are not defined by different conceptual presuppositions, but different affective orientations. What Sedgwick calls paranoid reading is defined by an insistence on slicing off the surface of an object under study in order to augur its secret intentions. “The first imperative of paranoia,” Sedgwick writes, “is there must be no bad surprises” (Sedgwick 2003, 130). Sedgwick proposes that paranoid reading is a methodology constituted by a particular “cognitive/affective theoretical practice” (Sedgwick 2003, 126). That slash between “cognitive” and “affective” is what puts this point of view in alignment with the other thinkers we’ve explored. Sedgwick characterizes paranoid reading as a particular way of entering an affective relationship with an object of study. What makes this configuration of affect and knowledge distinctive is the specific way that it makes knowledge feel. “Whatever account it may give of its own motivation,” Sedgwick writes, “paranoia is characterized by placing, in practice, an extraordinary stress on the efficacy of knowledge per se—knowledge in the form of exposure” (Sedgwick 2003, 138). Paranoid reading is about becoming the agent of unmasking, the director of surprises, rather than their victim. As one of the founders of queer theory based in the storied literature program at Duke University, Sedgwick’s milieu is the world of what is sometimes called “high theory” or the “high humanities,” the first wave of interpretations in the American adoption of poststructuralism in the 1970s and 1980s. Sedgwick’s account of paranoid reading, first offered in the mid-1990s, is a critique of this phenomenon. She laments that “to theorize out of anything but a paranoid critical stance has come to seem naïve, pious, or complaisant” (Sedgwick 2003, 126). She proposes that the adoption of certain strands of Marxism, deconstruction, and the hermeneutics of suspicion have turned into an orthodoxy of an imperative to expose the secret machinations of every piece of culture. By way of example, she puts forward queer theorist Judith Butler’s analysis of drag performances, or camp. Butler believes that drag performances are designed to show the fungibility of masculinity and femininity, and thereby destroy gender. “By this account,” Sedgwick writes, “the x-ray gaze of the paranoid impulse in camp sees through to an unfleshed skeleton of the culture; the paranoid aesthetic on view here is one of minimalist elegance and conceptualist economy” (Sedgwick 2003, 149).

The Wild Experiment: Emotion, Reason, and the Limits of Science

75

Sedgwick counterposes paranoid reading with another approach, what she calls reparative reading. If paranoid reading is the analytics of eyerolling, the x-ray stare, the hermeneutics of the smirk, reparative reading is eager, wide-eyed, agape, a passion for encountering texts that anticipates compassion and nurturance. It cultivates a strategic vulnerability, a willingness to allow oneself to be surprised. “The desire of a reparative impulse,” she writes, “is additive and accretive. Its fear, a realistic one, is that the culture surrounding it is inadequate or inimical to its nurture; it wants to assemble and confer plenitude on an object that will then have resources to offer to an inchoate self” (Sedgwick 2003, 149). Reparative reading is the antithesis of paranoid reading, a cultivated optimism, a passion for one’s own bedazzlement rather than an insistence on being immune to surprise. This spills over into Sedgwick’s reparative interpretation of camp. Rather than seeing it as an act of sabotage, she invites us to be open to the possibility that drag performance is an act of love, admiration, celebration, or joy (Sedgwick 2003, 149). Sedgwick’s pairing of these two cognitive-affective theoretical practices depicts the tight overlap between affect and the production of discourse. “What does knowledge do,” Sedgwick asks: “the pursuit of it, the having and exposing of it, the receiving again of knowledge of what one already knows?” (Sedgwick 2003, 124). This is what Lauren Berlant has called Sedgwick’s “sensualized epistemology,” the way that forms of knowledge are valued and circulated not or not only for their epistemic utility, but for their affective capacities (Berlant 2011, 158). For affect theorists, bodies seek out opportunities to learn and explore not because they provide truth per se, but because they make our affects shimmer. Sara Ahmed proposes that we “refuse to see the ambivalence of affect as pointed: maybe the point is that there is no point that points to some future horizon. Feelings may be perverse because they don’t always have a point” (Ahmed 2010, 177). As I have argued elsewhere, where rhetoric theory asks “how are emotions marshaled in the service of arguments?” affect theory asks the reverse: “how are arguments marshaled in the service of affect?” (Schaefer 2015). Feminist philosopher of religion Kimerer LaMothe gives us a microlevel view of this process. She writes: “The sensory patterns that guide our study can be subtle indeed. A tingle of awe in response to a text or ritual or event impels us to learn more. A wrinkle of confusion troubles us to revisit a mark that remains and ask new questions, gather more data, consult other experts. A surge of excitement pushes us on as pieces of our research start falling into shapes of understanding we have learned to recognize. A sense of ease lets us know

76

Chapter Four that our interpretations are objective or empathetic enough, or will be well received. Such micro-moments in our lived experience, more than our stated principles or procedures, come to define our way of studying religion” (LaMothe 2008, 589).

This passage is a catalogue of the subtle affective mechanisms of knowledge-production, a theme presented by all of the thinkers we have considered so far. It is a diagram of both the possibilities of science— broadly understood—and its ineradicable limits. But as affect theory in particular emphasizes, we know in ways that have specific configurations of feeling.

Conclusion Christopher Southgate describes, in poem form, his ecstasy at being given the opportunity to spend a fortnight at Fred Sanger’s MRC Lab: “I want the fortnight to last forever”. My Oxford colleague Andrew Briggs once described his process of working on a physics problem, how he experienced it as an actual physical pain that could only be relieved when he arrived at the solution. In her biography of A Feeling for the Organism, Evelyn Fox Keller quotes the corn geneticist Barbara McClintock as saying: “I start with the seedling, and I don’t want to leave it…. I know every plant in the field. I know them intimately, and I find it a great pleasure to know them” (Keller 1993, 198). Darwin’s own ecstasy in the game of knowledge production is wellknown. “Delight,” Darwin journaled while on the Beagle, “is a weak term to express the feelings of a naturalist who for the first time has wandered by himself in a Brazilian forest… such a day brings a deeper pleasure than he can ever hope to experience again” (Browne 1995, 211). George Levine writes that the “affect of [Darwin’s] writing has little to do with the wondrous nature of the phenomena science investigates, everything to do with the moral urgency of pursuing rational truth above any other forms of desire, and with the scientist’s power to do just that” (Levine 2006, 100, emphasis original). Science is laced with emotion, not just as its fruits, as Dawkins claims, but in its everyday fabric. Everyone feels this. Everyone knows it. But we are held back by a structure of common sense—again typified by Dawkins—that says that knowledge is emotionally inert. In this, he is insensate to the internal limit of science: its total dependence on our irrevocably sensualized epistemology. This is not an anti-science position. It is, instead, a way of amplifying an eminently Darwinian motif, spotlighting the animality of science—the animality of knowledge production itself. This motif—the way that

The Wild Experiment: Emotion, Reason, and the Limits of Science

77

science is done by bodies, rather than detached minds—has been developed extensively (though seldom with reference to affect) in the field of Science and Technology Studies. As Bruno Latour writes, the project of Science and Technology Studies “was never to get away from facts but closer to them, not fighting empiricism but, on the contrary, renewing empiricism” (Latour 2004, 231). By tracing the emotional seams of knowledge production, we find ourselves back in Darwin’s hands, exorcising the pre-Darwinian, Enlightenment logic that tries to disconnect human minds from the world. But even though my theme here is the internal limits of science—its inability to ever jump outside human feeling—I want to conclude by indicating a way that this perspective rearranges science’s limits, rather than merely fortifying them. Setting science against the broader context of human knowledge production dissolves the boundaries dividing science from other intellectual disciplines. As Peter Harrison reminds us, repeatedly, science has not always been “science.” In the time of Aquinas, there were seven “liberal sciences”, grammar, logic, rhetoric, arithmetic, astronomy, music, and geometry (Harrison 2015, 14). In the first edition of the Encyclopedia Britannica in 1771, science is defined as “any doctrine, deduced from self-evident and certain principles, by a regular demonstration” (Harrison 2015, 15). Science means any form of systematic knowledge, hence the literary periodical the Athenaeum, in 1828, “divided the sciences into ‘exact, experimental, speculative and moral’”. William Whewell considered natural theology to be an inductive science, as did others who still referred to theology as “the first and greatest of the sciences” (Harrison 2015, 147). Science as we now imagine it—test tubes and soil samples, equations and telescopes—was once only a fraction of the whole. But seen from the perspective of emotion, we realize that the more ecumenical definition of “science” was the correct one—all endeavors of human knowledge production, the natural and social sciences, the humanities, poetry, art, and religion itself, are dimensions of an emotionally animated way of encountering the world. In this sense, the meaning of the word “science” must dilate to include all modes of knowledge production, outstripping artificial barriers between, for instance, the natural sciences, social sciences, and humanities. They are all ventures, all games of chance, all ways of throwing ourselves into the wild experiment of life.

78

Chapter Four

References Ahmed, S. 2014. “Not in the Mood”, New Formations 82, 13-28. Asad, T. 2003. Formations of the Secular: Christianity, Islam, Modernity. Stanford: Stanford University Press. Ayer, A. 2000. Hume: A Very Short Introduction. Oxford: Oxford University Press. Berlant, L. 2011. Cruel Optimism. Durham and London: Duke University Press. Browne, J. 1995. Charles Darwin: Voyaging: A Biography. Princeton, NJ: Princeton University Press. —. 2002. Charles Darwin: The Power of Place. New York: Knopf. Damasio, A. R. 1994. Descartes’ Error: Emotion, Reason, and the Human Brain. New York: Putnam. Darwin, C. 1898. The Life and Letters of Charles Darwin. Darwin, F. (ed.), New York: D. Appleton and Co., 1897. —. 2009. The Expression of the Emotions in Man and Animals. Cain, J. and Messenger, S. (eds.), London: Penguin Books. Dawkins, R. 1998. Unweaving the Rainbow: Science, Delusion and the Appetite for Wonder. Boston: Houghton Mifflin Co. —. 2008. The God Delusion. New York: Mariner Books. Dennett, D.C. 1995. Darwin’s Dangerous Idea: Evolution and the Meanings of Life. New York: Simon & Schuster. Felski, R. and Fraiman, S. 2012. “Introduction: In the Mood”, New Literary History, 43(3), v-xii. Hamner, M. G. 2003. American Pragmatism: A Religious Genealogy. Oxford: Oxford University Press. Harrison, P. 2015. The Territories of Science and Religion. Chicago and London: The University of Chicago Press. Heidegger, M. 1962. Being and Time. Macquarrie, J. and Robinson, E (trans.). Malden, MA, Oxford, UK, and Victoria, Australia: Blackwell Publishing Ltd. Hume, D. 1960. A Treatise of Human Nature. Selby-Bigge, L.A. (ed.), Oxford: Oxford University Press. —. 2007. An Enquiry Concerning Human Understanding. Milligan, P. (ed.), Oxford: Oxford University Press. Jakobsen, J. R., and Pellegrini, A. 2008. “Introduction: Times Like These”, in Jakobsen, J.R. and Pellegrini, A. (eds.), Secularisms. Durham & London: Duke University Press, 1-35.

The Wild Experiment: Emotion, Reason, and the Limits of Science

79

James, W. 1907. The Will to Believe and Other Essays in Popular Philosophy. New York, London, and Bombay: Longmans Green & Co.. —. 1961. The Varieties of Religious Experience: A Study in Human Nature. New York: Collier Books. Keller, E.F. 1993. A Feeling for the Organism: The Life and Work of Barbara McClintock. Tenth Anniversary Edition. New York: Owl Books. LaMothe, K. 2008. “What Bodies Know About Religion and the Study of It”, Journal of the American Academy of Religion, 76, 3, 573-601. Latour, B. 2007. “Why Has Critique Run Out of Steam? From Matters of Fact to Matters of Concern”, Critical Inquiry, 30, 2, 225-248. Levine, G. 2006. Darwin Loves You: Natural Selection and the Reenchantment of the World. Princeton, NJ: Princeton University Press. Panksepp, J. 1998. Affective Neuroscience: The Foundations of Human and Animal Emotions. Oxford and New York: Oxford University Press. Rosfort, R. and Stanghellini, G. 2012. “In the Mood for Thought: Feeling and Thinking in Philosophy”, New Literary History, 43, 3, 395-417. Schaefer, D. O. 2015. Religious Affects: Animality, Evolution, and Power. Durham, NC: Duke University Press. Sedgwick, E.K. 2003. Touching Feeling: Affect, Pedagogy, Performativity. Durham and London: Duke University Press. Shusterman, R. 2012. “Thought in the Strenuous Mood: Pragmatism as a Philosophy of Feeling”, New Literary History, 43, 3, 433-454. Svendsen, L. 2012. “Moods and the Meaning of Philosophy”, New Literary History, 43, 3, 419-431. Tomkins, S.S. 1995. Shame and Its Sisters: A Silvan Tomkins Reader. Sedgwick, E.K. and Frank, A. (eds.), Durham, NC and London: Duke University Press.

CHAPTER FIVE WHAT SHOULD CHRISTIAN THEOLOGY (NOT) LEARN FROM SCIENCE? THE CASE OF THE HUMAN BRAIN NEIL MESSER

Introduction Recent scientific developments seem to promise new insights into many aspects of human nature, including religion and morality. Neuroscientific studies of religion, alongside evolutionary and cognitive science accounts, generate diverse claims about the neural mechanisms involved in religious belief and their evolutionary origins. Brain imaging studies of moral decision-making suggest that our moral judgements are much less reasoned than we usually imagine, and some authors have tried to build normative ethical proposals on the back of these empirical studies. Research like this might seem to have far-reaching implications and raise troubling questions for Christian theology and ethics. Yet some of the bold claims made for it reflect confusion about the kinds of intellectual work that science is capable of doing. So, the cognitive neuroscience of religion and morality offers a valuable test case of the contributions science can and cannot make to the understanding of theology and ethics – the limits of science in understanding these aspects of human life. It therefore also serves as a test case of the opportunities and the pitfalls for Christian theologians who wish to engage with scientific accounts of what it means to be human. In this chapter I shall try to develop this test case in three stages. First, I shall briefly survey some evolutionary, cognitive and neuroscientific accounts of religion. Next, I shall sketch out various ways in which Christian theology might engage with such accounts, and assess their strengths, weaknesses and hazards. I shall use this assessment to propose some methodological conclusions about how the science-theology

What Should Christian Theology (Not) Learn from Science?

81

dialogue should be conducted. In the final part of the paper I shall use the cognitive neuroscience of morality as a kind of worked example to illustrate my methodological proposal.

1. Scientific studies of religious belief and experience I begin with the scientific study of religion; specifically with cognitive, evolutionary and neuroscientific approaches.

a. Cognitive science of religion (CSR) In broad terms CSR is the attempt to identify cognitive mechanisms that give rise to religious or supernatural beliefs. Its basic premise is that human minds consist of many specific mental “tools” or “devices” – cognitive systems for processing different kinds of information. Though CSR accounts are diverse, they have a widely shared core of ideas that some commentators call the “standard model” (Murray and Goldberg 2009, 183-89). The first concerns what are called “minimally counter-intuitive” (MCI) concepts. These are concepts that violate a few common-sense expectations but not too many: for example, the concept of a human being who can pass through solid walls, but otherwise resembles everyone else, would be minimally counter-intuitive. It turns out that such concepts are more memorable and easily transmitted than either fully intuitive or highly counter-intuitive concepts (e.g. Boyer 2001, chp. 2). CSR scholars argue that concepts of supernatural agents such as gods are MCI, and therefore tend to persist in our minds and spread in human cultures. Second, we have a “hypersensitive agency detection device” (HADD): a cognitive system that attributes events to the intentional activity of agents, even when it has little information on which to base such a judgment (Barrett 2007, 773). It is argued that this device predisposes us to attribute unexplained events to the activity of supernatural agents. Authors vary, however, in the weight they give to the HADD in explaining religious beliefs (compare Boyer 2001, 144-48, with Barrett 2007, 77273). Third, we have a theory of mind, which leads us to attribute minds to other human and human-like beings, and therefore to assume that they have cognitions, beliefs, desires, motivations and so forth rather like ours. This predisposes us to regard supernatural beings as “minded” (Murray and Goldberg 2009, 187).

82

Chapter Five

Fourth, concepts of supernatural agents are “inference rich”, enabling our minds to construct narratives about them that enhance their memorability and importance to us (Murray and Goldberg 2009, 187-88). For example, invisible, powerful and all-seeing agents may know what we are up to even when we think no-one is looking (cf. Johnson and Bering 2009), and they could be in a position to affect our lives for good or ill (Boyer 2001, 155-60). CSR is by no means without its critics. One general criticism is that it represents religion in simplistic or caricatured ways. For example, the catch-all category of “supernatural agents”, often used to include ghosts, spirits, zombies, demons, gods and sundry other beings, seems rather distant from the God whom Christians, for one, believe they encounter in their prayer, worship and service (Deane-Drummond 2009, 63-68). Sometimes such simplistic and caricatured presentations reflect a sceptical or anti-religious stance (Murray 2009, 168-69). This is not inevitable: Justin Barrett, for example, strongly denies that anti-religious agendas need be a part of CSR, though even he acknowledges that it is often used “rhetorical[ly]” to promote such agendas (Barrett 2007, 779). More specifically, the theologian Markus Mühling criticizes CSR for the assumption that the paradigm religious phenomenon is some kind of exceptional or extraordinary experience. If this assumption is made, he says, a good deal of Christian faith and practice tends to disappear from view (Mühling 2014, 92-97). Mühling is surely correct to criticize this assumption, but it is a moot point whether it is a general problem with CSR approaches. It is expressly denied, for example, by Pascal Boyer (2001, 307-11). The criticism may be nearer the mark in relation to neuroscientific studies, as we shall see later.

b. Evolutionary approaches While CSR is a distinct field from evolutionary theories of religion, the two are closely linked in much of the literature, for fairly obvious reasons. Two broad approaches to the evolutionary explanation of religion can be found in the literature: one is to account for religion as an evolutionary byproduct of other adaptive features; the other is to argue that religion itself is or has been adaptive in our evolutionary history. Evolutionary explanations that make use of CSR concepts tend to account for religion as a “spandrel” (Gould and Lewontin 1979): a nonadaptive by-product of an adaptive trait. For example, a mental device that enabled our ancestors to recognize other agents with desires and intentions could have conferred an obvious evolutionary advantage. Moreover, it

What Should Christian Theology (Not) Learn from Science?

83

would have been adaptive for this device to be hypersensitive, because false positives would be less costly than false negatives. Likewise, it is easy to see how the other mental devices in the standard model could have been adaptive in human evolutionary history. The interaction of these various devices might then generate religious concepts and beliefs even though those concepts and beliefs themselves conferred no selective advantage. Other evolutionary explanations do treat religion as adaptive, often on the grounds that it promoted group co-operation among our ancestors. One of the more sophisticated accounts of this sort is offered by David Sloan Wilson in his book Darwin’s Cathedral (Wilson 2002). Emphasising that evolution operates at various levels of selection, he gives a groupselectionist account of religions as adaptations for promoting group cohesion. At the heart of his book, a detailed account of John Calvin’s Geneva is used as a test-case. Wilson argues that Calvinist belief in all its aspects, as reflected in the first Genevan catechism of 1538, represents a pro-social adaptation. He reads Calvin’s catechism as fundamentally concerned with two things: the “people-people relationship” and the “Godpeople relationship” (Wilson 2002, 94). Concerning the former, “Calvin’s catechism provides a blueprint for human conduct that makes sense from a group-level functional perspective” (Wilson 2002, 98). This blueprint includes the Decalogue, the Golden Rule, and more contextually specific injunctions about the church and the civic community. Wilson treats Calvin’s account of the God-people relationship as “a fictional beliefsystem” designed to motivate believers to conform to the requirements of the people-people relationship (Wilson 2002, 98-105). Wilson more or less takes it for granted that Calvin’s theology is fictional, a point to which I shall return. His way of accounting for theological beliefs also becomes something of a Procrustean bed. He emphasizes aspects of Calvin’s theology, like awareness of human depravity and gratitude for God’s mercy, which can easily be fitted into his explanation, but more or less ignores other aspects like the Trinity, creation and Christology, which are less convincingly explained as fictions designed to encourage group co-operation.1 The problem this gives Wilson is that (in contrast to some other approaches, including CSR) he has 1

His Table 3.2, ‘Elements of the God-people relationship specified by Calvin’s catechism of 1538’ (Wilson, 2002, 101) lists ‘Forgiveness of sins’, ‘Faith’, ‘Internalization’, and ‘Preparing for the second coming’, but neither the table nor the following discussion even mentions the discussions of the Trinity, creation or Christology found in article 20 of the catechism, which concerns the Apostle’s Creed.

84

Chapter Five

committed himself to explaining the whole package of Calvinist belief and practice, and has deliberately chosen the 1538 catechism as his dataset. But at best he can only account convincingly for some of his chosen data. A further kind of evolutionary explanation emphasises cultural rather than genetic evolution. A well-known example of this type is found in Daniel Dennett’s book Breaking the Spell (Dennett 2006, 179-88). Dennett understands religious beliefs and practices as “memes,” or units of cultural inheritance, which promote behaviours that enable them to persist and spread in the population. The concept of “memes”, however, is itself contentious (Midgley 2000; Schloss 2009, 24), and this is just one of the objections that have been levelled at Dennett’s account (for other criticisms, see Geertz 2008). Of course, not all these evolutionary explanations need be mutually exclusive. If human evolution operates at multiple levels of selection, it would not be surprising if an area of human life as complex and diverse as religion required a variety of evolutionary (and other) explanations (Richerson and Newson 2009).

c. The neuroscience of religion Although the brain’s role in religious belief has been of interest at least since the days of William James, the experimental neuroscience of religion is a far younger enterprise: Uffe Schjoedt remarked in a 2009 review that “the list of actual experiments [was] surprisingly short” (Schjoedt 2009, 333). An early – and controversial – episode in this story was Michael Persinger’s “God helmet”, a device that he claims can induce the “sensed presence’ of another being by using a weak magnetic field to induce temporal lobe activity (e.g. St-Pierre and Persinger, 2006). It would be fair to say that this claim is regarded with scepticism by others in the field (Schjoedt 2009, 320-22).2 Various investigators have studied the neural correlates of religious or mystical experiences. Among studies of Christian mysticism, one of the best known was by Mario Beauregard and Vincent Paquette, who carried out fMRI and EEG studies on the mystical experiences of a group of Carmelite nuns (Beauregard and Paquette 2006; 2008). They found that these nuns’ experiences were associated with several regions and systems in the brain: there is no single “God module” or “God spot”, as others have 2

For one particular line of critique and Persinger’s responses, see Granqvist et al. (2005), Persinger and Koren (2005), Larsson et al. (2005), and St-Pierre and Persinger (2006). For further critique of Persinger’s work, see Beauregard and O’Leary (2007, chp. 4).

What Should Christian Theology (Not) Learn from Science?

85

claimed (Beauregard and O’Leary 2007, 274). In a popular account of this work, Beauregard and his co-author Denyse O’Leary have also argued that neuroscientific studies of religious and mystical experience do not favour a materialist view of religion and the brain, but rather support – though they cannot prove – the view that such experiences refer to a divine or supernatural reality beyond the individual subject (Beauregard and O’Leary 2007, chps. 9, 10). Some researchers have proposed particular models or mechanisms for religious experience. For example, in the 1990s Eugene d’Aquili and Andrew Newberg (1999) developed a complex model of the neural mechanisms underlying religious experience. At its heart was a proposed mechanism for the experience of “absolute unitary being” (AUB), said to be a common feature of mystical experience in diverse traditions. The proposed mechanism involved the blocking of input to the posterior superior parietal lobe, and more recently Newberg has offered some functional imaging evidence in support of this proposal (Newberg et al 2001; 2003; 2006; 2010; Wang et al 2011). Schjoedt acknowledges that such models may be useful in generating testable hypotheses. However, he is critical of theories that (as he sees it) depend on conjectures about distinctive brain systems or mechanisms for which there is not yet strong evidence. More generally he has concerns about the methodology and design of many studies in this field prior to his own. He is also critical of approaches that focus on exceptional experiences and expert practitioners – echoing from a very different standpoint the criticism by Markus Mühling noted earlier. In his own studies Schjoedt adopts a piecemeal approach, focuses on quotidian experiences and practices – such as different forms of prayer used by lay Christians – and works on the assumption that the neural mechanisms associated with religion will be the same as those involved in other aspects of human life (Schjoedt 2009, 327-32). In this section I have briefly surveyed three distinct – though related – scientific approaches to the study of religious belief and experience. Each is internally diverse and has its fair share of controversies. There is no shortage of theories and proposals in any of these fields, though few seem to be very firmly established. In a forum concerned with the interactions of science and religion, what should we make of these rather tangled discussions? Do these scientific studies of religion have anything to say about the validity or the content of the beliefs and practices they study? In particular (since I am a Christian theologian): what if anything should Christian theology be

86

Chapter Five

willing to learn – and what should it not be willing to learn – from studies such as these?

2. Theological implications and a methodological proposal Answers to the questions I have just posed are diverse, sometimes confusing, sometimes – perhaps – confused. I can best sort through the tangle and propose my own answer by using a typology of approaches that I first developed a few years ago for a different (though related) discussion (Messer 2007, 49-62). Those familiar with the work of Hans Frei will recognize my indebtedness to his Types of Christian Theology (Frei 1992; see also DeHart 2006, chp. 5) – though my typology is not simply equivalent to his, having been developed for a different purpose to answer different questions. If Christians wish to understand what it is to be human, what contribution to that understanding should they expect from theology and what contribution from evolutionary, cognitive and neuroscientific accounts? Broadly speaking, five types of answer can be distinguished: (1) theology makes no contribution to an understanding derived from science alone; (2) both science and theology contribute, but science is the dominant partner in determining the shape and content of the understanding; (3) both science and theology contribute by means of what Frei (1992, 38 and passim) called “ad hoc correlations”, neither side dominating; (4) both contribute, but theology is the dominant partner, shaping the understanding and critically appropriating insights from science; (5) only theology contributes, science having nothing to offer to a Christian understanding of humanity before God.

Type 5: Only theology contributes, science having nothing to offer to a Christian understanding To begin with type 5: one form this type can take is the late Stephen Jay Gould’s proposal that science, on the one hand, and religion and ethics, on the other, are “non-overlapping magisteria” (NOMA): each is valid in its own realm, but they ask distinct and different questions and neither has a bearing on the other (Gould 2002). In that case, if Christians wish to understand themselves as moral beings in relation to God, one another and the world, their understanding must be derived entirely from their Scriptures and traditions of faith. Scientific accounts of religion or morality may be interesting in their own right, but will be simply irrelevant to such a Christian self-understanding.

What Should Christian Theology (Not) Learn from Science?

87

This position might seem to gain some support from the wide consensus that scientific explanations of religion are neutral with respect to religious truth-claims. Most people who work in this area understand well enough that it would be a kind of genetic fallacy to draw conclusions about the truth or falsity of religious beliefs from a causal explanation of how those beliefs come to exist. This might initially seem an attractive way to keep the peace between science and theology, but in the fields we are considering, the border between Gould’s two “magisteria” turns out to be difficult, maybe impossible, to mark out and police. One aspect of the difficulty is that, as Jeffrey Schloss (2009, 10-12) observes, neutrality is not always what it seems. In some cases it may reflect a commitment to strict objectivity in the scientific study of religion, and a concern that either pro- or antireligious agendas may compromise that objectivity. Thus Armin Geertz, for example, is hostile to neurotheology because he discerns in it a religious agenda that risks compromising the scientific quality of its studies, and he exhorts neuroscientists who study religion: “Please keep your religious agendas out of your work … Otherwise we can only conclude that your results indicate nothing other than teleological confirmation of religious claims” (Geertz 2009, 324). But he is equally critical of the anti-religious polemic in Dennett’s Breaking the Spell (Geertz 2008). In other cases, talk of neutrality goes along with a general scepticism towards religious truth-claims. Paul Bloom, for example, maintains that while scientific explanations cannot refute theistic beliefs, they can nonetheless “challenge the rationality of those who hold such beliefs’ by showing that the route to those beliefs is epistemically unreliable (Bloom 2009, 125). In short, many evolutionary and cognitive scientists of religion, though they profess some kind of neutrality, are better described by type 1 than type 5. Moreover, for a faith that regards the physical world and its living inhabitants as God’s good creation, there are theological reasons for being uneasy with the NOMA solution. Science cannot tell us that the world is (or is not) God’s creation; but it would be odd if science had nothing of importance to teach us about the world that we know theologically to be God’s creation.

Type 1: theology has nothing to contribute to an understanding derived from science alone At the other end of the scale is type 1. This is hardly a promising option for those wanting to articulate a Christian understanding of what it is to be

88

Chapter Five

human, since it is in effect a rejection of that project: a call for theological understandings to be displaced by those based on science alone. Whether such an aim is possible or even coherent is open to question, but at any rate views like this are quite frequently encountered in the literature on scientific approaches to the study of religion. As I have already observed, most authors in these fields sign up to neutrality with respect to the truth of the beliefs they study; but that neutrality is often associated with some kind of naturalistic perspective. Depending on the form this takes, people of faith need not necessarily have any quarrel with it. It is widely agreed that natural scientific research is quite properly methodologically naturalist: aiming to account for natural phenomena by natural cause and effect, without invoking supernatural causes to fill explanatory gaps. Moreover, there could in a paradoxical way be theological reasons for doing natural science etsi deus non daretur, “as if there were no God” – to re-borrow a phrase that Bonhoeffer borrowed from Grotius.3 Presumably, if explaining the existence of religious belief and practice and the forms they take is a valid area of natural-scientific enquiry, it should be no exception. It is a further step, however, from this kind of methodological naturalism to the ontological naturalism which claims that “[n]aturalistic explanations of supernaturalistic belief … are ‘all the explanation there is’” (van Inwagen 2009, 134). In the literature, this kind of ontologically naturalistic claim may be related to methodologically naturalistic studies of religion in either of two different ways. For some authors it is an implicit or explicit prior assumption. It seems to function in this way, for example, for David Sloan Wilson in Darwin’s Cathedral, since he describes the theological claims of Calvin’s catechism as a fictional belief system designed to encourage group co-operation, but does not offer reasons for thinking that they are fictional (Wilson 2002, chp. 3). Elsewhere, in a critique of Christian Smith’s book Moral, Believing Animals (Smith 2003), Wilson does in effect try to justify this stance by attempting what he describes as a “demolition with respect to theism” (Wilson 2009, 338). However, to my mind he does not do it at all 3

Bonhoeffer’s use of the phrase is found in the prison letters, in his reflections on the abandonment of ‘God as a working hypothesis’, and the well-known dark phrase, ‘The God who lets us live in the world without the working hypothesis of God is the God before whom we stand continually. Before God and with God we live without God’ (Bonhoeffer, 1971, 360). We should however take care not to mis-read Bonhoeffer’s point the way his first generation of English-speaking readers did. For an illuminating reflection on the context and significance of the etsi deus in Bonhoeffer’s prison writings, see Lenehan (2013).

What Should Christian Theology (Not) Learn from Science?

89

convincingly. For example, he gives a list of supposedly empirical claims said to be generated by theism and allegedly refuted by science; but his list is oddly chosen, and each of his examples is either not straightforwardly an empirical claim, or not obviously refuted by science, or not entailed by theism, or has been contested within Jewish and Christian traditions since biblical times. For other authors, ontological naturalism seems to be a conclusion that they think their studies suggest or support. Few are naïve enough to think that scientific explanations of religious beliefs refute those beliefs; but more seem to think they render them epistemically doubtful, and thereby lend support to a naturalistic view – as in the remarks of Paul Bloom cited earlier. What, though, does this epistemic challenge to belief really amount to? At most, as Gregory Peterson (2010) puts it, it is a “hermeneutics of suspicion”. In other words, if there are neuropsychological factors that motivate us to hold certain beliefs, this might lead us to suspect that the reasons we give for those beliefs could really be rationalizations. It could be that our reasons for believing seem persuasive to us, not because they really are good reasons, but because we have psychological mechanisms that motivate us to find them persuasive. This “hermeneutics of suspicion” hardly justifies ontological naturalism, but it might turn out to have some theological uses of its own, as I shall suggest shortly.

Type 2: Both science and theology contribute, but science is the dominant partner With the two extreme positions in the typology ruled out, the three intermediate types remain to be considered. The kind of neurotheology proposed by d’Aquili and Newberg in the 1990s is a clear example of type 2, in that they aspired to a “metatheology” that would regulate the practice of specific theologies and a “megatheology” that would incorporate, unify and extend their insights (d’Aquili and Newberg 1999, 195-201).4 More broadly, any apologetic strategy that relies heavily on scientific evidence to make the case for the existence of God or spiritual reality can be considered an instance of type 2, because to a large extent the theological claims will stand or fall by the scientific data and their interpretation, and scientific evidence is likely to have a major influence 4

In his more recent Principles of Neurotheology, Newberg has rowed back somewhat from these bold claims, but still maintains that neurotheology should be both a meta- and a megatheology (Newberg, 2010, 64-66).

90

Chapter Five

on what theological claims are made. Among the science and theology pioneers in the second half of the twentieth century, for example, this was true of Arthur Peacocke: he was convinced that to recover any intellectual credibility, theology must reform itself to conform to the natural sciences both in its methods and in its substantive content (e.g. Peacocke 2000). Here, however, Peterson’s hermeneutic of suspicion comes into play, because scientific studies of religion not only offer naturalistic explanations of religious phenomena; they also suggest naturalistic reasons why we will be predisposed to find non-naturalistic explanations of those phenomena more plausible. In other words, a theology that relies heavily on science to support its claims and to guide its methods and content will risk being exposed as epistemically unreliable in the way described by Paul Bloom (2009, 125). It will then be in danger of collapsing into type 1, in which methodologically naturalistic scientific accounts of religion generate the ontologically naturalist claim that scientific accounts are “all the explanation there is”. On closer inspection, then, type 2 does not seem such a promising strategy for theological engagement with evolutionary, cognitive or neuroscientific research. Moreover, by enlisting science to do theological work, this type of engagement may also hinder it from doing its own proper work. If science is indeed best done in methodologically naturalistic ways, then using it to support non-naturalistic claims or investigate non-naturalistic questions risks distorting either the scientific investigation or the answers it supplies. In that sense Armin Geertz (2009) may be right to insist that religious agendas should be kept out of scientific research – with the proviso that this also applies to atheistic agendas: these too are, in a negative sense, religious agendas that science qua science is not competent to support. Ironically, a type of theological engagement like type 4, which sets narrower limits on the part natural science should play in shaping a Christian understanding, may thereby show more respect for the integrity of science as an intellectual practice.

Type 4: Both contribute, but theology is the dominant partner This brings us to type 4, and it is worth recalling again Peterson’s scientific hermeneutic of suspicion. Like others, he makes the connection with earlier “masters of suspicion” (as Paul Ricoeur called them), remarking that the cognitive science of religion is often “seen as an inheritor of the Feuerbachian project to reduce religion to something else, whether it be psychology, class struggle, or will-to-power” (Peterson 2010, 547). Christian theology, of course, has been here before. Although the

What Should Christian Theology (Not) Learn from Science?

91

Feuerbachian critique has often been thought to threaten Christian faith, it is well known that for Karl Barth (for one), Feuerbach turned out to be an ally (Barth 2002, chp. 18). By exposing the human tendency to project our own experience onto a transcendent screen and call it “God”, a Feuerbachian type of critique forces us to acknowledge that “one can not speak of God simply by speaking of man in a loud voice” (Barth 1957, 196). Barth made this famous remark about Friedrich Schleiermacher in an early essay, and there is a growing consensus that it is not only unfair to Schleiermacher, but unrepresentative of Barth’s mature evaluation of him (McCormack 2002). However, it seems all too apt an assessment of the kind of neurotheology which claims to learn from the scientific analysis of different mystical experiences that God has “anterior” and “posterior” natures (d’Aquili and Newberg 1999, 201). Against such transcendental projections of human experience, Christian theology is recalled to the only ground on which it is entitled to stand, understanding itself as a faithful and rigorous response to God’s self-disclosure in Christ. If that is the central task of Christian theology, how should it interact with the natural sciences in articulating a Christian understanding of what it is to be human before God and in the company of others? Barth himself looks an unpromising example, since he sometimes seems to deny that the natural sciences have any contribution to make to Christian doctrine. In the preface to the first part of his doctrine of creation, for example, he anticipates Gould’s NOMA: There is free scope for natural science beyond what theology describes as the work of the Creator. And theology can and must move freely where science which really is science … has its appointed limit. I am of the opinion, however, that future workers in the field of the Christian doctrine of creation will find many problems worth pondering in defining the point and manner of this twofold boundary (Barth 1956-75, vol. 3.1, x).

This appears to restrict the interaction of theology with natural science to the negotiation and policing of borders. However, elsewhere in his work Barth shows himself quite ready to engage creatively with non-theological disciplines, and to appropriate their insights critically for his theological work (Biggar 1993, chp. 5), so there seems no reason in principle why a theological investigation done somewhat after the manner of Barth cannot engage critically and constructively with evolutionary, cognitive and neuroscientific perspectives on human nature (see further Chapman 2014).

92

Chapter Five

Type 3: Both science and theology contribute, neither side dominating Type 3 remains to be considered. While this type of theological engagement resists any dominance of science in shaping a Christian understanding of humanity, and thus allows space for theology to operate according to its proper sources and methods, I have in the past questioned how stable it is. I have speculated that this type of engagement is always liable to drift towards type 2 or 4, so that either science or theology comes to dominate the encounter, and the real methodological choice in science and theology is between those two types. Moreover, in this field a drift to type 2 is perhaps the more frequent temptation. This is not to say, however, that it always happens. For example, Markus Mühling (2014) offers a highly sophisticated and genuinely mutually critical encounter between theology, phenomenology, neuroscience and evolutionary theory, which does seem to avoid the temptation. What conclusions does all this suggest about how theology should engage with science in the neuroethical arena? Although my typology is not simply equivalent to that of Hans Frei, some recent reflections by Paul DeHart on Frei’s typology may by analogy prove instructive for mine. On DeHart’s reading, Frei’s third and fourth types (ad hoc correlation and theological subordination of external discourses) represent the range of legitimate possibilities for theology, and they form what he calls a “mutually stabilizing pair” (DeHart 2006, 217), enabling the third to remain clearly distinct from the second type and keeping the fourth from drifting towards “the theological dead end represented by Type Five” (DeHart 2006, 216-17). Something comparable could be said about my types 3 and 4.5 I have suggested that when science plays a dominant role in shaping a Christian understanding of human being (type 2), this may have a distorting effect on the science as well as the Christian understanding. I have also suggested that type 3 is often at risk of drifting towards type 2. On the other hand, a hazard for type 4 is to drift towards type 5, separating theological and scientific discourses and denying any relevance of one to the other. Barth’s remarks about the “twofold boundary”, which I quoted a moment ago, seem a case in point. Perhaps, therefore, my types 3 and 4 also make a “mutually stabilizing pair”, in DeHart’s phrase, each drawing the other away from its characteristic pitfall. If so, Christian theological 5

This represents a shift of emphasis from my previous use of the typology in Messer (2007).

What Should Christian Theology (Not) Learn from Science?

93

dialogue with the natural sciences needs both types; but more of the work in science and theology resembles type 3 than type 4. One reason for attempting a type 4 engagement in this paper is to redress that balance: to test the capacity of a Christian theology firmly rooted in its own distinctive sources and methods to appropriate evolutionary, cognitive and neuroscientific insights critically, to incorporate them in a theologically-formed vision, and so to articulate a distinctive and compelling theological anthropology.

3. A worked example: the neuroscience of morality6 I have made a proposal about the limits of science in its capacity to inform theological enquiry, and have argued for the kind of science-theology dialogue in which theology sets the agenda and critically appropriates insights from science. In the final part of the chapter I shall try to show what that might look like in practice, with reference to a related field of research: the neuroscience of morality. From the early 2000s, Joshua Greene and his colleagues conducted a series of functional magnetic resonance imaging (fMRI) studies on subjects who were asked to respond to the moral thought-experiments known as “trolley problems” (Greene 2014, chp. 4). In one of these, a runaway railway truck is about to collide fatally with five workers on the track. You have no way of stopping the truck, but you can change the points to divert it from the main line to the siding. The five workers on the main line will then be safe, but unfortunately the truck will kill another worker who is at work on the siding. Should you change the points? In another of these scenarios, there is no siding. You are on a footbridge above the railway line with the runaway truck, and next to you is a large man (or in some versions, a man wearing a heavy backpack). If you push him off the bridge into the path of the truck, he will stop it hitting the five workers, but he will surely die. (You are not heavy enough to stop the truck yourself, so self-sacrifice is not an option.) Should you push the man to his death to save the other five? Most people answer Yes to the first of these, and No to the second: they make a moral distinction between directly killing an innocent person to save more lives, and taking action to save several lives knowing that one person will die as an unintended consequence. A minority answer Yes to both. Greene interprets the majority view as deontological and the minority view as utilitarian: the majority judge that directly killing an 6

An earlier and longer version of the argument in this section can be found in Messer (2016).

94

Chapter Five

innocent person, even for the greater good, is simply wrong; the minority weigh up the number of lives saved and lost, and they opt for the best outcome in both cases. (It is worth noting that some of his critics dispute this interpretation: e.g. Kahane 2012; 2015.) Greene has found that these different scenarios are associated with different patterns of brain activity detectable by fMRI. He maintains that these imaging data support a dual-process theory of moral cognition, similar to the one proposed earlier by the cognitive psychologist Jonathan Haidt (2001). Non-utilitarian moral judgements are intuitive, generated by unconscious automatic processing. Utilitarian judgements are the products of conscious reasoning. Now of course, as every student of ethics knows, there is plenty of non-utilitarian moral reasoning around; but Greene, like Haidt, thinks that much of this is the rationalization of intuitive judgements after the fact. Indeed, Greene makes the bold claim that deontological theories from Kant to the present day are simply elaborate post hoc rationalizations of moral intuitions (Greene 2008; 2014). There has been considerable criticism of the conceptualization, design and interpretation of Greene’s studies (e.g. Kahane 2012; 2015), so his findings should be handled with care. But let us accept for the sake of argument some broad conclusions that a reasonable number of commentators seem prepared to allow: that our moral judgements are more the products of intuition and automatic processing, and less of conscious reason, than we often imagine; that conscious moral reasoning often comes into play not to inform the making of our moral judgements, but to justify intuitive judgements post hoc; and perhaps that these conclusions apply to non-utilitarian judgements in particular. Do these conclusions have any normative ethical significance? In common with others like Peter Singer (2005), Greene (2014) thinks that in an indirect way, they do. Our systems of moral intuition, he argues, are the product of our evolutionary history: an adaptation to foster prosocial behaviour. The intuitive (non-utilitarian) judgements they generate are effective heuristics to help us co-exist in relatively homogeneous social groups. But in the modern world we face a different kind of challenge that our ancestors did not: many different social groups or communities (“moral tribes”, as Greene calls them) have to co-exist in complex modern societies. Different “moral tribes” have different sets of group norms, and in modern societies these rival norms will often clash. Since these norms are deeply held and based on intuitive rather than reasoned judgements, conflicts between them will prove very difficult to resolve – the more so because we so habitually use our moral reasoning to rationalize and defend our intuitions, not to examine them critically or dispassionately. To resolve

What Should Christian Theology (Not) Learn from Science?

95

intractable conflicts of this kind, we need what Greene calls a “metamorality”: a common moral currency for settling the conflicts that arise between our different tribal moralities. And the metamorality that he pulls out of the hat is utilitarianism. Leaving aside philosophical difficulties with Greene’s argument, what relevance if any does it have for Christian theological ethics? I want to suggest that to the extent that they are scientifically robust, Greene’s findings can be critically appropriated to play a useful – if subsidiary – role. The contribution they might make begins with something akin to Peterson’s hermeneutic of suspicion. I shall try to explain. One fragment of Dietrich Bonhoeffer’s unfinished Ethics begins like this: “The knowledge of good and evil appears to be the goal of all ethical reflection. The first task of Christian ethics is to [overcome] that knowledge” (Bonhoeffer 2005, 299-300).7 Bonhoeffer is here articulating a profound theological suspicion of “ethics”, understood as the human project to know about good and evil. This suspicion has its roots in his reading of the Genesis 3 “Fall” narrative, in which the knowledge of good and evil is at the heart of what it means to be “like God” (sicut deus) in the sense promised by the snake. As Karl Barth later put it, in this story, “what the serpent has in mind is the establishment of ethics” (Barth 1956-75, vol. 4.1, 448). In the story the humans succeed in gaining that knowledge and becoming “like God”; however, this is not a triumph, but a disaster. Their bid to become gods amounts to a refusal of their true creaturely being. Their godlike knowledge of good and evil divides them to their very core, alienating them from God, one another, themselves and the earth. If scientific studies of moral judgement give reason to think that our moral judgements are a source of deep divisions, and our moral reason is prone to be used in rationalizing and self-deceiving ways that replicate rather than healing those divisions, this should not come as a surprise to anyone who has learned Bonhoeffer’s theological suspicion of ethics-as-humanproject. If it is news to theologians, perhaps Greene’s work is serving as a useful reminder of things they should have known all along. However, theological suspicion of ethics is only the first move in its theological reconstruction. The knowledge of good and evil, says Bonhoeffer, must be “overcome” by our reconciliation in Christ. Then our divisive and alienating quest to be “like God, knowing good and evil” is 7

In the second sentence of the quotation, I have amended ‘supersede’ in the published translation to read ‘overcome’, which to my mind gives a better sense of the German ‘aufheben’ and is indeed routinely used by the translators in the remainder of the fragment (311ff.).

96

Chapter Five

replaced by what he calls our “[election] to do the one will of God in simplicity” (Bonhoeffer 2005, 320). In place of our anxious attempts to discriminate between right and wrong and appraise our own and others’ goodness, there is the call (in Paul’s words) to be transformed by the renewing of our minds, and the daily task of discerning the “good and acceptable and perfect” will of God (Romans 12:1, 2). But this daily task and calling emphatically do not bypass our rational faculties. Bonhoeffer insists that the whole array of human reason and experience should come into play in such discernment. “Under no circumstances [he says] must one count on … unmediated inspirations, lest all too easily one fall prey to self-deception” (Bonhoeffer 2005, 323). In this task, insights from the scientific study of moral judgement – to the extent that they are scientifically robust – may have a useful ongoing critical and perhaps also constructive part to play. They may for example be very helpful in drawing attention to the forms of distortion and selfdeception that afflict, not only our moral intuitions and “unmediated inspirations”, but also our use of moral reason. It is also worth saying – though space does not permit me to elaborate the point – that theologians engaging with this research will also feel free to turn the same critique back on Greene and his advocacy for utilitarianism. They may conclude that this looks less like a “metamorality”, enabling different “moral tribes” to coexist peaceably, than a land-grab by one of those “moral tribes”. Theologians can be grateful to Greene for the critical perspective of his scientific studies without being committed to his normative conclusions.

Conclusion I have offered a proposal and an example of how Christian theology should engage with scientific studies of human nature – among other things, a proposal about the scope and the limits of what science can contribute to theological understanding. My proposal can be summed up by Karl Barth’s remark that accounts of the “phenomena of the human” are “like an interesting commentary on a text which must first be known and read for itself if the commentary is to be intelligible and useful” (Barth 1956-75, vol. 3.2, 122). As any preacher knows (or should know), commentaries are fallible tools to be used with all due caution; but used in this spirit, the commentary offered by the cognitive neurosciences may be of real value to theological readers of the “text” of human creaturely being.

What Should Christian Theology (Not) Learn from Science?

97

References Barrett, J.L. 2007. “Cognitive Science of Religion: What Is It and Why Is It?”, Religion Compass, 1.6, 768-86. Barth, K. 1956-75. Church Dogmatics. Bromiley, G.W. and Torrance, T.F. (eds), 13 vols., Edinburgh: T & T Clark. —. 1957. The Word of God and the Word of Man. Smith, D.H. (trans.), New York, NY: Harper. —. 2002. Protestant Theology in the Nineteenth Century. Cozens, B. and Bowden, J. (trans), Grand Rapids, MI: Eerdmans. Beauregard, M. and O’Leary, D. 2007. The Spiritual Brain: A Neuroscientist’s Case for the Existence of the Soul. New York, NY: HarperCollins. Beauregard, M. and Paquette, V. 2006. “Neural Correlates of a Mystical Experience in Carmelite Nuns”, Neuroscience Letters 405, 186-90. —. 2008. “EEG Activity in Carmelite Nuns during a Mystical Experience”, Neuroscience Letters, 444, 1-4. Biggar, N. 1993. The Hastening that Waits: Karl Barth’s Ethics. Oxford: Clarendon Press. Bloom, P. 2009. “Religious Belief as an Evolutionary Accident”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 118-27. Bonhoeffer, D. 1971. Letters and Papers from Prison. Bethge, E. (ed.), Fuller, R. et al. (trans.), London: SCM Press. —. 2005. Ethics (Dietrich Bonhoeffer Works, vol. 6). Tödt, I. et al. (eds.), Krauss, R. West, C.C. and Stott, D.W. (trans), Minneapolis, MN: Fortress. Boyer, P. 2001. Religion Explained: The Evolutionary Origins of Religious Thought. New York: Basic Books. Chapman, P. 2014. “Barth and Darwin: What Is Humanity?”, Theology and Science 12.4, 362-77. D’Aquili, E. and Newberg, A.B. 1999. The Mystical Mind: Probing the Biology of Religious Experience. Minneapolis, MN: Fortress. Deane-Drummond, C. 2009. Christ and Evolution: Wonder and Wisdom. Minneapolis, MN: Fortress. DeHart, P.J. 2006. The Trial of the Witnesses: The Rise and Decline of Postliberal Theology. Oxford: Blackwell. Dennett, D. 2006. Breaking the Spell: Religion as a Natural Phenomenon. New York, NY: Viking. Frei, H. 1992. Types of Christian Theology., Hunsinger, G. and Placher, W.H. (eds.), New Haven, CT: Yale University Press.

98

Chapter Five

Geertz, Armin W. (2008), “How Not to Do the Cognitive Science of Religion Today”, Method and Theory in the Study of Religion 20, pp. 7-21. Geertz, A.W. 2009. “When Cognitive Scientists Become Religious, Science Is in Trouble: On Neurotheology from a Philosophy of Science Perspective”, Religion 39.4, 319-24. Gould, S.J. 2002. Rocks of Ages: Science and Religion in the Fullness of Life. London: Vintage. Gould, S.J. and Lewontin, R.C. 1979. “The Spandrels of San Marco and the Panglossian Paradigm: A Critique of the Adaptationist Programme”, Proceedings of the Royal Society B 205, 581-98. Granqvist, P. et al. 2005. “Sensed presence and mystical experiences are predicted by suggestibility, not by the application of transcranial weak complex magnetic fields”, Neuroscience Letters, 379.1, 1-6. Greene, J.D. 2008. “The Secret Joke of Kant’s Soul”, in SinnottArmstrong, W. (ed.), Moral Psychology, Vol. 3: The Neuroscience of Morality: Emotion, Brain Disorders, and Development, Cambridge, MA: MIT Press, 36-79. —. 2014. Moral Tribes: Emotion, Reason, and the Gap between Us and Them. London: Atlantic. Haidt, J. 2001. “The Emotional Dog and Its Rational Tail: A Social Intuitionist Approach to Moral Judgment”, Psychological Review 108.4, 814-34. Johnson, D. and Bering, J. 2009. “Hand of God, Mind of Man: Punishment and Cognition in the Evolution of Co-operation”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 26-43. Kahane, G. 2012. “On the Wrong Track: Process and Content in Moral Psychology”, Mind and Language 27.5, 519-45. —. 2015. “Sidetracked by trolleys: Why sacrificial moral dilemmas tell us little (or nothing) about utilitarian judgment”, Social Neuroscience 10.5, 551-60. Larzson, M. et al. 2005. “Reply to M.A. Persinger and S. A. Koren's response to Granqvist et al. ‘Sensed presence and mystical experiences are predicted by suggestibility, not by the application of transcranial weak magnetic fields’”, Neuroscience Letters 380.3, 348-50. Lenehan, K.A. 2013. “Etsi Deus Non Daretur: Bonhoeffer’s Useful Misuse of Grotius” Maxim and its Implications for Evangelisation in the World Come of Age’, The Bonhoeffer Legacy: Australasian Journal of Bonhoeffer Studies 1.1, 34-60.

What Should Christian Theology (Not) Learn from Science?

99

McCormack, B.L. 2002. “What Has Basel to Do with Berlin? Continuities in the Theologies of Barth and Schleiermacher”, Princeton Seminary Bulletin 23.2, 146-73. Messer, N. 2007. Selfish Genes and Christian Ethics: Theological and Ethical Reflections on Evolutionary Biology, London: SCM Press. —. 2016. “Cognitive Science, Moral Reasoning, and the Theological Suspicion of Ethics”, Journal of the Society of Christian Ethics, 36.1, 51-68. Midgley, M. 2000. “Why Memes?”, in Rose, H. and Rose, S. (eds.), Alas Poor Darwin: Arguments against Evolutionary Psychology, London: Jonathan Cape, 67-84. Mühling, M. 2014. Resonances: Neurobiology, Evolution and Theology. Evolutionary Niche Construction, the Ecological Brain and RelationalNarrative Theology. Göttingen: Vandenhoeck and Ruprecht. Murray, M.J. 2009. “Scientific Explanations of Religion and the Justification of Religious Belief”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 168-78. Murray, M.J. and Goldberg, A. 2009. “Evolutionary Accounts of Religion: Explaining and Explaining Away”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 179-99. Newberg, A.B. 2010. Principles of Neurotheology. Farnham: Ashgate. Newberg, A.B., et al. 2001. “The measurement of regional cerebral blood flow during the complex cognitive task of meditation: a preliminary SPECT study”, Psychiatry Research: Neuroimaging, 106, 113-22. —. 2003. “Cerebral blood flow during meditative prayer: Preliminary findings and methodological issues”, Perceptual and Motor Skills, 97.2, 625-30. —. 2006. “The measurement of regional cerebral blood flow during glossolalia: A preliminary SPECT study”, Psychiatry Research: Neuroimaging, 148.1, 67-71. —. 2010. “Cerebral blood flow differences between long-term meditators and non-meditators”, Consciousness and Cognition, 19.4, 899-905. Peacocke, A. 2000. “Science and the Future of Theology: Critical Issues”, Zygon, 35.1, 119-40. Persinger, M. A., and Koren, S.A. 2005. “A response to Granqvist et al. ‘Sensed presence and mystical experiences are predicted by

100

Chapter Five

suggestibility, not by the application of transcranial weak magnetic fields’”, Neuroscience Letters, 380.3, 346-47. Peterson, G.R. 2010. “Are Evolutionary/Cognitive Theories of Religion Relevant for Philosophy of Religion?”, Zygon 45.3, 545-57. Richerson, P.J., and Newson, L. 2009. “Is Religion Adaptive? Yes. No. Neutral. But Mostly, We Don’t Know”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, pp. 100-17. St-Pierre, L. S., and Persinger, M.A. 2006. “Experimental Facilitation of the Sensed Presence Is Predicted by the Specific Patterns of the Applied Magnetic Fields, Not by Suggestibility: Re-Analyses of 19 Experiments”, International Journal of Neuroscience, 116, 1079-96. Schjoedt, U. 2009. “The Religious Brain: A General Introduction to the Experimental Neuroscience of Religion”, Method and Theory in the Study of Religion 21, 310-39. Schloss, J. 2009. “Introduction: Evolutionary Theories of Religion. Science Unfettered or Naturalism Run Wild?”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 1-25. Singer, P. 2005. “Ethics and Intuitions”, Journal of Ethics, 9, 331–52. Smith, C. 2003. Moral, Believing Animals: Human Personhood and Culture. Oxford: Oxford University Press. van Inwagen, P. 2009. “Explaining Belief in the Supernatural: Some Thoughts on Paul Bloom’s ‘Religious Belief as an Evolutionary Accident’”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 128-38. Wang, D.J. J., et al. 2011. “Cerebral blood flow changes associated with different meditation practices and perceived depth of meditation”, Psychiatry Research: Neuroimaging, 191, 60-67. Wilson, D.S. 2002. Darwin’s Cathedral: Evolution, Religion, and the Nature of Society. Chicago: University of Chicago Press. —. 2009. “Evolutionary Social Constructivism: Narrowing (but Not Yet Bridging) the Gap”, in Schloss, J. and Murray, M.J. (eds.), The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origins of Religion. Oxford: Oxford University Press, 319-38.

PART TWO

CHAPTER SIX PHYSICS: THE ENDLESS FRONTIER NATHAN AVIEZER

What are the limits of science? This question has been widely discussed, and by 1900, a definitive answer emerged, based on the enormous progress that had taken place in physics. The answer given was that the power of science is limitless; science is capable of explaining everything regarding the development and functioning of the physical world. According to the laws of nature, the future behaviour of all physical objects is already fixed in the present. Stated simply, the present determines the future. In other words, if one knew all the details of the universe at the present time, and one were clever enough to solve the known equations, one could predict in detail the entire future of the universe. This picture of the physical world was enshrined in the “clockwork universe.” Given the present positions of the hands of a clock, the cogs, wheels and gears of the clock mechanism determine where the hands of the clock will be at any instant in the future. In exactly the same way, the present situation and the laws of nature completely determine the detailed situation of the entire universe at every instant in the future. As we shall see, this picture of physical reality has been completely overturned by the revolutionary discoveries in physics of the twentieth century. In previous years, a distinguished physicist would sometimes declare that we now know all basic laws of physics. A famous example is the statement by Lord Kelvin (William Thomson), one of Britain’s greatest scientists, in his 1900 address to the British Association for the Advancement of Science. Lord Kelvin declared the following: “There is nothing new to be discovered in physics now. All that remains is to make more and more precise measurements”.

Physics: The Endless Frontier

103

This extreme statement may seem quite naïve today. However, if one examines the situation in physics in 1900, it seems that Lord Kelvin actually had very good reasons for his bold statement.

Fundamental Discoveries in Physics Before 1900 The universe consists of particles which range in size from the very small (single atoms) to the very large (planets). One of the main tasks of physics is to explain how these particles move. Explaining planetary motion was one of the basic problems in physics/ astronomy throughout the Middle Ages. Nicolas Copernicus and Johannes Kepler finally succeeded in describing how the planets move. Planetary motion is enshrined in Kepler’s laws, formulated in 1609, which state that each planet rotates around the Sun in an elliptical orbit, and the planet sweeps out equal areas in equal times. However, no one could explain why planets move in this fashion. This problem was solved by Isaac Newton in 1687 in his famous book, Principia, probably the most important book of science ever written. Newton formulated the fundamental law of motion, F = Ma. This law, known as Newtonian mechanics, states that if force F is exerted on a particle of mass M, the particle will move with acceleration a. This law of motion requires one to know the force that is exerted on the particle. Newton solved this problem for the planets by formulating the law of gravity, which is the force that the Sun exerts on each planet. Combining the law of motion with the law of gravity fully explained all aspects of planetary motion. Moreover, these same laws explained many other phenomena as well, such as the tides. Clearly, great progress had been made. However, important problems remained, such as explaining the electrical and magnetic forces, and the phenomena of light, heat, sound, and thermodynamics. Many of these problems were solved in the 1860s by James Clerk Maxwell, whose equations showed that the electrical force and the magnetic force were really two aspects of the same force, now called the electromagnetic force. His equations also showed that if a charged particle is accelerated, it will generate energy in the form of electromagnetic waves. Moreover, he discovered that light is an example of such electromagnetic waves. In summary, Maxwell’s research showed that the electrical force, the magnetic force, and light were all different aspects of the same phenomenon. This was an enormous simplification. It was also discovered during the nineteenth century that both sound and heat were the result of molecular motion. If the motion is coherent, it

104

Chapter Six

produces sound; if the motion is incoherent, it produces heat. In addition, the laws of thermodynamics were established, including the law of the conservation of energy. In view of all this progress by the end of the nineteenth century, it indeed seemed that Lord Kelvin was right; there was nothing fundamental left to discover in physics.

Fundamental Discoveries in Physics Since 1900 In fact, Lord Kelvin was completely wrong. Many fundamental discoveries occurred in physics during the twentieth century. We shall describe the principal ones.

Relativity In 1905, Albert Einstein developed the theory of relativity (called the “special” theory), which showed that the motion of a particle is described by relativistic mechanics, rather than by Newtonian mechanics. Among the startling predictions of this theory are the following. (i) No particle can move faster than the speed of light. (ii) The rate at which time passes is not the same everywhere in the universe, as Newton has assumed. Rather, the rate at which time passes depends on the motion of the clock and of the observer. (iii) The velocity of light will always be observed to be the same, regardless of the motion of the source of the light or of the observer of the light. (iv) Energy can be converted into matter and matter can be converted into energy, according to Einstein’s famous equation E = Mc2. The reason that these dramatic new predictions were not discovered earlier is that these predictions differ significantly from those of Newtonian mechanics only when the particle is moving very fast. In this context, a speed of 1000 kilometres per second is considered slow! When a particle moves slowly (at only 1000 kilometres per second!), the laws of Newtonian mechanics are perfectly adequate to describe the motion. Only when the speed of the particle approaches the speed of light (300,000 kilometres per second) must one use relativistic mechanics to describe the motion.

Gravity It follows from Einstein’s research that every theory of physics must be consistent with the theory of relativity. Maxwell’s theory of electromagnetism is consistent with the theory of relativity, but Newton’s

Physics: The Endless Frontier

105

theory of gravity is not. Therefore, there was need for a new theory of gravity. Einstein himself set out to develop such a theory. It took him ten long years, from 1905 until 1916, to develop a new theory of gravity (which he called “general relativity”) that is consistent with the theory of relativity. Einstein’s new theory of gravity was conceptually very different from Newton’s theory, but in practice, its predictions were almost indistinguishable from those of Newton’s theory. For this reason, for over 200 years, no one had realized that something was wrong with Newton’s theory of gravity. In a few special cases, Einstein’s theory of gravity did make different predictions from those of Newton’s theory. It was found that in each case, Einstein’s theory agreed with the observed results. The differing predictions include the following. (i) According to Einstein’s theory, a ray of starlight should bend slightly when the light ray passes near a large body, such as the Sun. This prediction was tested and it was found that the ray of starlight was bent exactly as much as Einstein had predicted. (ii) According to Einstein, there should be a slight change in the orbit of a planet when the orbit is near the Sun. The orbit of Mercury, which lies nearest the Sun, was observed to be very slightly different from the ellipse predicted by Newton’s theory but in exact agreement with the prediction of Einstein’s theory.

Radioactivity Early in the twentieth century, an entirely new phenomenon was discovered, called radioactivity. It was discovered that certain atoms are not stable, and the unstable atoms radiate particles, hence the name “radioactivity.” The radiated particles consisted of three types, named alpha, beta, and gamma particles, since no one knew what these radiated particles were. It was later discovered that alpha particles are helium nuclei, beta particles are electrons, and gamma particles are highly energetic photons. In order to explain the phenomenon of radioactivity, physicists postulated the existence of a new particle, called the neutrino (subsequently discovered), and a new force, called the weak nuclear force (whose existence was subsequently confirmed).

106

Chapter Six

Quantum Theory As experiments and calculations became more accurate, a number of problems arose. These included the following. (i) When a piece of metal is heated, it radiates energy, with the amount of energy radiated depending on its temperature. A calculation produced the absurd result that at all temperatures, an infinite amount of energy should be radiated by every piece of metal. Something was clearly wrong. This problem was solved by Max Planck in 1900 by assuming that light (the radiated energy) is not a wave phenomenon, as previously assumed. Rather, light consists of a stream of particles, which he named photons or quanta (singular: quantum). Planck’s assumption produced perfect agreement with the radiation data. However, Planck’s assumption that light is a stream of particles contradicted a wide variety of experimental evidence that seemed to have established that light is a wave phenomenon. So, what is light? Is light an electromagnetic wave or a stream of particles? Each of these contradictory opinions was supported by much experimental data! (ii) The experiment of Rutherford clearly showed that an atom consists of a central nucleus of positively charged protons, around which rotate a number of negatively charged electrons. This model is very similar to the solar system, which consists of a number of planets that rotate around the central Sun. The electrons correspond to the planets, and the protons in the nucleus correspond to the Sun. Just as the force of gravity keeps the planets constantly rotating around the Sun, so the electrical force should keep the electrons constantly rotating around the protons in the nucleus. But there is a very serious problem that destroys the analogy. Unlike planets, electrons are charged particles. According to Maxwell’s equations, a rotating charged particle radiates energy. Therefore, the rotating electrons should radiate energy and soon fall into the nucleus, causing the atom to collapse. Since it is obvious that atoms are perfectly stable and they do not collapse, it is clear that something was very wrong with one’s understanding of Maxwell’s equations. The resolution of these and many other problems lay in a totally new theory, called quantum theory. Quantum theory is the strangest theory ever to be introduced into science. It took 30 years, from 1900 to 1930, for quantum theory to be firmly established.

Physics: The Endless Frontier

107

The most astonishing aspect of quantum theory is that it is a probabilistic theory of nature. This means that for any physical system, the most that can ever be known are the probabilities that certain events will occur in the future. Through the Schroedinger equation, one can calculate the probability for the occurrence of each of possible event. However, which of the various possible events will actually occur in practice, can never be known before the measurement is carried out. The probabilistic nature of quantum mechanics leads to an important conclusion: the present does not determine the future. This nondetermination of the future is called quantum indeterminacy and is enshrined in Heisenberg’s famous uncertainty principle. It is important to emphasize that quantum indeterminacy is not due to lack of knowledge. That is, it is not correct to state that the future is already determined in the present but no one is able to predict it because of technical difficulties, as is the case for Newtonian mechanics. Quantum theory states that the future cannot be known even in principle, because it has not yet been determined. This can be illustrated by the following example: If one performs the same experiment twice, with the two experiments being absolutely identical in every respect, one may nevertheless obtain different results in the two cases. In other words, the same present (the same experiment being simultaneously performed twice) has led to two different futures (different results in the two cases). This scenario would be quite impossible according to classical science. In fact, this phenomenon violates the very essence of pre-quantum physics. The reader may be wondering how such a dramatic phenomenon (the present does not determine the future) was not noticed earlier by Newton and other great scientists. More to the point, our everyday experience tells us just the opposite. Throughout our lives, we observe that the present does indeed determine the future. Every football player knows that if he kicks the ball in the right direction (the present), in a few seconds (the future) the ball will enter the goal to the roar of the crowd. Why do athletes, as well as all the rest of us, remain unaware of quantum theory in our daily lives? The answer is that the effects of quantum theory are significant only for very minute particles. For macroscopic objects, such as footballs, the difference between the quantum prediction and the classical prediction is completely insignificant. (A tiny speck of dust weighing less than a trillionth of a gram is considered large in this context.) When the football is kicked in the right direction, classical science predicts a goal with 100% certainty, whereas the quantum prediction is that the chances of the ball

108

Chapter Six

entering the goal are 99.99999999....%, with only an extremely small chance of the ball missing the goal. Since the difference between these two predictions is unmeasurably small, an athlete need not be aware of quantum theory to become a football star. However, when dealing with microscopic particles, such as electrons and protons, the predictions of quantum theory are very different from those of relativistic mechanics. Quantum theory explains, for example, why atoms are stable. Quantum theory is of utmost importance for understanding the universe. In fact, the existence of a stable universe would be impossible if the classical laws were correct.

String Theory Every theory of physics must be consistent with quantum theory. However, Einstein’s theory of gravity is not consistent with the quantum theory. Moreover, it was proved that it is impossible to modify Einstein’s theory to make it consistent with quantum theory. This is a very serious problem indeed. Quantum theory is certainly correct, and Einstein’s theory of gravity is also certainly correct. How can it be that two correct theories of nature cannot be made compatible with each other? Why is it impossible to formulate a consistent theory of “quantum gravity”? The contradiction between quantum theory and Einstein’s theory of gravity is resolved by string theory. String theory is a totally new conceptual framework for describing the physical universe. It was previously thought that the basic entities of the universe are particles – electrons, protons, etc. String theory asserts that the basic entities of the universe are tiny “strings”. These strings vibrate (like a violin string) and the energy of vibration appears to us as a “particle” through the Einstein relation between energy and mass (E = Mc2). String theory does not yield a theory of quantum gravity. However, if the universe were to contain ten spatial dimensions, then within the framework of string theory, Einstein’s theory of gravity can be made compatible with quantum theory. Since these two theories must be consistent with each other, it follows that the universe must contain ten dimensions. How can one reconcile the ten-dimensional universe of string theory with our everyday experience of a three-dimensional universe? What is the meaning of the “missing” seven dimensions? Why do we not experience them?

Physics: The Endless Frontier

109

String theory answers that the seven “missing” dimensions have become “compacted” to such a small size that it is impossible to experience them. Each of the three usual dimensions of space extends to infinity, whereas the other seven dimensions extend for only a very minute distance because they have become “compacted”. According to string theory, the extent of a compacted dimension is a billionth of a billionth of a billionth of the radius of an atom. Such a small size can never be detected. Although compacted dimensions are far too small to be measured directly, they have very significant indirect effects on the universe. The most important indirect effect of the compacted dimensions is that they lead to a consistent theory of quantum gravity.

Chaos It was discovered that many systems are unbelievably sensitive to the slightest change in conditions. These systems, called chaotic systems, include the atmosphere that produces the weather. This phenomenon is known as the “butterfly effect”, because if a butterfly flaps its wings in Tokyo, this tiny change in the atmosphere can have an appreciable effect on the weather in London within about two weeks! Since no calculation can include the effect of all the butterflies in the world, one will never be able to calculate the weather more than two weeks in advance. The atmosphere is just one system out of many that display chaos. Our ability to predict the future behaviour of systems that display chaos is thus extremely limited. It should be made clear that chaos is not a quantum phenomenon. These are two separate and independent phenomena. Our inability to predict the future behaviour of systems that display chaos is a technical limitation – one is simply unable to include the effects of all the butterflies in the world into the calculation of the weather. But the future behaviour of a system displaying chaos is completely determined in the present. In practice, of course, both chaos and quantum theory are operative, with the former effect limiting our ability to predict the future on technical grounds, and the latter effect limiting our ability to predict the future in principle.

The Expanding Universe It was discovered in the twentieth century through astronomical observations that the universe is expanding. Previously, it was universally believed that

110

Chapter Six

the universe is static. Even the great Einstein thought so. Therefore, when his theory of gravity predicted that the universe is expanding, Einstein thought that there must be something missing from his equations. He found that by arbitrarily adding a term (called the “cosmological term”) to his equations, they had a solution that implied a static universe. When Edwin Hubble showed that the universe is not static, but expanding, Einstein removed this term from his equations. He later referred to his addition of the cosmological term as “the biggest blunder of my professional career”.

The Big Bang Theory A theory for the origin of the universe has been developed, known as the big bang theory. This theory makes a number of dramatic predictions, all of which have been confirmed. The theory states that the universe began with the sudden appearance of an enormous ball of light, known as the “big bang”, and hence the name of the theory. The remnant of this original light was discovered in 1965, a discovery that merited the Nobel Prize. The big bang theory predicts that universe had a beginning. Recent measurements have enabled scientists to date the origin of the universe at about 14 billion years ago. The big bang theory also predicts that the universe is expanding, thus explaining the result that had been observed previously.

Elementary Particles and Forces A comprehensive theory has been developed for the elementary particles that comprise the universe and the forces between them. This theory explains all experimental data, and is known as the “standard model of elementary particles”. The standard model predicts the existence of 12 fundamental particles, four of which are stable and comprise all known matter. The other eight fundamental particles decay rapidly after their formation. All these 12 particles have been discovered. In addition, the standard theory predicts the existence of one additional particle, called the Higgs particle, after the physicist who predicted its existence. The Higgs particle was produced in 2013 by the Geneva accelerator, thus leading to the final confirmation of the standard model of elementary particles. According to the standard model, there are four forces in nature. Two of these forces were known previously (gravity and the electromagnetic force) and two additional forces were discovered in the twentieth century.

Physics: The Endless Frontier

111

The latter two are the weak nuclear force and the strong nuclear force. The nuclear forces are short-ranged forces, meaning that they are operative only within the nucleus. At distances greater than the radius of an atomic nucleus, the nuclear forces are completely negligible. By contrast, the effects of gravity and the electromagnetic force extend to infinity.

Unsolved Problems in Physics One should not think that all fundamental problems in physics have now been solved and Lord Kelvin’s prediction has finally come true. This is far from the case. I will conclude with a brief survey of some of the outstanding new problems that have arisen in physics.

Dark Matter It has recently been discovered that there is another type of matter in the universe, whose existence was hitherto completely unsuspected. This matter is known as “dark matter”. The discovery came about as follows. Very accurate measurements have recently been performed of the rotation of the galaxies. Galaxies rotate because of the mutual gravitational attraction of the stars within the galaxy. However, it was found that the rate of galactic rotation is much too fast to be explained by the stars. Therefore, there must be additional matter in the galaxies that causes the speedy rotation. This additional matter is called “dark matter”. It was shown that this dark matter must be different from any known particle. There is currently no known explanation for the particles that comprise dark matter. The most surprising feature of dark matter is its large amount. The rotation of the galaxies is so rapid that dark matter comprises nearly 80% of the matter in the universe! Thus, for hundreds of years, scientists have been studying only 20% of the universe, with the vast majority of the universe being completely unknown.

Dark Energy Because of the big bang, all the galaxies are flying apart from each other at a constant rate. Since the galaxies are very far apart, their mutual gravitational attraction is slight and their rate of separation is virtually constant. However, the slight gravitational attraction of the galaxies for each other should cause their rate of separation to slightly slow down. Therefore, it came as a great surprise when careful measurements in 1998

112

Chapter Six

showed that the rate of galaxy separation was increasing, rather than decreasing. It is as if there is some kind of negative gravity that pushes the galaxies apart. This phenomenon is known as “dark energy”. There is currently no explanation for the “dark energy” that accelerates the rate of the separation of the galaxies.

Difficulties with the Standard Model of Elementary Particles The main problem with the standard model of elementary particles is the large number of parameters involved. Each of the 12 fundamental particles and the Higgs particle has a mass which is known only through experiment. Also, the strength of each force is known only through experiment. Thus, the standard model contains about 20 parameters and there is no indication of why each of these parameters has the value that is observed. A fundamental theory of nature cannot have so many different parameters. There must be some underlying theory that explains the values of these many parameters. Although scientists are trying hard to understand what the underlying theory might be, there has not been any significant progress.

Additional Unsolved Problems One should not think that these are the only unsolved fundamental problems in physics. There is also the problem of explaining the origin of cosmic rays, the extreme energy of gamma ray bursts, and many other problems. But we will not describe them all here.

Conclusion A century after Lord Kelvin’s optimistic prognosis, physicists are still very busy trying to determine the secrets of nature. It appears that this task will occupy physicists for at least another century and probably much longer. As each fundamental problem is solved, another problem is discovered. This is what makes physics so exciting. One of the most interesting and important aspects of these recent discoveries is the new assessment of the limits of science. It has become clear that there are definite limits to the explanatory power of science. What these limits might be continues to be a subject of heated debate among leading scientists. But it is universally admitted that the “clockwork universe” of Newton, Galileo and Laplace is not consistent with the modern scientific principles of quantum theory and chaos theory. We still have much to learn about our wonderful universe.

CHAPTER SEVEN THE CASE OF THALES’ OX: AN HISTORICAL LIMIT TO SCIENCE (AND ITS RELATION TO RELIGION) MICHAEL FULLER

One limit on science is its historical limit – the extent to which we can trace its historical origins. These are sometimes held to go back to the very foundations of classical thought: to the earliest Greek philosophers of the Ionian school, Thales of Miletus and his younger contemporaries Anaximander and Anaximenes. These thinkers were concerned about the basic principles of things, for which they used the term “archƝ”. This term was apparently first used by Anaximander, who used the term “apeiron” (the boundless, or infinite) for this first principle. It is their location of this archƝ in the natural world rather than in the world of the gods (as is sometimes anachronistically said, the world of reason rather than the world of myth) that has led to these thinkers being widely hailed as protoscientists. This short paper focusses on the senior member of this trio, Thales. It must be stated at the outset that there are serious difficulties in ascertaining exactly what Thales thought and did. Ancient sources differ on the question of whether or not he wrote anything down himself; but if he did, his writings have not survived. Instead, we have testimonies concerning him in the work of later writers, which inevitably reflect the biases of those writers, and of the sources which they themselves used. These may tell us something about how Thales was perceived by later generations (which is in itself interesting), but it means that they need to be treated with a degree of scepticism regarding their historical accuracy. Similarly, as we shall see, more recent commentators have sometimes generated, from the scant material we have about Thales, inferences which support their particular understandings of him, and of his achievements.

Chapter Seven

114

With these caveats in place, what can be said about Thales from ancient sources? It is recounted that he was a man of Miletus; that his family hailed originally from Phoenicia; and that he spent some time studying in Egypt (Kirk et al. 2007, 76-79). We can glean some ideas about his activities, and his views. Some sources bear testimony to Thales as a practical man, who made empirical observations of the world around him:1 x

x x

x x

“It is said that Croesus was at a loss how his army should cross the river … and that Thales, who was present in the army, made the river, which flowed on the left hand of the army, flow on the right hand also. … he dug a deep channel … the result was that as soon as the river was divided it became fordable in both its parts” (Herodotus). “Some think he was the first to study the heavenly bodies and to foretell eclipses of the sun and solstices” (Diogenes Laertius. Thales’ predication of an eclipse is also related in Herodotus and Dercyllides). “Thales was clever in judgment, not least because he was said to have measured the little stars of the Wain, by which the Phoenicians sail” (Callimachus). “Hieronymus says that [Thales] actually measured the height of the pyramids by their shadow, having observed the time when our own shadow is equal to our height” (Diogenes Laertius). “Thales is traditionally the first to have revealed the investigation of nature to the Greeks; he had many predecessors … but so far outpassed them as to blot out all who came before him” (Simplicius).

This last quotation is interesting in that it reminds us that although these ancient thinkers may be the oldest ones of whom accounts have come down to us, they were – perhaps inevitably – themselves building on traditions of thought that pre-dated them. Other testimonies speak of Thales’ beliefs about the natural world, most famously of his understanding that the most primitive of all the elements is water: x

1

“Others say that the earth rests on water. For this is the most ancient account we have received, which they say was given by Thales the Milesian, that it stays in place through floating like a log or some other such thing … as though the same argument did not apply to the water supporting the earth as to the earth itself” (Aristotle).

These quotations, and all subsequent ones from ancient sources relating to Thales, are taken from Kirk et al. 2007: 76-99 (unless otherwise indicated).

The Case of Thales’ Ox x x

115

“[Thales] said that the world is held up by water and rides like a ship, and when it is said to ‘quake’ it is actually rocking because of the earth’s movement” (Seneca). “Thales … says that it [the original source of all existing things] is water (and therefore declared that the earth is on water), perhaps taking this supposition from seeing the nurture of all things to be moist, and the warm itself coming-to-be from this and living by this (that from which they come-to-be being the principle of all things) – taking the supposition both from this and from the seeds of all things having a moist nature, water being the natural principle of moist things” (Aristotle).

Some of Thales’ speculations about the natural world suggest a dualistic outlook. He clearly believed in the existence of souls, and ascribed these even to what we would consider to be inanimate objects, like magnets: x x x

“Thales seems … to have supposed that the soul was something kinetic, if he said that the [Magnesian] stone possesses soul because it moves iron” (Aristotle). “Aristotle and Hippias say that he [Thales] gave a share of soul even to inanimate objects, using Magnesian stone and amber as indications” (Diogenes Laertius). “Some say that [soul] is intermingled in the universe, for which reason, perhaps, Thales also thought that all things are full of gods” (Aristotle).

There are also some personal anecdotes about Thales, the historical nature of which is questioned: x

x

x

“A witty and attractive Thracian servant-girl is said to have mocked Thales for falling into a well while he was observing the sky and gazing upwards; declaring that he was eager to know the things in the sky, but that what was behind him and just by his feet escaped his notice” (Plato). After being mocked because of his poverty he correctly predicted there would be a large olive crop one year, so the previous winter he bought up interests in all the local olive-presses, which he was then able to hire out on his own terms; “thus demonstrating that it is easy for philosophers to be rich, if they wish, but that it is not in this that they are interested” (Aristotle). “[Thales] was the first to inscribe a right-angled triangle in a circle, whereupon he sacrificed an ox. Others tell the tale of Pythagoras” (Diogenes Laertius).

116

Chapter Seven

This is not an exhaustive compendium of testimonies, but it gives a flavour of the person and thought of Thales as these have been transmitted to us; and it is on the basis of such tantalising fragments as these that it has been suggested that Thales was a proto-scientist. Thus Bertrand Russell (1979, 44-45) is clear that Thales may be regarded “rather as a man of science than as a philosopher in the modern sense of the word”, adding that “The statement that everything is made of water is to be regarded as a scientific hypothesis”. W. K. C. Guthrie (2013, 22) writes that the Ionians “were absorbed in the first attempts at a scientific explanation of the Universe”, this being “an attempt to solve the problems of the Universe by reason only, as opposed to the acceptance of purely magical or theological explanations”. Barnes (1987, 16) comments of the Pre-socratics in general that they “invented the very idea of science and philosophy. They hit upon that special way of looking at the world which is the scientific or rational way”, adding that “they offered reasons for their opinions, they gave arguments for their views” (Barnes 1987, 24). Extending these arguments further, Julian Baggini (2003, 74) maintains that “atheism can be understood … as a self-contained belief system if it is seen as a commitment to the view that there is only one world and this is the world of nature”, from which he infers that “to understand the origins of atheism you have to understand the origins of naturalism. And naturalism starts with the pre-Socratic Milesian philosophers of the 6th century BCE, Thales, Anaximander, and Anaximenes”. So, it would appear that Thales was not only a proto-scientist: he was a proto-atheist, too. Now, there were of course atheists in the ancient world: Plato devotes book 10 of his “Laws” to arguing against them, and this includes an attack on those who “maintain that fire, water, earth and air owe their existence to nature and chance, and in no case to art, and that it is by means of these entirely inanimate substances that the secondary physical bodies – the earth, sun, moon and stars – have been produced” (Plato 1975, 416 (889A)). But there are three important questions here which commentators tend to confuse. First, to what extent were Thales and his followers truly “scientists”, in our modern sense of the word? Is this a word which it is in any way appropriate to use of them? Second, to what extent were they atheists spurning belief in a God or gods? And third, is there any sense at all that their “scientific” approach to world compelled them to atheistic positions? Clearly, there are real dangers of introducing anachronisms when considering a figure so remote from us and from our world as Thales. Trivially, he was clearly not a “scientist” in any sense congruent with our modern understandings, since the word “scientist” was first used by

The Case of Thales’ Ox

117

William Whewell in 1834 (Ross 1963, 71-72). Waterfield (2000, xvii) cautions that “they [i.e. the Pre-socratics] should not be described as scientists in the modern sense of the word”. Might it be said, though, that Thales was in any sense a practitioner of science? Care is again necessary here. Brooke (1991, 6-7) observes that “There have been so many definitions [of science] offered by philosophers, and by scientists themselves … Many refer to some unique ‘scientific method’ to which exemplary science is supposed to conform. But, as the Cambridge philosopher William Whewell observed, almost a hundred and fifty years ago, the history of science already showed that each new branch of scientific enquiry had required its own distinctive methodology.”

From the testimonies concerning him that have come down to us, we may perhaps go so far as to say that Thales was someone who observed things in the natural world (like the stars, and the length of shadows) and made predictions and inferences on the basis of them (concerning eclipses, and otherwise unmeasurable heights). He appears to have hypothesised about the nature of things in a way that went beyond what we would now consider to be the available evidence (his idea that earthquakes might be explained by the rocking of the earth as it floats on water is clearly a scaled-up analogy with something that might be observed (like a boat on the sea), rather than anything empirically testable). However, he does appear to have sought out reasons for phenomena that go beyond resting content with explanations taking the form “the gods are responsible for it”, and this may suggest a kinship with the naturalistic methodology of modern science. To this extent, it might appear reasonable to think of Thales as a proto-scientist. But does this observation mean that Thales should be considered to be a proto-atheist, too? “Atheism”, of course, is as problematic a word as “science”. To quote Brooke again (1991, 194), “‘atheism’ often takes its character from the particular form of theism it rejects”. Clearly, then, there will be different forms of atheism appearing at different times in history. Thales does not appear to have suffered any opprobrium for countering the theistic beliefs of his contemporaries. It would appear that such opprobrium could be aroused by those seeking natural causes for phenomena which others attributed to the gods: the later philosopher Anaxagoras of Clazomenae is said to have been put on trial in Athens “on a charge of impiety, because he declared the sun to be a mass of red-hot metal” (Diogenes Laertius 1972, 143). However, Diogenes Laertius notes that such charges may have been brought as a consequence of the politics of the day, rather than relating to Anaxagoras’s beliefs per se – an interesting reminder of the way in which orthodoxy and heterodoxy may

118

Chapter Seven

often be bound up with political, as much as religious, questions. Additionally, as we have seen, Plato wrote against “atheists” who saw everything in terms of nature and chance, and no such label appears to have been attached to Thales in antiquity. Further evidence against the notion that Thales espoused ideas which might have been considered atheistic is the statement that he believed all things to be “full of gods”. Another testimony, which is downplayed by scholars to the extent that it is dismissed as “entertaining, if not convincing” by Kirk et al. (2007, 86) – presumably on the grounds that it acknowledges itself that it may be misattributed – is that anecdote that Thales “was the first to inscribe a right-angled triangle in a circle, whereupon he sacrificed an ox. Others tell the tale of Pythagoras” (Diogenes Laertius 1972, 27). But can we dismiss the tale of Thales’ ox as readily as this? We may also note Thales’ attribution of soul to magnets and to amber, on the grounds that they have the power to move other things. All of this may make us inclined to agree with Peter Harrison (2015, 24), that the testimony we have of Thales suggests that his “were not the actions of a hard-nosed scientific naturalist”. What, then, about the idea that in Thales and the other Ionian philosophers we may see the origins of a split between “science” and “religion”? It is fascinating that some classical scholars have been keen to emphasise such a split, painting a vivid contrast between “science” and “religion” in these thinkers. Thus Guthrie (2010, 29) writes that with the Ionian philosophers, “For religious faith there is substituted the faith that was and remains the basis of scientific thought with all its triumphs and with all its limitations: that is, the faith that the visible world conceals a rationality and intelligible order, that the causes of the natural world are to be sought within its boundaries, and that autonomous reason is our sole and sufficient instrument for the search.”

This reads very much like an example of the danger highlighted by Catherine Osborne (2004, 34), who writes that “Historians of philosophy have always looked back at the past to reassure themselves that earlier thinkers were on the way to discovering the things that we now believe”. More judicious, perhaps, is the conclusion of Thrower (2000, 16), who writes, “if we look to the Ionian philosophers for a wholly naturalistic explanation of the world then we look in vain”, adding: “if [the Ionians] are the first natural philosophers, they are the first natural theologians as well. It is perhaps an injustice to them that the purely naturalistic side of their thinking should subsequently have exercised the greatest influence” (Thrower 2000, 17).

The Case of Thales’ Ox

119

Perhaps some may find it strange that a proto-scientist could be anything other than an atheist; but it is clear that the term “atheist” can only be applied to Thales in defiance of such evidence as we possess about his beliefs. In any case, why should it be the case that seeking out natural explanations for natural phenomena must necessarily go hand-in-hand with a rejection of any belief in God? This would appear to be a peculiar modern prejudice required of those buying into what Harrison refers to as “the mythical origins of science”, which he describes as follows (Harrison 2015, 22): “Science is said to have its origins in Greek antiquity when philosophers first broke away from the myths of their forebears and sought rational explanations for natural phenomena. Science subsequently suffered a setback with the advent of Christianity … But it then emerged triumphant with the scientific revolution of the seventeenth century when it finally broke away from religion and set out on its progressive path to the present.”

This has been a pervasive view in some quarters, but it is clearly unsustainable (cf. Lindberg 2009, Osler 2009). In seeking out the historical limits of science, with the earliest thinker of whom we have written testimony – and who has been claimed by many as an originator of scientific ways of thinking – we are therefore brought to the following conclusions. Insofar as we can reconstruct what Thales of Miletus did, thought and believed, he appears to have interrogated the natural world in a way which resonates with those of a modern scientific temperament. He fashioned hypotheses about the natural world which did not require assumptions about a God, or gods. But he also appears to have espoused a kind of pantheism and to have attributed souls to objects we would consider inanimate; and he may have participated in religious rites that went beyond what was required of him (why else would that ox be thought worthy of mention?). He was certainly not subject to any of the disapproval that could be directed against those who were believed to be irreligious in the ancient world. At the historical limits of science, then, we see what we would now consider to be scientific ideas and practices sitting happily alongside what we would now consider to be religious beliefs and practices. And perhaps we should be asking ourselves why we might think this to be unexpected, or why we should find it in any way remarkable. The myth that sets science against religion cannot be sustained by any impartial analysis of the testimonies we have about Thales and about the society in which he lived. This historical limit of science would strongly suggest it to be perfectly compatible with religion.

120

Chapter Seven

References Baggini, J. 2003. Atheism: A very short introduction. Oxford: Oxford University Press. Barnes, J. 1987. Early Greek Philosophy. London: Penguin. Brooke, J. H. 1991. Science and Religion: Some historical perspectives. Cambridge: Cambridge University Press. Diogenes Laertius. 1972. Lives of Eminent Philosophers vol. 1, trans. R. D. Hicks. Cambridge, Mass.: Loeb Classical Library. Guthrie, W. K. C. 2010. A History of Greek Philosophy vol. 1: The earlier Presocratics and the Pythagoreans, revised edition. Cambridge: Cambridge University Press. —. 2013. The Greek Philosophers: from Thales to Aristotle. Abingdon: Routledge Classics. Harrison, P. 2015. The Territories of Science and Religion. Chicago: University of Chicago Press. Kirk, G. S., J. E. Raven and M. Schofield. 2007. The Presocratic Philosophers: A critical History with a selection of texts, second edition. Cambridge: Cambridge University Press. Lindberg, D. C. 2009. “Myth 1: that the rise of Christianity was responsible for the demise of ancient science”, in R. L. Numbers (ed.), Galileo Goes to Jail, and other myths about science and religion. Cambridge, Mass.: Harvard University Press. Osborne, C. 2004. Presocratic Philosophy: A very short introduction. Oxford: Oxford University Press. Osler, M. J. 2009. “Myth 10: that the Scientific Revolution liberated science from religion”, in R. L. Numbers (ed.), Galileo Goes to Jail, and other myths about science and religion. Cambridge, Mass.: Harvard University Press. Plato. 1975. Laws, trans. T. J. Saunders. London: Penguin. Ross, S. 1962. “Scientist: The story of a word”. Annals of Science vol. 18 no. 2. Russell, B. 1979. History of Western Philosophy. London: George Allen and Unwin. Thrower, J. 2000. Western Atheism: A short history. Amherst, NY: Prometheus Books. Waterfield, R. 2000. The First Philosophers: The Presocratics and the Sophists. Oxford: Oxford University Press.

CHAPTER EIGHT JOHN SPENCER AND THE LIMITS OF NATURAL CAUSATION IN EARLY MODERN ENGLAND PETER N. JORDAN

For some members of modern western society, it is an article of faith that science will ultimately be able to understand and explain every phenomenon in nature. This belief is usually accompanied by the conviction that the development of a scientific understanding of nature comes at the expense of religious or theological claims about the existence of God and about what happens within the natural world. Because of this perceived zerosum relation between science and religion, scientists, as sociologists Elaine Howard Ecklund and Elizabeth Long (2011) explain, are often seen as purveyors of a “secularist impulse” to the extent that they are seen as “undermining religious authority by their success in deciphering the mysteries of the natural order without recourse to supernatural aid or guidance” (Ecklund and Long 2011, 254). Although this view—that theological understanding or conviction must necessarily lose out to naturalistic exploration and explanation as science continues to gain insight into the world’s operations—is often regarded as self-evidently true by those who hold it, history shows that it is by no means inevitable. Much recent historical scholarship has demonstrated that the perception that science and religion are necessarily at odds with one another is a modern one that emerged with vigour only during the nineteenth century (Lindberg and Numbers 1986; Brooke 1991; Lindberg and Numbers 2003; Harrison 2015). Rather than seeing them as opposed to each other, pre-modern authors typically sought to integrate naturalistic and theological elements into a single holistic account of the world, with each regulating the other in crucial ways. Indeed, as the writings of an array of figures from seventeenth-century England in particular—a time and place of great importance in the emergence of recognisably modern science—reveal, the two were often so intimately intertwined that scholars speak of the “integrated religioscientific philosophical outlook” of the

122

Chapter Eight

period (Burns 1981, 12), and of the “comprehensive scientific theism” developed at the time (Force 1990, 143). The present chapter focuses on one author from early modern England whose work typifies this desire to hold together these two elements that in recent years have so often been thought of as opposed to each another. In his 1663 Discourse Concerning Prodigies, the Cambridge theologian and Hebrew scholar John Spencer (1630–1693) demonstrates a staunch commitment to natural-philosophical study and explanation of phenomena, while at the same time seeing God as sovereign over and acting within the created order. Far from the former occluding the latter, Spencer thinks natural philosophy has a crucial role to play in distinguishing between better and worse theological positions, and his Discourse indicates how natural philosophy can help to overcome what he deems to be a flawed understanding of God’s providential activity in the world. Spencer’s treatise therefore represents an ideal case study both for appreciating how naturalistic and theological perspectives could fit together in the seventeenth century, and for understanding how theology could (uncontroversially) set limits on what natural causes can explain. *** Spencer wrote his Discourse in response to the problem of prodigy belief plaguing his native country. England, in Spencer’s view, has “grown Africa,” by which he means that members of the public all too readily see “monstrous and strange sights” as bearing some sort of divine meaning or significance (Spencer 1663, B1v). This tendency, one that signifies England’s degeneration into an “age of wonders,” was producing interpretations of unexpected phenomena that saw them as presages of “strange and unusual plagues” approaching individuals or groups in the future—in other words, as prodigies (Spencer 1663, B1v). As Alexandra Walsham’s comprehensive study of providential thinking in early modern England suggests, Spencer is hardly exaggerating how common such views were. The notion that God communicates with human beings through unexpected occurrences in the created order was widespread in the sixteenth and seventeenth centuries. Not confined to popular religiosity, such providentialism exerted “practical, emotional and imaginative” effects on men and women from all walks of life (Walsham 1999, 3). Yet for all its popularity, Spencer saw numerous problems with the more intense forms of providentialism in which God is seen as communicating with humanity with great frequency, and as regularly acting directly within the created order. Chief among his concerns are its

John Spencer and the Limits of Natural Causation

123

theological and political implications: prodigy belief leads to superstitious and enthusiastic expressions of Christian piety and practice, and it has the potential to be politically destabilising. Given especially the theological errors that inform it and the impious forms of Christian life that it inspires, Spencer sets himself the tasks of ridding England of the scourge of prodigy belief and of turning his native land towards a more mature form of Christian life, one that derives from what he regards as a more faithful and sedate understanding of providence. Although eliminating prodigy belief is in Spencer’s view a necessary undertaking, it is also a difficult and risky one, and not only because of the extent to which prodigy belief has captured the imaginations of his contemporaries. Purging prodigy belief is a perilous undertaking because one risks throwing out the baby with the bathwater. Uncertainty about whether any given unusual occurrence is in fact miraculous—divinely caused—or the product of natural causes, or whether a specific event is divinely communicative or not, means that unless one is careful a real miracle may erroneously be taken as a natural event, and a real divine missive ignored. The challenge facing one who would overturn the intense providentialism that serves as the basis of prodigy belief is to recognise those events that really do convey something of God’s intentionality, or which really do reflect God’s direct agency, while not imagining that one sees either of these in situations where they are not actually present. Spencer offers a smorgasbord of reasons why unexpected events should not be regarded as messages from on high, with arguments like the following put forward to challenge the flawed view of providence associated with prodigy belief and the sacrilegious forms of piety that accompany it: (i) The signs that God supposedly sends to humanity are too “dark and obscure” and fail clearly to convey the message that God wants to communicate. As a result, it is impossible to determine definitively what message a particular occurrence portends, or to identify which person or persons the message is intended for (Spencer 1663, 6-7); (ii) According to Spencer’s chosen sources, prodigy belief was a major feature of preChristian Egyptian and Roman religion, a period characterised by the “greatest ignorance … in matters of religion” (Spencer 1663, 8). Given Spencer’s assumption that the character of the historical period in which something originates or becomes commonplace says a great deal about that thing, the fact that prodigy belief emerged during a period of religious philistinism does little to recommend or legitimate it; (iii) The frequency with which people see divine intentionality and agency in the created order tends to increase in times of political instability, or more generally during “times of publick fears, troubles, [and] confusions” (Spencer 1663, 9); (iv)

124

Chapter Eight

During earlier periods of history, the state of affairs between God and humanity—the “condition and temper of the oeconomy we are now under”—may have been such that God needed to address himself to the “lower faculties” of the human soul. Since the Incarnation, however, God manages the world in a more “sedate, cool, and silent manner” that needs few, if any, dramatic declarations from heaven (Spencer 1663, 11-12); (v) The guilt that human beings feel when we do wrong makes us expect punishment even though none may be in the offing. Because of this, even when we cannot conclusively determine the meaning of God’s “strange characters” in nature, we nevertheless conclude that the “general intention” must be “wrath and judgement” (Spencer 1663, 21); (vi) The “general temper” of the human soul is such as to make us expect parallels between what happens in history and what happens in nature. We are, according to Spencer, “easily inclin’d to believe great and mighty changes in states, usher’d with the solemnity of some mighty and analogous changes in nature” (Spencer 1663, 26). As if these and other similar observations and arguments were not enough to bring prodigy belief into disrepute, Spencer also attacks it from a completely different angle by appealing to the insights and explanatory power of natural philosophy. The study of created causes in the early modern period took place within the realm of natural philosophy, a subject that constituted a central element of the university curriculum (Blair, 2006). With a university education a reality for only a small portion of the population, however, many of Spencer’s contemporaries were ignorant of natural philosophy and what it teaches about the natures and causal powers of created entities. As a result, they tended to see the world as a theatre of constant divine activity and communication oriented toward human beings. Such a “religion of prodigies”, conceived in large part “in the womb of gross ignorance” (Spencer 1663, 9), fails to appreciate how the regular operations of nature—the “harmonious consort of uniform agents” (Spencer 1663, 10)—can fully account for practically every phenomenon (unusual or not) that occurs in nature without needing to appeal to God’s miraculous work above and beyond what created causes can themselves produce. Without natural philosophy, all one is left with is “deformed thought of God” (Spencer 1663, 53), superstition based on a mistaken view of providence, and a distorted religious devotion characterised by “slavish fears” (Spencer 1663, 23), “bondage of spirit”, and “perpetual fear and astonishment” (Spencer 1663, 24). What can be done to combat this ruinous disorder? In addition to recommending that his peers change how they imagine God and how they habitually respond to the world—tempering their fears and hopes

John Spencer and the Limits of Natural Causation

125

regarding future things, and cultivating a “generous indifferency and deadness of minde as to the good and evil things of this world” (Spencer 1663, 76)—Spencer unequivocally endorses the study of natural philosophy. If human “fancy” has a tendency to “suggest very monstrous and superstitious notions” or interpretations of those occurrences whose causes and natures we are unsure of, the study of natural philosophy will remedy this by revealing the “methods of nature” responsible (Spencer 1663, 76). The illumination that this knowledge provides is transformative both affectively and theologically: those who comprehend the orderliness of the world “smile at their former superstitions, as men in the day laugh at those dreams and visions which frighted them in the night” (Spencer 1663, 9). To demonstrate the power of natural philosophy for debunking prodigies, Spencer uses the example of comets. During the early modern period comets were frequently deemed to be prophetic symbols (Genuth, 2000), “beacon[s] fir’d from heaven to alarm the world, and to give intimation of an approaching evil” (Spencer 1663, 13). In Spencer’s view there are two problems with the prevailing position. First, people consistently misunderstand comets’ true nature and composition. Second, they misunderstand what comets are really made of because they let their understanding of what comets are made of be driven by two assumptions: first, that comets represent a foretaste of a coming judgement, and second, that comets’ nature reflects the judgement that they foretell. According to this perspective, comets are “hot and sulphureous exhalations, set on fire, which (as ‘twer the feavers of nature) prey upon the …moysture of it’ because they augur fevers in men and women and ‘choler in princes and nations’” (Spencer 1663, 14). Prodigy belief thus leads to erroneous judgements about matters that properly should lie within the purview of natural philosophy. Against this “vulgar conceit” Spencer offers a range of observations of comets which together suggest that they do not consist of fire, and should in no way be taken as signs of impending punishment. Spencer points to the constant size and shape of the comet, along with the “direct and uniform ray of light in the tayl or train” (Spencer 1663, 14-15). Such constancy would be highly unlikely if the comet were made of fire, because the size of the fire would change as the size of its fuel source changed. He also highlights the uniformity and constancy of comets’ motion, as opposed to the typically “giddy and casual” course that fire follows (Spencer 1663, 15). Combined with other common-sense arguments—for example, the fact that comets typically pass over numerous countries makes it almost impossible to know at which population its message of coming judgement (if it indeed bears one) is

126

Chapter Eight

targeted (Spencer 1663, 17)—Spencer uses natural philosophy to repudiate the claim that comets function as predictions of impending judgements whose form reflects comets’ nature. One can only know what a comet is by carefully observing its behaviour in the night sky, and by drawing conclusions from those observations that are untainted by superstitious theological presuppositions. Unlike twenty-first century thinkers who might conclude from empirical observations and scientific theorising that science rids the world of God and renders religious notions vacuous, Spencer sees such knowledge as serving a purifying or corrective function within religion. Only a proper understanding of irregular occurrences like comets, for example, based on sound natural philosophical study can generate a true sense of the “being, providence, and greatness of God” (Spencer 1663, 42) in place of the unsatisfactory versions of these notions associated with prodigy belief. Unusual events do indeed stir up awareness of God, but when rightly understood they point to God’s greatness and inspire reverence rather than distrustful fear (Spencer 1663, 42). If an unusual event does genuinely serve as a sign, it signifies not an individual’s punishment in the immediate future, but rather the dissolution of the entire world and God’s judgement of all humanity on the last day (Spencer 1663, 42). And in their exceptionality and unusualness, rare occurrences remind those who observe them that God is sovereign over his creation and in control of all things, and can “command nature into a quite different order and posture from what our shorter thoughts can reach into” (Spencer 1663, 45). Unusual occurrences in nature may be a little less exotic and less interesting after natural philosophy has finished with them, but what is left is a religion freed from those elements that make it readily susceptible to critique and rejection. In proclaiming its explanatory potential Spencer is more hopeful about natural philosophy and its power to reveal the extent of the causal powers inherent to created entities than many before him had been. Following Francis Bacon (1561-1626), Spencer thinks that Aristotelian natural philosophy has limited itself to studying the surface of nature, and has not properly probed nature’s depths because it too quickly skips past the rare and the unusual—in the early modern period, those things often categorised as “preternatural”—in pursuit of understanding that which happens usually or for the most part (Daston 1991; Daston 1998). By first generating a Baconian natural history of the “anomals in nature”, Spencer believes that one might then use natural philosophy to “see beyond the surface of those things, which as yet seem plac’d in the world, but to confound and pose us” (Spencer 1663, 45; Daston 1991; Harrison 2006).

John Spencer and the Limits of Natural Causation

127

The fact that there is “more of nature” in rare occurrences than in more common ones makes the former the perfect place for natural philosophy to demonstrate its capacity to reveal how the created order works (Spencer 1663, 10). Through such study one will “discover [nature’s] silent processes, and more cryptick methods in the building up and compleating of her works” (Spencer 1663, 45). For all his interest in natural philosophy, however, Spencer is wary of under- or overemphasising its importance. Just as too strong a focus on God at the expense of created causes has led countless of his contemporaries to stray into delusion and impiety— “an utter ignorance of causes natural, leads … to superstition, a slavish observance of and blind devotion toward God” (Spencer 1663, 104)—so the wrong level of focus on created causes involves its own hazards. Insufficient attention to natural causes leaves one at risk of becoming an atheist: “As a distinct and full view of second [i.e., created] causes, begets religion, (because necessarily directing the eye, at last, upon the First [i.e., God]) a superficial and imperfect notice, [begets] Atheism, (which like the bat, is noted to flie abroad in the twilight, in a kind of middle state between the darkness of ignorance and light of knowledge)” (Spencer 1663, 104).

Too much attention is similarly detrimental: “neither must we loose our religion in philosophy, by dwelling on second causes, till we quite forget the first, and become profane” (Spencer 1663, 43). Both the right degree of attention to nature and the right reverence toward God are crucial for Christianity to flourish. As if these complications were not enough, further challenges present themselves when one tries to determine in any given case whether created causes are sufficient to explain what is observed. As we have seen, Spencer’s critique of prodigies rests on the assumption that the vast majority of unusual occurrences can be explained naturalistically by identifying the created causes underlying them. There are times, however, when nature appears to go so far beyond its regular course that one cannot help but regard an occurrence as supernatural or miraculous. The providential context within which miraculous episodes may occur thus sets a clear outer limit on what natural philosophy will be able to understand or explain: if an effect genuinely transcends the power of created causes to generate on their own, natural philosophy will not be able to account for it. While this distinction between that which is caused by the intrinsic causal powers of created entities (i.e., the natural or the preternatural) and that which transcends the power of created causes (i.e., supernatural) is straightforward in theory, Spencer concedes that it can be rather difficult

128

Chapter Eight

to identify where the boundary lies in practice. It should come as no surprise by this point to discover that Spencer’s tendency is to err on the side of caution when identifying causes, it being a “great example of rashness” to too quickly label something a miracle (Spencer 1663, 66). Among the reasons for his tentativeness Spencer points to the lack of clear understanding of the “extent and compass” of created causes, and the degree to which those causes may “exceed the lines of common and ordinary operation” and give the appearance of a miracle without a divine hand genuinely helping them to do so (Spencer 1663, 66). Repeating a claim often made during the early modern period, Spencer also blames the “prince of the powers of the ayr”—that is, the devil—for confounding attempts clearly to distinguish between natural and supernatural events (Spencer 1663, 67). As a created being, the devil falls squarely on the natural side of the natural/supernatural boundary (Clark 1984; Daston 1998), yet his extraordinary powers within the created realm mean that he can “ape” miracles “by those wonderfull impressions, he can make upon natural bodies” (Spencer 1663, 67). To cope with the ambiguities of the conceptually clear but practically murky distinction between the natural and the supernatural, Spencer offers a couple of criteria for how past miracles have been discerned. First, he claims that in the scriptures God always produces miracles at special times for particular purposes—that is, miracles “attend some great and excellent ends and occasions” (Spencer 1663, 67). Following this biblical precedent, one should be very wary of labelling something a miracle when similar conditions are not obviously apparent. And second, Spencer claims that the meaning of past miracles is never mysterious, because God always grants someone the ability to declare the meaning that God intended for a given occurrence. As he puts it, “miracles are Gods seal, and therefore some writing and evidence they must be affixt unto, for as the writing without the seal wants authority, so the seal without the writing, certainty and significancy” (Spencer 1663, 68). It is unclear how readily these criteria can be applied to future miracles, but at the very least they demonstrate Spencer’s awareness of how difficult it can be to discern what kind of causes ultimately are responsible for unusual events. *** According to the theological and philosophical traditions that had been handed down to scholars in early modern Europe, investigation of the natural world was subject to certain limits. Aristotelian-inspired natural philosophy was limited in one respect by the conviction that true

John Spencer and the Limits of Natural Causation

129

knowledge could only be obtained for that which happened usually or for the most part. When undertaken within a Christian view of the world, natural philosophy was understandably limited because it could not explain miraculous occurrences. As we have seen, the first of these limits was directly challenged during the early modern period. Francis Bacon and those inspired by him wanted to expand natural philosophy’s scope by collecting a natural history of unusual phenomena and anomalies in nature, and subjecting those to natural philosophical scrutiny. The second limit, however, one put in place by theology’s understanding that God could act within the created order, remained firmly in place. As Spencer’s treatise suggests, the line dividing what could be explained through natural causes and what must be explained by appealing to divine action was not always easy to discern, but it did exist. In light of these limits, one of the more instructive aspects of Spencer’s Discourse for twenty-first-century readers is the surprising fact that expansion of what was regarded as properly within the purview of natural philosophy—a key precursor to modern science—not only occurred within the limits provided by theology, but that this expansion was seen to be a religiously useful development. Early modern scholars like Spencer saw a potential in natural philosophy that many do not see in modern science: that knowledge of the inner workings of nature need not challenge or undermine religious belief and practice, but can instead refine them.

References Blair, A. 2006. “Natural Philosophy”, in Park, K. & Daston, L. (eds.), The Cambridge History of Science, Volume 3: Early Modern Science. Cambridge: Cambridge University Press, 365-406. Brooke, J.H. 1991. Science and Religion: Some Historical Perspectives. Cambridge: Cambridge University Press. Burns, R.M. 1981. The Great Debate on Miracles: From Joseph Glanvill to David Hume. Lewisburg: Bucknell University Press. Clark, S. 1984. “Scientific Status of Demonology”, in Vickers, B. (ed.), Occult and Scientific Mentalities in the Renaissance. Cambridge: Cambridge University Press, 351-374. Daston, L. 1991. “Marvelous Facts and Miraculous Evidence in Early Modern Europe”, Critical Inquiry, 18, 93-124. —. 1998. “The Nature of Nature in Early Modern Europe”, Configurations 6, 49-172.

130

Chapter Eight

Ecklund, E.H. & Long E. 2011. “Scientists and Spirituality”, Sociology of Religion 72, 253-74. Force, J.E. 1990. “The Breakdown of the Newtonian Synthesis of Science and Religion: Hume, Newton, and the Royal Society”, in Force, J.E. & Popkin, R.H. (eds.), Essays on the Context, Nature, and Influence of Isaac Newton’s Theology. Dordrecht:Kluwer, 143-163. Genuth, S.S. 2000. “Comets and Meteors”, in Ferngren G.B. (ed.), The History of Science and Religion in the Western Tradition: An Encyclopedia. New York: Garland, 371-374. Harrison, P. 2006. “Miracles, Early Modern Science, and Rational Religion”, Church History, 73, 493-510. —. 2015. The Territories of Science and Religion. Chicago: University of Chicago Press. Lindberg, D.C. & Numbers R.L. (eds.) 1986. God and Nature: Historical Essays on the Encounter between Christianity and Science. Berkeley: University of California Press. —. 2003. When Science and Christianity Meet. Chicago: University of Chicago Press. Spencer, J. 1663. A Discourse Concerning Prodigies. London. Walsham, A. 1999. Providence in Early Modern England. Oxford: Oxford University Press.

CHAPTER NINE NATURALISM, THE LIMITS OF SCIENCE AND THE CASE FOR NON-SCIENTIFIC KNOWLEDGE EMMANUEL NARTEY

Alex Rosenberg defines naturalism as a philosophical theory that treats “science as our most reliable source of knowledge and the scientific method as our most effective route to knowledge” (Rosenberg 2011, 247). The claim that all truths are discoverable by hard science—paradigmatically physics—is predicated on four hundred years of scientific success in prediction, control, and technology. Rosenberg argues that since the triumphs of science show that physics has made a good start, we should be optimistic that “it will do better than any other approach at getting things right” (Rosenberg 2011, 248). But are those “things” confined to the questions of physics or do they encompass all questions? In what follows, I begin by examining the nature and varieties of philosophical naturalism. Next, I discuss the intrinsic limits of the scientific method and its presuppositions. I argue that science does not have an answer to all questions, and that there is a strong case to be made for an irreducibly nonscientific knowledge. I conclude by exploring why scientists and nonscientists cannot duck the ultimate questions about meaning, values, and purpose. Varieties of naturalism I take scientific naturalism to be a conjunction of the following claims: (1) The natural world described by the natural sciences is all that there is, and no other ontology is credible; supernatural or spooky entities do not exist. (2) Science is “the measure of all things”. The scientific method is the one and only legitimate mode of investigation and knowledge and all other forms of knowing are either candidates for elimination or reducible in

132

Chapter Nine

principle to natural science. Claim 1 is often characterized as ontological, and claim 2 as methodological naturalism.1 The scientific method is considered the final arbiter of truth, as it is authoritative, embodying a spirit that accentuates and privileges empirical observation, accuracy, honesty, rigor, precision, testing, and curiosity—all of which are essential to discovering truths about the world. Naturalism is predicated on scientific progress and works on the assumption that the natural world is knowable and if we want to really understand the world and our place in it, we need only look to scientific findings. Centuries of remarkable scientific progress in the fields of physics, medicine, technology, psychology, and neuroscience has revolutionized and deepened our understanding of the world, shedding light on things previously considered miraculous or unknowable. Much still remains inexplicable to science, but the belief is that these current shortcomings are not an indication of how things will turn out in the future. Science is on the march and our past and current successes are a reliable indication for the future. Scientific success and optimism inspires us to dismiss ontologies that fail to meet the scientific test. The key to explanatory potency is to accept only explanations of natural phenomena that rely solely on natural causes. While Aristotle famously held that in our very human desire for intelligibility and meaning we must invoke both the physical world and the metaphysical realm, scientific naturalists think that the only things that can be truly said to exist are the constituents of the natural world. In our quest to explain and understand reality we recognize that science is fallible and amenable to self-correction, and a resource of tremendous epistemic value. Rosenberg claims that, unlike other disciplines, science is strongly committed to providing a unified theory of everything. But achieving the scientific ideal of this unification project entails that “science can’t accept interpretation as providing knowledge of human affairs if it can’t at least in principle be absorbed into, perhaps even reduced to, neuroscience. That’s the job of neurophilosophy” (Rosenberg 2011, 324–5). It is clear that this form of extreme naturalism is reductive. It would be hard to challenge the basic picture of reality presented by the hard sciences—the main programmatic orientation of Anglo-American philosophy. Its picture of reality is at odds with interpretation, the dominant mode of explanation in the humanities, theology, and in our own inner psychological monologues. In Rosenberg’s words,

1

See Papineau, D. 2015. ‘Naturalism’, The Stanford Encyclopedia of Philosophy, Zalta, E.N. (ed.), URL = .

Naturalism, Limits of Science, Case for Non-scientific Knowledge

133

“The only solution to the problem faced by the humanities, history and (autobiography) is to show that interpretation can somehow be grounded in neuroscience. That’s job no. 1 for neurophilosophy. And the odds are against it. If this project doesn’t work, science will have to face plan B: treating the humanities the way we treat the arts, indispensable parts of human experience but not to be mistaken for contributions to knowledge.” (2011, 325–6)

This picture of science appears to leave no room for non-scientific knowledge2, but does a proper understanding of science really eliminate the need for non-scientific knowledge? While reductionists such as Philip Pettit claim that all sciences are reducible to microphysics, non-reductionists such as Hilary Konblith (1994, 39–52) and Philip Kitcher (1992, 53–114) deny this claim. “Disenchanted naturalists” like Rosenberg (2009; 2011; 2013) argue that the so-called “fundamental questions of life” disintegrate once they are properly framed within the scientific worldview, while “optimistic naturalists” like Kitcher contend that fundamental questions of life are legitimate and can be construed and even answered with scientifically legitimate conceptual tools.3 But serious problems confront scientific naturalists when they endeavour to naturalize the most relevant features of the common-sense view of the world. Huw Price (2004, 187) has identified a “placement problem”: “If all reality is ultimately natural reality, how are we to ‘place’ moral facts, mathematical facts, and so on? How are we to locate topics of this kind within a naturalistic framework, thus conceived?” P.F. Strawson, John McDowell, Jenifer Hornsby, Hilary Putnam, and Barry Stroud have proposed a different version of a more liberal naturalism (De Caro and Macarthur 2010) with the aim of accounting for the common-sense features of the world taken at face value, while maintaining compatibility with the scientific worldview. What is clear in all these approaches is the indisputable fact that naturalism is wanting and naturalists are confronted with a deep problem, namely how to account for the significant and arguably irreducible features of the

2

Here I work on the broad definition of ‘knowledge’ as warranted true belief. See Rosenberg, A. ‘The Disenchanted Naturalist’s Guide to Reality.’ In On the Human. http://onthehuman/2009/11/the-disenchanted-naturalists-guide-to-reality/.; Rosenberg, A. 2011a. The Atheist’s Guide to Reality: Enjoying Life without Illusions (New York: W.W. Norton); Rosenberg, A. 2013. ‘Disenchanted Naturalism,’ in The Metaphysics of Evolutionary Contemporary Philosophical Naturalism and Its Implications, Bashour, B & Muller, H.D. (eds), New York: Routledge.

3

134

Chapter Nine

world and experience that elude naturalization.4 The ontological unacceptability of such phenomena may point to the limits of science, to which I now turn.

The limits of science and the case for non-scientific knowledge In his Religion and Science, Bertrand Russell concludes that, “… while it is true that science cannot decide questions of values, that is because they cannot be intellectually decided at all, and lie outside the realm of truth and falsehood. Whatever knowledge is attainable, must be attained by scientific methods; and what science cannot discover, mankind cannot know”5 Russell (1953, 243).

With this in mind, I will focus on the two-pronged question: “What are the limits of scientific inquiry and does it leave any questions for nonscientists to answer?” Science deals with highly quantified variables and yields precise and reliable results. However, as I will show, when it comes to human inquiry, the scientific method is only one of many. Consider the following statements: “lying is wrong” and “honesty is better than dishonesty”.6 The warrant condition for such statements is not satisfied by scientific

4

Baker has recently argued that there are genuinely real first-person perspectives of the world. The first-person perspective is indispensable for agency, moral responsibility, practical reasoning, deliberation, self-evaluation, selfunderstanding, and self-consciousness, and cannot be explained or dismissed by science. This is not because science adopts a third-person perspective, as is commonly thought, but rather because ‘There is no third-person perspective at all. The so-called third-person perspective is centerless; it is the ‘view from nowhere.’’ (Nagel 1986, quoted in Baker 2013, xix). Lynne Rudder Baker defends ‘nearnaturalism’, a view that she thinks adequately accounts for the first-person perspective, denies the claim that science is the exclusive arbiter of reality but is not committed to supernaturalism. Baker, L.R. 2013. Naturalism and the FirstPerson Perspective (New York: Oxford University Press), xix. See also De Caro’s interesting review of Baker in De Caro, M. 2015. ‘Naturalism and the First-Person Perspective’, Philosophical Review, 124(1), 156-158. 5 My emphasis. 6 I do not take these to be universal rules that are invariably true as one can certainly make a convincing case for justified and rational exceptions in some situations.

Naturalism, Limits of Science, Case for Non-scientific Knowledge

135

investigation. This applies to knowledge of mental states, but also to moral truths, and truths about oneself. Alister McGrath is right in thinking that science, although in some regards vital, “illuminates only part of the picture [and] we need that whole picture if we are to live authentic and meaningful lives” (McGrath 2015, 182). Physics, for instance, tells us about the dispositional properties of physical objects but it is unable to disclose the intrinsic nature of physical objects, that is, the categorical properties of those objects. To say that an object has a dispositional property only tells us what the object in question is disposed to do, and what physical law dictates will obtain spatio-temporally once certain antecedent conditions are satisfied. An object—a vase, say—is fragile if it is disposed to break under certain conditions. But saying that a vase is susceptible to being easily broken does not say anything about the molecular constitution or structure of the vase. This structure, the categorical base, is what furnishes the object with its dispositional properties. So we cannot know why the vase, or any fragile object, is brittle without knowing the categorical base of the object. Even if we learn that electrons are disposed to do such-and-such given certain conditions, we still do not know anything about the intrinsic nature of the object that makes this the case. Physics only gives us relational information about how things are nomologically connected and how a modification in one area produces a change in another. It defines how electrons, protons, or quarks fit into and play a causally defined role in this structure of causal connections. Physics tells us about function and what plays what role without saying much about the role-filler. Now, we have no reason to think that nothing else holds true of the role-filler beyond its abstract functional characterizations and causal position. In the words of Bertrand Russell: “The physical world is only known as regards certain abstract features of its space-time structure—features which, because of their abstractness, do not suffice to show whether the physical world is, or is not, different in intrinsic character from the world of mind” (1992 [1948], 240). Physics cannot pretend to give us a complete picture of reality if it bottoms out at dispositional properties; if it offers only relational information and none regarding the intrinsic character of the object it describes and explains. We are thus able to state that science does not give a complete account of reality; it does not even seem to offer a decent account of matter, its own subject matter. This enervates the claim that the hard sciences offer a complete picture of reality leaving no questions for non-scientists to answer, and indicates limits to scientific inquiry. As noted above, scientific naturalists claim that science is the measure of all things and that

136

Chapter Nine

the scientific method is the one and only legitimate mode of investigation and knowledge. What they fail to acknowledge is that science counts on the success of logic and mathematics, which derives its results not from empirical verification but from pure reasoning. The fact that empirical investigations cannot establish the truths of logic and mathematics give further evidence of the limits to science. Heisenberg, Gödel, and others have argued that physics, math, and the humanities all converge on the idea that we cannot know truth or communicate it with total confidence. Knowledge, as warranted true belief, consists in the ability to provide a true answer to a question. Scientific knowledge is knowledge whose warrant derives from, or is due to, scientific inquiry. But there is irreducibly non-scientific knowledge whose warrant is not due to scientific inquiry. As mentioned above, these include knowledge of one’s own mental states, moral knowledge, religious knowledge, and mathematical knowledge. Russell (1912) distinguished between two types of knowledge: “knowledge by acquaintance” and “knowledge by description”.7 The former is knowledge achieved through direct awareness. We are directly aware through sense perceptions (sound, sight, taste, smell, and touch), memories, introspections, universals (e.g. circles, numbers, brotherhood) and—possibly—ourselves. Knowledge by acquaintance makes possible knowledge by description. The latter is either definite (the such and such, e.g. the cat) or indefinite (a such and such, e.g. a cat) in which “such and such” stands for a word or a group of words denoting things we know by acquaintance. For Russell, to know a person is to have knowledge by description, since what we are directly aware of is not a person, but sense perceptions of a body, a voice, and so on. Thus, when we say “We know the President”, then “President” is a kind of shorthand for a description that picks out only a single entity and no other: “the man with black and grey hair we see at a news conference every week”. This description contains only sense perceptions and universals (black, grey, man, etc.), and is therefore knowledge by acquaintance. Through these, we can gain knowledge by description. When the descriptions are definite rather than universal, we can replace the description above with a proper noun: “President Obama”. Conversely, there is a vast difference between experiencing the taste of wine and knowing facts or propositions about wine, e.g. whether it is Italian or French. Likewise, you can know many facts about Abraham Lincoln without being acquainted with him. While 7 See Chapter V of Russell, B, 2009. The Problems of Philosophy, Sioux Falls, SD: NuVision Publications.

Naturalism, Limits of Science, Case for Non-scientific Knowledge

137

objects of propositional knowledge have truth value, objects of knowledge by acquaintance do not. Science operates on a propositional level; it does not give us knowledge by acquaintance. Physicalism starts from the assumption that everything is ultimately physical and can be exhaustively described and explained in the vocabulary of the natural sciences, paradigmatically physics. However, there is reason to suspect that there are limits both to the standard scientific worldview and to physicalism as a theory of mind. Frank Jackson’s (1982 & 1986) so-called “knowledge argument” shows that it is possible for someone to have all the physical facts about x without having all the facts about x: since our experience is not entirely physical, there are facts that are not strictly physical. Jackson asks us to imagine Mary, a super-scientist who knows “everything in completed physics, chemistry, and neurophysiology, and all there is to know about the causal and relational facts consequent upon all this, including of course functional roles” (Jackson 1982, 51). Mary has spent all her life in a room painted entirely in black and white, and has thus never experienced colour. Suppose that she one day leaves her sequestered environment and perceives a red rose for the first time. Given that she already has all the available scientific facts about “red”, does she learn anything new? Most people would respond in the affirmative: Mary didn’t know what it is like to see red before, or to perceive the velvety redness of a scented live rose. She has now, for the first time in her life, had the visual experience of the colour red. Mary thus acquires new information outside the black and white room, and this new information is non-physical. Even though Mary already had full descriptive knowledge of “red”, she lacked the qualitative dimension of what it is to see red. Scientific knowledge is propositional, while experiential or qualitative knowledge defies scientific description and explanation. There thus exist ways of knowing things that can’t be acquired through the scientific method, which is an indication of the limits of science. Adding to this, the philosophical naturalist David Chalmers (1996) argued that the conceivability of the idea of having a twin zombie8—like you in every physical sense, save a lack on consciousness—guarantees its metaphysical possibility, which, in turn, means that there are facts about consciousness and the human dimension that transcend physical reality. How can we explain the fact that your zombie twin shares your complete genetic and physical constitution without sharing your experiences, other than by admitting that physical science does not encompass all the facts 8

It is conceivable that God could have created a world containing zombies.

138

Chapter Nine

about human constitution? The zombie argument thus attacks the physicalist claim that there exist no facts over and above the physical by suggesting that consciousness and experience are not exhausted by physical facts. Explanations are important because they yield understanding. Scientific explanations are carried out by subsuming phenomena under general laws and initial conditions. Science discovers and identifies these general laws of nature but is unable to explain why the laws of nature exist. If science can explain the occurrences of one state of affairs by a previous state of affairs but not why there are states of affairs in the first place, then this reveals a further limitation of science. One of the characteristics of scientific explanation is that it does not refer to acts and intentions, thus omitting irreducibly personal and intentional explanations. However, if personal explanations make sense in everyday life, then there must be a sense in which they have legitimacy and explanatory power. According to Swinburne (2004), there are two kinds of explanation: scientific explanation (in terms of scientific laws) and personal explanation (in terms of personal free, conscious choice). He writes: … in scientific explanation we explain particular phenomena as brought about by prior phenomena in accord with scientific laws (and perhaps also particular phenomena)…[yet] from the very nature of science it cannot explain the highest level laws of all; for they are that by which it explains all other phenomena.9

If we are to explain the fact that there are such laws, then a different type of explanation must be invoked. The forgoing arguments indicate that there are limits to science. There exist both scientific and non-scientific knowledge. The scientific method is good at getting certain things right, but this does not mean that science can explain everything. There are questions that physics simply does not address. Neuroscience will surely be better than any other discipline in answering questions about the brain and how it works, but it cannot explain what happened on September 11, 2001. That’s now part of historical knowledge, as are the results of past experiments and observations in the sciences.

9

Richard Swinburne, ‘A Philosophical defense of the argument from design.’ Accessed August 24, 2016 at http://www.orthodoxytoday.org/articles2/SwinburnDesign.php.

Naturalism, Limits of Science, Case for Non-scientific Knowledge

139

Further, if reality consists only of things recognized by science, would that not make science both the arbiter and jury in its own case? Williamson articulates the fundamental contradiction in the naturalist claim that all truths are discoverable by hard science as follows: “If it is true that all truths are discoverable by hard science, then it is discoverable by hard science that all truths are discoverable by hard science. But it is not discoverable by hard science that all truths are discoverable by hard science” (Williamson 2016, 253). From this, it is only a short step to a modus tollens with an unambiguously clear conclusion: it is false that all truths are discoverable by hard science. When science is broadly construed in the sense of its etymological origin, the Latin scientia, it means any body of accurate and organized knowledge, including philosophy, mathematics, history, theology, natural science, and so forth. The meaning has since seen the term narrow to refer only to the hard sciences, but we still have no good reason to assume that reality has followed suit, and includes only discoveries that are provisional, exclusivist, and amenable to being superseded and discarded in the future.

Naturalism, theism and meaning We know we are here on planet Earth, but who are we and why are we here? And why is there something rather than nothing? These are profound questions that human beings have always grappled with. We tend to transcend ourselves and our immediate environment, and express feelings of mystery, awe, and existential anxiety about the meaning and significance of life. Naturalism is motivated by, or depends on, empiricism—the view that the ultimate source of all our concepts and knowledge is sense experience. A. J. Ayer (1946) and other logical positivists tie meaning to empirical verification and argue that religious beliefs are meaningless or nonsensical because they are not empirically verifiable. Advocates of logical positivism endorse the verification criterion of meaning, asserting that only those propositions that are empirically verifiable have meaning. This criterion excludes all ethical, aesthetic, theological, and incorrigible discourse as meaningless. Meaningful statements consist of mathematical statements (e.g. 2+3=5); logically necessary statements (e.g. “P and not-P cannot both be true”), and factual statements, which can be confirmed by e.g. physics, chemistry, and biology applied to sensory data. Sense and meaning has come to be located alongside experience. The verification criterion of meaning is, however, incoherent in that it is, as Williamson pointed out, itself empirically unverifiable, and therefore meaningless by its own criterion.

140

Chapter Nine

Thomas Nagel (2012) argues that the scientific revolution, notwithstanding its remarkable progress, is predicated on a limiting objective stance, which leaves behind our subjective experiences and the various perspectives from which the world appears to the knowing subject. In Nagel’s words, the physical sciences “depended on subtracting from the physical world as an object of study everything mental—consciousness, meaning, intention or purpose” (Nagel 2016, 255). He contends that, in order to transcend the inherent limits of contemporary science, the current scientific outlook “must expand to include theories capable of explaining the appearance in the universe of mental phenomena and the subjective points of view in which they occur—theories of a different type from any we have seen so far” (Nagel 2016, 256). Possible responses to this may be to, (1) endorse some form of psychophysical reductionism by claiming that mental states are just physical states; they are not an irreducible aspect of reality, or, (2) endorse eliminativism: the mental is not part of reality but some kind of illusion. Second, one may claim that the mental does not require scientific explanation by some novel and expanded conception of the natural order because, (3) the mental can merely be regarded as an inexplicable accident of biological evolution, or, (4) (endorsing a theistic outlook) the mental is explicable, but its explanation pertains to the province of theology. However, Nagel, who says he lacks the sensus divinitatis that compels others to accept the theistic outlook, concludes that someone who finds 1, 2, & 3 rebarbative need not endorse 4, even though some versions of 4 are compatible with the available scientific evidence. Nagel thinks we should seek a deeper and broader form of scientific understanding that includes the mental: “So my speculations about an alternative to physics as a theory of everything do not invoke a transcendent being but tend toward complications to the immanent character of the natural order” (Nagel 2012, 12). One need not endorse Nagel’s atheistic conclusion in order to agree with him that physicalist accounts lack the resources to explain consciousness, reason, and value. Owen Flanagan (2007) is probably right in thinking that if explaining how consciousness emerges from brain activity is “the hard problem” in mind science, then the really hard problem is explaining how meaning is possible in the material world.10 Flanagan argues that meaning is possible 10

Philip Kitcher defends a secular humanist viewpoint that whereas religious belief has been important in the past, evolution from religion is now essential and that without a belief in the transcendent, we can still live lives that are rich and meaningful, because they are thoroughly committed to the pursuit of what is good and beautiful and true. See Kitcher, P. 2014. Life After Faith: The Case for Secular Humanism. New Haven, CT: Yale University Press.

Naturalism, Limits of Science, Case for Non-scientific Knowledge

141

for naturalists, proposing that eudaimonics—a naturalistic investigation into the nature, causes, and conditions of human flourishing that are consistent with science—enables us to live a meaningful life and advance human flourishing without invoking the supernatural. In The Really Hard Problem (2007) Flanagan argues that naturalism is “plenty broad enough to make room for robust conceptions of the sacred, the spiritual, the sublime, and of moral excellence” (Flanagan 2007, 189– 190). Distinguishing between assertive theism and expressive theism, he understands the former to be the kind of theism that “takes certain texts as authoritatively true” and “asserts that certain facts that cannot possibly be known by humans to be true are uncontrovertibly true” (Flanagan 2007, 190). Expressive theism embodies “various extra-mundane impulses, feelings, emotions, and expansive not-humanly-possible love” (Flanagan, 2016). Flanagan endorses the former and rejects the latter. In the end, science and religion are two explanatory worldviews available to us providing different but complementary levels of explanation of reality. They answer different but important questions, and the conflict between science and religion, exaggerated and marketed over the years, is now widely recognized as untenable. Science’s objectifying stance offers access to part of the story of our world and our place in it, leaving out deep issues of subjectivity, meaning, and purpose. Limiting our inquiries about life and reality to scientific inquiry would thus be tantamount to ducking the ultimate questions about meaning, values and purpose, and to do so would be epistemically and existentially irresponsible.

Conclusion We have seen that the natural sciences offer objective facts while leaving out the subjective dimensions of life, such as meaning, purpose, values, and consciousness. The richness and complexity of life cannot be reduced to objective facts. Different ways of knowing offer multiple levels of explanation and perspectives that enrich our understanding of reality and how we make sense of our lives. The method and presuppositions of scientific inquiry indicate that it is not equipped to answer all questions. We have seen that it is untenable to claim that science lends no warrant to religious belief, and there is no gainsaying that there is irreducibly nonscientific knowledge that we should treasure, even if part of this knowledge invokes the transcendent dimensions of life and the explanatory power of the supernatural.

142

Chapter Nine

References Ayer, A.J. 1946. Language, Truth and Logic. New York: Dover Publications, Inc. Baker, L.R. 2013. Naturalism and the First-Person Perspective. New York: Oxford University Press. Chalmers, D. 1996. The Conscious Mind. Oxford: Oxford University Press. —. 2002. “Consciousness and Its Place in Nature”, in Chalmers, D. (ed.), Philosophy of Mind: Classical and Contemporary Readings. Oxford: Oxford University Press, 247-72. De Caro, M. 2015. “Naturalism and the First-Person Perspective”, Philosophical Review 124, 1, 156–158. De Caro, M. and Macarthur, D. (eds.), 2010. Naturalism and Normativity. New York: Columbia University Press. De Cruz, H. Prosblogion, http://prosblogion.ektopos.com/2016/06/30/philosophers-and-theirreligious-practices-part-22-comparative-philosophy-the-unforcedmoral-consensus-and-the-charms-of-expressive-theism/. Jackson, F. 1982. “Epiphenomenal Qualia”, Philosophical Quarterly 32, 127–36. —. 1986. “What Mary Didn’t Know”, Journal of Philosophy 83, 291–95. Flanagan, O. 2007. The Really Hard Problem: Meaning in a Material World. Cambridge, MA: The MIT Press. Hume, D. 1999. An Inquiry Concerning Human Understanding. Beauchamp, T.L. (ed) New York: Oxford University Press. Kitcher, P. 1992. “The Naturalists Return”, Philosophical Review, 101, 1, 53–114. —. 2014. Life After Faith: The Case for Secular Humanism. New Haven, CT: Yale University Press. Korblith, H. 1994. “Naturalism: Both Metaphysical and Epistemological”, in Philosophical Naturalism, French, P.A., Uehling, T.E. & Wettstein, H.K. (eds), 19, 39–52. Notre Dame: Notre Dame Press. McGrath, A. 2015. The Big Question: Why We Can’t Stop Talking about Science, Faith and God. New York: St. Martin’s Press. Nagel, T. 2012. Mind and Cosmos: Why the Materialist Neo-Darwinian Conception of Nature Is Almost Certainly False. Oxford: Oxford University Press. Papineau, D. 2015, “Naturalism”, In The Stanford Encyclopedia of Philosophy, Zalta, E.N (ed). URL = http://plato.stanford.edu/archives/fall2015/entries/naturalism/>.

Naturalism, Limits of Science, Case for Non-scientific Knowledge

143

Peterson, M. and Michael R. 2017. Science, Evolution, and Religion: A Debate About Atheism and Theism. New York: Oxford University Press. Price, H. 2004. “Naturalism without Representationalism”, in Naturalism in Question, De Caro, M & Macarthur, D. (eds), 71–88. Cambridge, MA: Harvard University Press. Rosenberg, A. 2016. “Why I am a Naturalist”, in The Stone Reader, Catapano, P. and Critchley, S, (eds), New York: Liveright Publishing Corporation, 247–250. —. 2009. “The Disenchanted Naturalist’s Guide to Reality”, in On the Human. http://onthehuman/2009/11/the-disenchanted-naturalistsguide-to-reality/. —. 2001a. The Atheist’s Guide to Reality: Enjoying Life without Illusions. New York: W.W. Norton. —. 2013. “Disenchanted Naturalism”, in The Metaphysics of Evolutionary Contemporary Philosophical Naturalism and Its Implications, Bashour, B. & Muller, H.D. (eds), New York: Routledge. Russell, B. 1953. Religion and Science. London: Oxford University Press. —. 1992 [1948]. Human Knowledge: Its Scope and Limits. London: Routledge. Swinburne, R. 2004. The Existence of God. New York: Oxford University Press, 2nd edition. Williamson, T. 2016. “On Ducking Challenges to Naturalism”, in The Stone Reader, Catapano, P. and Critchley, S. (eds), New York: Liveright Publishing Corporation, 251-54 —. 2016. “What is Naturalism?”, in The Stone Reader, Catapano, P. and Critchley, S. (eds), New York: Liveright Publishing Corporation. Van Fraassen, B. C. 2002. The Empirical Stance. New Haven, CT: Yale University Press.

CHAPTER TEN THE LIMITS TO SCIENCE: A BARTHIAN APPROACH PHILIP CHAPMAN

My own reasons for saying that such limits exist derive from the twentieth century Swiss Reformed theologian Karl Barth. In this chapter, I briefly summarize Barth’s approach and sketch those aspects of his thought that touch on our theme including his attitude to science. Some comments on the scientific task follow. I shall then explain how I think that the course of scientific investigation in biology, and especially cosmology and physics over the last hundred years is consistent with Barth's point of view.

Some of Barth's principle themes Barth's major work is the “Church Dogmatics” (Barth 1956-77). It was unfinished at his death in 1968, but it occupies more than half a metre of shelf space. In it he expounded what he believed the Christian Church should be preaching. Legend has it that he once summarised its content in the words of the old Sunday school chorus: “Jesus loves me, this I know, for the Bible tells me so”. Be that as it may, he held, controversially, that the witness to Jesus Christ set out in Holy Scripture is our only genuine source of knowledge of God. Barth's stance has often been called into question. It differs from the widely held Christian position, known as “Natural Theology”, which holds that such knowledge is also revealed through God's creation, and that such knowledge can and should be upheld by reasoned argument (Mangina 2004, 59-63). The discussion about natural theology is still going on. However, without settling that question, I maintain in this paper that to argue on the basis of science about the existence or non-existence of God, or the qualities that God might possess, is inadmissible. That is because even well based current physics with usable practical consequences is provisional, incomplete and full of

The Limits to Science: A Barthian Approach

145

unanswered questions. It cannot be depended upon as a source of faith or theological knowledge. As I shall explain below, in spite of his opposition to some of the prominent features of “liberal” and indeed Catholic theology, Barth was not a fundamentalist. Moreover, his theology is Nicene in character and he acknowledges debts to the great teachers of the Church, both Roman Catholic and Protestant, including Augustine, Aquinas, Calvin and Luther. In Barth's thought God's own self is the primary and originating reality. According to Barth that self cannot be the object of study or investigation: “God is known by God and by God alone” (Barth 1956-77, 2.1 179). Consequently we can know nothing of God unless God explains God's self to us. Revelation is the event of that explanation. In Barth’s understanding, revelation takes place in and through Jesus Christ, God's Word; He is the pioneer human person. The human authors of Holy Scripture bear witness to that Word who is infallible, though the human witnesses of the Word are not (Barth 1975, 1.1, 112-115). They understand themselves as seeing God at work in human lives, but their judgements can be mistaken. Christian theology, as exemplified by Barth, makes use of two millennia of human commentary and interpretation, along with the insights of contemporaries, as it tries to declare the witness for us in our time in what he called “The strange new world of the Bible” (Barth 1935, 28-50). Theology's primary task is thus to discern, to hear as it were, God's speech that grounds Scripture. Scripture itself knows nothing of science as now practised. Its concern is God's active relationship to the human race as part of that secondary reality which is the created world. The communication of its witness to us is constrained by the limits of human language. Barth claimed not to be concerned with philosophy. However he is obliged to say something about what we can know and how we know it. Crucially for our enquiry he made a similar distinction to that of Kant, drawing a line between reality, that is to say things as they are in themselves and, on the other hand phenomena, which are the appearances available to human senses. Kant states that “The question, ‘What are the objects considered as things in themselves?’ remains unanswerable even after the most thorough knowledge of appearances” (Kant 1991 A42, B5960). So for Barth himself, phenomena (which must include those mediated by scientific instruments and described mathematically or otherwise) are to be understood as symptoms of the genuine reality known only to God.

146

Chapter Ten

Barth and science Barth values all “sciences”, that is to say, in his context, academic disciplines are sacred and secular. Each has its own methods and none can dictate to theology how it should pursue its work. From some of them theology can learn “praiseworthy fidelity” (Barth 1956-77, I.1 7-8). Barth would thus have agreed with philosophers of science like Feyerabend (2011), Chalmers (1999, 62) and, from a different standpoint, Blackburn (2006, 175-190) who hold that there is no universal method for acquiring scientific understanding. Consequently, each science adopts procedures appropriate to the object under study. Theories or laws follow from data acquired in such appropriate ways. They are to be judged not in terms of method but by consistent applicability to what we know of the physical world. That can be understood as providing a demarcation criterion between science and other areas of study. For Barth the findings are “relative”. That is to say we never understand more than a part, nothing we might call absolute truth is available. It seems to me that he is justified in thinking that science possesses only a measure of objective truth. He says that a scientific theory is a “working hypothesis” and thus it is always open to correction. Consistent with the observed phenomena it offers a provisional explanation of how the physical created world functions. Barth further contends that even if it is successful in resolving an immediate puzzle, such a theory cannot answer what he calls “ultimate questions”. The context here suggests that he is thinking about matters of existence and final purpose, which are not susceptible to empirical tests (Barth 1961, IV3, 146-7). A consequence would be that eminence in a particular field does not of itself give a scientist any special authority to pronounce on the existence or non-existence of God.

How Science moves forward Science searches for explanations of phenomena and proposes to test them by experiment and observation. Where the tests are successful, the resulting working hypothesis can be applied to further aspects of the world of sense experience or what can be deduced from it. Scientific propositions are held to be true in some sense if observation of the world justifies them. Different experimenters, in other times and places, must find unambiguously corresponding results. But the story of science contains many twists and turns and even in modernity science has not displayed a tidy or invariable path from empirical results to a general theory considered to be justified

The Limits to Science: A Barthian Approach

147

(Olby et al. 1996). The propositions science arrives at, by whatever method, at least attempt to set in order what we know of the world. Gaps and anomalies in that suggested order point to further investigations. That is how science moves forward.

Relative nature of scientific findings The actual state of science supports Barth's contention that any completeness ascribed to it is no more than relative. I offer two important instances. Terrestrial biological science has described fewer than 2 million living species out of a likely total of 10 to 15 million. That range may represent a mere 2-4% of what has existed since life began on this planet (Benton 2009, 84-85). We simply do not know how many other places there are in our universe where life could be or is supported. But even within the limited domain of the physical planet we inhabit with all its forms of life, a total inventory of items would not tell more than a tiny fraction of what we would like to know. Completeness would require knowledge of relationships and interactions between complex systems as well as among their components. Complexity is a subject in itself (Mitchell 2009). What and where is the agency in a complex system such as a living organism? Does causation operate downward as well as upward? Living things as individuals and communities increase the complication by several orders of magnitude. What does complexity have to do with consciousness? Science with all its power barely skims the surface even of the familiar physical world. Cosmology is my second instance of incompleteness. According to one estimate there may be at least 1022 stars in the observable part of our universe (Carroll 2016, 50). In addition, there are an unknown number whose light has not had time to reach us since the epoch of the big bang. But speculations of some scientists further envisage an infinite number of universes possibly each with their own physical constants and laws of nature (Barrow 2002, 275-292). Brian Greene offers a table of nine different categories of current speculation about the structure of the whole cosmos within which our own partially knowable universe may be contained. It is beyond our scope to examine them, but the last speculation asserts what he calls “the principle of fecundity” (Greene 2011, 309): a proposal that everything that can happen does happen. The principle is already widely held within the predominant version of quantum mechanics (Cox and Forshaw 2011), but Greene does not confine his consideration to

148

Chapter Ten

the laws of nature as presently understood. Once that limitation is dispensed with we do not know where to stop. A second requirement for a body of knowledge about the physical world is that it should be coherent. It might seem obvious that the various propositions stated in empirical science should agree with each other. We are not here talking about Hegelian dialectic or even about some aspects of Barth's own theology, where contradictions, real or apparent are in the realm of ideas. Indeed if we examine our own notions of God we are sure to find inconsistencies. However we are surprised to hear of accepted but formally contradictory propositions concerning the nature or whereabouts of a physical object. It happens that established scientific theories that do vital work in the world give rise to contradictions. This increases suspicion that we do not have access to reality but only to appearances, especially in some foundational matters we cannot get beyond the symptoms. I offer two well-known interlinking examples from physics. (1) Quantum theory and General relativity together form the basis of physical science. Quantum theory determines the design and functioning of the ubiquitous microchips in electronic devices. It now proposes to extend its grasp to fundamental biological processes (AlKhalili and McFadden, 2014). General relativity enters into the calculation of position for satellite navigation used by ships, aircraft, drones, terrestrial transport and increasingly people navigating their way on foot in city or in the countryside. Thus in a few decades quantum mechanical and relativistic applications have transformed computation, communication, travel and consequently social interaction. But their theoretical bases have not yet been reconciled with each other. In quantum theory matter-energy is grainy, for Einstein it is continuous (Rovelli 2016, 40; Penrose 2004, 934-955). String theory with its 10 or 11 dimensions is one attempt to bring them together (Randall 2005). Another is loop quantum gravity (Rovelli 2016, 40-42). Those two are reasoned mathematical speculations which also enter into the cosmological speculations mentioned above. However, neither they nor their competitors can as yet predict any measurable or even observable consequences. (2) Quantum theory itself is indispensable to physics and in daily use for practical calculations. Paradoxically it implies either that elementary particles can be in two places at once or that the universe, and every item it contains, are repeatedly splitting into an infinite

The Limits to Science: A Barthian Approach

149

number of diverging entities (Alkhalili 2003, 1-15 and 134-7). Yet, quantum theory very accurately predicts the experimentally measured physical properties of each item, 16 in all, including electrons and photons, in what is collectively referred to as the “Standard Model”. Scientists believe now that the model is likely to be a complete catalogue of the fundamental particles that make up the world we are familiar with. Beyond those, the detection in 2012 of the Higgs Boson, which is believed to be responsible for mass, leads physicists to think that they are on their way to a complete understanding of the matter that makes up our own universe. However, the experimentally determined mass of the Higgs itself does not match expectations. Further, the status of certain neutrinos remains puzzling and the model does not account for dark matter which comprises a large proportion, on one estimate 85%, of the mass in the cosmos (Billings 2016, 910). Dark matter is known by its gravitational effects, but it is invisible and transparent without action on light or other electromagnetic radiation. There are many other examples of unresolved questions in science. Some certainly will be resolved-but we don't know which. In the two cases above we cannot yet even imagine what a solution would look like.

Scientific propositions as knowledge Indeed there is no absolute certainty about any of the propositions of natural science. Neither is there agreement that a number can be arrived at strictly representing the probability that a given empirical general proposition is true1 (Chalmers 1999, 191-192; Cushing 1998, 361). Probability is always on the basis of evidence. So we have the problem of “unknown unknowns” (Rumsfeld 2007)2. Further, theories, however useful, arrived at through the natural sciences, are liable to correction. Even when we believe we have solid empirical results there are still problems. What is the logical relation between observation statements and general laws? How it is that general laws come to be formulated modified or replaced? However, the work of science does not wait for its philosophical difficulties to be resolved. Scientists must apply other criteria than those of logical purity. 2

‘Reports that say that something hasn't happened are always interesting to me, because as we know, there are ‘known knowns’; there are things we know we know. We also know there are ‘known unknowns’; that is to say we know there are some things we do not know. But there are also ‘unknown unknowns’ - the ones we don't know we don't know.’ Donald Rumsfeld quoted by the BBC news website 30/11/2007, see also www.c-span.org, 12/02/02.

150

Chapter Ten

Empirical science is successful because it contains within itself the devices for correcting mistakes (Brooke and Cantor 1998, 45). That is to say that even a well-established scientific theory remains a working hypothesis. Recognition of that leads to a fallibilism like that of Peirce, which insists not that scientific claims are false but rather that they are tentative (Gallie 1952, 72, 106-107; Popper 1968, 27 and passim). The above factors are at the heart of our attempts to understand the physical world. Perhaps our evolutionary origin forever constrains human intellectual capacity? We must live with what we cannot yet understand. From a Darwinian perspective Stephen Pinker acknowledges that state of affairs: We are so-so scientists....our brains were shaped for fitness, not for truth. Sometimes the truth is adaptive, but sometimes it is not. Conflicts of interest are inherent in the human condition....and we are apt to want our version of the truth, rather than the truth itself, to prevail (Pinker 1998, 305).

Pinker is not on his own in that view: according to the biologist Lewis Wolpert “scientists are concerned, not with absolute truth, but with theories that provide understanding of the phenomena involved” (Wolpert 1992, 103). It does seem plain that scientists should be aspiring to something called “truth”. However, we cannot judge the worth of a scientific theory by requiring “absolute truth”. We must judge by its conformity to our present knowledge of the physical world. We might speculate beyond that but the most immediate use of scientific propositions is in their organization and explanation of what the senses give us. Usually that is via apparatus designed by people according to pre-existing theories. Natural science, it would seem, cannot outgrow its roots in finite humanity. In any case natural science cannot be, in any fundamental sense, “objective”. The world as seen through human eyes is one constructed by the eye and the brain from data selected by them (Frith 2007). In the best case, human scientific understanding of the empirical world is an elaborately agreed perspective constructed from the evidence supplied by many observers. But any or all of them may be mistaken. Even if humans thought we were approaching finality, we cannot know how much there is that we do not know. The twenty first-century physicist, several revolutions later, is astonished that his or her late nineteenth-century counterpart might have thought his science nearly complete3. Since then

3

‘There is nothing new to be discovered in physics now. All that remains is more and more precise measurement’ William Thomson, Lord Kelvin (1824-1907),

The Limits to Science: A Barthian Approach

151

quantum theory and relativity have revolutionised physics. The development of both continues. The evidence above suggests that there may be no finality in empirical science. If we follow Barth we learn that theology is in the same situation since finality is forbidden by its subject matter, the nature and acts of the God whom God alone, by God's own free decision, reveals. This is not scepticism about the existence of the real world, but doubt about human ability to know it. Thus the sovereignty of natural science is not as clear as an apologist for it might hope: Barth is right, pragmatically and strategically not to base any conclusions on scientific findings.

Conclusion In this paper I have assumed Barth's theological starting point that the one primary reality is creator God's own self, and that this gives a limit to science though we may not know in advance where that limit will be. On this view the created world is the secondary reality. Natural science investigates the physical phenomena arising in it. As I have noted above, it has learned to explore and even manipulate the physical world, using for instance quantum theory, to an extent beyond what it understands. Consequently a certain modesty is appropriate for the scientific community. Useful though they may be, scientific results are always provisional since comprehending the nature of things in themselves, is not possible for us finite creatures. Our knowledge is relative to our human condition. We cannot see with God's eye. Theology is in a different case but nonetheless limited and, according to Barth, provisional. It must hear and expound for its own time what the eternal God may be saying through the witness of Scripture alone. That human task, also, will never be finished. Given those limitations, Barth thinks it is impossible to combine theological with scientific and other secular knowledge to bring about the total world view, he calls a “cosmology” . That is consistent with his prohibition of natural theology and the picture he offers us of himself being laughed at by the angels as he trundles a barrow load of his books up the road to heaven (Mangina 2004, 25). Perhaps, already knowing how truly inadequate our knowledge is, this is our best preparation for receiving what we shall know when we come “face to face” (1 Corinthians. 13:12).

Address to a meeting of physicists of the British Association for the Advancement of Science, 1900. Cited in Randall 2005, 85.

152

Chapter Ten

References Al-Khalili, J. 2003. Quantum-A Guide for the Perplexed. London: Weidenfeld and Nicolson. Al-Khalili, J. and Mcfadden, J. 2014. Life on the Edge: The Coming of Age of Quantum Biology. London: Transworld. Barth, K. 1935. The Word of God and the Word of Man. Horton, D. (trans), London. Hodder and Stoughton —. 1956-1977. Church Dogmatics, Bromiley, G.W. and Torrance, T.F. (eds), Edinburgh: T&T Clark. —. 1958. Dogmatics in Outline, Thompson, G.T. (trans.), London: S.C.M. Barrow, J.D. 2002. The Constants of Nature. London: Vintage. Billings, L. 2016. In the Dark about Dark Matter, Scientific American, 315, 4, 8-10 Blackburn, S. 2006. Truth. London: Penguin. Brooke, J.H. and Cantor, G.W. 1998. Reconstructing Nature, The Engagement of Science and Religion. Edinburgh: T&T Clark. Carroll, S. 2016. The Big Picture. London: One World. Chalmers ,A.F. 1999. What is this thing called Science? Milton Keynes: Open University. Cox, B. and Forshaw, J. 2011. The Quantum Universe: Everything that can happen does happen. London: Allen Lane. Cushing, J.T. 1998. Philosophical Concepts in Physics: The Historical Relation Between Philosophy and Scientific Theories. Cambridge: Cambridge University Press. Frith, C.D. 2007. Making up the Mind: How the Brain Creates Our Mental World. Oxford: Blackwell. Feyerabend, P. 2011. Against Method. London: Verso. Gallie, W. B. 1952. Peirce and Pragmatism .London: Penguin. Greene, B. 2011. The Hidden Reality. London: Allen Lane. Kant, I. 1991. Critique of Pure Reason. Politis, V. (ed.), London: Everyman. Körner, S. 1968. The Philosophy of Mathematics. London: Hutchinson. Lane, N. 2015. The Vital Question-Why is Life the Way it is? London: Profile. Mangina, J.L. 2004. Karl Barth. Aldershot: Ashgate. Mitchell, M. 2009. Complexity: A Guided Tour. Oxford: Oxford University Press. Olby, R.C., Cantor, G.N., Christie, J.R.R. and Hodge, M.J.S. 1996. Companion to the History of Modern Science. London: Routledge. Pinker, S. 1998. How the Mind Works. London: Penguin.

The Limits to Science: A Barthian Approach

153

Penrose, R. 2004. The Road to Reality: A Complete Guide to the Laws of the Universe. London: Jonathan Cape. Popper, K.R. 1968. The Logic of Scientific Discovery. London: Hutchinson. Randall, L. 2005. Warped Passages: Unravelling the Universe’s Hidden Dimensions. London: Penguin. Rovelli, C. 2016. Seven Brief Lessons on Physics. London: Penguin. Wolpert, L. 1992. The Unnatural Nature of Science. London: Faber and Faber.

CHAPTER ELEVEN WAYS TO DEVELOP STUDENTS’ APPRECIATION OF THE POWER AND LIMITATIONS OF SCIENCE BERRY BILLINGSLEY AND MEHDI NASSAJI

1. Introduction This chapter describes the rationale for and impact of a workshop for teenagers called, “Can a robot hear?” The workshop was designed to help students aged 14-16 to progress in their appreciation of the power, relevance and limitations of science when addressing big questions in multidisciplinary arenas. Robots that seem to be context-aware and autonomous are becoming increasingly present in homes, hospitals, care homes and other institutions. Perhaps unsurprisingly, given the pace of change, there is much research to do to discover how schools can best prepare young people for the questions they are likely to encounter as the co-workers, consumers and inventors of increasingly humanlike machines. The design of this workshop was prompted by research conducted by LASAR (Learning about Science and Religion). LASAR is a research and dissemination enterprise which seeks to understand how questions bridging science and religion are managed at school. For our team, the opportunity to develop a workshop on the advance of robotics stood out as one that can capitalise on the research we already have underway. Firstly the prospect of finding out about humanoid robots is one that appeals to most teenagers (Billingsley 2016); secondly the theme raises philosophical questions such as whether humans and artificial entities are on a continuum or whether a person will always be a step beyond anything a robot can become; thirdly there are cognitive, sociological and pedagogical questions to address relating to how young people reason

Ways to Develop Students’ Appreciation

155

about this question and what opportunities they have in school to develop more scholarly responses. Calls for schools to give young people more teaching to help them with multidisciplinary and real world contexts have been made for some time. Teaching through subjects is valued within education as a way to immerse students into each of the disciplines. Today, however, subject compartmentalisation has become entrenched and pervasive across the stages of education – each subject tends to work as a silo, with its own curriculum, text books, subject examinations, subject-specific teacher training and – in secondary schools – its own specialist teachers and classrooms. These boundaries are detected by students who then constrain their questions and thinking in the way they suppose their teacher prefers (Billingsley, Brock, Taber, & Riga 2016). Thus and for example, David (not his real name) was one of many 14 year old students who explained that students hold back questions they perceive as “off-topic” and/or culturally sensitive (Billingsley, Taber, Riga, & Newdick 2012). With this in mind, we anticipated that a series of workshops on the theme of humanoid robotics could provide ways to raise teenagers' appreciation of what it means to be human and deepen their understanding of how science, religion and other disciplines inform our thinking about these questions. For example the puzzle of whether and why a person has a capacity for self-determination is longstanding. Today we are not only searching for answers for ourselves to satisfy our philosophical curiosity but are also trying to work out what we should be saying about the progress of increasingly humanlike machines. If having free will and a moral code and a sense of responsibility are aspects of humans that we can understand mechanistically then arguably we can build robots that also have these capacities. Alongside these intriguing questions, we envisaged that these workshops would provide a space in which teachers could draw students' attention to careers in engineering and in the caring professions and the notion that there are careers that would focus on both.

2. Aims and development of the workshop The robotics workshops we are designing form part of a research journey that has employed surveys, focus group discussions, interviews, and intervention studies with more than 4000 school students. The research has identified a number of gaps in the education children currently receive to do with developing students’ appreciation of how to manage so-called Big Questions – which is to say – questions which most scholars agree cannot currently or potentially ever be resolved using

156

Chapter Eleven

science alone. For the first workshop in the series we focus on the aim of developing students’ appreciation of the nature, power and limitations of science. The key ideas we address in the workshop emerged from some of our exploratory focus group and interview studies (Billingsley 2013, 2016; Billingsley et al. 2016). We also consult regularly with philosophers and theologians, biologists, engineers and philosophers and, on occasions, our expert scholars have joined us to run workshops in schools. In all, developing and fine tuning the workshop has taken place over a period of about a year with our team delivering the workshop many times with small changes each time. These changes were partly to become more effective at achieving the objectives and partly on the basis of student and teacher feedback. The primary aim of the workshop is to develop ideas and pedagogies which teachers can use to introduce students to the ideas that (a) some questions are more amenable to science than others; and (b) that some questions are more metaphysically sensitive than others. The rationale for these objectives is as follows. This period of schooling (for students age 15-16) is the last stage in which students are required to study both science and non-science subjects and so the last opportunity to ensure that all students experience the specified teaching. In our proposed schema for progression, these two objectives provide, we argue, all students with some essential and key ideas about the ways in which scholars approach big questions and the power and limitations of science within those approaches. We have noted when designing our schema for progression that there is already an objective in the National science curriculum in England for this age group which says that students should develop an appreciation of “the power and limitations of science” (DfE, 2014, p. 5). This corresponds to the objective in our schema that students should appreciate that some questions are more amenable to science than others. We recommend (however) that teaching relating to this objective begins in lower secondary school. We deem it to be important for students in this final stage of statutory education to also know that there is a diversity of scientific opinion on the extent to which so-called Big Questions can eventually be resolved scientifically. For this reason we have added the objective that “some questions are more metaphysically sensitive than others”. Given that the aims and approach are novel, we have been particularly pleased to find when we review feedback from teachers and students that participants seem to be picking up the objectives we are endeavoring to cover. For example we ran the day at a school event where there were many workshops on different themes with different presenters.

Ways to Develop Students’ Appreciation

157

The event organizers picked out one piece of feedback for each workshop. For our workshop the feedback was: That depending on your metaphysical position, your opinions and furthermore answers to certain questions vary. It was awesome, my favourite by far. Was really interesting and the class had really good discussions. It opened my eyes a lot to the opinions of others and as to what metaphysics is about.

In addition to testing the efficacy of the workshop itself, we are also keen to develop a workshop that can be carried out by staff in the school and so incorporated more widely into schools’ curriculum planning. When we run the workshop in schools, we invite teachers from the relevant subjects (RE, computer science and science) to assist and/or watch. We have noticed in so doing, the value of interdepartmental collaboration and also the importance of recognising that the expertise and pedagogies needed to teach this workshop are likely to go beyond the expertise and experience that is usually found with one subject teacher. In particular our experience has been that RE teachers frequently have the expertise needed to help students with the philosophical aspects of the workshop while computer science teachers frequently help when we move to technology-related questions. Thus, for example, an RE teacher worked with students to help them to understand that the question of “how many neurons are in the brain” is less metaphysically sensitive than “can a robot have a mind?”. In another workshop, a computer science teacher reinforced a teaching point about the greater complexity of attempting to build a robot that can hear over a robot that can respond to sound. A third aim underpinning the research is to look at whether interdisciplinary workshops such as this one can increase the size and diversity of the cohort of students who feel attracted to the idea of a science-related career. This possibility is motivated by the circumspection that young people who enjoy multidisciplinary ways of thinking may feel more positively towards science if they are given examples of the ways in which science can inform our thinking about a big question without requiring a commitment to the stance that science is sufficient on its own. The workshops have provided some support for this idea, with teachers often commenting that the workshop is particularly well received by girls who (as our own and other research indicates) are more likely than boys to favour teaching which makes links between different subjects. As a result of many cycles of trials and we feel we are now in a position to offer teachers a workshop with pedagogies and assessment tools that they can use themselves to develop and assess students’

158

Chapter Eleven

capacities to reason about the power and limitations of science. The next section of this chapter sets out the details of the workshop including the overview, objectives, activities and the concluding remarks from the workshop. We will also introduce some of the assessment tools that we use to examine to what extent the workshop attendees have met the teaching objectives. Finally in the chapter we set out and discuss the findings from the assessment carried out by students.

3. Workshop outline-Can a robot hear? a. Overview In the opening of the workshop, students are presented with a headline which claims that a robot has been invented which can hear. They are asked to consider what criteria they feel should be used to decide whether or not the robot really can hear. Is it sufficient if the robot responds to a sound? Should the robot also demonstrate a level of understanding and if so how and what level? Finally students compare the difficulties of addressing and assessing each of these challenges-a robot that can respond to sound / a robot that can hear; a robot that can understand / a robot that appears to understand. The workshop helps students find the critical questions to ask when reporters use words associated with human experiences and capacities when talking about technology. It also introduces the idea that some questions are more amenable to science than others.

b. Objectives The objectives are for students: x to be able to critically analyse the language used to describe human/robot behaviours, to draw attention to linguistic distinctions that enable further discussions about the progress of robotics towards humanlike machines (Ep LO1); x to appreciate that some questions are more amenable to scientific methods than others (Ep LO2); x to appreciate that some questions are more metaphysically sensitive that others (Ep LO3); x to appreciate that the scientific community is diverse and have a range of metaphysical positions on whether human experience and

Ways to Develop Students’ Appreciation

159

behaviour can be reduced to properties that are amenable to science (Ep LO4).

c. Session structure Introducing the terms: “bridging questions” and “interdisciplinary questions” The workshop begins by explaining that different subjects such as history, economics, science, philosophy, art and computer science, investigate different questions, and students may have particular interest in one or more of them. (The facilitator may ask students, “who likes history? Who likes science? Who likes engineering” and so on). Then it is explained that there are some questions that students might be interested in, but that could not be investigated by any one of these disciplines alone. These are called bridging questions or interdisciplinary questions. (The facilitator asks students if they could suggest some interdisciplinary questions.) Then the facilitator explains that in the workshop they are going to think about some questions that robot engineers in particular are interested in, and that these questions are better understood and possibly better answered if we bring several disciplines into the discussion. Workshop activity: can a robot hear? The facilitator asks students to give their opinions about whether we can design and build a robot that can hear. There is a work sheet with these two questions: 1) Suppose you were designing a robot that can hear – how would you address that challenge? 2) How would the robot demonstrate that it can hear (if it can hear)? Then the facilitator demonstrates a robot that starts and stops moving on the sound of a clap and again asks the question, “Can this robot hear?” The aim is to help students consider whether there is a distinction between “hearing” and “responding to sound”. Students are asked whether there is a difference between a person hearing and a robot hearing. Pupils may suggest that “understanding” or “emotions” are involved in the person hearing. The list of the differences between a robot hearing and a person hearing is written on the board by the facilitator. (Participants may suggest

160

Chapter Eleven

that hearing for a person is more complicated than just a responding to a clap. In that case, the facilitator may ask what about a Siri: Does Siri hear what the user says and if this is hearing how this is different from a person hearing?) During the discussion among students, the facilitator should try to highlight two different answers that students may give to the question of whether hearing is the same as responding to sound. One view is that “hearing and responding to sound are the same”, the other is that “a robot responding to sound is different from a human being hearing” (the facilitator refers back to this distinction later). Workshop activity: the difference between a humanoid robot and a human being The facilitator explains what a humanoid robot is (perhaps showing some interesting photos or video clips). She/he asks the group of students to do a simple task (such as raising their hands a couple of times). Then the facilitator asks the students to imagine that there is a group of humanoid robots in one room and a group of students in another room, and that both groups have been asked to follow the same instruction (raising their hand). The facilitator asks, what are the similarities and what are the differences between these groups and what they are doing. The point is to discuss the difference of “rule following” between programmed humanoid robots and the human beings. These are the questions for thinking and discussion: o Would the robot get tired if we asked them to do this for many, many times? Would that be a difference between a humanoid robot and a human being? (If students say robots never get tired in the way that a human being gets tired, the facilitator may ask them to list the signs of tiredness in humans and say, “How about if I give this list to an engineer and ask for a group of robots that show all these signs after repeating the job for a certain number of times? Does this reduce or even fill in the gap between the robots and the human beings?”); o Do you think that any of the humans or robots or both would start to get cross if they are asked to do this for several times? (The facilitator can then say that the engineers will be asked to address this gap in their design.); o Do you think that any of the humans or robots or both would refuse to follow the instruction after a while? (The facilitator can again say that this will be addressed in the design of the robots.); o Does the robot group understand what they are doing?

Ways to Develop Students’ Appreciation

161

The facilitator broadens the question and asks whether, in general, engineers can fill the gap between humanoid robots and human beings? Then the facilitator explains that before answering this question she needs to introduce some terms. Metaphysically sensitive questions The facilitator introduces the term “metaphysically sensitive question” by saying that whether scientists can fill the gap between a humanoid robot and a human being is an example of a “metaphysically sensitive question”. This is because the answer might be different based on our understanding of “what it means to be human”. To explain this further, the facilitator gives a definition of “metaphysics”: Metaphysics is the study of any of the most fundamental concepts and beliefs, on which many other concepts and beliefs rest. Or, metaphysics discusses the question of what is real.

The facilitator then goes back to the question of hearing and the list of the differences between hearing and responding to sound and explains that the question of “can a robot hear?” is also a metaphysically sensitive question because how we answer depends on what we mean by “a human being hearing”. If a person is just molecules and atoms and an assembly of mechanisms then we are well on the way to having a robot that can behave like a human. If however, hearing is more than a mechanical process and also requires some kind of personal experience and subjective response then this will be more difficult to achieve in a robot and, further, we may never know for sure if a robot is merely displaying the signs of hearing or if there is truly hearing happening too. In order to help students better understand the idea of “metaphysically sensitive” questions, the facilitator then gives them some examples of less metaphysically sensitive questions: • • • •

Does this robot respond to sound? How many batteries does this robot need? What is the maximum weight that this robot can pick up? How many neurons are in an adult brain?

The facilitator highlights that if a question is not metaphysically sensitive then it is likely that scientists will agree that the question is amenable to

162

Chapter Eleven

science. In contrast, scientists who hold different metaphysical positions are likely to disagree over how to answer a metaphysically sensitive question – and in particular may not agree about the amenability of the question to science. Workshop activity: ordering questions from amenable to science to more metaphysically sensitive The facilitator gives 8 cards with questions that are more or less amenable to science on each and asks them to categorise them into (a) very amenable to science; (b) partly amenable to science; (c) not very amenable to science – but there may be smaller scientific questions that we can usefully explore. A worksheet for this and other activities can be found on the LASAR website – as LASARcentre.com and also our newly developed site for teachers looking to develop students’ epistemic insight at epistemicinsight.com. Workshop concluding remarks The facilitator sums up by highlighting the following points from the workshop: 1. In this workshop, we discussed what an “interdisciplinary question is” and how science and engineering interact with other disciplines such as psychology and philosophy. 2. In this workshop, we discussed the meaning of the word “hearing” and the importance of being clear about what we mean when we use it to describe what a robot and a person do when they respond to sound. 3. We discussed whether it would be possible in the future for science and engineering to fill the gap between a humanoid and a human being. We also asked, if not, what might be special about being human that cannot be produced by science and engineering? 4. We discussed what a “metaphysical view” is and what it means that a question is “metaphysically sensitive”. 5. We discussed that the question of hearing is a “metaphysically sensitive question”, we also discussed other examples of metaphysically sensitive questions. 6. We discussed that scientists cannot fully answer a metaphysically sensitive question.

Ways to Develop Students’ Appreciation

163

7. We discussed that scientists with different metaphysical views may not agree on metaphysically sensitive questions.

4. Findings The data collected from the survey conducted before the workshop revealed that two thirds of students believe that “One day there will be robots that are as intelligent as humans” while about a third agree that “One day there will be robots that have minds”. Only 6% agree that “One day there will be robots that have souls”. In the pre-workshop and post-workshop survey we asked students, “Which of the following questions is more metaphysically sensitive ‘Why does my pen still exist?’, or ‘Why does my pen fall to the ground?’”. They also had the option to choose, “I don’t understand the question”. In the pre-workshop survey only one in three of the respondents gave the correct answer, while in the post workshop survey about 70% gave the correct answer. Similarly in another question in the pre-workshop and post-workshop survey we asked students, “Which of the following questions is more metaphysically sensitive: ‘Can a robot talk?’, or ‘Can a robot make sounds?’”. Again they had the option to choose, “I don’t understand the question”. In the pre-workshop survey 62% gave the correct answer, while in the post workshop 88% gave the expected answer. In the post-workshop survey we also asked students how their thinking has changed and below is a sample of comments from students attended the workshop: x I have questioned the difference between hearing and responding which is particularly significant in terms of understanding of robot. x I was made to think about inter-disciplinary questions and about hard questions. x I can appreciate the difference between hearing and responding and it has developed my ethical views about robots. x I am thinking more metaphysical. Science is not all about grades. x Now I think there is a way bigger question and meaning to think about with robots and humans. x My thinking has changed by me now knowing what counts as hearing and listening compared to responding to sound. x It has enabled me to think about the source of our mental thoughts and if it is possible to implement senses and the power of thoughts into machinery/robots

164

Chapter Eleven

5. Conclusion In this chapter, we described a workshop on robotics designed for secondary students. The workshop aimed to help students develop their epistemic insight and particularly their appreciation that some questions are more amenable to science while others are more metaphysically sensitive. The other learning objective of the workshop was to draw attention to attributions such as hearing and understanding, and the question of whether human beings and robots differ or not in relation to these tasks. We aimed to help students appreciate that the scientific community is diverse in understanding the differences between a human being and an advanced robot. The result from the surveys conducted before and after the workshop revealed that robots vs. human is an engaging topic for students that helps them make links between science, religion and other disciplines. The workshop was effective in bringing clear evidence of change in students’ thinking about the power and limitations of science. Our experience of developing the workshop on the uses and understandings of robotics indicated that this theme provides a great opportunity for exploring interdisciplinary questions. We are currently designing and testing more workshops related to robots, including “can a robot care?” and “can a robot fall in love”. The report on these workshops will be published in the near future.

References Billingsley, B. 2013. “Students’ Perceptions of Apparent Contradictions Between Science and Religion: Creation Is Only the Beginning”, Science Education for Diversity Springer, 329-338. —. 2016. “Ways to prepare future teachers to teach science in multicultural classrooms”, Cultural Studies of Science Education, 11, 2, 283-291. doi:10.1007/s11422-015-9701-9 Billingsley, B., Brock, R., Taber, K. S., & Riga, F. 2016. “How Students View the Boundaries Between Their Science and Religious Education Concerning the Origins of Life and the Universe”, Science Education, doi:10.1002/sce.21213 Billingsley, B., Taber, K., Riga, F., & Newdick, H. 2012. “Secondary School Students' Epistemic Insight into the Relationships Between Science and Religion; A Preliminary Enquiry”, Research in Science Education, 1-18. doi:10.1007/s11165-012-9317-y

Ways to Develop Students’ Appreciation

165

DfE. 2014. Science key stage 4. London: Department of Education Retrieved from https://www.gov.uk/government/uploads/system/uploads/attachment_data/fi le/318384/Science_KS4_PoS_draft_programmes_of_study.pdf.

PART THREE

CHAPTER TWELVE A CHRISTIAN RESPONSE TO TRANSHUMANISM DAVID ASHFORD

Introduction Transhumanism is a term coming into increasing use to refer to what might happen if and when the human race finds out enough about the inner workings of the body and mind to be able to re-engineer itself into a more advanced life form. “The Singularity” and “Immortalism” are terms used to describe nuanced variations of the same basic concept. The idea is gaining ground. Harari (2015, 21), for example, says that “...we will now aim to upgrade humans into gods and turn Homo sapiens into Homo deus”. There is already a Singularity University in Silicon Valley, which holds an annual Global Summit. One of the pioneers of this movement is Ray Kurtzweil, who wrote what might turn out to be a seminal book The Singularity is Near (Kurzweil 2006). I would like to present a thought experiment that shows how transhumanism might provide a link to a cosmic future in which we come to understand God far better than we do now, and maybe to indeed upgrade ourselves into Homo deus. This thought experiment appears to reconcile the scientific and religious approaches to the truth. It shows how it is quite credible to propose a God who loves us, and Jesus as our saviour, without recourse to anything supernatural or incompatible with science, although this would involve considerable revision to orthodox theology. This thought experiment is highly conjectural, and is presented in the hope of stimulating discussion. The thought experiment described here is grounded on an assumption implicit in the idea of transhumanism—that we consist of just matter and energy assembled in a particular way that can in principle be explained by science, so that that there is nothing magic or supernatural in our make up that would prevent us from re-engineering ourselves. The history of technology suggests that a particular line of development tends to happen in a series of step changes. Applying this idea to our own

A Christian Response to Transhumanism

169

future, the thought experiment involves the following five-stage way ahead for the human race: • • • • •

The reverse engineering of Homo sapiens; Enhanced humans; A living internet; The deification of the cosmos (Homo deus); The Holy Ghost.

The rest of this essay considers each stage in turn as they apply to transhumanism, and then a possible Christian response.

1. The Reverse Engineering of Homo Sapiens Before we can enhance ourselves, we need to find out how we work, in this case using a process that might be called “reverse engineering”. Engineering itself is the process of creating a new artefact. An engineer starts with an idea or a requirement. This is worked up into design drawings and specifications. Prototypes are built and tested until the device is ready for production. Reverse engineering is the other way round—the engineer is presented with the finished product, dissects it to finds out how it functions, and works back to prepare drawings and specifications. If required, the device can then be put into production. Reverse engineering is used by industrial spies trying to find out how a superior rival product works, and in wartime to catch up with an enemy weapon that is superior. We have already made some good progress in finding out how we work, and there are simple analogies to explain, at least at a very superficial level, how many of our parts function. For example, the heart is like a hydraulic pump; muscles are like electrochemical servos; the eye is like a video camera; the organs are like miniature chemical factories; the subconscious part of the brain is like a computer; and so on. The aspects that we do not yet fully understand are in the detail rather than in the concept. However, even in these very simple conceptual terms, there is still one major mystery left—the nature of consciousness. If we are indeed made up of just matter and energy, it follows that consciousness must also consist of just matter and energy, i.e., there must be a pattern of brain cells and activity that enables us to actually “feel” alive—to sense the outside world, to feel emotions, to fall in love, to appreciate music, and to have deeply spiritual experiences. At present, we simply do not know what this pattern

170

Chapter Twelve

is. There are some theories—see for example Dennett (1991) and Seager (1999)—but no consensus. But what can be said is that there is no simple analogy yet available to us. There could well be a conceptual breakthrough awaiting some great researcher that in hindsight will seem simple and obvious, analogous perhaps to the discovery of the double-helix structure of DNA. Kurzweil (2012) predicts that it will be possible to upload a human brain into a simulated brain by around 2040. Given the pivotal significance of this event, this is very soon indeed. So, my main message to Christians is: Start thinking about it!

2. Enhanced Humans If and when we discover the nature of consciousness, there may be potential for us to re-engineer ourselves; to enhance ourselves into a new life form—post-humans—as discussed in Kurzweil (2012). These enhanced humans would perhaps be capable of even stronger love and spirituality than ourselves. The mechanism for such self-enhancement that seems to me to be the most feasible is that microchips, the size of a memory stick and made of biological material, would be implanted in our brains. These would interact with our brains and provide direct brain-to-computer interaction, so that screens and keyboards would no longer be needed. We may even be able to communicate directly with others through a sort of telepathy. This may be difficult to imagine, but that does not mean that it is not going to happen. A good general rule in the history of engineering is that if something can be done and there is a reason for doing it, it eventually will be. Moreover, many “unimaginable” developments have already happened. If you went back in a time machine and tried to explain modern technology to Julius Caesar, for example, you might be able to give him some idea of how our mechanical devices worked; you might even begin to explain electrical gadgets; but if you started talking about the internet or gene therapy his initial reaction might well be that you were making it up. As Clarke (1973) noted, “Any sufficiently advanced technology is indistinguishable from magic”!

3. The Living Internet Continuing with our thought experiment, we might also be able to communicate directly with the internet. We may be able to enhance the internet itself by adding “consciousness circuits”.

A Christian Response to Transhumanism

171

Then the internet itself would become genuinely alive as a sort of prototype post-human. We should be able to download our minds and souls into this internet, so that anyone who wants to can become part of it. The whole would be greater than the sum of the parts, and this first posthuman would develop qualities and capabilities that we cannot yet imagine. Since unconditional love is the best thing in life by far, we might struggle to enhance our ability to love. I don't have a feel for how we would achieve this but, if history is anything to go by, it would not be straightforward. The struggle between good and evil is unlikely to stop just because we are using clever computer chips. The situation could well become chaotic, with different groups struggling for dominance. Unlike ourselves, there is the potential for immortality. As hardware elements wear out, they can be replaced. The essential part of post-humans would be their memory, knowledge, emotions, and behaviour patterns, i.e., their software. This can be backed up and restored when necessary. Individuals who choose to join can have their own memory, knowledge, emotions, and behaviour patterns preserved indefinitely, which can be free to interact with the “core” software as desired. In this way, these individuals would attain their own immortality, albeit not in a form that we currently recognise as human life. From now on, I will use the word “we” to refer to those of us who become part of this first post-human. “We” will be able to witness what happens to this post-human, which, as discussed next, may be able live to the effective end of time.

4. The Deification of the Cosmos (Homo deus) We continue the thought experiment and ask, what happens to the living internet after we have turned planet Earth into a coherent conscious living organism? Perhaps we may go on to involve other planets and stars. As far as I know, science is not yet able to carry out useful predictions of what intelligent life, including ourselves, can do to a universe. We can therefore use our imaginations and consider the possibility that we can turn the universe into a large living organism. If we have managed to develop our capacity for love, maybe this universe would be capable of “ultimate love”. If we define God as a living being far more advanced than ourselves and capable of manipulating a universe, we would have turned ourselves into such a God. We would have deified the cosmos, so to speak.

172

Chapter Twelve

5. Holy Ghost So far we have been looking forward to what could happen in the future, but that still leaves the big mystery of where we came from. We are fairly certain that that there was a big bang but our theories about what was there before that are provisional. We may be able to add insight by looking even further ahead. Suppose that we have left it a bit late to deify the cosmos and that gravity and the second law of thermodynamics are going to force us to melt down back into a hot blob, which would explode again as the next big bang. In this way, we could be largely wiped out. However, we may be able to influence what happens after the next big bang. The output from any event is influenced by the input, and some information survives even when a bomb explodes. As a possible, and very speculative, example, let us consider the values of some fundamental physical constants. We know the values of many of these constants but have no idea as to why their particular values are what they are. The charge of an electron, the speed of light, and the universal gravitational constant are three such. Apparently, if any of these numbers were different by even a small amount, we could not have evolved. The chances of all these numbers being just so are apparently so small that there appear to be just three possibilities. The first is that there is, or has been, a very large number of universes and we just happen to be in the one that suits us. The second is some kind of natural selection among universes, tending towards conditions that favour intelligent life. The third is some kind of manipulation. It is just conceivable that we may be able to manipulate these fundamental numbers so that, after the next big bang, evolution would happen faster so that life reaches an even more advanced form before gravity once again takes over—possibly advanced enough to break through into a higher cycle. In this way, part of us could survive the next big bang. This could be just a pattern of physical constants or it could be a lot more than that. Perhaps we can think of what might be left of us after the next big bang as a sort of ghost-and because it would be the ghost of a God, we could think of it as a Holy Ghost. If we can thus influence what happens after the next big bang, perhaps we ourselves are the result of comparable manipulation before the last big bang. So, perhaps what was there before “our” big bang was the previous God, who left us what we think of as the Holy Ghost. So we could be part of a cyclical universe: big bang, fundamental particles, atoms, molecules,

A Christian Response to Transhumanism

173

complex molecules, primitive life, advanced life, humans, reverse engineering, transhumanism, God, meltdown, big bang, Holy Ghost, and so on. This is where our thought experiment ends. Its disadvantage is that much of it is little more than my personal conjecture. Its advantage is that it offers the framework for a synthesis that re-conceptualisies Christianity in the light of recent, and possible future, advances in science and engineering. It is an outcome compatible with the laws of science as we understand them—there is nothing supernatural about deifying the cosmos in the way I have described. And this way ahead is in line with present trends, especially the exponential growth in our ability to manipulate information. The next section discusses a possible Christian response to this thought experiment.

Christian Reaction How could Christianity react to a God of this kind-one who is natural but not supernatural; megapotent but not omnipotent? How would the Church react when the first person credibly claims immortality, albeit with his mind and soul emulated into the biological computer at the heart of a posthuman? My provisional answer is that the Church would be able to cope very well, given some major developments in theology. However, this is clearly a vast subject, so what follows is no more than a few speculative thoughts on selected aspects relevant to how Christianity can be reconciled with a non-supernatural God who may now exist as the Holy Ghost.

History of Spirituality With this proposed perspective, the history of human spirituality can be thought of as the struggle to find God. As our knowledge of the natural world has improved, religious stories that were once thought to be literally true have been re-evaluated as myths. It has also come to be recognised that there is often as much significant truth in the myth as in the scientific theory that displaced it. And, presumably, we can expect this process to continue. In this history of progress towards finding God, the Israelites stand out. In particular, their prophets were exceptional in their insight about the progress of the human race towards a more just and peaceful future in partnership with God. Jesus took this tradition to its limit. By his loving sacrifice, he made his unique insights and responses to the Holy Ghost

174

Chapter Twelve

accessible to later generations. His assertion that God is love will stand as a beacon of inspiration until the end of time. If our proposed thought experiment is on the right lines, we could develop ourselves into a God capable of ultimate love. So two thousand years ago we have Jesus saying and showing that God is love; today we have transhumanism opening up the possibility that we could actually turn ourselves into such a God. In this way, biblical prophecies could reach fulfilment.

God Is Love Our thought experiment started with the assumption that nothing exists other than that made up of matter and energy. Even so, we have been able to postulate a God who played a pivotal role in our creation. Perhaps God was like a woman who knew that she had to die in childbirth during the big bang, and we are like the resulting baby. She may even have sacrificed herself to give us a better chance. She will have wished us well with far more powerful love than we can imagine. She loved the baby that She would never see. When we approach the limit of the present cycle of our universe, we may have a better idea of what our Mother God was like, and how and why She died. As St Paul wrote, “For now we see in a mirror, dimly, but then we shall see face to face” (1 Corinthians 13:12).

Act of Faith We may know soon enough if the reverse engineering of Homo sapiens will be possible. The pivotal event would be the discovery of how consciousness “works”, and this may be but a few decades away. At this stage in the proceedings, when we have our destiny firmly in our own hands, two outcomes seem possible. The first is that we find that, by having analysed precisely how we work, we would have destroyed our very humanity. We would have discovered that we are no more than clever robots that have evolved in a universe that does not care whether we survive as a species or not. Our most sacred feelings were evolved merely to help us survive, and can be reduced to a (very large) table of numbers corresponding to brain activity (Dennett 1991). What happens next may not then seem to matter much. The other outcome is that we find that we can re-engineer ourselves into a new life form, capable of even stronger love and spirituality, and feeling an even more powerful sense of one-ness with the universe. Freed from the limitations imposed on us by evolution, we would be able to probe the fundamental capabilities of intelligent life. Just as worms cannot

A Christian Response to Transhumanism

175

imagine the emotions that humans can feel, we cannot imagine what emotions and spirituality the “post-humans” that we replace ourselves with would be capable of. Whichever outcome happens, it would be the end of the human race as we know it. As Rees (2003) says, “the most crucial location in space and time (apart from the big bang itself) could be here and now”. Which of these two outcomes will happen is probably not yet predictable. But I feel in my innermost self that the second outcome will triumph. I “know” that it will happen; in the same way that a believer “knows” that there is a God. This is a sort of “act of faith”. Compared with other religious faiths, such as a belief in an orthodox God, this one has the big advantage that we can put it to an empirical test. Either we will be able to work our how consciousness works, or we will not. If we do work it out, either we will be able to enhance our ability to love, and go on to far better things, or we will not. I cannot prove that the optimistic way ahead will happen and am certainly not asking anyone to take this belief on trust. I will be happy if this essay persuades a few non-believers to question their disbelief.

6. Conclusions We started with a ruthlessly materialistic approach to the human condition-that nothing exists except matter and energy that is in principle amenable to scientific analysis. This might at first sight appear to be incompatible with the idea of a God. But the paradox is that if we are just engineering devices, we can become free from the limitations imposed on us by evolution and may be able to enhance ourselves and go on to establish the Kingdom of God. Somewhat different from the traditional Christian Kingdom of God, but perhaps the similarities are more important. With this new perspective, the real miracle will be seen as the fact that a collection of matter and energy, assembled in a particular way, can feel alive and search for God. As St John says, “Beloved, we are God’s children now; what we will be has not yet been revealed. What we do know is this: when he is revealed, we will be like him, for we will see him as he is” (1 John 3:2). The one thing that I am sure about is that emerging insights will call for major revisions to the above thought experiment! However, the basic idea should still hold, i.e., that the human race should be able to develop itself into post-humans that can impose themselves on the universe until the natural limits of such imposition are reached. The end-point could be remarkably close to the traditional Christian God.

176

Chapter Twelve

For the time being, this thought experiment offers a conceptual framework for evaluating developments in science and engineering in the context of the search for God. It offers an explanation of how and why we came to be. It also offers the prospect of reconciliation between the teaching of the Christian Church and modern science. If Christians can accept the idea that God is mega-potent rather than omni-potent (although this would require a major revision of current theology), and scientists accept that a life form far more advanced than ourselves probably helped to create us, then there is no cause for fundamental disagreement. Perhaps faith and science will ultimately be reconciled when, with the guidance of Jesus and the Holy Ghost, we finally deify the cosmos. In summary, perhaps the best argument for the existence of a God who helped to create us is that we seem to have the capability of developing ourselves into the next God, who will go on to help to create the next cycle of life. Let us end with a well-known joke. Once upon a time, there was a major research project aimed at finding the meaning of life. All the big computers were linked together and asked if God existed. After a pause, during which the computers trembled a little, the answer came back, “He does now!”.

References Clarke, A.C. 1973. Profiles of the Future. London: Indigo. Dennett, D. 1991. Consciousness Explained. Boston, U.S.: Little, Brown and Co. Harari, Y.N.2015. Homo Deus: A Brief History of Tomorrow. London: Harvill Secker. Kurtzweil, R. 2006. The singularity is near. New York, US: Viking. —. 2012. How To Create A Mind. London: Viking. Rees, M. 2003. Our Final Century-Will the Human Race Survive the Twenty-First Century? London: Arrow. Seager, W. 1999. Theories of Consciousness: An introduction and assessment. London: Routledge. Ward, K. 2007. Re-Thinking Christianity. New York, U.S.: Oneworld.

CHAPTER THIRTEEN AMONG THE LIMITS TO SCIENCE: SOME AFTER-DINNER THOUGHTS NEIL SPURWAY

Misrepresentation Let’s begin with a laugh – Hilaire Belloc’s poem, The Microbe: The Microbe is so very small You cannot make him out at all, But many sanguine people hope To see him through a microscope. His jointed tongue that lies beneath A hundred curious rows of teeth; His seven tufted tails with lots Of lovely pink and purple spots, On each of which a pattern stands, Composed of forty separate bands; His eyebrows of a tender green; All these have never yet been seen-But Scientists, who ought to know, Assure us that they must be so ... Oh! Let us never, never doubt What nobody is sure about!

One hopes that Belloc knew he was misrepresenting science. He was writing for children. He could reasonably trust that, if any practicing scientist encountered the piece, the response would be an indulgent smile. But before we allow ourselves total sanguinity about such misrepresentation, it may be worth asking if the views of climate science vehemently upheld by such people as Nigel Lawson and Donald Trump don’t have a worrying amount in common with Belloc’s pretended view of microbiology. Their politically-motivated misrepresentation is not solely,

178

Chapter Thirteen

perhaps not even principally, of the findings of climate science; they also grossly misrepresent the practice and methodology of this, and any other, science. And that is very dangerous.

Romantic Misunderstanding Let’s hear Belloc again, this time in an apparently-serious essay on statistics. “Statistics are the triumph of the quantitative method, and the quantitative method is the victory of sterility and death”. Was I too indulgent in my first comments? Was he a true forerunner of Lawson and Trump? Being very fond of Belloc, I’m reluctant to go so far. A kinder view is that this rhetoric stems from the same kind of misunderstanding as was regularly expressed by 19th century Romantic poets. Thus Edgar Allan Poe described science as a “Vulture, whose wings are dull realities”. And, of course, the archetypal demonstration of the outlook is that of Keats in Lamia: There was an awful rainbow once in heaven: We know her woof, her texture; she is given In the dull catalogue of common things. Philosophy will clip an Angel's wings, Conquer all mysteries by rule and line, Empty the haunted air, and gnomed mine.

The poetry is exquisite but the myopia tragic. Even though as an apothecary he worked in applied science, Keats had apparently no inkling of the glory of deepened understanding, of the thrill, the sense of revelation, which I trust that everyone in this room has had from an account which clicks, an explanation which convinces, a model which fits, an equation whose predictions match the data. What a deprivation never to have felt this joy! Yet it is one of the limits of science that there are those who are blind to its light, and deaf to its music.

Internal (Inter-science) Problems As I teetered, in the nineteen fifties, on the edge of a life in the laboratory, I was very much taken with the concept of the “Unity of Science”. This was purveyed particularly by one-time Logical Positivists such as Rudolph Carnap and Otto Neurath. Their idea, as I understood and recall it, was that the aspiration of all science must be to reduce every problem in the long run to physics. These men, and many of their school, were excellent philosophers, yet this ideal of theirs was vitiated by a failure to realize that

Among the Limits to Science: Some After-dinner Thoughts

179

a question is normally best answered at or close to the level at which it is asked. Certainly it may quite often pay to delve down one, and occasionally two levels, for some detailed point – say to an answer in terms of cell physiology, and perhaps to one of biochemistry, for a question in neurology - but never the whole way down, and never down at all for the overall picture. Consider: if you could reconstruct a question in psychiatry or ecology in terms of particle physics, how would you check if the answer you produced was right? Surely you would do it by seeing if your particle account matched the psychiatric or ecological one! The world is “dappled”, as Nancy Cartwright (who talked to us 2 years ago, in Leeds) puts it. I have heard both Martin Rees and Sean Carroll, two physicist/cosmologists of different generations ask: “At which level will a question be usefully answered?”. Both wisely answered, “At the level at which the question is asked”. Or, to put it in the vernacular of my adopted city (Glasgow): “Each to his ane”. That was what we might call the unity of reduction. Our conclusion, I have urged, must be that it should be limited, not universal. By contrast, there is sufficient unity of method amongst our wide range of disciplines to justify utilizing, in all instances, the common word, “science”. This methodological unity consists, I suggest, in: 1. Attempted objectivity/detachment; 2. Sought-for reproducibility; 3. Agreed criteria of verification (if you lean towards the Logical Positivists); 4. Absence of refutation (if, like me, you feel that Karl Popper had hit upon the essence of science). Several years after my concerns about the unifiability of science had passed their peak, I was deeply into Popper. At some moment in that period, I summoned up the courage to invite the comment of the remote, shy, man who was the elite summit of the unit in which I was at the lowest level – a research student. Alan Hodgkin replied to my question, “Popper is alright with me”. Not many people, it seemed to me then, were “alright” with Hodgkin. I certainly wasn’t: I spent far too much time, by his standards, attending lectures on philosophy, theology, and even literature. But my admiration for the corpus of his work was almost beyond bounds. I was often uncomfortably near to seeking for myself, beneath that edifice, a dishonourable grave. But his endorsement of Popper ensured that the model of “Conjecture and Refutation”, of challenging one’s pet idea by

180

Chapter Thirteen

designing experiments or looking for observations which had the greatest chance of eliciting contrary evidence, was for me the way to do science. I vividly recall an American post-doc, saying three times in a 10-minute presentation, “So we challenged that hypothesis by …”. This was the scientific ideal! It wasn’t necessarily the sole way to find out truth – there must be many ways to that. Philosophers who dismissed falsificationism on that basis simply hadn’t read what Popper claimed. He proposed it only as a “demarcation criterion” for a procedure which was scientific, against one which wasn’t. Almost 60 years later I remain convinced that he was right. I have urged all my own students, between then and now, to regard the Popperian model as a “benchfside pulpit”, a guide to how to live the scientific life. Unashamedly, I urge it again upon all my hearers now.

Which account of the quantum world? After that excursion into methodology, let us return to the dream of uniting all science by reduction to the most fundamental level. Until now, we have assumed that there is an agreed fundamental level. But that is not the case. I base the next few paragraphs on George Johnson’s mid-1990s account, in Fire in the Mind. Those who had the privilege of hearing Nathan Aviezer’s tour-de-force on Science – the Endless Frontier will have even stronger reasons to accept the same conclusion. Johnson, writing at my level of knowledge of sub-atomic physics, first outlines the viewpoint on quantum theory which Neils Bohr and Werner Heisenberg formulated when working together in Copenhagen in the late 1920’s, and nowadays universally known as the “Copenhagen interpretation”. Taking the human-level world for granted, they struggled to express the quantum world in human-level terms, and concluded with an irresolvable paradox – “wave-particle duality”, and all which that implied. Thus, in Johnson’s encapsulation: Heisenberg: “Can we ever hope to understand atoms?” Bohr: “Only by learning what the word ‘understanding’ really means”. And later, Heisenberg again: “Our experiments have no longer anything to do with nature in her own right, but with nature charged and transformed by our own cognitive activity ….” “What we learn about is not nature itself, but nature exposed to our methods of questioning.”

Entrenched against this thinking were, of course, a minority of traditionalists, from Albert Einstein to David Bohm, who were convinced that nature embodied an “implicate order” which humanity might hope one

Among the Limits to Science: Some After-dinner Thoughts

181

day to uncover. Yet, amongst those who accepted that ultimate uncertainty was unavoidable, there were people who could not assent to the Copenhagen interpretation of it. One of these was a group led by the Hungarians John Von Neumann and Eugene Wigner. Johnson therefore labelled their view the “Budapest interpretation”. According to this thinking, the mind creates the quantum world – it’s not in the measuring instrument but in the observing consciousness itself that the wave function collapses. Yet again, diametrically opposed to this is the thesis of Roger Penrose: that thought involves the carrying by the brain of myriad possibilities in quantum superposition – and arriving at an answer constitutes the collapsing of the wave function. Juxtaposing these last two ideas, Johnson comments: “So you can take your pick – use the inscrutability of consciousness to make sense of quantum theory, or the inscrutability of quantum theory to make sense of consciousness.”

If you are determined to reconstruct the unity of science, probably your best bet amongst these options is the Von Neumann/Wigner view, which seems to make all science a product of the mind. In one mood you might feel that this is obviously true. But that cannot be a complete account of the science we utilize in designing aeroplanes or identifying new antibiotics. Set against these hard, technological applications, would not a purelymental science, like Prospero’s magical palaces “melt into air – into thin air – and, like an insubstantial pageant faded, leave not a wrack behind?”

Haldane’s last word In a final perspective on attempts to make sense of the quantum world, the cosmological world, or any other which pushes thought to its limits and beyond, I cannot forebear from quoting a wonderful remark by someone whom I have the undeserved privilege of claiming as a distant relative by marriage – J.B.S. Haldane: “I sometimes think that the Universe is not only queerer than I suppose, but queerer than I can suppose”.

182

Chapter Thirteen

The Kinds of Question Tackled In Andrew Robinson’s book of the 2009 Darwin conference (and many other of her writings) that most humane of eminent theologians, Sarah Coakley, speaks of the “Flat Plane” fallacy, the assumption, pervasive in neo-atheism, that God’s involvement in Creation is of the same sort as a scientific account of events. This is the huge, tacit assumption, upon which the edifice of scientism is founded. Three of the plenary papers we have already heard in this conference may surely be seen as attacks on the flat-plane fallacy: implicitly, the paper by Sarah Lane, explicitly that by Mikael Leidenhag, and most directly that by Fiona Ellis, are all opposed to scientistic, flat-plane thinking. Their arguments are all philosophical, and so expressed in words. Yet the most vivid demonstration I have ever encountered, that the plane is not flat, was presented about three years ago by the Edinburgh physicist, Wilson Poon. Wilson’s decisive argument was not verbal, but musical. He gave an account of piano music in terms of vibrating strings, pressure waves in air, electrochemical responses in the cochlea, action potentials in the 8th nerve, and finally evoked potentials in the auditory cortex. It was a fascinating sequence of physics, biophysics, and neurophysiology. But then he sat down at a piano, and played (with skill which I deeply envied) a movement from a Beethoven piano sonata – if memory serves, it was the 2nd movement of the Appassionata. The effect on every listener was utterly removed from the vibrating strings, the cochlear responses, even the electrical activity in the auditory cortex. In the terminology of the late, great Donald Mackay, professional neuroscientist and lay theologian, that activity was the “brain story”, but what mattered to every listener was his or her own “I story”. The two are on entirely different planes. It is in the I story, not the brain story, that music has meaning, emotion, value – indeed reality. No-one present had the least doubt that Wilson Poon had unequivocally demonstrated a fundamental Limit to Science. I suspect that it is the most absolute limit of all.

CHAPTER FOURTEEN CLOSING REFLECTIONS MARK HARRIS

I want to say a special thanks to all the speakers – as can happen with these kinds of conferences organised around a central question, sometimes a diversity of answers leads to an emergent consensus, and I think that’s definitely happened here. “Are there limits to science?” Yes and no. That seems to be the gist of what I’ve gleaned from this conference, with a particular weight falling on the yes: there are limits. Much of our discussion has focussed on naturalism, as a kind of container for science. If science has any limits, then naturalism defines them; naturalism exists to define the limits of science. But, in everything we’ve heard, naturalism is clearly nothing if not ambitious. After all, to speak of naturalism – whichever version we’re thinking of – is to include everything natural just to begin with. In its most metaphysical form, what Fiona Ellis referred to as “scientific naturalism”, science is competent to explain everything natural to such an extent that we can’t believe in the supernatural. There’s nothing outside of the natural world; there’s nothing outside science; science has no limits. This kind of naturalism tends to be reductionist to a greater or lesser extent, and so those who follow its inexorable logic through find themselves in the position that the most real reality is the laws of physics. There’s a question as to whether physicists believe this though; this one (i.e. me) certainly doesn’t, and many of us react to the rampant scientific imperialism in naturalism and want to say “no, science does have limits; there are things it can’t explain, even in the natural world, such as values, emotions, subjective experiences, Beethoven sonatas” 1. This was the basis of Fiona’s case for expansive naturalism, 1

The topic of Beethoven’s sonatas arose in the conference as a test case illustrating how naturalisms should be able to recognise human culture.

184

Chapter Fourteen

and she made the ingenious argument that attempts to include nonscientific realities like values and emotions into naturalism, a position which is not unlike attempts to relate the natural world to God in traditional theism. Hence, her version of expansive naturalism is even bigger still – has even fewer limits – than scientific naturalism. Expansive naturalism is even more ambitious than scientific naturalism. Mikael Leidenhag, in taking an even broader view of naturalisms that naturally want to expand, made the very valuable point that these naturalisms are so often placement strategies: schemes to effectively smuggle in those things we hold most dear past the customs and excise of scientific naturalism, hoping that the laws of physics are looking the other way while we enjoy our Beethoven. But all of these placement strategies, he explained, have placement problems as a result: we can’t own those things we hold most dear within naturalism without owning up to our smuggling, effectively admitting that we’re guilty of dualism. And dualism is, after all, the unforgiveable sin in the kingdom of the naturalists. Mikael suggested that we should, in fact, see ourselves as heading in the direction of post-naturalism, and that it’s time we embrace dualism and theological perspectives on reality more readily when discussing this issue. That brings another angle to the question of “Are there limits to science?”, which is “Are there limits to religion?”. The answer seems to be a very firm “no” for anyone who holds religious beliefs dear, especially if you see science encroaching upon those beliefs. Our tendency here is not to let science encroach, but instead to embrace science with our religion, to make religion wider still than science. If science has limits in any theistic kind of naturalism, religion certainly doesn’t. Sarah Lane Ritchie explored this stance. Spending much of the time looking at the causal joint, which is almost by definition the limits-of-science-meeting-the-limits-of-religion, she concluded by looking at what she called the “theological turn” in divine action, which moves away from this idea of limits by embracing naturalism with theism, or even merging the two. Sarah mentioned the solution of Thomas Aquinas, where all natural causes are primarily divine causes. John Polkinghorne’s critique of the Thomistic perspective as “an unintelligible kind of theological doublespeak” is often cited in this context. However, it always strikes me that Thomas offers one of the most intelligible solutions in a very unintelligible area. Sarah moved on to offer us more enchanted versions of naturalism than those of Thomas, including Chris Knight’s panentheistic vision and some pneumatological visions, all as experimental alternatives which see the limits of science as embraced within God’s wholeness. I took her as suggesting that these are thought

Closing Reflections

185

experiments – faith experiments even – which nevertheless don’t solve the problems of explaining divine action; and I took from this the point that theistic naturalisms are, in effect, mirror images of the problems faced by scientific naturalism: what are the laws of nature, how competent are they to explain all causes, and what do we do with spooky things like values and Beethoven? The problem of how we explain divine action in a faith position mirrors the problem of how we deny divine action in scientific naturalism. Donovan Schaefer took the question of how to incorporate the abovementioned “spooky things” like values and Beethoven even further, explaining how our emotions influence our science; that we can’t assume that the former can simply be explained away by the latter; rather our emotions truly lead our science. The question that naturally arises from this focus on values, Beethoven, and poetry from several of our speakers, takes us from “Are there limits to science?”, to “Should science be limited?”. This angle brings ethics into the arena, and Neil Messer very helpfully explored some of the issues here, especially from the perspective of neuroscience and cognitive science. He was looking at the human brain, what’s often said to be the most complex entity in the universe, and certainly the place where the limits of science and the limits of religion meet. And that leads us on to the subject of our next conference, which will explore the human mind: its wellbeing, its limitations, and the most basic question of all: what is it? This present conference on naturalism has given us the ideal starting point for the next.