106 30 4MB
English Pages 346 [347] Year 2020
The Material Image
The Material Image Reconciling Modern Science and Christian Faith Donald H. Wacome
LEXINGTON BOOKS/FORTRESS ACADEMIC Lanham • Boulder • New York • London
Published by Lexington Books/Fortress Academic Lexington Books is an imprint of The Rowman & Littlefield Publishing Group, Inc. 4501 Forbes Boulevard, Suite 200, Lanham, Maryland 20706 www.rowman.com 6 Tinworth Street, London SE11 5AL, United Kingdom Copyright © 2020 by The Rowman & Littlefield Publishing Group, Inc. All rights reserved. No part of this book may be reproduced in any form or by any electronic or mechanical means, including information storage and retrieval systems, without written permission from the publisher, except by a reviewer who may quote passages in a review. British Library Cataloguing in Publication Information Available Library of Congress Control Number: 2020945935 ISBN 978-1-9787-0390-2 (cloth: alk. paper) ISBN 978-1-9787-0391-9 (electronic) TM
The paper used in this publication meets the minimum requirements of American National Standard for Information Sciences Permanence of Paper for Printed Library Materials, ANSI/NISO Z39.48-1992.
For Karen
Contents
Preface Acknowledgments
ix xiii
1
Christianity, Naturalism, and Science
2 3 4 5 6 7 8 9
Knowledge Miracles Origins Mind Freedom Morality Religion Resurrection
23 49 83 123 163 193 239 277
Conclusion Bibliography Index About the Author
311 315 327 331
vii
1
Preface
This book is firmly grounded in the historical Christian faith that “in Christ, God was reconciling the world to himself.” 1 Yet it advances views that many of my fellow Christians will find surprising, counterintuitive, and perhaps not just at odds with some things Christians have traditionally believed, but incompatible with genuine Christian belief itself. For this reason, I believe I owe the reader an explanation that locates The Material Image in the context of the journey that led to it. In retrospect, I realize that the roots of this project lie deep in my upbringing. The faith community in which I was raised embodied a conservative, Evangelical expression of Christianity. Many of the congregants, including my father, worked in the sciences, in industry, or in the universities in nearby Cambridge, Massachusetts, and Boston. They took for granted that their scientific work was not merely compatible with their faith, but an expression of it, done for the glory of God. I recall hearing, as a young child, my Sunday school teacher, a distinguished biochemist, explaining the wonderful way in which God created the plants and animals. 2 Years later, I realized that he had been describing biological evolution. So far as I remember, until I went to college I had little inkling that what science tells us about human beings deeply troubles many Christians. I learned that there was a “war” between faith and science. It was, I saw, a war with many casualties. I encountered those who, sorrowfully or gladly, had surrendered their faith in favor of the science they were learning in their classrooms and labs. Others, in what they imagined a heroic act of faith, simply denied science in favor of what they had grown up believing. And still others secured their science and their faith in separate compartments, evading scientific challenges to traditional beliefs, but forgoing the opportunity for their fruitful interaction. I was grateful I had been spared all this, having been given confidence early on that the God we meet in Jesus Christ is the Creator of the world that science portrays. Later, as a professor, I came to know many more young victims of the unnecessary conflict between faith and science. I also had the privilege to work with colleagues in various disciplines who have a deep respect for the sciences and a firm Christian faith, and believe, or at least hope, the two can be reconciled. I sensed, however, that they felt the lack of an intellectually honest, comprehensive account of how this might be done. ix
x
Preface
Often, they were particularly challenged by the current encroachment of the natural sciences into the human domain, such as the portrayal of the mind’s full materiality on offer from the cognitive sciences, and evolutionary accounts of morality and religiosity. It is principally with these folks in mind that I have written this book. However, in my philosophical work, and in my explorations in science, theology, and biblical studies, I realized that making peace between science and the Christian faith cannot be cost-free. Contemporary science undermines much of the traditional human self-image, a conception of our origins and nature long integrated into, and, for some, virtually identified with, the Christian faith. To reconcile Christianity and science calls for some significant revisions in how we conceive ourselves as God’s creatures. I hope to show that these departures from tradition do not attenuate Christian belief, but enhance it, affording a clearer vision of who God is and what God has done for us in creation and salvation. This is only to be expected if not only the essential Christian confession, but the well-confirmed theories of science, are true. The origin of my approach to our traditional beliefs, like my assumptions about the consistency of science and faith in Christ, lies in the faith community into which I was born. It embodied a solid commitment to the essentials of the Christian faith, to the Bible as God’s infallible word, to the holy and undivided Trinity, and to Jesus as God incarnate, fully human and fully God, the crucified and resurrected savior of the world. I am thankful that it did; this is what I still believe. Yet, as part of the Plymouth Brethren movement, it deviated from the Christian tradition in significant ways. It eschewed professional, ordained clergy in favor of the equal ministry of all believers, and it rejected liturgical worship in favor of the spontaneous guidance of the Spirit. My community held fast to the essential core of Christian faith, respected the long tradition that has grown around it, but was willing to give up long-held beliefs on secondary matters, in favor of what to it seemed greater fidelity to the essentials. This clear sense of the difference between essential and inessential, even if often still important, matters carried over to the many issues that in those days divided the Church: questions of eschatology, modes of baptism, and who should share in the Eucharist; of politics, such as the civil rights movement and the Vietnam War; and of personal behavior: drinking, smoking, dancing, card playing, movie going, and so on. Debate on such things was welcomed and I am thankful I grew up immersed in it. I learned a healthy distrust of anyone quick to deny the authenticity of someone’s faith because of disagreement on such secondary matters. What was of ultimate importance was faith in Jesus Christ, leaving the faithful a good deal of latitude on lesser matters. I am grateful to have imbibed the attitude: “in essentials, unity, in non-essentials, liberty, in all things, charity.” 3
Preface
xi
That I was shaped by this approach to tradition, including that of the Plymouth Brethren themselves, was, perhaps ironically, manifest as later I came to love liturgical worship and was delighted when my wife answered a call to become an Episcopal priest. The Episcopal communion into which I was confirmed is rich in tradition, yet it is a tradition in which the individual is free to work out his or her faith as reason and one’s understanding of Scripture lead. Belonging to the wider Protestant community, I accept that the effort to be faithful now may lead us to give up some well-established traditional beliefs, as it has in the past. My core conviction is that God, decisively revealed in Jesus Christ, is properly conceived and addressed as a personal agent. God is not only the ultimate object of intellectual inquiry, but a subject who calls for personal commitment on our part. The Material Image attempts to show that this personal God’s quest for fellowship with created persons is the key to reconciling the naturalistic implications of science with the Christian faith. In chapter 3, the reasons I offer to justify belief that God exists are reasons that involve God being a rational free agent, acting on desires and intentions, and ipso facto a person. 4 However, in the spirit of full disclosure, I report that I have given up various other beliefs held by Christians over the centuries, in favor of what to me seems better supported by Scripture and truer to the Gospel. I no longer think that God is timeless, immutable, and without passions. On the contrary, I believe that God created us to share in the everlasting divine life, and this communion of persons presupposes that what we do affects God, moving God to sadness or to joy. I no longer conceive our being made in God’s image as involving some kind of similarity to God. Instead I see our being imago Dei as our vocation to represent God in creation and share in the divine creative work. I also hold that divine omniscience does not imply that God has exhaustive foreknowledge, for I believe that God intended this world as a home for free persons, facing an open future they would, together with their Creator, bring about. I believe that God became a human being not just as a response to our fallen condition, but that this was God’s original intention, in fact the reason for creation. I deny that we possess immaterial souls, and that heaven is a parallel realm to which we go after death. Instead, I believe that our hope for the life to come rests entirely upon the bodily resurrection we will share with Christ, and that the future holds the new heavens and the new Earth, creation redeemed. I do not believe that those outside the Church, those not so blessed as to know Christ in this life, are damned to perdition. There is, I now believe, no hell as a venue of punishment, but that the hells Christ harrows are made of the pride, hate, and fear in which we immure ourselves and others. I do not see God’s justice as a cosmic accounting of who deserves what, but as God’s utter faithfulness to those God has created. Jesus’ death saves us not from divine justice, but delivers us from the power of sin and death.
xii
Preface
All these ideas are very important to me, but I would not pretend that these ways of understanding Christianity are normative, essential to an authentic faith. Several emerge explicitly in The Material Image; a few more are discernable in the background. None are unique to me and some are, I believe, if not already mainstream views in Christian theology, on their way to becoming such, and thus are components of tomorrow’s tradition, to be subject to faithful scrutiny by future believers. This is not the case with my contention that the traditional assumption that creation entails design should be put aside, in favor of the idea, central to this book, that God has designed a world that in due course gives rise to creatures that are not designed. God created us, but did not design us. This too, I hope, will become part of the ongoing Christian tradition. Having said all this, I am confident that the theology on which the reconciliation of science and the Christian faith this book proposes adheres to orthodox Christianity. Each Sunday, I recite the creed in good faith. Nothing in this book is at odds with it. In fact, I take the book to be quite conservative in the wide context of contemporary theology. The pivotal event in the personal journey that eventually brought me to write The Material Image was the experience that enabled me to internalize, in a vivid and life-changing way, the reality of God’s unconditional love. With this in sight, we have nothing to fear, no worry that by asking hard questions, or questioning tradition, we will get something wrong and lose God’s favor. Wrong as I often am, God loves me no less. Whatever challenges to faith we find in science or anywhere else we can meet boldly, knowing that when we look into the abyss we see, not, as Nietzsche said, merely ourselves looking back, but the face of our loving God. 5 My hope is that those who read this book will experience this. NOTES 1. 2 Corinthians 5:19. This, and all subsequent biblical citations are from The New Revised Standard Version Bible (New York: National Council of the Churches of Christ in the United States, 1989). 2. Dr. Robert L. Herrmann, then of Boston University. Much later, he co-authored books on faith and science with Sir John Templeton, including The God Who Would Be Known: Revelations of the Divine in Contemporary Science (West Conshocken, PA: Templeton Foundation Press, 1989) and Is God the Only Reality? Science Points to a Deeper Meaning of the Universe (New York: Continuum, 1994). 3. Ascribed to the Lutheran theologian Rupertus Meldenius (1582-1651). 4. Those of us who believe in the Triune God encounter the ancient problem of explaining how the Father, the Son, and the Spirit is each a distinct person, that God is a person, yet that three, not four, persons comprise the Godhead. For that reason, and many others, God is the proper object of intellectual inquiry. For an engaging examination of the issue, see William Hasker, Metaphysics and the Tri-Personal God (Oxford, UK: Oxford University Press, 2013). 5. Beyond Good and Evil, Walter Kaufman, trans. (New York: Vintage Books, 1966), 89.
Acknowledgments
The idea for this book took root in the course of a sabbatical year granted by my employer, Northwestern College, spent at the Center for Theology and the Natural Sciences (CTNS) in Berkeley, CA. I am grateful for the hospitality of physicist-theologian Robert John Russell, Director of CTNS, and the rest of its helpful staff. And I am grateful to Northwestern College for that sabbatical and, beyond that, for being a unique institution in which firm faith and bold intellectual inquiry nourish one another. Friends and colleagues have read and discussed all or parts of the manuscript as it evolved, sometimes agreeing, sometimes disagreeing, always helping and encouraging. For their critical acumen and support I thank Dave Arnett, Daniel Berntson, Ralph Davis, Col. Fred Chesbro, Laird Edman, Randy Jensen, Brian Jones, Mike Kugler, Sam Martin, Jim Mead, Ryan Pendell, Rein Vanderhill, and Joel Westerholm. Many students in my philosophy of science courses also read portions of the manuscript and responded with valuable questions and criticisms. I am also grateful for the wise editorial advice I received from Neil Elliott, senior acquisitions editor for Fortress Academic, and for an anonymous reviewer’s challenging responses. As this book approached completion, an emergent medical situation put its appearance in doubt. The compassionate application of scientific expertise made it possible for me to bring the project to its conclusion. I have abiding gratitude to Dr. Nigel Millard, Dr. Miroslaw Mazurczak, Dr. Lori Maness, Dr. Mary Finnegan, and Dr. John Halgren, as well as my Case Manager, Diane Hill-Porecky. I am also happy to express my thanks to the American Cancer Society, and in particular to Lisé Anderson, for facilitating my access to medical treatment. Writing The Material Image over the years, I have anticipated being able to dedicate it to my priest, best friend, and wife, Rev. Dr. Karen Wacome. She is my life’s great blessing.
xiii
ONE Christianity, Naturalism, and Science
FAITH AND SCIENCE The project of reconciling religion and science is increasingly urgent as contemporary science moves rapidly into domains once sacrosanct, offering compelling naturalistic explanations of human origins, of the mind and behavior, and even of morality and religiosity that challenge our long-held image of ourselves. Old and dear ideas of human uniqueness and value disappear under the rising waters of natural scientific explanation. For many this is no less a challenge to their faith, which for them is tied to that traditional image. However, the attempt to integrate faith and science is too often ineffectual because it proceeds as though the particularities of faith can be safely set aside. For theists the temptation is to allow convictions about the nature, character, history, and creative purposes of God to recede into the background as we ask how to relate contemporary science to a merely generic deity. For Christians specifically, the further risk lies in proceeding without making God’s Triune nature, incarnation, death and resurrection central to the undertaking. What we need to know is what we can reasonably expect the world to be like if its creator is the God specific to the Bible and Christian faith, and thus how the world as science describes it might fit into God’s creation. Framing our inquiry this way we can ask how the marvelous but unsettling account of the world on offer from contemporary science can be reconciled with faith in the biblical God. Conceivably, confidence in contemporary science can be reconciled with a vague and general notion of deity, but this is of little concern for those of Christian faith. What matters is the particular character of the God we encounter in the biblical revelation and in the life of the Church: can we faithfully and honestly welcome the scientific view of the world 1
2
Chapter 1
and our place in it without attenuating Christian belief, or is the old conflict of faith and science inescapable? Discussions of how faith in God coheres with science often invoke general features of the world such as its orderliness, intelligibility, predictability and unity, and thus its (perhaps surprising) openness to scientific inquiry. This in turn is seen as pointing to a wise and powerful creator. So far so good, but not good enough. 1 The deity it gestures toward can seem austerely remote, the God of the philosophers, not the God of Abraham, Isaac, and Jacob, a deity whose proper task is just to create and oversee a well-ordered universe. 2 Yet this is a selective description of the creation. It is also a world in which some things happen by chance, a world that harbors chaos: tiny initial differences in simple beginnings culminate in profligate unpredictable consequences. It is a messy world, burgeoning with contingency, in which things that most matter to us, like human persons, cannot be smoothly mapped onto its fundamental constituents, but are superficial, dwelling in the shallows of reality. This is a creation with an open future, home to real risk and genuine fears and hopes. Any serious reconciliation of faith and science begins with acknowledging the actual and surprising Deity who could have created this vast, ancient, recalcitrant, dangerous, and beautiful world. 3 The task calls us not to fear or flee, but firmly grasp, the scandal of God’s particularity, focusing on who the Deity really is and does, however counterintuitive and out of kilter with our natural expectations for a god. The God of the Bible, the God of Christian faith, is no remote and solitary being but the holy and undivided Trinity, the everlasting community of Father, Son, and Spirit. This God eschews a “God’s eye point of view,” the neutral view “from nowhere.” 4 Never serene and disinterested, this God is committed, passionate, persistently pursuing aims for creation. The biblical God creates freely, but for good reasons and in keeping with the divine nature. “Because God is God, God is not all there is.” 5 The creation reflects the love between persons that is the essence of the tri-personal God. From mere matter the Creator brings forth persons called to serve as imago Dei, created persons to know and love, and who can know and love the one who made them. They are truly distinct from their Creator who endows them with the freedom and responsibility possible for mere material beings. God wants them to be distinct from, but not separate from, their maker. God reaches out to the other, reaching as far down into the depths of creation as possible, to call created persons to see things from their Creator’s point of view, to share in the Triune everlasting life, to take up their share of the Trinity’s ongoing joyful creative work. The Creator’s self-revelation takes the form of human words the lowly can understand. God is diffident, respecting creatures’ fragile identities. God seeks trusting responses but gives created persons space to be themselves. The Creator cloaks infinite power in utter humility, creating
Christianity, Naturalism, and Science
3
a world for the sake of incarnation. The Deity we encounter in the Bible becomes a creature among creatures, to share in their lives and to draw them to share in the everlasting life and work of their God. When things go wrong in the space made for us, this God does not abandon us. God patiently persists, entering upon the long work of salvation, dealing with the damage we cause in the best possible way. Our Creator lets us do our worst to ourselves, to one another, and to our world, and turns it to our healing. God condescends to our natural morality and religiosity, bending them to salvific purposes. In the fullness of time God comes to us incarnate, not as exalted Lord, as might have been, had things gone differently, but as crucified Savior. As a human being, God submits to being taken into the hands of angry sinners and accepts humiliating execution for blasphemy and sedition against creatures’ arrogant authority. And then this God is revealed not only as the one who gives life to all, the Creator’s life given for the creature’s, but who resurrects the dead, bringing us out of the nothingness of death to everlasting life in a restored creation. For Christians, there is no meaningful reconciliation of faith and science unless the world science portrays can vividly be seen as the world in which the three-person “God was in Christ, reconciling the world to himself” (2 Corinthians 5:19). This book’s focus is the creation, the object of scientific inquiry, but Christian faith views creation through the incarnate life, death, resurrection, and ascension of Jesus Christ. 6 The distance that separates this strange, awe-inspiring story of a God who overturns our assumptions and shocks our sensibilities, first told in the concepts and language of ancient, pre-scientific cultures, from the seemingly so different picture of the world on offer from contemporary science, might seem too great to traverse. It might appear that we can accept only one or the other. But this is not the case. This book’s task is to make the crossing. If we keep the old but always new story in clear view we can see how the marvelous account of the world on offer from contemporary science becomes part of the all-but-too-good-to-be-true good news about God. However, a travelers’ advisory is in order: the required strategy brings us into territory that will be unfamiliar and sometimes disturbing for persons of faith. We will find it leads to embracing the naturalistic implications of contemporary science. THE AIM OF THIS BOOK Carl Sagan wrote: In some respects, science has far surpassed religion in delivering awe. How is it that hardly any major religion has looked at science and concluded, “This is better than we thought! The universe is much
4
Chapter 1 bigger than our prophets said, grander, more subtle, more elegant. God must be even greater than we dreamed.” 7
Part of an answer to Sagan’s question is that along with delivering awe, science has delivered a series of blows to the human self-image, an image familiar, cherished, and entrenched in our natural religiosity. Religion, by and large, looks at science and finds a naturalistic picture of human beings that it is unwilling to accept. What science delivers is often not awe, but fear and loathing. Persons of faith feel compelled either to reject science altogether, or to make a futile attempt to deny its naturalistic implications. Contemporary science increasingly supports a naturalistic picture of human knowledge, human nature, and human origins, and almost everyone sees this as profoundly at odds with belief in God. The Material Image challenges this assumption and contends that the naturalistic account of human beings emerging from contemporary science is entirely consistent with the Christian faith. This is not a matter of a bare logical consistency but consonance, accord, or harmony. 8 Science, with its implications about human nature, origins, and limitations, can be well integrated with Christianity. Not mere peaceful coexistence, but a true marriage. Such a union is possible only if the partners are free to be themselves: science with no evasion of its naturalistic implications and the Christian faith in its historical, biblical particularity. NATURALISM AND CHRISTIANITY Naturalism can mean a variety of things, and I do not claim that Christianity coheres with everything the term might denote. For some, the term simply functions as a synonym for atheism, or at least as a label for a view that immediately implies that there is no God. One widely used reference work defines naturalism as “the view that everything is natural, i.e., belongs to the world of nature.” 9 God is not a physical thing, not part of nature, and not subject to scientific explanation, so these definitions imply that there is no God. Others make a connection of naturalism to atheism explicit. Philosopher Kai Nielsen, a well-known proponent of atheism, asserts, “Naturalism denies that there are any supernatural or spiritual realities.” 10 The theologian Ted Peters describes naturalism as equivalent to secular humanism and as involving the view that there is nothing but the finite reality of nature. 11 The Christian philosopher Alvin Plantinga writes, “The basic idea of philosophical naturalism (which from now on I’ll just call naturalism) is that there is no such person as God, or anything at all like him.” 12 The authors of a recent book critiquing naturalism from a theistic point of view tell us that it is the belief that everything that exists is a part of nature, and assert that what all naturalists have in common is the rejection of theism. 13
Christianity, Naturalism, and Science
5
Nothing of ultimate importance attaches to matters of definition. Those who see fit to use naturalism as a synonym for atheism are free to do so. What is objectionable is using words in ways that render certain points of view invisible. The term atheism can be deployed to straightforwardly express the claim that God does not exist, without making any further claim about the nature of the world or human beings. 14 The rationale for using naturalism rather than atheism is to point to further claims about the world and its inhabitants. Consider three such naturalistic claims: 1. Science is the most reliable means for human beings to know themselves and the world. 2. Human beings are material things: the immaterial soul does not exist. 3. Human beings are the product of unguided Darwinian evolution. It is possible that God does not exist and these claims are true. And it is possible that God exists and these claims are false. However, there is a third possibility: the God of Christian faith exists and these claims about our most reliable source of knowledge, what we are, and how we came to be are essentially correct. The automatic equation of naturalism with atheism hides this third option, obscuring the possibility that the world as described by contemporary science is the world that God created. That is unfortunate since this is, in fact, true. Science is the endeavor to explain what we observe by subsuming it under the causal laws of nature. Science seeks natural, not supernatural causes. Many Christian thinkers acknowledge this truism, labelling it methodological naturalism, in contrast to ontological naturalism, which they reject. 15 Insofar as the latter means that nature is all there is and thus amounts to atheism, this is a useful distinction. With it in view, I will use scientific naturalism for the conjunction of the preeminent reliability of science as a source of knowledge, a materialist view of human nature, and an evolutionary account of human origins. However, scientific naturalism, so defined, is not merely methodological: that we are entirely material beings obviously is an ontological claim. But for a Christian, it is a claim not about all of reality, but about the human beings God created. Scientific naturalism embodies no general claim about reality that implies that there is no God, but some may contend that the step from it to atheism is unavoidable. Arguments to this effect are readily available, but they invariably depend upon philosophical or theological premises that cannot simply be taken for granted; they need to be defended. For example, an essential Christian belief is that there is life after death, but naturalism implies that human beings lack immaterial souls that continue to exist beyond the body’s demise. 16 Almost everyone, whether or not they believe in life after death, assume that its possibility depends on there being immaterial souls, but the issue is whether this widespread assump-
6
Chapter 1
tion is true. Does the possibility of life after death require disembodied existence, and thus a dualistic conception of human nature that conflicts with naturalism? Or can it be grounded in an account of personal identity consistent with scientific naturalism’s materialist view of human beings? 17 Perhaps no such account succeeds, but this is a matter to be resolved by argument, not simply assumed. Someone might admit that it is possible that both naturalism and Christianity are true, but insist that it is highly unlikely. She might, for instance, admit that it is possible that God created the human species by means of the unguided, aimless processes of biological evolution, but that it is not at all plausible that God would do so. However, when it comes to such alleged improbabilities, the lack of fit of Christian faith and naturalism is not self-evident. Such assertions rely upon assumptions that must be made explicit and defended. It might seem otherwise, in virtue of deeply entrenched associations of ideas. Naturalistic ideas have for centuries been intimately associated with hostility to Christianity. We should, however, recall that the Copernican theory of the solar system, the idea of inertia, and the great age of the Earth were once among the scientific findings rejected as implausible in light of Christianity, if not overtly inconsistent with it. Today, the naturalistic results of science might seem just as improbable from the perspective of Christian faith. Yet history teaches that combinations of ideas that once seemed unlikely companions ultimately can be brought together without intellectual discomfort once underlying assumptions are brought to light, challenged, and discarded. We are entitled to hope for a similar outcome with Christianity and the naturalistic implications of contemporary science. Another line of reasoning sees a lack of fit between Christian faith and scientific naturalism as a matter of a competition Christianity has long since lost. Much that God was once invoked to explain is now adequately explained by science. For instance, the diversity, complexity, and adaptedness of living things are now adequately explained by the mindless mechanisms of natural selection, with no overt recourse to a divine intelligence. For many, such discoveries reveal that there is no longer any reason to believe that there is a God. This point of view is abetted by many persons of faith, who appear to stake their belief in God on what they regard as the failure of evolutionary theory and other well-confirmed scientific theories. Here too, the shared assumption needs to be made explicit. Does reasonable belief in God depend upon explicitly theistic explanations of natural phenomena? If it does, then naturalism and Christianity are irreconcilable competitors and we cannot accept both. The implicit assumption that it does, although common among both atheists and believers, is far from obviously true. Rather than accepting the idea that theistic belief is justified because it explains general features of the natural world that lie in the purview of science, Christians may appeal to other evidence, such as to human history, preeminently to the life,
Christianity, Naturalism, and Science
7
death, and resurrection of Jesus of Nazareth, as making belief that there is a God reasonable. If we do, our faith does not depend on a forlorn hope for the failure of well-confirmed scientific theories, and we have no reason to see Christian belief and science as competing explanations. Christian theology was first articulated in late antiquity, a culture innocent of scientific knowledge in the modern sense. Christian faith challenged some of that culture’s most important assumptions, but it also took for granted much that is false yet in that pre-scientific era was eminently reasonable. Assumptions relied on for centuries in the articulation of Christianity have become so closely associated with it that to reject them can seem like the rejection of the faith itself. When, for instance, it seemed obvious that life could not have arisen from lifeless matter, or when the idea that complexity cannot arise naturally out of simplicity was incontestable common sense, Christians reasonably concluded that only divine intervention could account for the existence of living things. Belief in immaterial immortal souls was widely accepted in the late ancient world and many—but not all—early Christians relied on it in defending hope for everlasting life with God beyond death. Science today challenges ideas that Christians have held for centuries, but we need not suppose that there is any necessary connection between Christianity and these ideas. Scientific naturalism does not call for abandoning beliefs unique to Christians; they are common to cultures not informed by modern science. There is no guarantee that these ancient beliefs best fit with the Christian faith. A significant example is the idea of heaven. In the late-ancient and medieval world, the heavens were the concentric crystal spheres in which the Sun, Moon and planets were embedded. The Earth was motionless, at the center—really the bottom—of this spherical cosmos. Beyond the outermost sphere, the sphere of the fixed stars, God dwelled in the highest heaven. The demise of this cosmological picture was a crisis for Christendom. Modern science came to birth as it constructed an account of a world that seemed to have no place for God. Theologian Robert Jenson reminds us that this had a “powerful and destructive impact on the actual theology of believers.” 18 If there is no heaven, where is God? Yet, in retrospect, the loss of heaven was not the disaster for the Christian faith it seemed at the time. Forced to abandon the idea of a preferred place in the cosmos where God dwells, Christian thinkers were in the long run able to arrive at a conception of God’s presence to the creation less indebted to Greek philosophy and more faithful to the biblical revelation. Heaven is not, ultimately, an alternative reality, superior to this mundane realm, but the future of God’s creation, redeemed and restored. Today no serious theologian would welcome the opportunity to return to the Ptolemaic cosmos and its timeless heaven. Early modern science forced Christians to revise and refine their theology for the better. Contemporary science, capturing human beings themselves in the net of naturalistic
8
Chapter 1
explanation, holds even greater promise. THE CHRISTIAN FAITH It might seem that the only hope of reconciling Christian faith with naturalism is to surrender its traditional supernatural claims and avoid all possible conflict with science. 19 One might dispense with supernaturalism altogether, interpreting talk about God not as referring to a transcendent personal being, but as a means of expressing various worthy values, attitudes and commitments about human life and its meaning in the universe. 20 Or one might conceive of religious belief as commentary on subjective experience, involving no assertions about objective reality that science might confute. Alternatively, one might retain a realist understanding of theological language, but move toward deism, abandoning Christianity’s claim that God not only created, but miraculously acts in, the world. I will not pursue strategies of this type. The Christianity that I claim strongly coheres with scientific naturalism is the traditional, orthodox faith articulated in the historic creeds. This faith presupposes belief in a transcendent God who is an everlasting community of loving persons: the Father, the Son and the Holy Spirit. This Triune God is the Creator. Things other than God exist only because God freely acted to bring them into existence. God is utterly free, yet a rational agent. The divine wisdom far surpasses human comprehension, but God’s actions are not arbitrary or capricious, nor are they completely inscrutable; the reasons for which God acts can be communicated to us. We can know at least some of the purposes for which God created. A central Christian confession is that God created this universe with the aim of bringing into being persons who can be known and loved by God, persons called to share the everlasting, communion of the Triune life. 21 This conviction is this book’s guiding theme: knowing the goals God pursues in creating helps us understand what was created and how it was created. Knowing this, we ask whether the world science reveals is that kind of world and find that it is. God can, and on occasion does, miraculously intervene in the created world, making knowledge of their Creator available to created persons. God is self-revealed in the history of Israel. The Scriptures of the Old and New Testaments are an inspired, reliable witness to that revelation. 22 Christian theology is the attempt to systematize, formalize, and explore the presuppositions, content, and implications of this revelation. It is not a kind of philosophy, our most general reflections upon the meaning of life and the world, though it has implications for such matters. God became incarnate as Jesus of Nazareth and it is by means of his life, death, resurrection, and ascension that God is bringing about reconciliation with creatures. The good news Christians proclaim is that God
Christianity, Naturalism, and Science
9
acted, and continues to act, in this decisive way, calling everyone to faith in Jesus Christ. Our most secure and accurate knowledge of God is our knowledge of the man Jesus of Nazareth, witnessed to by the Christian Church. Knowledge of God is, as Archbishop Rowan Williams writes, principally “not knowing propositions about an object, but sharing in God’s life.” 23 Christian faith is trust in, and love for, the God revealed in Jesus Christ. In a secondary sense, it is the body of beliefs about God, the creation, and ourselves that this faith presupposes. It is incumbent upon Christians to inquire into the various relations the beliefs that comprise Christian theology have to other things we know, and to do so in a way that is at once faithful and intellectually honest. Among the demands of that honesty is recognition of the superior reliability of science as a source of knowledge and an openness to its claims. I thus write from the standpoint of a traditional, biblically grounded Christian faith. I assume that the core of essential Christian belief expressed in the ancient creeds of the Church and adhered to by most Christians over the past twenty centuries is true. There are, within the wide scope of historical Christianity, matters that remain unsettled. Would God have become incarnate if the Fall had not taken place? What can we mean when we say that Jesus Christ’s death atoned for our sins? Is every human being ultimately reconciled to God or are only some saved? Does God possess exhaustive knowledge of the future? Is God eternal, in the sense of being timeless, or does God exist everlastingly in time? What does it mean to confess that God is omniscient, or providentially governs creation? I do not imagine that scientific naturalism is consistent with every combination of answers to questions of this kind. I will sometimes opt for an answer in part because it coheres with the naturalistic picture of human beings produced by science; but only in part: in all such cases I regard the theological view I espouse as warranted on biblical, theological, or philosophical grounds independent of science and its naturalistic implications. As will become obvious, some of the ways of interpreting the Christian faith I offer depart from tradition, but I trust that none depart from the essentials of the faith. This book asserts that setting out from a broadly Christian starting point, we might well have expected to find that God created a world of the sort science portrays, particularly its depiction of humans. It is specifically the Christian faith that I contend is consonant with scientific naturalism. I do not suppose that what I say about the fit between Christianity and the naturalistic views supported by contemporary science holds for every construal of Christianity, or for other varieties of theism. Neither do I suppose that naturalism can be reconciled with any particular nontheistic religious perspective. I suspect that other religious beliefs may well be as difficult to reconcile with the naturalistic account as so many take Christianity to be. But my understanding of other faiths is too fragmentary and shallow to pronounce on the matter.
10
Chapter 1
RATIONALE My principal reason for writing The Material Image is to persuade my fellow Christians that they should embrace the naturalistic picture of human beings on offer from contemporary science. It is, I believe, the only way faith and science can be reconciled. Most Christians take it for granted that their faith commits them to believing they have immaterial souls, an idea that is already untenable and will only become less plausible as scientific understanding of the mind/brain deepens. Among Protestants, millions of fundamentalists and conservative Evangelicals emphatically reject the scientific theories that underlie naturalism. They regard the evolutionary origins of the human species, and even the Big Bang theory, as patently false, embraced by scientists not in virtue of the evidence, but because they are dishonest or deluded, in the grip of an ideological commitment to an atheistic naturalism that systemically distorts scientific inquiry. 24 This is troubling. Many Christians are alienated from the scientific community and from the broader intellectual culture it informs. Too many of the devout fall into the intellectual dishonesty that they project onto secular science. They are tempted to dismiss subtle and complex theories without understanding and without regard to the reasons for which science accepts them. Many Christians place their hopes on the failure of science, setting up an unnecessary obstacle for honest and reasonable persons who seek the God who resurrected Jesus Christ. If, as is likely, the incoming evidence will only further confirm the naturalistic picture they reject, the divide between the faithful and science will widen. Contemporary science presents a picture of the world that is awe inspiring in its subtle intricacies and beautiful in its elegance; ancient pictures of the world pale in comparison to it, but many Christians cannot see in it cause to praise the Creator. For them, it occasions only fear and rejection, defensive arrogance and evasion. Many Christians do not, of course, share the antipathy to contemporary science common among conservative Protestants. Rejecting Darwinian evolution and the Big Bang are not articles of faith for mainstream or liberal Protestants, nor for most Roman Catholics. Christians in these traditions are less likely to focus on these matters and, when they do, they are much less willing to impugn the honesty or competence of scientists. Yet here too there is unease with the emerging scientific picture. These Christians also seek to avoid the disquieting implications of what science is teaching us about ourselves. They hope that accepting the results of contemporary science does not commit us to naturalism’s starkly materialist picture of human beings. They hope for a kind of freedom somehow ungoverned by the laws of nature. Some hedge their acceptance of Darwinism by holding that God guides the evolutionary process. Some endeavor to blunt the impact of contemporary science on the hu-
Christianity, Naturalism, and Science
11
man self-image by drawing on such possibilities as emergence and topdown causation, quantum mechanical effects in the brain, chaos theory and complexity theory, non-linear dynamics, self-organization, non-reductive physicalism, or even by embedding the scientific account of what we are in some broader, but ultimately dubious, philosophical scheme, such as process philosophy or the speculations of Teilhard de Chardin. Legend has it that the comic actor W. C. Fields lay on his deathbed, paging through a Bible. A surprised friend asked the ailing reprobate what he was doing; Fields muttered, “Looking for loopholes, looking for loopholes.” Persons of faith, concerned that the rising tide of scientific inquiry threatens to inundate humanity, sometimes seem intent on finding loopholes in the “book of nature” that will make the world safe for traditional conceptions of what we are. My strategy involves no loopholes, no appeal to speculative or exotic possibilities that might soften the blow science delivers to pleasing conceptions of human nature. Whatever value any of these ideas have in their own right, they do nothing to mitigate the news from science. I propose the marriage of a robust scientific naturalism with a robust Christian theology, tempted neither by denial nor evasion. That union relies not on loopholes, but on close attention to the actual implications of both the Christian faith and scientific naturalism. To explore these implications I will often rely on ideas recruited from contemporary philosophy, which seeks to find a place for human persons, for mind, consciousness, rationality, meaning, value and purpose, freedom, responsibility and moral truth in the physical universe we know by way of the sciences. If the theology and the science are this book’s poetry, the philosophy is the prose, the undramatic making of careful distinctions and cautious inferences. What I take from contemporary philosophy is its naturalizing project, a sober appraisal of the preeminence of science as a way of knowing, a reliabilist account of knowledge, the computationalist theory of mind, a compatibilist conception of freedom, personal identity through time as a matter of causal continuity. All face any number of objections and difficulties, as befits philosophical theories, even those fairly widely held. I do not want to suggest that the naturalistic project is triumphant, but I am confident that it will eventually prevail. My aim is not, in general, to defend the various naturalistic theories in their own right, but to set forth the possibility of an account of the human condition that is at least plausible, thoroughly naturalistic, and which is consonant with the Christian faith in all its particularity. If I am right in thinking that it is naturalism that best fits with Christianity, then adherence to a competing, pre-scientific image of what we are threatens to distort the Christian faith. Many Christians are beguiled by an inflated conception of human nature, one difficult to reconcile with our status as God’s creatures and one which makes little sense in light of the good news of Jesus Christ. The transition to a naturalistic view of human knowledge, human origins, and human nature enhances our abil-
12
Chapter 1
ity to articulate an authentic Christian account of the world and our place in it. The Christian faith contains a liberating critique of human pride, self-righteousness, religiosity, and of the “idols,” the many and varied substitutes for God devised by human beings. Science, challenging the traditional human self-image, is a valuable ally in this project, but only if we embrace its naturalistic implications. I address The Material Image principally to my fellow Christians, but a secondary aim is to nudge those who do not share that faith away from the conviction that one cannot be a Christian and reasonably accept what science tells us about human nature, origins, and knowledge. 25 Christianity has become associated in the public mind with hostility to science, even to reason in general. Many scientifically educated persons, aware of the credibility of science and the compelling evidence for naturalism it produces, regard the Christian faith as too implausible to take seriously. Christians appear willfully ignorant or irrational, and Christianity seems no more worthy of serious consideration than Zeus or Zoroastrianism. If there is no good reason to regard Christianity as at odds with the naturalism implicit in contemporary science, this is a sad state of affairs that should be rectified. Sadder still, what Alister McGrath refers to as the “rhetoric of dismissal” indulged in by those who see faith as an object of ridicule, rather than a target for rational criticism, is not entirely unfair. 26 However, this is not a work of apologetics. My aim is not to convince those who think Christianity is false to change their minds. The fact that science tells us that this is the sort of world we might have expected the God of Christian faith to create does not in itself count as a reason to accept Christianity. The goal of The Material Image is not to invoke the findings of science as evidence that Christianity is true. To remove an objection is not to produce a positive reason to believe. The book’s more modest hope is that those who reject Christianity will see that they must do so on grounds other than its alleged incompatibility with a naturalistic conception of human beings and the science that sustains it. We believers sometimes lose sight of the audacity of Christianity’s core claims about God becoming a particular human being whose execution and subsequent resurrection somehow saves the world from evil and death. We believe things whose sheer implausibility to plain common sense cannot easily be underestimated. Undermining the almost universal supposition that Christianity is incompatible with contemporary science merely removes a spurious obstacle to faith. More than enough genuine difficulties remain. Considering how they might be overcome is beyond the scope of this book. Yet I cannot avoid hoping that the case I make for the idea that this, the real world that scientific naturalism describes, is just the sort of world we might expect, given the truth of Christianity, will motivate the secular reader to see the Christian faith as worth investigating.
Christianity, Naturalism, and Science
13
Richard Dawkins, in the God Delusion, takes pains to show that the belief that there is a God is a scientific hypothesis. 27 In this he is, I believe, mistaken: science is the project of delineating the laws of nature and using them to explain observed reality. Science as such makes no claims about God or any other non-empirical reality. 28 What Dawkins does make plausible is something more general: belief in God has empirical content. As he said elsewhere, “A universe with a God would look quite different from a universe without one. A physics, a biology where there is a God is bound to look different.” 29 From this he concludes that there is no God. The further premise, left implicit, perhaps because he finds it too obvious to make explicit, is that this universe, the one portrayed by contemporary science, does not look like it was created by God. Dawkins’ two premises, if true, entitle us to conclude that this is a universe that has no God. But is the premise left implicit true? The Material Image brings this unexpressed premise into view and challenges it. It aims to show that the universe revealed to us by science is what we could reasonably have expected if it is God’s creation. 30 This challenge is narrowly focused. Dawkins was right to judge that the universe known to us by means of modern science does not look much like the creation of God as generally conceived. Human religiosity, in some of its prevalent theistic manifestations, conceives a deity who creates for the sake of command and control, bringing into being a universe in which everything is a product of divine design, and who relates to personal creatures largely with a view to punishing the wicked and rewarding the good in a disembodied afterlife. With respect to such hypothetical creators, there is little reason to disagree with Dawkins. This world we know most reliably through science does not look like something such a deity would have created. However, what might be unlikely or simply imponderable on generic theistic assumptions, e.g., material persons created by indeterministic means, need not be unlikely in light of the specifics of the Christian confession. This is the kind of universe Christians could have reasonably expected to discover. STRATEGY The Material Image asks, “Given that the God who created the universe is the God revealed in Jesus Christ, the God who seeks a relationship of trust and love with creatures, what might we reasonably expect creation to be like?” Bringing into focus specifically biblical, Christian claims about God’s desire for fellowship with created persons, we find it is the naturalistic, rather than the traditional, account that is most reasonable. An organizing theme of this book, and the source of its title, is the biblical idea that human beings are made in the image of God. The term
14
Chapter 1
material image focuses on the naturalistic claim, amply supported by science, that human beings are material things. 31 Acknowledging that this is what we are helps us make sense of ourselves as made in God’s image, and knowing that we are made in God’s image helps us make sense of ourselves as material beings. Many Christians suspect that much of what science tells us about what we are is opposed to our being created imago Dei. For them, this means we are not mere material beings, not products of blind evolution, not governed by natural laws, not completely open to scientific investigation, and that we are endowed with ways of knowing unfathomed by science. Christians, among others, have long been concerned to fend off what they perceive as an insufficiently exalted view of human nature drawn from science, one that renders us too much creatures of this material world, too similar to the animals and not sufficiently godlike. They think the attempt to make human beings the object of scientific inquiry subverts our true value and dignity. Christians often see themselves as holding off a scientifically led onslaught that threatens to dehumanize us. 32 Although not often well grounded in an understanding of the theories and research projects under attack, these fears are not groundless. It is possible for a scientific understanding of human nature to inspire a picture of what we are that fails to do justice to what is special and valuable about us. Christians, believing that humans are made in God’s image, are rightly sensitive to this possibility. However, our image of ourselves can be amiss in another way. We can have too high a view of human nature, one that aggrandizes our selfimage. The concern to avoid this, not to make more of human beings than we actually are, is also a theme in the biblical tradition. Indeed, the idea that we are so proud as to forget that we are not gods but God’s, God’s creatures, lies deep in Christian theology. Its biblical roots go back to the story of the temptation in Eden, where the lying promise of becoming like gods sets us on our course of rebellion and alienation from our Creator. 33 Rowan Williams observes that we manifest a “deeply rooted aversion to our own creatureliness.” 34 I suspect that we are more prone to overestimate than to underestimate ourselves, more tempted by metaphysical self-congratulation than moved by proper creaturely humility. We can properly welcome the humbling revisions in the human selfimage due to contemporary science as more in keeping with our faith. PLAN OF THE BOOK In chapter 2, Knowledge, I refine and defend the first principle of scientific naturalism, which is that science is our preeminent way of knowing and as such constrains all other ways of knowing. I sketch a naturalistic account of human knowledge, one shaped by the fact that our minds are merely our brains, the products of mindless natural selection. Here, as in
Christianity, Naturalism, and Science
15
the later chapters, I set out not the only possible naturalistic account, but the one that seems to me the most emphatically naturalistic. In contrast to earlier conceptions of knowledge, naturalism regards knowledge as the product of reliable causal mechanisms. The most reliable way for us to acquire knowledge of the world and of ourselves is by using our senses. The methods of the sciences, embodying the most critical and careful use of our senses, constitute our most reliable way of knowing. Beliefs that conflict with the well-confirmed theories of science are unreasonable. However, science is not our only way to know. Nor is it an adequate way to know many things that matter to us. We know much by less reliable means. Naturalism’s claim for the status of scientific knowledge does not preclude the possibility of knowing the God who transcends the natural world accessible to scientific inquiry. The naturalistic view of knowledge abandons the ancient quest for certitude. We cannot prove that the neural machinery that produces belief in us is reliable. Science, relying ultimately on the senses, cannot prove their reliability. There is no first philosophy, no place to stand to establish the reliability of science. There is no foundation for knowledge of ourselves and the universe more secure than the reliable yet fallible practices of the sciences. As disconcerting as this might be for our traditional self-image, it is consonant with our being creatures called not to strive for certitude, but to live by faith in our Creator. Christian faith depends crucially on the belief that God has intervened in the creation, bringing about events that would not have occurred in the natural course of events. Belief in miracles is generally regarded as blatantly incompatible with a naturalistic point of view. However, in chapter 3, Miracles, I argue that a belief in the miraculous is entirely consistent with a robust naturalism. I will show that the reasoning that can justify belief in divine intervention is analogous to a type of reasoning crucial to progress in science: the reasoning that leads us to suspect that a hitherto well-confirmed theory is, after all, false. Belief in the miraculous is compatible with scientific naturalism, but only so long as we make a principled distinction between God’s action in nature and in history. Christian theology offers good reasons to make this distinction, and to believe that God intervenes in human affairs. The constraints naturalism imposes on what human beings can know leave open the possibility of justifying Christian belief, although they make it difficult, and perhaps impossible, to construct a decisive defense that compels all reasonable persons to agree. The humility this calls for is entirely appropriate for those whose faith is in the God so humble as to become one of us. Chapter 4, Origins, turns to the fact that human beings are the product of biological evolution in contrast to the still widely held belief that God miraculously intervened in the creation to bring human beings into existence. Evolutionary biology’s explanation of human origins challenges entrenched convictions about how God created the human species, as
16
Chapter 1
well as cherished ideas about human uniqueness and what it means to be created in the image of God. On reasonable assumptions, Darwinian evolution implies that God did not specifically intend the existence of the human species. It is incompatible with our Creator having designed us. Divine design is relegated to the most primal initial conditions and most fundamental laws of the creation, not to its specific contents, not to the persons who inhabit it. This separation of creation from design, though unfamiliar, is all but inevitable in light of crucial Christian claims about God’s creative aims. God did not design the human species, but our existence answers to the purposes, revealed in Scripture, for which God created the world. Indeed, given those purposes, it is reasonable to expect that God would create by means of the sort Darwinism describes. Chapter 5, Mind, sketches the naturalized understanding of the human mind as the functioning, embodied, socially situated brain. My aim here is to undermine the dualist conception of human beings still closely associated with Christian faith and to show the reasonableness of a materialist theory of the human person. Despite its association with the idea that we are spiritual beings, there is little in the metaphysical theory that our minds are immaterial to recommend it to Christians. Attempts to defend belief in immaterial minds or souls draw upon a conception of what it means to be made in the image of God that we should reject on theological grounds. Scientific exploration of human cognition and consciousness deflates the pretension that we are godlike, transcendent beings, and it coheres with the confession that we are the material image of the immaterial God. I criticize the hope that some concept of emergence can mitigate the impact of current science on traditional conceptions of the human person. Some thinkers concerned with faith’s relation to science make much of the fact that the human mind is not reducible to the brain, but no safety for the traditional human self-image is to be found in this irreducibility. It points instead to the ontological superficiality of the human person. We are not found among the fundamental constituents of the world but exist only when contingent, indeed chancy, processes assemble the right components in certain ways. This is not to denigrate human beings, but to see our special place in creation not as a matter of how we are made, but of who made us and why. The sixth chapter, Freedom, pursues the implications of our being material beings for freedom and responsibility. Whatever freedom and responsibility we possess must be consistent with our being completely embedded in the natural world and governed by its causal laws. The compatibilist approach, which supposes that our choices can be both free and caused, allows us to save much of what matters in our image of ourselves as free and responsible agents. However, compatibilism cannot fully reconcile our being material beings with common conceptions of responsibility. Indeed, no created thing, material or otherwise, could be
Christianity, Naturalism, and Science
17
responsible in the way the traditional notion envisages, for it requires that we be agent causes, something no material thing can be. However, naturalism together with a compatibilist conception of freedom affords a kind of freedom and responsibility appropriate to creatures insofar as it leaves open the possibility that we are, or can become, material simulations of agent causes. This modified conception is, I argue, consonant with the Christian faith’s confession that God addresses humans as responsible persons yet assumes final responsibility for creation. Some theorists look to the quantum domain, or to downward causation, as a ground for a more radical freedom, but I suggest that these strategies hold little promise. The indeterminism of nature’s fundamental quantum mechanical laws is of no use in salvaging traditional conceptions of freedom and responsibility, but this is a matter of crucial significance for understanding God’s action in the world by way of natural, secondary causes. Authoring indeterministic laws, the Creator makes possible creatures who are not mere extensions of their maker, but distinct persons, capable of responsibly responding to and interacting with God. Chapter 7, Morality, outlines the scientific explanation of morality as originating in evolutionary adaptations to social life. Again, we see scientific naturalism at odds with views long associated with Christianity. Here it challenges the idea that in our moral experience we hear divine commands, or make contact with a transcendent realm of values, exercise some sort of pure, disembodied rationality, or at least encounter a realm of moral fact ultimately independent of human beings. I consider the question of whether science threatens to debunk morality, denying its objectivity and authority, and contend that while this scientific explanation humbles our conception of ourselves as moral agents, it does not abolish it. There are good theological reasons to accept this chastened conception of human morality. Current science gives us no reason to see the moral truth as relative to the individual or to society, but it implies it is less objective than most Christians assume it must be. Science puts morality in its place, characterizing it ultimately as human, not divine, and this the Christian should endorse. The implied conception of morality, as something merely human to which God condescends, and makes critical use of, is to be recommended on theological grounds. This chapter also addresses the biblical idea of the Fall, and the correlative idea of original sin, which have long been understood in moralistic terms. I speculate on better ways to construe these doctrines and I argue that an understanding of them consistent with a naturalistic account of human origins is reasonable from the point of view of a Christian faith that rejects popular, but biblically and theologically untenable, conceptions of sin as moral wrongdoing. Chapter 8, Religion, describes the current attempt to explain human religiosity as a by-product of the mind’s evolved cognitive architecture.
18
Chapter 1
In contrast to the naturalistic explanation of morality as an adaptation to social life, the explanation of religion on offer here is a debunking explanation, one that threatens to undermine the beliefs it explains. Can Christians consistently accept this general explanation without claiming an unjustifiable exception on behalf of our own beliefs? I suggest that we can, but only if we see the Christian faith’s relation to religion as analogous to its relation to morality. God condescends to our natural religiosity, shaping it for redemptive purposes, but also subjecting it to critical scrutiny and finally putting it, along with morality, in its proper, merely human place. The ninth, and final, chapter, Resurrection, enters into eschatological matters, both personal and cosmic. The principal objection to the materialist view of the human being, put on hold in chapter 5, is that the Christian belief in life after death requires that we have immaterial souls that survive the destruction of our bodies. I present ways to understand the identity through time of material persons that make reasonable the Christian hope for their miraculous resurrection for an everlasting future with God. There is no need to resort to an immaterial soul and its theologically problematic natural immortality to defend the possibility of resurrection. The Christian hope is not, I point out, threatened by the alleged tedium of immortality or the finitude of resurrected persons. Turning to cosmic considerations, I speculate on the vocation of God’s material image as the created co-creator. WITHOUT ILLUSIONS Alex Rosenberg, in his delightful book, The Atheist’s Guide to Reality, seeks to show his fellow atheists the path to a life free of the illusions that science dissipates. 35 For him, the greatest illusion is belief that God exists. This conviction is grounded, at least in part, in the conviction that science renders that belief unreasonable. My task is to show that whatever the ultimate case may be with respect to justifying belief in God, Rosenberg’s assumption is itself unreasonable. If belief in God is unreasonable, it is not in virtue of conflict with the news from science about what humans can know, how we came into existence, or what we are. I do, however, share Rosenberg’s view that we ought to welcome the scientific unmasking of old and comforting illusions about our origin, nature, and capabilities. Too often, the faithful resist, evade or deny what science reveals. The case I make in this book is that this is misguided, and that Christian faith has no good use for the illusions science prompts us to discard.
Christianity, Naturalism, and Science
19
NOTES 1. John Polkinghorne, the physicist and Anglican priest, expresses the concern when he writes, “Broad general ideas are attractive (a Divine mind behind the order of the universe), but I believe that theism only becomes persuasive when it is elaborated in greater detail and when it is anchored in the experience and interpretation that are preserved and propagated within a religious tradition.” Science and the Trinity: The Christian Encounter with Reality (New Haven, CT: Yale University Press, 2004), xii. 2. Blaise Pascal made the contrast on a scrap of paper sewn into his coat and found after his death, his “Memorial” of his profound religious experience of 1654. 3. George Santayana describes the world as vast, ancient, and recalcitrant in Scepticism and Animal Faith: Introduction to a System of Philosophy (New York: Dover Publications, 1955), 191. 4. Thomas Nagel, The View from Nowhere (New York: Oxford University Press, 1986). I believe that the term “God’s eye point of view” originates in Hilary Putnam, Reason, Truth and History (Cambridge, UK: Cambridge University Press, 1981), 49. 5. The earliest appearance of this sentence I know of is in Jacques Pohier’s God-In Fragments (London: SCM Press, 1985), where he uses it as a title of one section of the book. 6. In turn, faith understands the incarnation, resurrection, and ascension of Christ, as well as creation, in light of the cross. In the crucifixion of Jesus God’s nature is decisively revealed. In it who God freely chooses to be in relation to us is enacted. “Crux probat omnia:” the cross tests everything. 7. Pale Blue Dot: A Vision of the Human Future in Space (New York: Random House, 1995), 52. 8. Ted Peters uses consonance to describe the relation of faith to science in his preface to Ted Peters, ed. Cosmos and Creation: Theology and Science in Consonance (Nashville, TN: Abingdon Press, 1989), 13–17. 9. Alan Lacey in Oxford Companion to Philosophy. Ted Honderich, ed. (Oxford, UK: Oxford University Press, 1995), 604. 10. Naturalism Without Foundations (New York: Prometheus Books, 1996), 25. 11. Ted Peters, ed. Cosmos and Creation: Theology and Science in Consonance (Nashville, TN: Abingdon Press, 1989), 14. 12. Alvin Plantinga and Michael Toole, Knowledge of God (Malden, MA: Blackwell, 2008), 19. 13. Stewart Goetz and Charles Taliaferro, Naturalism (Grand Rapids, MI: Eerdmans, 2008): 6, 8. 14. In contrast the term naturalistic is commonly used in contemporary philosophy to refer to theories grounded in a scientific view of the world, rather than in a priori reasoning, without reference to the question of God’s existence, e.g., the idea that the human mind in some sense is the embodied, socially situated brain is a naturalistic theory. This book’s naturalism is consistent with this use. 15. Walter R. Thorson defends methodological naturalism from a biblical perspective in The Woodpecker’s Purpose: A Critique of Intelligent Design (Wenham, MA: Gordon College Center for Faith and Inquiry, 2014), 47–59. Methodological naturalism becomes controversial once we move beyond the default assumption, made even by those of us who have a firm belief in the miraculous, that events have natural explanations, to the stipulation that belief in the supernatural can have no legitimate effect on scientific inquiry. For discussion of methodological naturalism and its limits see Del Ratzsch, Science and Its Limits: The Natural Sciences in Christian Perspective (Downers Grove, IL: InterVarsity Press, 2nd ed. 2000), 122–29, and Alvin Plantinga, Where the Conflict Really Lies: Science, Religion, and Naturalism (Oxford, UK: Oxford University Press, 2011), 168–74. 16. Henceforth my use of the term naturalism abbreviates scientific naturalism. 17. The case for this appears in chapter 9.
20
Chapter 1
18. Robert W. Jenson, Systematic Theology Volume I: The Triune God (New York: Oxford University Press, 1997), 202. 19. Christians willing to embrace the term naturalism sometimes do so only after dispensing with supernatural elements of traditional Christian belief. For some, this involves rejection of a transcendent deity; see, e.g., Charley D. Hardwick, Events of Grace: Naturalism, Existentialism and Theology (Cambridge, UK: Cambridge University Press, 1996). For others, closer to traditional Christian views, embracing naturalism involves rejecting the miraculous while retaining belief in the God who transcends creation; see, e.g., Arthur Peacocke, All That Is: A Naturalistic Faith for the Twenty-First Century (Minneapolis: Fortress Press, 2007). 20. See, e.g., Willem B. Drees, Religion, Science and Naturalism (Cambridge: Cambridge University Press, 1996). 21. I follow, somewhat reluctantly, what has become the standard practice of refraining from the traditional use of using masculine pronouns to refer to God, hoping that the reader will not be tempted to conflate the quite particular Deity of Christian faith with something generic or impersonal. That God is a personal agent who interacts with created persons is this book’s central claim. 22. I do not imagine that the biblical narratives are typically history in the modern sense, but I do take the Bible to be an accurate, theologically interpreted, account of God’s action in history and of the divine nature, character, and purposes revealed in that action. For an examination of the methodological issues involved in this conception of Scripture see James K. Mead, Biblical Theology: Issues, Methods, and Themes (Louisville, KY: Westminster John Knox Press, 2007). 23. The Wound of Knowledge: Christian Spirituality from the New Testament to Saint John of the Cross (Cambridge, MA: Cowley Publications, 2nd ed. 1990), 23. 24. Phillip Johnson and William A. Dembski are probably the most well-known proponents of this view, e.g., Johnson’s Reason in the Balance: The Case against Naturalism in Science, Law and Education (Downers Grove, IL: InterVarsity Press, 1995), and Dembski’s No Free Lunch: Why Specified Complexity Cannot be Purchased without Intelligence (Lanham, MD: Rowman and Littlefield, 2nd ed. 2007). A more subtle version appears in Alvin Plantinga, e.g., “Methodological Naturalism?” in Robert T. Pennock, editor. Intelligent Design Creationism and Its Critics: Philosophical, Theological, and Scientific Perspectives (Cambridge, MA: MIT Press, 2001), 339–61. 25. The nudge I have in mind comes in two phases. First, the presentation of Christianity as something reasonable human beings could hope is true, something we can hear as good news. However, we cannot reasonably hope that something is true when it conflicts with what we are convinced is true, such as the well-confirmed theories of science; the most we can do is wish. Thus, the second phase is making the case that there are no such conflicts, indeed that the scientific picture of the world and humanity’s place in it is what we should expect if Christianity is true. Reasonable hope makes it reasonable to be carefully open to evidence that what one hopes for is true. I offer that evidence in chapter 3. 26. The Big Question: Why We Can’t Stop Talking about Science, Faith and God (New York: St. Martin’s Press, 2015), 20. 27. The God Delusion, chapter 3, “The God Hypothesis” (New York: Houghton Mifflin, 2006). 28. Conjoined with other, non-scientific claims about God, scientific claims do have implications about God, e.g., God created the world, conjoined with e= mc2, implies God created a world in which e = mc2. 29. Debate with Archbishop of York Rt. Rev. John Habgood at the Edinburgh Science Festival, 1992. Quoted in Mark Johnston, Saving God: Religion After Idolatry (Princeton, NJ: Princeton University Press, 2009), 46. 30. Is Dawkins’s other premise true? A sufficiently generic conception of deity might leave us with nothing to say about what sort of universe to expect. But so far as the God of Christian faith is concerned, there is a great deal to say, since that is a God
Christianity, Naturalism, and Science
21
who would create persons distinct from their Creator, choose to become empirically revealed to them, and enter into the creation as one of them. 31. On the contemporary scene it is common to speak of “physicalism” rather than “materialism,” in virtue of the worry that the latter term harks back to an outmoded scientific picture of reality as at bottom composed of tiny material objects, rather than the forces, fields, and particles of modern physics, which bear little resemblance to microscopic versions of macroscopic material objects. While “physicalism” is preferable to describe the overall nature of the created world, since physics fixes all the (natural) facts, I describe human persons as material objects, and humans as God’s material image, in view of the fact that we, like stones, cats, and trees, middle-sized improbable objects, are part and parcel of the natural world of human experience, not among the world’s basic constituents. 32. C. S. Lewis, still extremely influential in conservative Christian circles, was a mid-twentieth century example of this. 33. In light of what ensues for humanity, I suspect sad divine irony, mocking the Serpent’s empty promise, in Genesis 3:22: “Then the Lord God said, ‘See, the man has become like one of us, knowing good and evil…” What the story’s human couple learned was that they are naked, vulnerable creatures, not at all godlike (3:7). 34. “On Being Creatures” in On Christian Theology (Malden, MA: Wiley-Blackwell, 2000), 77. 35. The Atheist’s Guide to Reality: Enjoying Life without Illusions (New York: W.W. Norton and Company, 2011).
TWO Knowledge
THE FIRST PRINCIPLE OF SCIENTIFIC NATURALISM Scientific naturalism, as I will use the term, is the view that in a narrow but important sense science is our best way of knowing, and that by way of science we know that we are material beings, the result of unguided biological evolution. Of these three tenets of scientific naturalism, the primacy of science as a way of knowing underlies the other two and takes precedence. What we can call the first principle of naturalism is that what scientific inquiry tells us about the world and ourselves constrains all other claims to know. We ought to believe the well-confirmed theories of the sciences, and we ought not to believe things inconsistent with them. In Arthur Peacocke’s words, “the scientific account of the natural world, including human beings, is the best-established account of the realities in which we are embedded.” 1 That human beings are material things and the product of biological evolution, are (very general) scientific theories confirmed by a great deal of independent but converging evidence. Science is our most reliable way of knowing ourselves and the world we inhabit, but it is not infallible. Like any scientific theory, no matter how well-confirmed, these ideas about the nature and origins of human beings could be wrong. Therefore, naturalism exhibits a kind of contingency: were science to reject evolutionary theory or the materialist account of the human person, those of us who acknowledge the primacy of science as a way of knowing would abandon naturalism. However, neither theory is currently subject to actual reasonable doubt. We cannot in the long run rationally believe what conflicts with science. If Christianity were logically inconsistent with evolutionary biology, the materialist account of the human person, or any other genuine deliverance of scientific inquiry, that would be sufficient 23
24
Chapter 2
reason to conclude that it is false. Later chapters of this book seek to dispel the persistent illusion that Christian belief really does conflict with such well-confirmed scientific theories and to show that, in fact, naturalism is consonant with the Christian faith. The world science reveals is the kind of world Christians should expect to discover. Before moving to those issues, I address the relation of Christian belief to the first principle of naturalism itself. Christians claim to know things that are not known to science, and we must ask whether these claims are reasonable, given the constraints naturalism imposes on human knowledge. I devote this chapter to showing that, when properly qualified, the first principle of naturalism is obviously true, and Christians have no particular reason to object to it. OUR MOST SECURE KNOWLEDGE The best way humans have to acquire knowledge of the world and its contents, including ourselves, is science: physics, chemistry, biology, and the other natural sciences, as well as the social sciences to the extent that they are integrated into the overall project of the natural sciences. Science is not merely one way of knowing among others. Its methods are more reliable than those deployed in other areas of human life. Devising hypotheses with precisely testable empirical implications, science prods nature, reticent to disclose its secrets, to speak for itself, even if never absolutely definitively and unequivocally. Everywhere else, including most matters of human significance, we hear not nature, but one another, doing our best to justify what we say. With whatever care our questioning proceeds, we are always speaking for, not hearing, the reality that can make what we say true. Once we ascertain that a scientific hypothesis has become well-confirmed, that is, that there is, in scientific terms, adequate evidence for it, then we ought to believe it. For example, we ought to believe that there are exoplanets orbiting stars other than the Sun because we read in Scientific American that astronomers in the relevant subdisciplines have reached consensus that this is the case. This deference to science rightly extends to things we cannot imagine and can barely understand, to quantum tunneling, black holes, and gravity waves. It extends beyond belief to practice. We rely, in our daily lives, even in matters of life and death, on technologies that can be devised and reasonably relied on only in light of scientific theories. We send people to prison on the basis of DNA lab results, even in the face of otherwise credible exonerating eyewitness testimony. All this contrasts with our attitude toward the professional pronouncements of non-scientists. No one reasonably grants this degree of credence to the assertions of politicians, economists, generals, business
Knowledge
25
leaders, literary critics, theologians, or philosophers, irrespective of their honesty or their competence in their respective fields. We must of course distinguish what any particular scientist asserts from what we are rationally compelled to believe because it is what science says. The critical reasoning that constitutes scientific practice is embodied in social institutions. We ought to heed the voice of a community of primary researchers and peer reviewers speaking within its narrow area of expertise. The critical consumer of scientific knowledge takes care to distinguish well-confirmed theories, theories that have held up to rigorous testing to such an extent that they have won the allegiance of the appropriate community of investigators, from hypotheses that are merely speculative, or not yet well confirmed. The superiority of science as a source of knowledge is due to its methods, but its methods are due not to the unique virtues of scientists, but to the nature of its explanatory aims. Seeking the causes of natural phenomena, it advances hypotheses that can be cast as hypothetical propositions that have precise observational implications and are thus subject to systematic testing. Hypotheses are disconfirmed and discarded, or confirmed and eventually rendered too probable for reasonable disbelief; they become well-confirmed. The scientific project is our best attempt to let nature speak for itself, telling us what to believe about it. This is not to say that it necessarily speaks unambiguously; inescapably, fallible human judgment comes into play in the interpretation of experimental results. 2 Yet it is in science that we most closely approximate to the ideal of conforming belief to objective reality, rather than imposing our preconceptions and wishes upon it. For some, acknowledging the power of science represents a proud ambition to demystify, disenchant, and gain mastery over the world. But at its heart, science is an exercise in intellectual humility. NATURALISM, NOT SCIENTISM Naturalism’s high view of the authority of science so often elicits the charge of scientism, the notion that science is the only way of knowing, that it is worth dwelling on its denial. Bertrand Russell in 1935 asserted, “Whatever knowledge is attainable, must be attained by scientific methods; and what science cannot discover mankind cannot know.” 3 This bold assertion is consistent with scientific naturalism, but surely not required by it. This is fortunate, since it is plainly false. Science is not merely one way of knowing among others but it is one way of knowing among others. Most of what is worth knowing can be known only by means independent of science. We know, for example, that Vicksburg finally fell to Grant in 1863, that Mozart’s music is better than that of the Monkees, that it is morally wrong to suffocate your roommate to make him stop
26
Chapter 2
snoring, that Juneau is the capital of Alaska, that the mythological Juno was Jupiter’s wife as well as his sister, that Flannery O’Connor’s stories manifest a Christian perspective on the world, that Citizen Kane influenced film noir, that you should not eat yellow snow, that the cafeteria closes early on Sundays . . . and that we should not believe things that contradict well-confirmed scientific theories. We know all these things, but none is an instance of scientific knowledge. The last, the first principle of naturalism itself, is a philosophical view, not a scientific theory. Science has little or nothing of direct relevance to say about vast areas of human knowledge, knowledge without which human life would be neither possible nor worthwhile. Even when science overlaps other ways of knowing, we cannot assume that the most useful, interesting, or important knowledge can be acquired scientifically. We might learn things about the natural world by contemplating Cezanne's watercolors, reading the poetry of Denise Levertov, tending a kitchen garden, or by climbing Mount Katahdin that we cannot acquire from the natural sciences. More valuable insight into the workings of the human mind might be acquired by reading novelists like Jerzy Kosinski and Iris Murdoch, or by studying the history of the Chan Dynasty, or by reading Aristotle's Nichomachean Ethics, than one gleans from a scientific psychology whose explanatory theories uncover the causal mechanisms of the mind/brain. As many have observed, there is more wisdom about human behavior to be had from Henry James than from his brother William. Acknowledging the superiority of scientific methods for justifying belief does not commit us to the absurd claim that science can do the cognitive work done by other ways of apprehending ourselves and the world. It is easy to lose sight of how vast and varied the intellectual landscape is, and to forget how many different things about the world, and about human beings, there are to know and try to understand. We can hear the assertion that natural science can explain many things as the different, and false, assertion that it can explain most, or even all, of what needs to be explained. This is abetted by the tendency of those engaged in a scientific research program to pronounce hopefully—but not always realistically—on its potential while ignoring the many things they have no hope of explaining. There is much to learn by means of scientific inquiry, but also much we can learn only by other means. Scientific naturalism requires that genuine knowledge has a home in the world science describes, not that there is no knowledge other than science. However, this imposes a further constraint on what we can reasonably believe. While there are things we know by methods other than those of science, we reasonably expect science to be able to explain how humans, material beings produced by biological evolution, are capable of acquiring knowledge in these ways. The naturalist supposes, for instance, that there can be a scientific explanation of how human beings can know
Knowledge
27
that it is wrong to cause animals pain for fun, but he does not assume that this would render moral knowledge a kind of scientific knowledge. For the naturalist, science explains why we can have non-scientific knowledge, but it does not assimilate other ways of knowing. Naturalism does not claim for science a monopoly on knowledge, but it does insist that in the event of conflict, when a highly-confirmed scientific theory and some other belief cannot both be true, science trumps its competition. This is the practical import of science being our most reliable way of knowing. It is unreasonable to accept any belief, however obvious it appears, however cherished it is, that contradicts what scientific inquiry clearly confirms. Beliefs might be ancient, deeply entrenched, widely held, delivered by revered authorities, and profoundly comforting, but they should be rejected if they conflict with science. Even beliefs that are sheer common sense, or evident to the senses, even beliefs about one’s own mind acquired by introspection, are subject to revision and refutation as a result of scientific investigation. This claim about the primacy of science as a way of knowing is not extravagant. On the contrary, it comes close to being a truism. This is not because science embodies a unique kind of reasoning remote from our common cognitive practices. We acquire our knowledge of the world by using our senses and reasoning about what we sense. Scientific reasoning is doing this to the best of our ability. The claims of science deserve to be believed because in science we most squarely face our fallibility as knowers and in so doing come by our most reliable way of finding things out. As Albert Einstein once said, “all of science is no more than an elaboration of everyday thinking.” 4 It is in science as a social institution that candidates for belief are most carefully scrutinized and their credentials critically tested. The reliability of scientific methods contrasts with other ways of knowing, and this constrains what we can reasonably claim to know by other means. This is not because science is infallible; it is simply less fallible than other ways of knowing our world. The claim for the primacy of science as a way of knowing this world is problematic for some, but it should not be much more newsworthy than the fact that mathematics as a way of knowing trumps all competitors, including science. The naturalistic claim for the primacy of science is restricted in scope, applying only to this contingently existing, empirically knowable, world, not to the domain of mathematical truth, which is neither contingent nor empirically known. It is never reasonable to believe what is inconsistent with the theorems of mathematics. Were, for example, a theologian to devise an account of the Trinity that implies that 1 = 3, we would have no difficulty dismissing the claim as mistaken. We would not regard this as an illicit privileging of mathematics over faith or theology. Whatever understanding we might have of the Triune nature of God, or of anything else, is properly constrained by mathematics. This extends to the claims of science; any scientific hypothesis discovered to
28
Chapter 2
conflict with mathematical truth is ipso facto disconfirmed, with no need for empirical testing. The superiority of mathematical reasoning over other means of justifying belief is accepted with near-universal equanimity, even in the context of our most important beliefs. We take for granted the infallibility of mathematics and get on with our business, feeling no need to deny the obvious in order to safeguard our other, fallible, ways of knowing. The same should be true of the sciences: we should take their superior reliability for granted and go on, feeling no need to deny the obvious so as to safeguard other, less reliable, ways of knowing. RATIONAL FAITH We should avoid beliefs that are at odds with science. We can justify no exception when it comes to matters of faith. We have better reasons to accept what science says than we have to accept the doctrines of Christianity or any other religion. There is, for example, stronger justification for believing the Big Bang theory than that God created the universe. We have better evidence for the belief that the human species came into existence by way of biological evolution than for the belief that God created human beings imago Dei. The quality of the evidence for Christian belief is not as good as the evidence for the well-confirmed theories of science. One need only articulate one’s reasons for accepting, for example, the belief that God exists, that the Bible is God’s revelation, or that Jesus was resurrected, with the evidence that supports the atomic theory of matter, general relativity, or the theory that DNA is the vehicle of heredity, to be compelled to admit that so far as the rational consideration of evidence goes, scientific belief is more securely grounded. Reasonable persons who examine the evidence for well-confirmed scientific theories uniformly come to believe them, while reasonable persons can examine the evidence for Christianity and fail to accept it. Good scientific evidence compels belief in ways that even our best theological arguments do not. Crucially, this does not imply that the evidence for Christian belief is not good enough, but it does imply that it would be unreasonable to maintain Christian belief if it did conflict with the well-confirmed theories of science. If we were to encounter inconsistency between faith and science it would ultimately be incumbent upon us to believe the scientific results and give up our Christian beliefs. Admitting that Christian beliefs are less strongly warranted than scientific claims with which they could conflict might seem to belie a defective faith, as though faith properly involves not only having particular beliefs, but a further conviction about the relative strength of one’s reasons for them. I suggest that this rests upon the mistaken assumption that there is a connection between the importance of a belief and the strength
Knowledge
29
of its justification. A Christian’s theological beliefs are among the most important she has. They matter most to her. They are among those she is most concerned to understand, those whose grounds and implications she is most concerned to discern. They are central among the beliefs with which she organizes her experience and orders her life. They are among the beliefs that the loss of which would drastically change her life, casting her into confusion or despair. An encounter with even apparently falsifying evidence would be traumatic, placing serious demands on her intellectual integrity. It would be comforting to be assured that such important beliefs are the most strongly warranted, not vulnerable to being refuted by science, history or anything else. This is not true only of religious beliefs: a man’s conviction that his closest friends are honorable and that his wife loves him are much more important to him than his belief that the Earth is spherical, orbits the Sun, and is billions of years old. Nonetheless, given what he knows of the sad experience of human beings, honesty requires him to admit that it is the personally more important beliefs that are more likely to be overturned by experience. He knows that others, whose evidence and degree of subjective certitude were comparable to his own, turned out to be mistaken on such vital matters. We might well wish the beliefs Christian faith presupposes were more, rather than less, securely anchored than those of science but, while we might be so blessed as to escape an excess of actual doubt, we cannot reasonably hope to secure faith beyond the possibility of falsification by science. EMPIRICAL CONTENT There is a temptation to insulate faith from possible refutation by science by construing it as making no claim about what the world is like. Stephen J. Gould defended this approach in his book Rocks of Ages where, contending there can be no conflict of religion and science, he writes: Science tries to document the factual character of the natural world, and to develop theories that coordinate and explain these facts. Religion, on the other hand, operates in the . . . utterly different realm of human purposes, meanings and values—subjects that the factual domain of science might illuminate, but can never resolve. 5
This approach attracts those who are sympathetic to religion but convinced that as traditionally conceived it commits us to patently false beliefs about the world. However, to accept Christianity is not simply to embrace a certain set of values, or to adopt a certain attitude toward life, although it has practical implications of this sort. It makes factual claims about the world. It has empirical content. Christian faith “goes out on a
30
Chapter 2
limb,” committing us to beliefs vulnerable to empirical experience. Christianity is falsifiable. To acknowledge that a belief is empirically falsifiable is not to say that it will be refuted, or that it is likely to be refuted. It implies nothing about the quality of the evidence for it or about one’s confidence in its truth. In believing, for instance, that President Kennedy was assassinated in 1963 we believe something falsifiable, but to acknowledge this is not to express doubt. To believe that Kennedy was assassinated in 1963 is to be committed to the world being one way rather than any way that conflicts with his being murdered in 1963, and thus makes us vulnerable to being wrong. We can conceive of evidence that would show this belief to be false, as utterly surprising as that would be. Genuine beliefs about the world bear the theoretical risk of falsity, however negligible in practice. Most obviously, the Christian faith is subject to falsification by historical evidence: the bones of Jesus of Nazareth, unearthed in the environs of Jerusalem, would show that Christianity is false, at least the traditional Christianity that confesses Jesus’ bodily resurrection and informs this book. Christian faith is first and last trust in the God manifest in Jesus Christ, not a theory of the universe, but it has implications about how things are, claims vulnerable to being falsified by empirical experience in general and science in particular. REAL VULNERABILITY Reasonable Christians will, on reflection, accept conflict between their faith and science as an abstract possibility. What many find unsettling is the conjunction of this with the idea that we ought not to believe things that conflict with well-confirmed scientific theories. For many religious persons, not only Christians, to label a belief a matter of faith is to privilege it as somehow trumping all beliefs with which it conflicts. I see no reason to indulge in this. Christianity could conflict with science and if it did, we should believe what science says. However, one aspect of being a reasonable human being is a kind of intellectual conservatism. We ought not abandon important, long-held beliefs simply because we encounter a problem we do not know how to solve, or some evidence that appears to overturn them. All honest, informed persons, no matter what their beliefs, live with unresolved difficulties. Honesty and rationality demand that we admit, and so far as we are able attend to, the difficulties with our beliefs, but not that we hastily abandon them. Yet integrity ultimately could require us to admit that our most dearly held beliefs are simply false. The attempt to rule this out as impossible unmasks faith as ideology, a body of beliefs held without genuine regard to whether they are true.
Knowledge
31
In our cultural context many persons suppose that science has already produced knowledge that refutes essential Christian assertions. I continue to be a Christian, believing that while science could refute Christianity, as a matter of fact it has not and will not. Indeed, I believe that science increasingly portrays a world profoundly consonant with a Christian view of creation and of human creatures, and that this becomes clear when we embrace, rather than evade, the naturalistic implications of contemporary science. However, part of Christian faith is trust in God with one’s mind. This calls for recognizing the vulnerability of our belief in God and living with it, while at the same time being truthful about the quality of the evidence for our beliefs, even when it is not as good as we might like. Persons of faith sometimes oscillate between two errors. We are prone on the one hand to acknowledge that our theological beliefs call for rational justification, but then to exaggerate the quality of the evidence for them, portraying it as though it were as good as, or even better than, that for more mundane beliefs, such as those derived from scientific inquiry. When this strategy’s implausibility becomes obvious, we are tempted to reject the idea that our beliefs require evidence, and treat faith as somehow beyond the demand for rational justification. Alternatively, we suddenly find attractive a relativist conception of rational justification and decide that the community of faith is entitled to devise its own standards for what counts as good enough reasons to believe. In contrast, the approach that is at once rational and faithful is to hope, with humility, that the evidence, such as it is, is good enough. Faith does not have at its disposal the most effective means of securing our beliefs or compelling others to adopt them. Our faith, like much else that matters to us, is not fully in our control but at the mercy of realities that transcend us. For Christians, this is far from surprising, for we regard faith itself not as ultimately our doing, but a gift from the God we trust. SCIENCE SHAPES FAITH Even as we acknowledge the possibility of a fatal encounter of faith with science, we should recognize that in practice the sensitivity of Christian beliefs to scientific discovery is not a simple matter of either survival or falsification. Few conceivable scientific discoveries would decisively refute Christianity, rather than leading to modifications in it. 6 As with any ancient and complex system of belief, as new knowledge appears there is plenty of room for adjustment, for reinterpretation, and even for redrawing the line between the essential and inessential. Over the last several centuries Christian belief has changed as a result of its engagement with the emerging sciences. Views long associated with Christianity, or even integrated into its theological articulation, such as vitalism, Aristotelian
32
Chapter 2
metaphysics, and Ptolemaic cosmology, were discarded when the advance of science rendered them untenable. Beliefs that could not so easily be left behind were reinterpreted in ways consistent with science. For example, the idea of divine providence was reinterpreted by Robert Boyle and others in the late seventeenth century to cohere with the mechanical conception of a universe governed by inertial laws. 7 For many, this shows that Christianity has been in ignominious retreat before triumphant science for years. They regard the progress of science as having long since defeated a faith whose adherents are too obtuse to realize what has befallen it. Many Christians, even while agreeing that in past conflicts science was in the right, and that the Church erred in tying Christian faith so tightly to, say, the cosmology of the ancient Near East or the idea that the Earth is at most several thousand years old, are uncomfortable with the general trend and worry that today we too readily revise our beliefs to make them cohere with science. With Darwin, this concern intensified as science intruded upon the once inviolate domain of human origins. It is increasingly urgent today with the application of evolutionary theory to the human mind and behavior. Yet these concerns are misguided. The lesson to draw from Christianity’s encounter with early modern science is that we should seek and welcome opportunities for scientific discovery to reshape our theology, including our account of human beings as made in the image of God. If science conflicts with our theology, it is more likely that it is our theology, not science, that is in error. We can rightly see the advance of the sciences—along with other factors including advances in the historical, linguistic, and critical study of Scripture—as leading toward improved, refined and purified Christian belief, not to its erosion. Acknowledging the power of science to affect the content of Christian theology need not be problematic, even for those who, as I do, regard the principal source of our theological beliefs to be a trustworthy divine revelation, the Bible. How can mere science correct the word of God? It might seem incongruous to accept the Scriptures as revelation from God while letting the theology derived from them be modified as a result of scientific discoveries, but it is entirely reasonable to do so. Long before the birth of modern science, Christian thinkers recognized that along with Scripture God has given us the “book of nature” and the methods of science with which to read it, thereby enabling us to ascertain some of the truth about God’s world. Neither book is easily interpreted and we are prone to error. Over the past few centuries, is has become increasingly evident that scientific methods, while not infallible, are reliable means for acquiring truth. Christians who acknowledge the Bible as God’s infallible word cannot seriously imagine that the interpretive methods we apply to it are themselves infallible. They are at best reliable ways of ascertaining the truth contained in God’s written word. Crucially, the methods of science are generally more reliable than those of Biblical interpretation. Scientific
Knowledge
33
practice converges on truth, but the procedures that interpreters of the Bible employ produce a host of persistently divergent results. This does not mean that we are never in a position to be confident in a particular interpretation of a biblical text, or in arriving at a theological conclusion. Nor does it imply that there is no consensus, over appreciable spans of time, among serious students of Scripture. It does imply that when we find our theological beliefs at odds with the firm consensus of the scientific community, we should first suspect that it is our interpretation of the Bible, not science, that is in error, and seek to revise it accordingly. The example that most readily comes to mind is, no doubt, the creation story presented in Genesis, a text that has a theological significance undiminished for those of us who regard its truth as in various ways figurative, rather than literal. 8 When the Scriptures teach something that in light of science cannot be literally true, that is often good reason to think the proper interpretation of the text is a non-literal one. 9 Reading the book of nature enhances our ability to rightly read Scripture. The historical record does not undermine confidence in divine revelation, but it chastens overconfidence in our powers of interpretation. It is now obvious that sixteenth-century Christians should have corrected their ideas about how to read texts like Psalm 93:1—”Yea, the world is established; it shall never be moved”—rather than condemning as heretics the astronomers who said that the Earth moves around the sun. My hope is that there will come a time when Christian rejection of the contemporary naturalistic account of human beings will seem as misconceived as these earlier dismissals of science in favor of current theological conviction. Finally, though, this cannot properly be used as a roundabout way to render Christianity immune from falsification. Biblical interpretation, and the theology based on it, is not infinitely flexible. At some point it could become all too clear that what science confirms is contrary to what Scripture teaches. Faith, at least of the Christian variety, never escapes its vulnerability. FAITH SHAPES SCIENCE Scientific theories enjoy stronger evidential support than is available for theological beliefs. Therefore, Christian conviction, insofar as it is honest, rational belief about this world, is in the final analysis subject to modification, and even refutation, at the hands of science. However, this does not imply the relation between science and theology is only one way, so that while scientific discovery can inspire modification of our theological beliefs, our theological convictions have no legitimate impact upon science. Nor does it imply that the scientist who is a Christian is obligated to leave her faith at the door of the laboratory and engage in scientific in-
34
Chapter 2
quiry as though she has no theological beliefs or as though they are guaranteed never to be relevant to the scientific enterprise. It is commonly assumed, even by Christians working in the sciences, that if there is a role for faith, it is not in the actual practice of scientific investigation, shaping the reasoning that leads to scientific conclusions, but beforehand, as a source of guidance in the choice of research programs, or as a source of moral constraint on the treatment of experimental subjects, or afterwards, when it comes to the practical application of newly acquired knowledge. But faith is granted no proper role at the heart of scientific reasoning, in the evaluation of hypotheses. To give it entrance is to fall short of objectivity. This assumption is mistaken. It relies on a spurious notion of scientific objectivity. Objectivity is not neutrality. It can be entirely reasonable to take account of anything one believes, including one’s theological beliefs, when evaluating evidence for or against a scientific hypothesis. There is no obvious specific answer to the question of what is required for the confirmation of a scientific hypothesis. What sort of evidence must accrue, what tests, and how many, must a particular hypothesis be subjected to, before we are entitled to regard it as true? There is no algorithm that specifies what is needed for confirmation; evaluating a hypothesis calls for informed judgment that weighs a variety of factors. A mere catalogue of experimental results, the observational evidence, the empirical data, does not tell us whether we ought to believe a hypothesis. Other factors must have their due: What is the status of competing hypotheses? Is the hypothesis elegant, simple, and beautiful or is it aesthetically displeasing? How well integrated is it with the rest of science? What anomalies does it engender and how tractable do they appear to be? One important consideration is that the quantity and quality of evidence that make it reasonable to believe something depends in part on other things we already believe, on our background assumptions. Some hypotheses are plausible, having high prior probability relative to our other beliefs, scientific and otherwise. 10 We accept them on the basis of relatively less, and less good, evidence. Some hypotheses are implausible, antecedently improbable in light of our background beliefs. We should accept them only on the basis of more, and better, evidence. This maxim applies in general to reasoning about what to believe, not just to the evaluation of hypotheses in science. When Max tells us that he is from the seaside town of Neptune, New Jersey, we don’t demand more evidence. In most contexts the mere fact that he makes the assertion is good enough reason to believe him. However, if he claims to be from the planet Neptune, his saying so falls far short of sufficient reason to believe it. If it ever becomes reasonable to believe him, it will be on the basis of much better evidence. It would be daft to think we should evaluate the evidence for Max’s assertions from a position of neutrality, one that ignores our background beliefs about which parts of the solar system are likely places for
Knowledge
35
human habitation. We should evaluate the evidence objectively, and that requires that we pay heed to his assertion’s plausibility in light of what we already know, not that we ignore it. If I believe something, and also believe that it renders a hypothesis I am considering improbable, it would be irresponsible to ignore this when weighing the evidence. The only beliefs that should automatically be denied this sort of influence are those that lack rational justification, those we should not have in the first place. There are of course many members of the scientific community, and many others, who take it for granted that Christian beliefs are not rational, particularly if they are construed as factual claims about the world. If they were right, Christian faith should not influence the evaluation of the evidence for hypotheses in science. But Christian belief is rationally defensible and there is no reason to assume it should never influence the practice of science. If someone lacks good enough reasons to believe Christianity then she should not believe it and ipso facto should not weigh scientific hypotheses in light of it. But if she has what she regards as good enough reasons to think it is true, then she ought to believe it, and she ought to regard any hypothesis that it renders improbable as believable only on the basis of exceptionally good evidence. A rational person, having examined the evidence, might initially reject a scientific hypothesis because of its lack of fit with the Christian faith. 11 It is unreasonable not to believe the well-confirmed theories of science, but this does not imply that there can be no reasonable disagreement about what theories really are well-confirmed. That judgment depends on background beliefs and judgments about prior probabilities they sustain. Because there can be rational disagreement about how good the evidence must be for a theory to be well-confirmed, there can be rational disagreement within the scientific community, and between the scientific community and those outside it, on whether the evidence has adequately supported a theory. A Christian who judges that a theory is improbable in light of his faith might initially reasonably reject a scientific theory, contending that the evidence for it is not yet good enough, even as non-believing scientists reasonably regard the evidence for it as already good enough. 12 Is it paradoxical to say both that in the event of conflict science trumps faith, and that we might reasonably reject a scientific hypothesis in virtue of its lack of fit with faith? Not at all. Consider the analogous case of our commonsense beliefs: it is unreasonable to reject a well-confirmed scientific theory on the ground that it is at odds with common sense. We recognize profound differences between the world as it appears and the way scientific inquiry reveals it to be in reality. Science so long ago established that we live on the surface of a sphere of rock that spins rapidly on its axis and travels through space around the sun, and that solid objects are made of tiny particles in mostly empty space, that we forget how
36
Chapter 2
implausible these ideas are from the perspective of ordinary experience. More recent scientific findings, such as the Big Bang theory’s assertion that the entire universe was once smaller than an atom, special relativity’s claims about the relativity of simultaneity, or general relativity’s implication that the shortest distance between two points is typically not a straight line, conflict mightily with what we find intuitively obvious, but this is a reason to conclude that common sense is wrong, not to reject the scientific theories. Yet the fact that science can defeat common sense does not mean that it has no legitimate effect on the evaluation of theories. It is sometimes right to reject an hypothesis as implausible in light of our other beliefs even when there is evidence in its favor. Given what was believed at the time, when in 1912 Alfred Wegener promoted his Pangea hypothesis, it gained few adherents; it wasn’t until the 1960s that the evidence finally led geologists to embrace it, overcoming its contrary to common sense claim that the continents are adrift. The Big Bang theory, put forward by Georges Lemaître in the 1920s, was strenuously resisted by many who found implausible the idea that the universe is temporally finite. It was widely accepted only when the cosmic background radiation was detected in 1964. Or, to take an example of a hypothesis that outrages common sense and which has not, at least yet, been generally accepted: when Hugh Everett proposed his many-worlds interpretation of quantum mechanics in the 1950s it was generally dismissed as a weird curiosity, but more recently a significant number of physicists have accepted it, despite its bizarre implications. Common sense initially constrains scientific discovery, but scientific discovery overturns common sense when the evidence is good enough. We are right initially to discount an hypothesis that our other beliefs render improbable, but it is always possible that evidence will emerge that makes it unreasonable not to accept it, and requires us to abandon or modify our other beliefs accordingly. There is no evaluation of hypotheses free of presuppositions about what is probably true in light of what we already believe. This includes the scientific beliefs we already have, as well as beliefs of many other kinds. The competence of science does not extend to the question of the prior probability of a hypothesis in light of a theological, or any other non-scientific, assumption. What lies in its purview is establishing what the empirical evidence is and its bearing upon the probability of the hypothesis. Nonetheless, scientific inquiry makes it possible for the evidence to subvert expectations and to revise our view of the world. Advances in science sometimes involve the discovery of excellent reasons to believe very improbable things. Scientific discovery itself is colored by our presuppositions about the world, and it ought to be; nonetheless, it gives nature the chance to overturn even our deeply entrenched beliefs. Because Christianity has empirical content, we must accept the possibility that science will produce compelling reasons to
Knowledge
37
believe things that we as Christians were once entitled to dismiss as implausible. 13 Faith initially constrains scientific discovery, but scientific discovery can finally overturn faith. Of course, one need only accord a cherished belief a high enough prior probability to render it forever impervious to contrary evidence. No matter how much, and how good, the scientific evidence for an hypothesis is, someone can insist that it is still not good enough to accept it, because it would falsify the conviction he wants to safeguard. It is always possible to evade the import of the evidence, but we should not mistake heroic departures from reasonableness for intellectual faithfulness. 14 Christians have no reason to try to keep their faith and their science in separate compartments. Whether the evidence is good enough to make a theory well-confirmed depends on two things, not just the quality of the evidence itself but the prior probability of the hypothesis. We are entitled to demand more and better evidence to accept hypotheses that have low prior probability in light of our faith, and at the same time we are obligated to recognize that we might get it, and thus be forced to modify, or even give up, theological beliefs. So, it is important that our judgments about what is and is not plausible in light of the Christian faith be made with care, that we not simply take it for granted that a given hypothesis is antecedently improbable—or probable—from the perspective of Christian faith. In subsequent chapters I will indicate that current conflict between Christians and the scientific community, and ultimately Christian hostility to the naturalistic view of human beings, arises out of mistaken judgments about these matters. Christian critics of science and scientific critics of Christianity often make the same erroneous assumptions about what science tells us being antecedently improbable from the perspective of Christian faith. MATERIAL KNOWERS Scientific naturalism includes the view that we are material beings. Thereby it denies human transcendence. We are part of, not apart from, the physical world. Our intellectual capabilities make us unique among this planet’s creatures, but they are not the capabilities of an immaterial entity. There is no non-physical mind or soul standing outside the physical world, perusing it from without. The human person is a material object, a part of the physical world, inescapably enmeshed in the causal order of nature. We are capable of knowing only because the neural machinery located inside our heads reliably produces accurate representations of what is outside. Looking out my office window, I know that Mitch is walking across the parking lot because I see him there, which is to say that my visual
38
Chapter 2
processing circuitry generates sensations that cause me to believe that he is there, and this perceptual mechanism reliably produces beliefs that are true. When I see that he is wearing shorts, believing that he never wears shorts on days that he teaches, I conclude that he is not teaching today, thanks to brain circuitry that makes logical inferences and thereby reliably produces further true beliefs on the basis of what I already know. Reflecting on Mitch’s light teaching load in contrast to my onerous one, I realize I am envious, thus knowing something about my own state of mind by way of the causal mechanisms involved in introspection. Later this afternoon, when someone asks me if I saw Mitch today, I will know that I did, because I will remember that I saw him, and memory is a mechanism that reliably causes true beliefs about the past. A human being has knowledge when he has belief that is true and caused by a reliable mechanism. 15 He need have no awareness of how the mechanism works; so long as it reliably produces true beliefs he possesses knowledge. For some kinds of belief, the process that reliably produces true belief is the conscious evaluation of evidence, but much of our knowledge is the product of cognitive processes to which we have no conscious access. 16 This naturalized portrayal of human knowledge breaks with centuries of philosophical thought, for it dispenses with the quest for absolute certainty, with the demand for unqualified assurance that we know what we think we know. Prior to taking the naturalistic turn, philosophy sought epistemological foundations, and thus for knowledge that is on the one hand absolutely secure, safe from any skeptical challenge, and on the other something we can use to justify reliance on other ways of knowing, such as science. The most well-known example from philosophy’s long foundationalist history is René Descartes’s 1641 attempt, in his Meditations on First Philosophy, to employ his bare powers of reasoning to bring to light innate indubitable foundations for trust in our senses and reliance on the emerging science that relies on them. 17 Naturalism denies the possibility of, but also the need for, such first philosophy. No way of knowing the world and ourselves is more reliable than science itself, so the project of finding foundations for science is misguided and doomed to failure. The best answer to the question, “Why should we believe that the neural mechanisms that produce beliefs in us are reliable, that they give rise to knowledge rather than deceiving us?” is that scientific investigation endorses the conclusion that they are reliable. We have scientific knowledge of the workings of perception, memory, reasoning, and so on. And we have some scientific knowledge of the workings of science itself. This response is, of course, patently circular. But it is a circularity creatures inescapably inhabit. Scientific study of, for instance, the multiple levels of information processing by which we get from photons impinging on the surface of the eye to a mental image of a three-dimensional
Knowledge
39
world of moving colored objects is itself largely dependent on investigators’ use of their own eyes. Science can give us no reason to trust our senses that does not ultimately beg the question against the skeptical challenge. With respect to any specific scientific claim, we can ask whether it is really true, or whether a false hypothesis unluckily has been well-confirmed, but such questions are posed reasonably only from within the current general scientific picture of things. If reasonable, they are always requests for more scientific work, not demands for some external justification of science. Intended to call into question scientific reasoning as a whole, “What if our best way of finding things out fails us?” is a worry the naturalist sees as groundless. There are those who will contend that they are privy to some more secure way of knowing, some standard against which to measure the practices of science, but the candidates are unconvincing. There is no reasonable expectation of acquiring foundational knowledge. The best we can do is to take for granted what looks like our most secure knowledge, the well-confirmed theories of the natural sciences. Nothing in the naturalistic view calls us to accept every claim of science uncritically, without being aware that even the most rigorous application of scientific methods sometimes leads to error. The fallibility, along with the reliability, of scientific methods is endorsed by science itself. The abandonment of the foundationalist project does not mean we must succumb to skepticism, giving up on the possibility of knowledge. Nor, contrary to some post-modernist impulses, does it mandate relativism. Foundationalism’s demise does not imply that we cannot aspire to genuine knowledge; it shows us that knowledge does not require foundations. However, it does humble us, compelling us to acknowledge that even our best ways of knowing the world and ourselves are inevitably fallible. No matter how carefully we reflect upon our methods for gaining knowledge we cannot validate them as a whole from some perspective beyond the human condition. Any critical evaluation of our cognitive powers is necessarily piecemeal; we can challenge candidates for knowledge only as we rely on our general knowledge of things. We could be drastically, systemically wrong and never know it. We have no possibility of living beyond all imaginable error. Whatever rational confidence may, as a practical matter, be available to us when we rely upon what seem our best methods of achieving knowledge, our nature denies us the absolute certitude sought in the philosophical tradition. As human creatures, we can do no better than cautious trust in our natural constitution. For persons of faith, this is trust in our Creator.
40
Chapter 2
MUNDANE KNOWERS A naturalized conception of knowledge says that so long as a belief is true and the result of a causal mechanism that reliably produces true beliefs, it constitutes knowledge. The person who knows need not be able to supply reasons for her belief. Karen knows that she had eggs for breakfast because she remembers that she had eggs for breakfast, not because she can justify this belief by appeal to evidence. 18 She simply has this belief about the past, and thus knows that she ate eggs for breakfast because her memory involves mechanisms that, as a matter of fact, reliably produce true beliefs about the past. This contrasts with traditional conceptions of knowledge, which required that one knows only if she can supply good reasons for her belief. The task of identifying and articulating the evidence that justifies a belief calls for a cognitive sophistication unique to mature human beings. No human infant or non-human animal is capable of reflecting upon the reasons for belief. Thus, on the traditional account of knowledge as justified, true, belief they are incapable of knowing. The naturalized view, shifting focus from the consciously accessible reasons for which a belief is held to its causes, which might, but need not, involve consciously accessible reasons, rejects this demarcation between humans and other creatures. We are not unique, so far as being capable of knowing goes. Other animals have mental states that accurately represent how things are, not accidentally, but as the result of the perceptual and cognitive machinery with which natural selection has endowed them, machinery that reliably produces accurate mental representations. Human knowledge is, no doubt, different than the knowledge other animals possess in important ways, but it is continuous with it. Human cognition involves linguistically and socially mediated representation and metarepresentation—mental representation of mental representation—that affords us mental lives that are uniquely conceptually rich and that underlies the improvisational intelligence that freed us from being tied to any particular ecological niche. 19 Still, knowledge is a natural kind: the role that the concept of knowledge plays in explaining human behavior is essentially the same role it plays in the scientific explanation of the behavior of plovers, baboons, or dolphins. 20 Those seeking a deeper categorical difference between humans and other creatures will not find it. Descartes, in his mid-seventeenth century Meditations, taught that while the human mind is finite, and thus inevitably ignorant, it is never inevitably subject to error. Unlike God, we cannot know everything, but, like God, we can be free of falsehood. Edward Craig argues in The Mind of God and the Works of Man that in early modern times this idea was enlisted to safeguard a traditional conception of human nature from the new scientific culture that threatened to unseat humans from their privileged place in the cosmos. As finite creatures, we are not omniscient; we cannot
Knowledge
41
know everything the infinite Creator knows, but we are capable of knowing with the highest possible degree of certitude and understanding. In our best reasoning we are capable of an intellectual insight into objective reality which is qualitatively indistinguishable from that of God. Craig suggests that this conception of humans as God’s image was adopted as “a defensive action against the cultural effects of the new science,” a reaction to a diminished image of our nature in which we are “not unique in any way that mattered, a speck of dust in the encompassing vastness.” 21 The old, pre-scientific grounds for an exalted human self-image had been shaken, but there was comfort in this supposed likeness of created, finite minds to the infinite mind of God. In this elevated portrayal a human person is not a natural object, not a superior animal amongst the lesser animals, but a little god beside the great God. 22 Contemporary science deflates these pretensions and leaves no serious doubt that the human mind is anything other than the embodied, and socially and culturally situated, human brain, and it adds the understanding that this mind/brain is the product of mindless natural selection, built for reproductive, not intellectual, prowess. The presumption that human reasoning, like God’s, transcends the constraints of the natural, causal order faces an endless supply of countervailing evidence. The human image that science forces upon us is precisely that of a “superior animal,” not “a little god.” Should this discomfit the Christian? The fear, noted in chapter 1, that modern science foists upon us a view of human nature that is too low, one that renders us insufficiently similar to our Creator, too much like other creatures, has a strong hold on the imagination of many Christians. It is understandable that they find their faith in conflict with what contemporary science tells us about human beings as knowers. But is this concern legitimate? If the more significant problem is the perennial disposition to imagine ourselves as more like God than we really are, that is, if our concern ought to be pride and an inflation of the human self-image, then we may well welcome the deflationary effect of naturalized epistemology. UNBIDDEN BELIEF The naturalistic account of knowledge locates it within the realm of cause and effect, delineated in the computational architecture of the knower’s nervous system. To know is to be in a state of mind that stands in a particular kind of causal connection to the world. For an organism to know is to be in a mental state which accurately represents some aspect of reality, and thus, by figuring in the causation of appropriate bodily movements, enables it to navigate its natural or social environment. As genuine knowledge, these states of mind are themselves the effect of causal mechanisms that reliably result in accurate representations. When
42
Chapter 2
Karen knows that the cat is on the sofa this is because she is caused to know this: something causes her to believe that the cat is on the sofa, where this belief is true, an accurate rendition of the cat’s illicit location. In the simplest case the reliable causal chain involves a complex sequence of events: light reflecting off the cat and being focused on her retina, and in turn causing a cascade of firings of photoreceptive neurons, and then many further neural happenings in her occipital lobe and other parts of her brain. This conception of knowledge, in virtue of being essentially causal, focuses attention on a fact that is readily available to common experience and reflection but which contrasts with what many of us seem to believe. For we often speak of our beliefs as though we were not always caused to have them. We speak of choosing what to believe, where this means that we are not ultimately compelled to believe what we believe and to not believe what we do not believe. Yet, barring a drastic failure of rationality, we cannot choose our beliefs. As I write this, I do not believe that there is an armadillo in my office and I cannot believe this. I can choose what to do, but I cannot choose what to believe. I can choose to pretend to believe that there is an armadillo here; I can cry, “Lo! An armadillo!” I could, perhaps, choose to act in a way that causes there to be an armadillo in my office: I drive to the zoo, liberate an armadillo from its pen, and bring it here. Now that it is installed in my office I believe there is an armadillo in my office. Indeed, I cannot help but believe it; I cannot choose to disbelieve that there is an armadillo here. In this indirect way I might be able to choose my beliefs, but I lack the power to choose my beliefs in a direct way. I cannot believe at will. 23 Some might agree, but circumscribe the conclusion, contending for an exception in cases where the evidence is insufficient to compel either belief or disbelief. Yet in such cases it is no less clear that we can neither believe nor disbelieve. At the present moment I have some change in my left pants pocket. This amounts either to an odd or an even sum, and I believe that the sum is either even or odd, but I do not believe that it is odd, nor do I believe that it is even. Were someone to suggest that I simply start believing that it is an odd amount I could not comply. In the absence of more evidence I am forced to believe neither. In this case the matter is both trivial and easily resolvable—I could pull the coins out and count, thereby putting myself in a situation in which I would be forced to believe one or the other. Some find it plausible that, where the issue is significant but adequate evidence is not readily available, then it is possible and desirable to choose what to believe. The well-known candidate is religious belief. It might seem plausible that because the issue of God’s existence joins supreme importance with what many take to be inconclusive evidence we are called upon to choose what to believe. 24 Yet the prospect for believing at will seems no more likely here than in other contexts. For here, as in other cases, the
Knowledge
43
idea of believing something simply because one has decided to believe it approaches sheer incoherence. For me to believe something implies that I believe that it is true, and thus that there is some feature of reality, generally independent of me, my choices, and my beliefs, that makes it true. I do not rationally think that I make it true by believing it. If I believe that the cat is under the sofa, then I believe that it is true that the cat is under the sofa, and I believe that this is true just because of the relative locations of the cat and the sofa. Were I to have and exercise the power to believe at will, choosing to believe that the cat is under the sofa, the belief would disappear as soon as I reflect on the fact that my choosing this belief provides no reason to think it is true and, as soon as I realize I have no reason to think it is true, I stop believing it. 25 The naturalistic picture integrates belief and knowledge into the causal order of nature, a reality we cannot transcend but of which we are a part. This leaves us singularly vulnerable. On the one hand, we cannot believe at will. We believe only what we are caused to believe. The will is epistemically impotent. Belief, and thus knowledge, befalls us. On the other hand, though we are compelled to believe, we lack the comfort of absolute certitude. We can be caused to believe what is false and to be sure it is true. With regard to what we believe, and thus to what we know, we enjoy neither ultimate control nor final certitude. On reflection, this vulnerability is not something persons of Christian faith should deny. The naturalized conception of knowledge resonates with a Christian view of things insofar as it forces us to own up to the fact that we have no alternative but to trust, hoping that our cognitive capacities when used with care afford an accurate portrayal of the world and our place in it. Rather than wishful thinking, this is a reasonable hope: we have good reasons to believe in the reliability of our cognitive capacities. Yet their fallibility is evident, and we have no reasonable expectation of obtaining proof of their reliability. We assume that they are generally reliable and, in so doing, find various specific ways in which they are unreliable. We are self-reflective creatures, capable of asking about the quality of the means we employ in quest of knowledge, but we are finally material beings. We cannot extricate ourselves from our own skins. Everything we believe about how we arrive at our beliefs is itself something we are caused to believe; we never leave the realm of natural causality to evaluate our cognitive capabilities. We are not godlike beings, serenely and disinterestedly surveying the world from a transcendent perspective immune to error and delusion. Instead, we are mere creatures, objects of a particular sort assembled by incredibly complex and highly contingent natural processes, inhabiting a tiny part of a vast and ancient universe not of our making and which we are not optimally equipped to know. The Christian can and should receive this naturalistic news with equanimity. The Christian faith, and before it the faith of Israel, warned humankind away from the temptation to conceive of our-
44
Chapter 2
selves as transcendent entities, reminding us that we are not gods but God’s created image, called to live by faith, to put our trust in our Creator. Reflecting on the inescapable incompleteness and partiality of human knowledge, Salman Rushdie wrote that we are “not gods but wounded creatures, cracked lenses, capable only of fractured perceptions.” 26 Even with our uniquely human cognitive powers in view, it is evident that we must dispense with the conceit that we occupy a “God’s eye” point of view on reality, what Thomas Nagel once called a “view from nowhere,” a perspective not profoundly shaped by a particular history, not the product of a contingent array of commitments and interests. 27 Whatever its plausibility for theism generically conceived, it is clear that the God of Christian faith occupies no such perspective. Rather than revealing a God who has a God’s-eye point of view on things, the Scriptures portray a God who has quite definite aims, sees things in light of them, and calls creatures to see things the way their Creator sees them, not to adopt a neutral view from nowhere. We creatures are capable of knowing, and critically reflecting upon, the world and ourselves, but it is misleading to say, with Arthur Peacocke, that this means that “human self-consciousness transcends nature.” 28 We are a part of nature that is self-aware, and we are a part of nature that knows itself, but we are a part of nature. Only God transcends this physical world. Human minds, never disembodied, are fully immersed in nature’s causal matrix, knowing the world only from within. KNOWING GOD? In denying the transcendence of the human knower, does naturalized epistemology preclude the possibility of knowledge of a transcendent deity? Science is not our only, but it is our best, way of knowing. Not all knowledge is scientific knowledge, but all claims to know must cohere with science and abide by the limitations on human cognitive powers it discerns. The mere fact that a belief is not a product of science is not, from the naturalistic point of view, a reason to reject it, but science will not cohabit with every conceivable claim to non-scientific knowledge. The naturalist rightly rejects claims that compete with well-confirmed scientific theories. Not all knowledge is scientific knowledge, but this provides no comfort to one who claims to know, for example, that it was Noah’s flood, not eons of uplift and erosion, that carved the Grand Canyon. Equally, if a claim to know presupposes cognitive mechanisms science supplies us with good reason to think we do not possess, then naturalism denies it. The naturalist can heartily agree that there are more things in heaven and earth than are dreamed of in his philosophy, but nonetheless dismiss claims for, say, telepathic knowledge, knowing that the current
Knowledge
45
scientific picture of the mind leaves no plausible space for such a faculty. The idea that human beings possess knowledge of God by special, nonempirical means seems no more plausible. If there is rational belief in God, it arises from the encounter of the mundane powers of human reasoning with the evidence available to the material creatures we are. Possibly, rational reflection on experience gives us some reason to believe that there is a Creator, but any evidence for Christian belief most likely must be construed as the effect of divine intervention in the created world. Christian faith is trust in the God who resurrected Jesus Christ, and that resurrection was miraculous. If, as is widely supposed, scientific naturalism rejects belief in miracles, it is incompatible with the Christian faith. I criticize and reject this supposition in chapter 3, and contend that the scientific naturalist has no general viable objection to concluding, on the basis of empirical evidence, that a miracle has occurred, and thus that there is a God who acts in this world. NOTES 1. Arthur Peacocke, All That Is: A Naturalistic Faith for the Twenty-First Century, Philip Clayton, ed. (Minneapolis: Fortress Press, 2007), 4. 2. Empirical evidence and logic do not suffice to tell us that an hypothesis must be regarded as disconfirmed, and there is no obvious general answer to the question of how much, and what kind of, evidence renders an hypothesis well-confirmed and thus worthy of belief. These are matters of judgment socially embodied in the various disciplinary communities of the sciences. 3. Bertrand Russell, Religion and Science (London: Oxford University Press, 1935), 243. One wonders how seriously Russell intended this claim, as opposed to a weaker, but less rhetorically exciting, assertion about science being our best way to acquire knowledge. Among the difficulties it faces when taken at face value: those of us who are not scientists will know nothing, since even if a scientist knows something, we will not be able to employ scientific methods to know that she knows it. I know that such and such is a well-confirmed theory, but I do not know this by means of scientific reasoning; I know it by reading it in the New York Times. 4. Albert Einstein, “Physik und Realität,” Journal of the Franklin Institute 221(1936), 313–31. The original reads: “Alle Wissenschaft is nur eine Verfeinerung des Denkens des Alltags.” 5. Stephen J. Gould, Rocks of Ages: Science and Religion in the Fullness of Life (New York: Ballantine Publishing Group, 1999), 4. 6. One possibility is cosmology establishing that the universe exists as a matter of metaphysical necessity, rather than contingently. This would refute the Christian claim that this world exists as a result of God’s free act of creation. 7. See John Hadley Brooke, Science and Religion: Some Historical Perspectives (Cambridge: Cambridge University Press, 1991), 130–35. 8. In colloquial English, “literally” has come to mean “really,” with the unfortunate result that many now hear the statement that the biblical creation account is not literally true as tantamount to the assertion that it is not really true. But what I want to say is that this figurative text is really true, just as, during a downpour, it is really—but not literally—true that it is raining cats and dogs. 9. It is striking that some important secular critics of Christianity agree with Christian fundamentalists that views held almost universally among biblical scholars amount to a dishonest dodge to avoid having to take the text literally. See, e.g., Rich-
46
Chapter 2
ard Dawkins, The God Delusion (New York: Houghton Mifflin, 2006), 237–38, and Sam Harris, The End of Faith: Religion, Terror and the Future of Reason (New York: W. W. Norton and Company, 2004), 16–23. 10. Prior or antecedent probability contrasts with posterior probability, i.e., the probability of the hypothesis once all relevant evidence is taken into account. 11. There remain, of course, the genuine practical difficulties that arise when one’s scientific colleagues do not share the beliefs that influence her evaluation of the evidence for scientific hypotheses. 12. This explains some of the extreme rhetoric in current debates about evolution, in which some claim the evidence is nowhere near adequate, and others claim that there is a superabundance of evidence for it. 13. Within the Christian faith, particular beliefs have been empirically falsified, such as the traditional idea that humans possess immaterial souls. As will become apparent in later chapters, the particular structure of Christian beliefs set forth in this book depends on the falsifiable view that quantum mechanics reveals an ontological, not merely epistemic, indeterminacy in nature. However, perhaps God has the ability simply to choose not to know everything his initial act of creating a deterministic universe entails. 14. A vivid example is Kurt Wise, who, although trained as a paleontologist under Stephen J. Gould, is a “young Earth creationist” who asserts, “If all the evidence in the universe turns against creationism, I would be the first to admit it, but I would still be a creationist because that is what the Word of God seems to indicate;” John F. Ashton, ed. In Six Days: Why 50 Scientists Choose to Believe in Creation (Green Forest, AR: Master Books, 2000), 355. Does Dr. Wise believe that the evidence that the Bible is God’s word is better than the scientific evidence for evolution, indeed, that it is better than any possible future scientific evidence for it? The evidence for the Bible being God’s revelation is, I trust, good enough, but it is not reasonable to claim that it is better than the evidence for a well-confirmed scientific theory. Setting aside the question of how to justify the belief that whatever the Bible teaches is true, does he also believe that the hermeneutical methods that lead him to interpret particular passages in Scripture literally, and thus as in conflict with well-confirmed scientific theories, are more reliable than the methods that lead the scientific community to embrace them? 15. See, e.g., Alvin I. Goldman, Reliabilism and Contemporary Epistemology: Essays (Oxford, UK: Oxford University Press, 2012). It would be difficult to overstate the complexity of the causal pathways. The route, e.g., from photons impinging on the retina to experience of a three-dimensional world of dynamic, colored objects involves multiple levels of complex computation, classically described in David Marr’s Vision: A Computational Investigation into the Human Representation and Processing of Visual Information (Cambridge, MA: MIT Press, 2010). Further, the causation is not unidirectional, the external world imposing representations upon a passive recipient. Instead, the mind incessantly probes, predicts, and corrects expectations of the raw material of sensation; the same neural circuitry sustains both perception and action. See Andy Clark, Surfing Uncertainty: Prediction, Action, and the Embodied Mind (New York: Oxford University Press, 2015), chapter 4. Further, perceptual knowledge is profoundly mediated by language and culture; see Daniel Dennett, From Bacteria to Bach and Back: The Evolution of Minds (New York: W. W. Norton and Company, 2017), chapters 9–13. 16. Externalist epistemologies, of which the reliabilism described here is the simplest example, continue to be debated within philosophy, but in the cognitive sciences, it is a given that knowledge is explained by the mind’s causal interaction with the environment. Even if the internalist view that knowledge always requires justification was true, the process of consciously examining reasons results in knowledge only if the underlying neural mechanisms that embody this cognitive activity reliably cause true belief. 17. René Descartes, Meditations on First Philosophy, Donald A. Cress, trans. (Indianapolis, IN: Hackett Publishing Company. 3rd ed. 1993).
Knowledge
47
18. Of course, in some instances she might, say, examine the most recent stratum of dishes in the sink, observe traces of egg on them, and from this evidence infer that she had eggs for breakfast. 19. H. Clark Barrett, Leda Cosmides, and John Tooby, “The Hominid Entry into the Cognitive Niche,” in Steven W. Gangstead and Jeffry A. Simpson, eds. The Evolution of Mind: Fundamental Questions and Controversies (New York: Guilford Press, 2007), 241–48. 20. Hilary Kornblith, Knowledge and Its Place in Nature (Oxford, UK: Clarendon, 2002), 28–69. 21. Edward Craig, The Mind of God and the Works of Man (Oxford, UK: Oxford University Press, 1987), 16. 22. Craig, The Mind of God and the Works of Man, 70. 23. Belief is unbidden and lies outside our direct control, but this does not place it beyond critical evaluation. We rightly speak of what we ought, and ought not, to believe. We cannot choose our beliefs, but we choose what to do, and we are obligated to act in ways that make it more likely that we will be caused to believe what is true, and less likely to believe what is false. If, for example, I believe something but know of evidence against it that I am too lazy or too frightened to look at, I am irresponsible and not believing as I ought to believe. 24. William James famously advocated this possibility: “The Will To Believe” in The Will to Believe and Other Essays in Popular Philosophy (New York: Longmans, Green, and Company, 1897). John Bishop’s Believing By Faith: An Essay in the Epistemology and Ethics of Religious Belief (New York: Oxford University Press, 2007) is a sophisticated contemporary defense of the choice to believe, although its focus is the moral permissibility, not the possibility, of doing so. 25. This argument was put forward by Bernard Williams in “Deciding to Believe,” in his Problems of the Self: Philosophical Papers 1956–1972 (Cambridge, UK: Cambridge University Press, 1973), 136–51. 26. Salman Rushdie, “Imaginary Homelands,” in his Imaginary Homelands: Essays and Criticism 1981–1991 (London; Granta Books, 1991), 12. 27. Thomas Nagel, The View From Nowhere (New York: Oxford University Press, 1986) and Hilary Putnam, Reason, Truth and History (Cambridge, UK: Cambridge University Press, 1981), 49. 28. “Biological Evolution and Christian Faith—Yesterday and Today,” in his Evolution: The Disguised Friend of Faith? Selected Essays (Philadelphia: Templeton Foundation Press, 2004), 44.
THREE Miracles
MIRACLES DEFINED The Christian who embraces scientific naturalism encounters objections to miracles on two fronts. First, almost everyone finds it obvious that belief in a God who intervenes in this universe is at odds with the preeminence of science as a way of knowing. For many, banishing the miraculous in favor of scientific explanation is what anything worth calling naturalism is all about. Second, many Christians, although opposed to naturalism, find the idea of the Creator miraculously intervening in the creation theologically problematic. In this chapter I examine both types of objection. I first show that while some appeals to divine intervention are incompatible with the constraints science places on human knowledge, scientific naturalism warrants no blanket rejection of the miraculous. I go on to show that theological qualms about invoking the miraculous are not well founded. The miracles scientific naturalism leaves room for are indispensable to the Christian faith. It may well be difficult to justify a belief that an alleged miracle has occurred, but it is impossible to justify either type of a priori objection to the idea of God intervening in the world. The conception of miracle I employ is a traditional one essentially set out, for example, by John Macquarrie: it is an event that breaks into the order of nature. By this is meant that the so called laws or regular procedures of nature are on some occasions suspended, so that miraculous events take place without natural causes and as the consequence of supernatural agency . . . and the result produced is other than it would have been, had only the natural causes been at work. 1
49
50
Chapter 3
The crucial idea is that God acts directly upon the world, causing an event that would not have happened in the natural course of events. With this definition in view, I will use miracle and divine intervention interchangeably. Miracles contrast with what God does indirectly, by means of what are traditionally known as secondary causes, what God does by creating a world in which things happen as a result of the Creator’s freely authored natural laws. Miracles are at once similar to and different than divine action by means of secondary causes. John’s Gospel portrays Jesus miraculously changing water into wine at a wedding in Cana, a town in Galilee (John 2:1–11). 2 Six stone jars, each holding twenty or thirty gallons of water, suddenly no longer held water but wine, and this happened because God acted directly upon the world. In the biblical story, wine came into existence, not the way it does in the natural course of events, but because God intervened. Under the right conditions, the biochemical processes of fermentation occur and wine naturally comes into existence. Even when this natural process is helped along by human action, as typically it is, we can, like the psalmist, give thanks to God for wine that gladdens the heart (Psalm 104:15), acknowledging it as God’s creation. The existence of a particular merlot was not, presumably, a predictable result of the Big Bang. God did not specifically intend it, but brought the cosmos into being knowing its possibility and its probability. In this sense, God makes wine by way of secondary causes. The natural and the miraculous wine differ in their causal histories; one is very long and one is very short, but both track back to an act of the Creator. The ultimate distinction is not between what God does and what happens naturally, for if God is the Creator, what happens naturally is no less God’s doing than the miraculous. The distinction is between what God does by secondary causes and what God does immediately. MIRACLES DENIED Naturalism, employed as a synonym for atheism, obviously excludes even the possibility of miracles. If there is no God, there can be no miracles. Only the Creator can bring about events that would not have occurred in the natural course of events. Leaving aside the obfuscating definition, the possibility remains open. Scientific naturalism, understood as the primacy of science as a way of knowing, together with the physicalist understanding of human nature and the evolutionary account of human origins, is consistent with the existence, as well as the non-existence, of God. Miracles are possible if there is a God who created the world and selected its laws, and no well-confirmed theory of science makes any claim about
Miracles
51
God. There is, then, a straightforward but uninteresting sense in which scientific naturalism is consistent with the miraculous. The significant question is whether naturalism so construed is consistent with rational belief in miracles, or if belief that something miraculous has actually occurred implies that some well-confirmed theory of science is false. We should first note that naturalism does not imply that all scientific theories are true. Even the most highly confirmed theories can be false and, on occasion, this comes to light when they suffer disconfirmation. But naturalism does imply that science alone reveals where science errs. What in chapter 2 I called the first principle of naturalism, the primacy of science as a way of knowing, implies that it is unreasonable to believe what conflicts with science. Specifically, beliefs acquired by ordinary means, observation not constrained by scientific methods, are less well justified than highly confirmed scientific theories. When conflict arises science trumps its competition. Presumably, belief that a miracle has occurred does not result from scientific inquiry, but from ordinary empirical experience, either one’s own, or what others have reported. Therefore, if the belief that a miracle has occurred implies that science is mistaken about something, that well-confirmed theories are false, then belief in miracles is at odds with scientific naturalism and unreasonable. Miracles do not, however, necessarily conflict with science. Christian witness to God’s miraculous acts on behalf of human beings never involves claiming that science is mistaken in its account of what happens in the natural course of events. Suppose that a young woman, Mary, reports that she conceived while a virgin. If her account of her condition implies that well-confirmed theories of reproductive biology are false, then it is unreasonable to believe her. Those theories could be false, but only further scientific investigation could produce sufficient evidence to conclude that they are. Mary’s testimony is not adequate for disconfirmation. As David Hume would tell us, it is more likely that she is lying, deluded, mistaken, or whatever, than that what science tells us about the mechanisms of human reproduction is wrong. 3 If claims on behalf of miracles implied that well-confirmed scientific theories are false, then the secular naturalist is right and it is irrational to believe that a miracle has occurred. But to believe that Mary the mother of Jesus became pregnant miraculously is not to believe that biological science is mistaken in what it tells us about what causes pregnancy. True scientific theories describe what happens in the natural course of events and, by implication, what happens in the absence of supernatural intervention. Far from denying scientific theories about the mechanics of reproduction, Christians who believe in the virgin birth take for granted that there is a way things happen naturally and with that in the background assert that something happened that was not naturally possible. God’s intervention results in an event other than what biological theories predict, but this does not
52
Chapter 3
falsify them because their scope is limited to what transpires in the natural course of events. To claim on the basis of something other than scientific inquiry itself, that a hitherto well-confirmed theory is now disconfirmed is unreasonable. Ordinary, non-scientific experience cannot disconfirm scientific theories, but it can make it reasonable to wonder whether something outside their scope has occurred. This can, in turn, prompt us to ask whether the event in question is something we could reasonably consider divine intervention. If we reach that conclusion, we do not thereby abandon the primacy of science as a way to know; we do not discard a well-confirmed scientific theory on non-scientific grounds. Ordinary empirical experience can play a preliminary role in reasoning that leads us to conclude that a miracle has occurred. The secular naturalist may, of course, contend that there is no reasonable route to the final conclusion that a deity has intervened in the natural order, but he cannot do so merely on the ground that the reasoning was not initiated by scientific investigation. For he cannot deny that similarly ordinary experience, even the most casual observation, can play an analogous preliminary role in the reasoning that leads to the disconfirmation of wellconfirmed theories. Claiming, on the basis of something other than scientific reasoning itself, that a hitherto well-confirmed theory is now disconfirmed and thus false is unreasonable. Ordinary, non-scientific empirical experience cannot itself disconfirm a theory, but it can launch a process that eventuates in disconfirmation. Ordinary human experience can play a parallel preliminary role in the reasoning that leads to the conclusion that there has been a miraculous exception to a true scientific theory. Imagine, to modify an actual notorious episode, that a researcher assembles an apparatus in which current sent through palladium electrodes immersed in heavy water causes electrolysis. A visitor to the lab touches the container and mentions how warm it feels. A researcher, besotted with visions of a Nobel Prize, leaps to the conclusion that he has induced nuclear fusion at room temperature. Well-confirmed theory precludes this, implying what all experience has hitherto agreed with: fusion occurs only at extremely high temperatures. Noting that the apparatus produces an unexpected amount of heat, initially observed by the visitor to the lab, does not justify the conclusion that cold fusion actually occurred, or that physics has been mistaken in its understanding of fusion. In the actual case initial media hype, inspired by the prospect of limitless cheap energy, was followed by failure to replicate the results. The claim for cold fusion was discredited, but things could have gone differently. We may continue our imagined scenario to a happier conclusion: the initial observation, which did not justify believing room temperature fusion had occurred, reasonably motivated scientific investigation which revealed that it does occur, and that the erstwhile highly-confirmed theories that entail its impossibility are false. The evidence that gives birth to
Miracles
53
the researcher’s suspicion that well-established theories are false could have been the first step toward properly scientifically constrained inquiry that disconfirmed what had been taken to be scientific truth. Had the inquiry turned out this way, the visitor’s initial report would not have justified belief that room temperature fusion was underway, but it would have justified the scientific investigation that led to the conclusion that this is what was occurring. 4 Ordinary evidence, the kind we rely on in daily life, cannot disconfirm scientific theories, but it can make it reasonable to pursue scientific inquiry that might ultimately lead to disconfirmation, and thus to science revising what it says. Similarly quotidian evidence can play an analogous role in reports about the miraculous. It cannot justify believing that something has happened contrary to what the theories of science predict−in this case not something that falsifies a well-confirmed theory, but an instance of divine intervention, but it can give rise to reasonable suspicion that this has happened and make it reasonable to pursue further inquiry. In this case the justified inquiry is not scientific but theological. The fact that to ordinary observation a theory seems to be false is not a sufficient reason to believe that it has been disconfirmed. It is at most a good reason to pursue scientific inquiry guided by the idea that it might be false. Similarly, the fact that to ordinary observation it seems that a miracle has occurred is not a sufficient reason to believe that there has been a miracle. It is at most a good reason to pursue theological inquiry in which we ask whether this is the kind of thing that God might do. The Christian faith is trust in the God who miraculously resurrected Jesus of Nazareth. One morning in the fourth decade of the first century C.E., women and men who had been close to Jesus, and who had the best of reasons to believe that he had been put to death two days previously, seemed to see and hear him alive and well. They were unaware of today’s highly confirmed biological, chemical, and physical theories that imply that—and explain why—the truly dead stay dead, but they had no doubt that this is how things go in the natural course of events. It seemed to them that something extraordinary had happened, something that would not happen naturally. Their experience, extraordinary as it was, would not in itself justify the belief that a miracle had occurred, that God had intervened, causing something that would not otherwise have happened. If that’s all there was to it, the reasonable course would have been to take what they saw as having some natural, but unknown, explanation. 5 The conclusion that God did, indeed, act directly upon the world, bringing something about that would not otherwise have occurred, is reasonable only upon the further consideration that this is, in fact, something that God could have been reasonably expected to do in the circumstances. The mere encounter with something lacking known natural explanation does not make it reasonable to believe that God has performed a miracle, even for those convinced that there is a God who can, and sometimes does, do
54
Chapter 3
this. We cannot reasonably invoke divine intervention merely to close an explanatory gap opened by some strange and striking event. 6 For those who already believe that there is a Deity with various characteristics and purposes, it can be reasonable to conclude that an event for which no natural explanation is available was an exception to the laws of nature, something that would not have happened in the natural course of events, but is instead due to divine intervention. That the event was miraculous might be the best explanation. If it is not, because there is some natural explanation that is at least as good, then we ought not believe that it was miraculous. Scientific reasoning that confirms an explanation is more reliable than our reasoning about what God might have done. We ought not to conclude that what happened was miraculous unless there is no good natural explanation. This is only a necessary, not sufficient, reason to believe that it was a miracle. The best explanation, in virtue of being the only explanation anyone can think of, might be that it was a miracle, but this does not justify the conclusion that it was a miracle. The best explanation might not be good enough. One must also believe that there is, or that at least that there might be, a God who would intervene in this way. Without those antecedent beliefs, it is most likely more reasonable to assume that there is some unknown natural explanation. 7 Despite intimations in Jesus’ teaching that something like this was coming, his disciples did not expect him to be resurrected. On the contrary, when he was executed, they were thrown into despair and fear. They did not readily believe that Jesus had been resurrected even when they encountered him freed from the grave. This came about only when, with help from the resurrected Jesus himself, they made retrospective sense of Jesus’ life and work and saw that it was reasonable to believe that the faithful God of Israel would resurrect and vindicate its messiah. A theological justification of this kind is as necessary to the rationality of belief that a miracle has occurred as the initial observation of the event that appears to lack natural explanation. If a random human being were to get up after being to all appearances dead, there would be no particular reason to believe that God was miraculously at work; the reasonable conclusion would be that an anomalous, unexplained, but presumably natural event had occurred, or that appearances are deceptive, and the person did not really die. The lack of a reasonable natural explanation does not, on its own, justify the supernatural explanation. Belief in the miraculous does not challenge the primacy of science as a source of knowledge. Belief in miracles can be reasonable, even for those guided by naturalistic scruples, so long as belief in God is reasonable. Whether the belief that God exists can be justified when naturalistic constraints on human knowledge are taken into account is a further matter, to which we will return, but first, there are other objections to belief in miracles.
Miracles
55
THEOLOGICAL OBJECTIONS Secular naturalists are not alone in denying the miraculous. Miracles as traditionally conceived are disreputable in the eyes of some Christian thinkers. They object to the idea of God acting in this direct way as crude, meddling, arbitrary, incompetent, and as violating the integrity of creation. No doubt for some this arises from the idea that belief in miraculous intervention is discredited by modern science but, as we have seen, this relies on the mistaken assumption that a miracle would disconfirm an otherwise adequately confirmed theory. A different type of objection, voiced by Arthur Peacocke, is that the idea of God bringing about results that would not otherwise have happened makes sense only if the natural world is understood “basically as a mechanism, controlled by inviolable ‘laws of nature,’ deterministic and therefore, ‘in principle’ at least predictable.” 8 Many Christian thinkers hold mechanistic accounts of the world in low regard, attributing any number of evils to early modern science’s portrayal of the world as a great machine. There is, however, no necessary connection of divine intervention to this picture of nature. All it requires is that nature exhibits enough order for there to be matters of fact about how things naturally go, facts that generate expectations that God might confound by causing something else to occur. One need not have assimilated a mechanistic, deterministic, or any other scientific model of the cosmos to have reasonable expectations about what happens in the natural course of events. Consider again the New Testament account of Jesus’ miracle at the wedding in Cana (John 21–11). The guests surely took for granted that water does not, in the natural course of events, spontaneously turn into wine and that, if it were to do so, that would be contrary to what normally happens. The idea of God acting to produce an effect that would not naturally have happened requires only belief that there is some way things naturally go, and that there is a God who can make them go differently. Scripture does not describe pre-scientific people taking miracles for granted; despite their lack of scientific knowledge, they were as amazed as we would be today. Rather than making miracles plausible, the idea that creation is governed by mechanistic laws might make divine intervention seem implausible. If the world’s future is completely determined by its initial conditions and the causal laws that govern it, and those conditions and laws are freely chosen by the Creator, then it might seem that whatever the divine aims, they can be achieved naturally, by secondary causes. A God who, despite this, creates a world in which natural causes realize only some creative ends, but must later intervene, performing miracles to achieve the remainder, may seem inept or capricious. However, this objection has force only on the assumption that all God’s aims in creation can be achieved by natural causation. If there are goals that in principle
56
Chapter 3
cannot be achieved by secondary causes, and which can be achieved only if God intervenes, making things happen that would not happen naturally, then there need be nothing untoward in divine intervention. From the perspective of Christian faith, God has aims that can be achieved by secondary causes, but God also has creative aims that can be achieved only by miraculous intervention. With God’s fundamental creative purposes in view, we can draw a principled distinction between what God achieves naturally and miraculously. CREATION’S INTEGRITY Some see miracles as manifesting a practical inconsistency that would impugn God’s judgment. A ruler enacts a law requiring that all cats be kept on leashes at all times, but occasionally, upon encountering a cat, frees it. Does he want the cats leashed or not? Making the law, he acts on the desire for all cats to be leashed, but when he unleashes them, he acts on the desire for some not to be leashed. Human law-givers can, of course, have inconsistent or inconstant aims and thus behave capriciously, but we expect more of the Creator. Authoring nature’s laws, God acted on intentions for creation. If God now makes things happen other than what those laws call for, the divine intentions must be in conflict. A Creator whose aims are inconsistent or vacillating makes a world that lacks the coherence, intelligibility, wholeness, and beauty—the integrity—it would otherwise have. The objection concludes that God’s purposes are not inconsistent or vacillating so, having set out the laws that govern creation, God does not proceed to make things happen that would not have occurred in the natural course of events the laws govern. God has the capability to do anything that is absolutely possible, even if it is not naturally possible, but having once authored this world’s laws, henceforth causes only what it naturally possible. This objection to miracles overlooks the crucial fact that God authored nature’s laws not for their own sake, but as a means to the creation’s ultimate goals. Indirect divine action, secondary causation by way of natural processes, is one way that God pursues creative aims, but there is no guarantee that those goals do not also require direct action. A central claim of this chapter is that God’s ultimate goal, persons truly distinct from God yet enjoying everlasting fellowship with their Creator, cannot be achieved solely by indirect divine action. The communion of Creator and creatures, their personal interaction, is possible only if the Creator acts directly upon them. In the next chapter, where the issue is the origin of the human species, I will argue—in opposition to widespread opinion—that the Creator’s ultimate goals require that created persons not be brought into existence in some direct, miraculous way, but be created by
Miracles
57
way of indeterministic secondary causes. Here, the issue is how God deals with persons once they have been created. DIRECT DIVINE ACTION WITHOUT MIRACLES Some Christian thinkers recognize the indispensability of God’s direct action upon the creation, but reject the miraculous, on the ground that this means that God intervenes and thus violates creation’s laws, which God, the author of the laws, would never do. First, we should recognize that to speak of violating a natural law is at best misleading. The laws of nature, not being prescriptive, cannot in any literal sense be violated, the way moral laws, or the positive laws humans promulgate, are violated when they are disobeyed. Natural laws are not prescriptions and thus cannot be violated. Scientific theories can be shown to be false, but to disconfirm a theory is just to discover that a human belief is false, despite our having had very good reasons for it. If Wile E. Coyote runs off the edge of a cliff and, instead of plummeting to the ground, hangs in mid-air, he does not disobey the law of gravity. It might be that, despite appearances, nothing other than what our wellconfirmed theories call for has occurred, because some other natural force is in play. Maybe tiny rocket motors are embedded in his paws. If something other than what our theories imply will happen has actually happened, this might be because what we took to be a law of nature is not, after all, a law. In this case, something has occurred that implies that our theories are false, and this might be revealed by subsequent scientific testing that disconfirms them. The further possibility is that Wile E. is—at long last—the beneficiary of divine intervention, and something is going on that would not occur but for direct divine action. Wile E.’s suspension in mid-air would not happen in the natural course of events, but it happens because God intervenes. In none of these cases are the laws of nature broken. All the claim that God would not violate nature’s laws appears to amount to is the assertion that God would not cause events that are not naturally possible, even when acting upon the world immediately, rather than indirectly. So long as the basic laws of nature are indeterministic, there are events that would not naturally have occurred but which are not natural impossibilities. For instance, suppose that there is a low but non-zero probability that a relatively unstable atomic nucleus will emit a particular particle in the next 60 seconds. God immediately causes it to emit that particle, thereby initiating some desired cascade of causes to produce an intended effect. This is divine intervention, a miracle, despite the fact that the emission of the particle was consistent with the relevant natural laws. It was improbable, but naturally possible. The crucial con-
58
Chapter 3
sideration is that God’s intervention causes something that would not otherwise have happened, not whether it was a natural possibility. 9 It is possible that God sometimes acts directly upon the creation but never brings about events to which natural law assigns zero probability. This might be true, and while it has, perhaps, an aesthetic appeal, there seems no compelling reason to insist on it. If God can on occasion best pursue the divine aims by causing events that natural law implies do not occur, there is no guarantee that this will not happen. The laws are not sacrosanct, but means to God’s ends, ends sometimes best pursued in other ways. When it comes to prescriptive laws, we know that God does not hesitate to promulgate laws at one time but later suspend or abandon; the purity code of ancient Israel is a familiar example. The same might be true of the laws of nature. 10 DOWNWARD CAUSATION One strategy for construing divine action without miracles construes God as acting upon the world by means of downward or top-down causation. Its advocates offer this as a mode of divine action that has no effects inconsistent with natural law, indeed, no intervention of the divine in the creation. That there are phenomena in the natural world plausibly described as instances of downward causation is clear enough. The workings of the human mind/brain can be described in more than one way. There are, for instance, the lower-level processes described in terms of GABA inhibitors, neurotransmitter gradients in synaptic clefts, evoked potentials, and so forth, as well as higher level mental processes described in terms of thought, sensation, emotion, and so on. 11 Neural events cause neural events and mental events cause mental events, but we can also describe mental events as causing neural (physical) events: “Rocky’s synapses are firing rapidly because he sees Sally.” The plausible cases of top-down causation are cases in which higherlevel entities, properties, or events supervene upon the physical world. The mental supervenes upon the physical in the sense that while there can be no mental changes without changes in the brain, there can be changes in the brain with no corresponding mental changes. However, supervenience is a mark of ontological dependence. It is what we should expect if the mental were entirely dependent upon the physical, if mental things are in some sense just a kind of physical thing: arrange the atoms in such and such a way, and a conscious mind exists. 12 But God does not depend upon the creation in order to exist. God is its transcendent Creator. Whatever role there might be for supervenience in understanding the relation of the mental to the physical within the created world, what cannot be right is the idea that a transcendent God supervenes on the
Miracles
59
creation. From any physicalist perspective, the relation of the mind to the body is not a plausible analogue of God's relation to the world. Similar considerations count against the appeal to another version of downward causation as a model for divine non-interventionist influence on the world. The behavior of constituent parts of a whole is sometimes explicable, not simply by paying heed to their causal relations to one another, but to the nature of the whole of which they are parts. Arthur Peacocke suggests that when there are “complex systems constituted of complex subsystems at various levels of interlocking organization,” then “the state of the system-as-a-whole could itself properly be regarded as a causal factor influencing events in the ‘lower’ subsystems; constraining them to follow one course rather than another.” 13 To construe God’s action upon the world as the effect of the whole upon its parts, as Peacocke suggests, again portrays God as dependent on creation in a way at odds with the Christian belief that the world depends upon God for its existence, while God does not depend upon the world in this way. When the parts of a whole stop existing and are not replaced the whole no longer exists. In contrast, were all created things to stop existing, God would still exist. It seems unlikely that divine action can be construed as downward causation without moving beyond traditional Christian conceptions of God’s relation to the world in the direction of panentheism, pantheism, or process thought. The theologian Keith Ward appeals to divine action by means of topdown causation to close gaps in Darwinian explanations of the emergence of rational, conscious beings, as well as to make possible God’s interaction with those persons once they have been brought into being in this way. Construing top-down causation as the effects a complex whole has upon its parts, Ward states, “for a theist, the ultimate complex whole consists of the universe and God.” He infers that God has a top-down causal effect upon created things, distinguished from the creative act in which God established the laws and initial conditions of the universe as well as from miraculous interventions. 14 However, the idea that God and the universe comprise a complex whole is problematic. Is there some entity of which God and the creation are parts? There is the set of existent things that has God and the creation as its members, but sets, unlike wholes, are abstractions and have no effects on their members. Top-down causation seems to have no relevance to divine action upon the world so long as we suppose that God and the world are distinct entities. Maintaining this distinction is of the first importance for the Christian faith: the God revealed to us in Jesus Christ is the God who reaches out to, and loves, that which is other than God.
60
Chapter 3
THE NECESSITY OF MIRACLES If God and the world are distinct entities, as Christian faith traditionally asserts, then, unless God is to have nothing to do with the world except by way of what follows from how it is created—secondary causation— God acts upon the world, bringing about effects in it that would not have occurred had God not acted upon it in this manner. Insofar as this is a world governed by causal laws, this may involve events the laws preclude, or at least that they describe as improbable. So it is not surprising that those who have a negative view of miracles find attractive conceptions of God’s relation to the world that do not construe God and the world as numerically distinct entities, that is, to panentheism, if not pantheism. There is, as Colin Gunton noted, a “pantheistic undertow” in Christian theology, an unwillingness to take with full seriousness the fact that God creates things that are not God. 15 One manifestation of this is a tendency to find miracles unworthy of God, rather than essential to a Christian conception of God and creation. The God who creates is the God who is love, an everlasting community of persons—Father, Son, and Spirit. The Triune God of Christian faith created for the purpose of sharing life with persons distinct from their Creator. This God creates not to have something to contemplate or to control but to reach out to persons who are not God but who can come to participate in the divine life. The Creator’s intent is personal relationship with created persons. This, creation’s ultimate purpose, cannot be achieved by secondary causes alone. The omnipotent God can do anything possible, but entering into personal relationship with other persons without interacting with them is not a possibility. One person can interact with another person only by acting upon her, being acted upon by her, acting upon her in response, and so on. To act upon another person is to cause in her an effect that would not otherwise have happened. This is truistic with respect to relations among human persons: Martha and Marvin might have gone through life having no effect on one another, but if they enter into any sort of relationship, they interact. Martha will make things happen to Marvin that would not otherwise have happened to him. He will have thoughts and feelings, particularly about her, that he would not otherwise have had. Because of these thoughts and feelings he will do things that he would not otherwise have done, and some of them will have effects on Martha, causing in her what would not otherwise have occurred. Marvin and Martha are inhabitants of the natural world, so they have no effects on one another that would not have happened in the natural course of events. Neither acts from beyond the natural world upon the part of nature which is the other person. In contrast, when God, who transcends the natural world, acts upon human beings, things happen that would not have happened naturally: miracles occur. Further, human actions that bring about effects in God are analogous to the mira-
Miracles
61
cles God performs. Creating distinct persons, God becomes open to being changed by others. Had humans not come into existence the life of the Holy Trinity would have unfolded differently. It might seem that this is not necessarily true, because God—if there was some reason to do so—could act upon personal creatures only indirectly, by means of natural causes, without resorting to intervention. God does act upon human beings in this indirect way. God creates the universe with its initial conditions and authors its causal laws; thus causing all the effects that ensue, irrespective of how much time transpires. God causes creatures to exist, to have the characteristics they have, and whatever befalls them in the course of their lives is, ultimately, an effect of God’s initial act of creation. This is true even if, as I assume to be the case, the causal laws are indeterministic, in which case the divine intentions and foreknowledge are only general. The existence and characteristics of persons in creation answers only to God’s general intentions; there were no specific intentions. We were created but not designed. This comes fully into view in chapter 4, which addresses human origins. It is possible for God to act upon creation only indirectly, without miracles, but to do so in a way that creates the appearance of genuine interaction with creatures: in the beginning God freely selects the initial conditions and laws with the intention that Martha, a human being who will come into existence billions of years later, will pray for guidance, and soon thereafter will receive a call from her friend Marvin, who will supply the needed advice, which Martha will regard as an answer to prayer. At least if the laws are deterministic it is possible—even if far beyond what finite minds can imagine—for God to orchestrate things in this way. It seems to Martha that God responded to her prayer, intervening to prompt Marvin to call her, but this is not quite what happened. God really does act upon Martha (by way of acting upon Marvin). God does so out of love for her, but the intention to help her in this way precedes her request for help. God, by secondary means, acts upon the human being, but she does not act upon God; it just seems to her that she does. There is no interaction and no genuine personal relationship. God could create a world in which it appears that what creatures do affects God, making changes in and supplying reasons on which to act, but we have no reason to believe that God would do so. The divine creative aim is to bring personal creatures into real communion with their Creator. God’s aim is to interact with created persons, not merely for it to seem to them that they interact with God. Perhaps, on other conceptions of the deity, miracles can reasonably be dismissed as unnecessary and arbitrary, unworthy of the Creator and unintelligible in light of the divine aims, but this is not the case for Christianity. God responds. Among the conceptions of deity most difficult to reconcile with this core Christian conviction are those on which God is timeless (atemporal) rather than everlasting in time. A time-
62
Chapter 3
less God cannot change, cannot be changed by creatures, and thus cannot interact with them. 16 “HISTORY” NOT “NATURE” Christian faith is committed to miraculous divine intervention. God interacts with created persons and thus causes events that would not otherwise have occurred; they are not effects of the initial act of creation. Not every event in the universe has a natural cause. Whatever claims we make about the formal consistency of miracles with naturalism, it might seem that this respects the letter of naturalism but traduces its spirit. For some the essence of anything reasonably called naturalism is the repudiation of supernatural causes in favor of natural explanations. This is a legitimate, but misplaced, concern. Many Christians do manifest a conception of miracles that is seriously at odds with scientific naturalism, for they place divine intervention in competition with scientific explanation. They contend that general features of the world, the sort of things we expect science to explain as due to natural causes, are results of miraculous intervention. To claim, for instance, that there can be no adequate scientific explanation of the origin of biological life in the universe, or of the origin of species, including our own, or of rationality or consciousness, undoubtedly is to abandon the naturalistic perspective. It is to snatch from science its most significant challenges and to hand them over to faith. The task of science is to explain how the world works, to delineate its causal structure, and thereby to satisfy intellectual curiosity and give us powers of prediction and control of nature. True to the revelation that it has received of the Creator's character and purposes, Christianity has no reason to declare life, or humanity, or mind, or morality, or religion, or any other general feature of the creation, beyond the reach of scientific explanation. We have no reason to deny that general features of nature come about in the natural course of events, products of secondary causation. Hydrogen atoms forming in the cooling debris of the Big Bang were drawn together by gravity into stars and the stars into galaxies. The heavy elements were created in the hot depths of expiring massive stars. In their explosive deaths those stars flung their carbon, oxygen, silicon, iron and so on into space. From these materials planets formed around newly forming stars and as they cooled, organic chemistry gave rise to living things, capable of making not quite perfect copies of themselves. The blind but endlessly subtle and persistent engine of natural selection went to work, and life evolved in marvelously diverse and complex ways. In due time rational, conscious animals—persons—appeared: all this, the secular naturalist insists, takes place naturally, without miracles. All this happened in the natural course of events.
Miracles
63
The Christian naturalist enthusiastically agrees with all this, but of course sees in it the mighty acts of the Creator, at work by secondary causes. Beyond this, he discerns reasons to believe that God also acts upon the creation in an immediate way, making happen what would not otherwise have happened, causing events for which no scientific or other natural explanation is possible. But Christianity has no reason to seek those miraculous acts in the processes that structure the world at large, in what we might call nature. In contrast, we properly expect them in history, that is, in the individual and corporate lives of created persons. The God of Christian faith at times intervenes and at times refrains from intervening, in either case with the enduring aim of loving, and being loved by, created persons, pursuing one purpose by different means. Worries that in acting miraculously God is capricious, arbitrary, or inconsistent are groundless. Keeping God’s deep creative purposes in view profoundly shapes expectations about what the world is like. We have no reason to suspect that after the initial act of bringing the universe into being from nothing, but prior to the emergence of created persons, the Creator intervened. God could have, but we know of no reason for doing so. In light of God’s aim of creating persons distinct from their Creator, the reasonable expectation is that God would not intervene in the natural processes that bring them into existence: this is the principal claim about the origin of the human species I make in chapter 4. Once created persons appear, God acts freely and responsively, intervening in the creation in pursuit of fellowship with them. So far as we know, these are the only miracles. Christian faith offers no support for miracles that undermine the scientific project of explaining the world and the place of human beings in it. We may usefully say that God intervenes in history but not in nature. However, in deploying this terminology, two things should be kept in sight. First, when we describe God’s acts in relation to created persons as miracles in history, as opposed to nature, this does not preclude effects in the physical world in contrast to human subjectivity. If God, in response to prayers, were to intervene to make it start to rain, or to cause malignant cells in a human body to stop replicating, this would count as an intervention in history, not nature. These acts would be miraculous divine acts in relation to created persons. Second, if God acts upon a human mind, bringing about some effect in a person’s subjective experience or cognitive dispositions, this, though an intervention in history, is a miraculous event in the natural world, for the human mind is the human brain and thus part of the physical universe. For the materialist, all God’s miraculous acts in history are in this sense also acts in nature.
64
Chapter 3
NOT ALL OR NOTHING Christian faith has no legitimate stake in the failure of the project of scientific explanation, combating its extension to human nature, or seeking—vainly—a way to accept contemporary science without its naturalistic implications. The Christian can approach the world with the same general expectations as the non-believing naturalist, practicing the same methodological naturalism that assumes things typically have natural explanations. Nor should Christians find plausible an integration of divine intervention into a general scientific account of the world, as socalled “creation scientists” and “intelligent design” theorists advocate. Such projects misconstrue the character of God's work in the world, seeking intervention where what God reveals about divine purposes should not dispose us to expect it. Science strives to know what happens in the natural course of events which, by definition, is how things go when God does not intervene, when God is at work by natural means. Miracles are particular occasions when things go otherwise, in keeping with God’s purpose in creating. Discussion of the miraculous has long been afflicted with an “all or nothing” mentality. We require a principled distinction between contexts in which God does, and does not, miraculously intervene. Today, secular persons, properly dismissive of Christian claims about the need to invoke miracles in place of scientific explanation, take it for granted that all Christian assertions about miracles, including those on which faith actually depends, are equally unfounded. Many persons of faith, recognizing the crucial role of the miraculous in Christianity's witness to a loving, responsive God, take for granted that God must have intervened prior to the existence of human beings, and thus that much of science is in error. There is nothing new here. Charles Darwin accepted the uniformitarian view that geological changes are the result not of divine intervention, but the very gradual effect of familiar natural forces. Commenting on this, the great geologist Charles Lyell wrote that Darwin rejected miracles and reported that for him, “as the miracles went out the window, so did Christianity, never to return.” 17 It is disconcerting that one can move so quickly from a geological theory, even one that supports a natural, as opposed to a miraculous, origin of the human species, to the conclusion that Jesus of Nazareth was not resurrected. From the mid-nineteenth century to today many, on both sides of the debates between faith and science, have seen the issue of miracles in stark terms: either the origin of species is miraculous, or nothing is. Either God intervened in ways that doom much of scientific inquiry to failure, or God did not intervene in the ways that the Christian faith confesses. This sustains a universe of discourse in which conservative Christians, committed to the traditional idea that God miraculously intervenes in the creation, feel compelled to repudiate science, and in which other Christians, disinclined to reject
Miracles
65
science, are driven to recast divine action so it does not involve miraculous intervention. A theology that embraces the Christian claim that God intends to live in relation to created persons, and that this is the underlying rationale for creation, has no quarrel with a science that aims to explain the general features of the world, including the existence and characteristics of human beings, without recourse to the miraculous. Nor does it hesitate to confess a God who interacts with creatures, miraculously self-revealing, intervening in their lives, and responding to them. Whether intervening in history, or refraining from intervening in nature, God acts for the sake of created persons. Absent that underlying rationale for distinguishing contexts of divine action, it might well seem arbitrary to deny alleged miracles in nature leading up to the existence of persons while affirming belief in the biblical miracles and in ongoing divine activity in the world today. And it might seem reasonable to see the success of science in explaining what some Christians took to be miraculous as a reason to dispose of Christian belief in miracles altogether. Drawing on the essentials of Christian theology to make a principled distinction between divine intervention in nature and intervention in human history, we can at least insist that claims about the latter be evaluated in their own right, and not thrown together with spurious claims about the inadequacy of scientific explanation. From the perspective of Christian faith, when God acts in exceptional ways, doing so is neither arbitrary nor unnecessary; it is explained by God’s ultimate creative purposes. In the next chapter, where I take up the evolutionary origins of humankind, I will argue that attention to these purposes leads to the expectation that God refrained from intervention in the processes that brought our species into existence. There are various perspectives, religious and secular, from which all, not just most, miracles appear incredible, but scientific naturalism as such is not one of them. Neither naturalism's high view of science as a way of knowing, nor its claims about what sort of world this is and what kind of things we humans are excludes the possibility of rational belief in divine intervention. The naturalist who is an atheist finds miracles incredible, but no more so than any other atheist. The Christian who is a scientific naturalist finds most reports of miraculous occurrences incredible, but no more incredible than any other sensible Christian finds them. Views long associated with naturalism do, of course, make belief in miracles dubious, but the conviction that this is a physical world, best known by science, and that human beings are material things, like other creatures the product of unguided natural selection, provides no basis for rejecting the possibility of divine intervention.
66
Chapter 3
MIRACLES TODAY It would be a mistake to discount the extent to which God’s miraculous action upon human individuals is indirect, mediated by history, when considering the current effects of God’s past miraculous activity. The Church is an ongoing community of persons bearing witness to God’s long past miraculous acts, above all the life, death, and resurrection of Jesus Christ. When someone today bears witness to Christ and thereby guides someone to faith, this is an effect, though a remote one, of God’s crucial, miraculous intervention in human history, channeled, by way of the apostolic witness, the New Testament writers, and the Church that over the generations preserved their testimony to the present day. However, Christians cannot restrict miracles to the past. We believe not only in the biblical miracles of twenty centuries ago, but in a God who continues to interact with human beings, revealing the divine presence and concern, fostering faith, nurturing trust, answering prayer, guiding, comforting, and encouraging. This is what we mean when we speak of God’s active presence in the person of the Holy Spirit. God cannot be responsive to human creatures in the robust way interpersonal relationship requires without intervening, and no miracle performed by God two millennia ago serves as a response to what we do today. God’s miraculous acts occur in the present as well as the past. These acts are, paradigmatically, directly upon the minds of persons. God can, and on occasion presumably does, act directly upon the physical world outside the human mind/brain, but we reasonably assume that because God’s intention is not to act unilaterally, but to trust those who put their trust in their Creator and choose to share in the divine work in the world, the typical contemporary miracle is within persons of faith. When we ask where God works today, other than by way of secondary causes, the most important answer is that God acts not alone, but through the community of faithful persons privileged to share in God’s work in this world. The human mind is part of the physical world; in acting upon the human mind God brings about changes in the human brain. It is conceivable that the immediate effects of divine intervention occur solely at, or below, the quantum-mechanical domain, and that they involve nothing at odds with natural law. Perhaps they do not even typically involve effects that are extremely improbable in light of physical law. God can bring about any desired effect in a human mind, but given concern for the integrity of the person, for her freedom and responsibility in relation to God, we can conjecture that interventions are most often not invasive or disruptive, but subtle and economical. Marvin suddenly realizes that it has been a long time since he talked to his friend, and this leads him to give her a call. In retrospect, he and Martha might ponder the possibility that God intervened, prompting him to come to her aid, but they have no
Miracles
67
way to know. It may well be impossible to distinguish the miraculous prompting of God’s Spirit from the natural workings of our minds in a world in which the good news of Jesus continues to be propagated. Here, as in other contexts where we have only a murky grasp of the natural causes in play, we rightly thank God for the good that comes our way, not knowing whether to regard it as miraculous, or to attribute it to divine providence, God’s work in the world by way of secondary causes. ACTUAL MIRACLES The primacy of science as a way of knowing provides no basis for rejecting the possibility of miracles, and the Christian faith has no commitment to miracles that compete with science. However, the Christian faith depends, not on the bare possibility of divine intervention, but upon the stronger claim that certain miraculous events have actually occurred. Martha need not know whether God intervened to move her friend to help her, but her Christian faith presupposes the miraculous resurrection of Jesus. If the world is as the naturalist conceives it, can belief in specified actual miracles be reasonable? In an influential essay, Langdon Gilkey once pointed out the instability of a theology that speaks of God’s acts in history yet, because of its implicit commitment to a modern, scientific view of the world, is too ambivalent to admit any actual miracles. 18 This book’s aim is not to show that Christianity is true, but simply that an alleged irreconcilability with science is no reason to believe that it is false. A sustained effort to justify theistic belief in general, or Christian belief in particular, is not my aim here, but a fundamental problem needs to be addressed. Christian faith is trust in the God revealed in human history, in the calling of Israel to be God’s people, in God’s word spoken through the prophets, and definitively in the word made flesh, Jesus, God’s Son. 19 The God who transcends this world enters into it, miraculously becoming known to the creatures for whom it was created. If trust in this God, and the beliefs it presupposes, are rational, it is because we rationally recognize certain historical events as miraculous. As a Christian, my first thought is that I believe that God exists because I believe that Jesus of Nazareth was miraculously resurrected. However, on pain of circularity in reasoning, the justification of belief in God cannot be grounded solely in the belief that a miracle occurred. Recall that the mere fact that something happens that appears to have no natural explanation does not justify belief that it was miraculous. The conclusion that a miracle has occurred can be reasonable, but only for those who already believe that this is not merely the best, but a good enough explanation. Otherwise, one has simply come upon a strange, unexplained event and has no good reason to ascribe it to divine intervention. As central as the life, death, and resurrection of Jesus Christ are in the grounding of Christian
68
Chapter 3
faith, we also need prior reasons, independent of claims about miracles, to believe there is a God capable of acting in these ways. If the Christian faith is to be vindicated as reasonable, some sort of recourse to natural theology is inescapable. However, whatever value traditional attempts to defend belief in God by natural means, with no reliance on divine revelation, might once have had, much of that project seems forlorn in light of contemporary science. Consider, first, arguments from design. Despite their ongoing popularity, they deserve no allegiance from the Christian community. The design argument implicit in the intelligent design movement’s dubious criticisms of evolutionary biology relies on one claim that is clearly true: in a world governed by indeterministic laws natural selection cannot be relied on to achieve results specified in advance. 20 This has force not against agnosticism or atheism, but against any form of theistic evolution which says that the natural processes were the Creator’s means to achieve specific ends, such as the existence of the human species. But it has no cogency against those who do not share the assumption that the human species was specified in advance. 21 The existence of the human species, and, in general, the remarkable diversity, complexity, and adaptedness of living things is adequately explained without invoking a divine designer. On the supposition that nature’s laws are not deterministic, what science tells us of human origins is at odds with the human species being the product of God’s design. In the next chapter I will contend, however, that belief in divine design itself conflicts with a Christian conception of creation, failing to cohere with the aim of bringing into existence persons distinct from their Creator. Some theists pin their hopes on cosmological fine-tuning as proof of divine design, but this is not clearly superior to explanations that invoke an anthropic selection principle. Our universe might be just one of many; alternatively, it might be one part of some greater metaverse. In either case, the highly improbable combination of fundamental constants that characterize our universe might demand no further explanation, such as a Creator who selected the constants required for it to produce life. One of the multiple universes explanations may well carry the day, and this design argument will no more survive the advance of cosmology than the traditional argument from apparent biological design survived Darwin. Evolutionary psychology, uncovering the natural origins of human morality as an adaptation to social life, undermines familiar arguments that ascribe to morality an objectivity and authority that allegedly only the existence of a divine lawgiver can explain. As we will see in chapter 7, this scientific explanation implies that the facts about right and wrong are not deeply rooted in objective reality, independent of human beings, yet it does so without implausibly making the moral truth dependent on the preferences of human individuals or the norms of human societies.
Miracles
69
The cognitive explanation of human religiosity, the subject of chapter 8, portrays it as a byproduct of the interacting evolved structures of human cognition. Along with the evident threat to debunk religious beliefs by revealing that they can be explained without the need to suppose they are true, this theory undermines attempts to justify belief in God by appeal to religious experience. In light of it, explanations of subjective experiences as actual experiences of God face the same problem faced by attempts to explain events as miraculous. They succeed only if we already have reason to believe that there is a God who could, and would, choose to be revealed in these ways. AN ABSTRACT ARGUMENT Contemporary science renders the most common kinds of natural theological reasoning ineffective. What remains is one highly abstract argument. This species of natural theology is traditional—it is essentially the third of Thomas Aquinas’s Five Ways to prove that God exists—but it now appears in a new context at the speculative edges of scientific cosmology. 22 Above, we saw that the theist who points to the cosmological fine-tuning argument as evidence of a designing God is countered by the possibility that our universe is one of some vast ensemble of universes or one component of a vaster, more ancient metaverse. If some such explanation of the surprising values of the cosmological constants is plausible, the inference to design fails. However, this does not mean that those of us who believe there is a God who created our universe should oppose the many-worlds hypotheses. The fact that an argument has a conclusion we agree with is hardly a sufficient reason to think it is a good argument. 23 To accept a plurality of universes, and thus to forego the design argument, is not to abandon the idea of creation. Multiple universe theories imply that in asking why our universe, with its particular set of basic laws, exists, we have not reached the limit of scientific explanation, the place where a supernatural explanation becomes a viable option. It leaves open the question of why the ensemble of universes, or the multiverse, exists. These theories dispel surprise at our universe’s fine-tuning, but it does not explain the existence of the greater reality of which our universe is a part. The reasoning is abstract, but simple; an attempt to account for the fact that the observed universe—or the greater reality of which the universe we observe is a component—is both contingent and intelligible. This physical reality, whatever its nature, exists, yet we assume that its nonexistence is possible. And we assume that there is a rational explanation for its existence and its nature. The physical reality we observe, whatever its relation to the whole of physical reality, exists, but it could have failed to exist. When something
70
Chapter 3
exists, but does not have to exist, we reasonably ask why it exists. More generally, when something is true, but does not have to be true, we reasonably ask why it is true. One asks why there is a doghouse in the backyard. There does not have to be a doghouse there. Its existence is contingent. Its presence is not self-explanatory. In contrast, some truths are not contingent; they are true and could not have been false. The supposition that 1 + 1 = 2 is false leads to a logical contradiction; it could not possibly be false. Once one understands a necessary truth, there is nothing left to explain. To persist in asking why it is true that 1 + 1 = 2 manifests not intellectual curiosity but failure to understand. This is not the case with contingent truth. If the universe is, as it appears to be, a contingently existing reality, it is reasonable to wonder why it exists. If there is an explanation of its existence, then it is intelligible. The existence of things within the universe is both contingent and intelligible: the doghouse does not have to exist, but it does, because Sally built it last summer. However, challenges arise when we seek an explanation of the contingent universe itself. The quest for explanation, and thus for intelligibility, ultimately terminates with necessity. Therefore, a dilemma arises: either the universe (multiverse, ensemble of universes) has no explanation, its existence is an ultimately unintelligible brute fact, or it is explained by reference to something else that exists not contingently but necessarily. In the latter case, the universe is intelligible, explained by way of its dependence on something that exists necessarily, and that necessary reality in turn is intelligible in virtue of being self-explanatory. In this scenario, everything is explained: the world we observe is not an unintelligible, brute fact. This reasoning delivers an intelligible universe, but it does so at the apparent cost of its contingency. If something necessarily exists and its existence necessitates the existence of something else, that thing also exists necessarily. Whatever necessarily follows from what is necessarily true is itself necessarily true. The contingency of the universe can be preserved only if its existence depends on, yet is not necessitated by, the necessarily existent thing. How might the universe depend on some necessarily existing being in a way that preserves its contingency, but which also renders it intelligible? One possibility is that it is a probabilistic matter: it is a necessary feature of the necessarily existing thing that there is some non-zero probability that it will cause our universe to exist. As a matter of fact, it did cause our universe to exist. Our universe thus depends for its existence on the necessary thing, yet its existence, being a matter of chance, is contingent. The explanation affords a degree of intelligibility, but it is not satisfying. We know why our universe exists: because the necessary, perfectly intelligible being exists, but we also know things could have gone differently and, if they had, our universe would not have existed. There is no further explanation of why they did not go differently.
Miracles
71
A second possibility points to a theistic explanation. Our alternatives are not just brute fact, necessitation or chance. Choice is a further possibility. Some things can be explained as the result of the free action of a rational person. We might explain why the doghouse exists by ascertaining that Sally built it, and that she freely chose to do so for good reasons. Concurring with her judgment, we have an intellectually satisfying explanation. Imagine that Sally is not completely embedded in the causal matrix of nature but in some mysterious way the ultimate originator of her choices and actions. She is an agent cause: what Sally does is caused not by events in her, by her having such and such beliefs and desires, but simply by her, the agent. In this case, Sally has reasons for her choices, but they do not cause them. Her choice to construct something for her dog to sleep in is explained, in the sense that it is clear that she had very good reasons for doing it and chose to do it for those reasons. Yet she could have chosen differently, even if her beliefs and desires had been no different. Her having those reasons did not necessitate the choice. Her radical freedom introduces an ineliminable element of contingency into the world. Sally might have had equally good reasons to do something else, and there might be no explanation why she did not do that instead. Because she is rational, she typically does not make choices for which she lacks good reasons, but because she is free she could do so; nothing necessitates her choosing reasonably. When she chooses among equally good courses of action, there is no explanation of why she chose one over the others. Her choice did not have to happen, yet it is an intelligible contingency, even if it lacks the full intelligibility of what happens as a matter of necessity. Things explained this way possess a greater degree of intelligibility than what is explained as the effect of probabilistic causation. Explanations of this kind cannot actually be applied to human action. As I will contend in chapter 6, a human being is not an agent cause; she lacks libertarian free will. She can at best become the material image of this radical freedom. In contrast, Christian theology assumes that God is radically free, an agent cause transcending the realm of natural cause and effect. God acts for reasons, but those reasons do not necessitate God’s choices. God freely created the world, so its existence is contingent. God is also wise, and in choosing to create, God acted for good reasons. Those revealed reasons are central to this book’s portrayal of human nature and the human condition. Excellent as they are, these reasons did not necessitate God to choose to create. Creating a world that gives rise to persons distinct from their Creator is profoundly in keeping with God’s character, yet God could have chosen not to create this world, or anything at all, and have been no less wise and good. A universe created by a necessarily existing being as the freely chosen expression of superabundant love has a high, but not the highest conceivable, degree of intelligibility. 24 A world that necessarily exists, either be-
72
Chapter 3
cause a necessarily existing being necessarily brought it into being, or simply because, despite appearances, it is not contingent, leaves nothing beyond the reach of rational explanation. Should we consider these superior accounts, in virtue of ascribing greater intelligibility to the world? All things being equal, explanations that offer greater intelligibility are superior to others, but countervailing considerations can make it more reasonable overall to embrace an explanation that is incomplete, leaving something unexplained. The best explanation is the one that offers the highest degree of intelligibility consistent with whatever else we know. A well-known example of this transpired in mid-twentieth century physics, when the long debate between Einstein and Bohr culminated in Bohr’s indeterministic interpretation of quantum mechanics carrying the day even though, in an obvious sense, a deterministic world manifests greater intelligibility than one in which the basic laws are probabilistic. 25 The discovery that our universe is just one of many might signal that we have at last reach the limits of scientific explanation, where a theological explanation might be reasonable. Scientific naturalism has no quarrel with the idea that natural explanation comes to an end, just as it has no quarrel with the possibility of events occurring other than those that occur in the natural course of events. Inherent limits to scientific explanation are an unexceptional feature of science. Science explains by locating events in the causal structure of the world, which is ultimately a function of nature’s fundamental dynamical laws: given such and such conditions the event to be explained had to happen, or at least its occurrence had a fixed probability. While less basic causal laws can, in principle, be derived from more basic laws, and thereby explained, nature’s most basic laws cannot be scientifically explained. Yet, to all appearances, these laws are not self-explanatory. They could have been other than what they are, and it is natural to ask why they are what they are. When we ask why the most basic laws of nature are what they are, and why the most primeval initial conditions were what they were, we reach a limit that the naturalist can readily acknowledge. And we reach the ground on which the theist can start to construct an argument for the existence of God, for reasonable belief in actual miracles, and thus for the truth of the Christian faith. So long as there is a distinction between actuality and possibility, between what does exist and what could exist, the Christian can invoke God as the agent whose free choice makes some of the possibilities actual. As physicist-theologian Robert John Russell notes, “No matter how strong an explanation science can give ‘of the way things are,’ an element of the unexplained always remains; hence science will never eliminate the meaningfulness of contingency in the creation tradition.” 26 Conceivably, the apparent contingency of the natural world is merely an artifact of the limitations of the human mind. Maybe the fundamental laws and initial conditions are absolutely necessary and would be self-
Miracles
73
evident to intellects more profound than ours, and possibly to future human science. This might be true, but it incurs significant costs. It was the recognition, in the late medieval period, that nature’s governing principles are contingent, and thus not knowable a priori, but knowable only by empirical observation and experiment, that set the stage for the birth of science in the modern sense. 27 Those of us unable to take seriously the idea that empirically discovered natural laws such as f = ma or e = mc2 are not ultimately different than mathematical necessities like 2 + 2 = 4, seem left with no good alternative to regarding the existence of this world as at bottom unintelligible or seeing it as a result of rational free choice. This latter explanation maximizes the intelligibility of the world without demanding that we discard the formative assumptions of modern science. It is, therefore, reasonable to believe that this is a created world and thus that there is a wise and free agent, a person, who created it. It is at face value an explanation that maximizes intelligibility to the degree that we can do so without having to deny this world’s contingency. In this rarefied realm, we find reasons to believe that there is a God who created the universe. But caveats abound, and diffidence is called for. First, we must acknowledge that there is no guarantee that reality cooperates with human hopes for intelligibility. Our minds are the product of natural selection in an environment in which being sure that there are causes of, and thus explanations for, what happens enhanced the odds of being around long enough to reproduce and to send to the future genes that build neural circuitry that sustains the assumption that things don’t just happen, but have a rational explanation. The quest for intelligibility might, for all we know, be fruitless when directed to the universe at large because reality is not, at bottom, intelligible, even though the ancestral human environment was a close enough approximation to intelligibility to select for creatures wired to make it their default assumption. I cannot imagine that this is true, and cannot help but proceed on the assumption that it is false, but there is no way to prove that this predilection fastens onto how things really are. Scientific naturalism rightly looks askance at the idea that in philosophical reasoning, such as we indulge in when we reflect on such matters as possibility, necessity, contingency, and intelligibility, we possess a way to know the world on a par with science. Even the most intuitively plausible results here are, in contrast to the well-confirmed theories of the sciences, tentative. This inherent uncertainty about the conclusions reached by this kind of reasoning may increase when we note that in describing the universe as a contingent entity, and postulating a necessarily existing person to explain its existence, we presuppose the legitimacy of applying the concepts of contingency and necessity to things, rather than seeing them merely as features of our language or of our conceptual schemes. This assumption is controversial; in fact, many philosophers
74
Chapter 3
whose naturalistic inclinations I share think it is a mistake to ascribe these modal properties to things. 28 The theistic explanation of the contingent world implies that there is a necessarily existent being about which there are contingent facts. God necessarily exists, is necessarily rational and necessarily makes free choices, but which choices is a contingent matter. It is, to say the least, not obvious how this can be true. More generally, I do not know how to explain how a necessarily existent entity can be a concrete particular, as I assume God is, as opposed to an abstract object, let alone to explain how it can be a personal being, in fact one with a very distinct character, beliefs, and aims which the exercise of a radical free will achieves. 29 In light of these issues, I do not suppose that this reasoning compels rational belief. There is enough in it that someone not already inclined to find theism plausible will find hard to swallow to prevent this. What it does is indicate is that the idea of a personal Deity who is a rational free agent is not absurd, but is in keeping with our deepest intellectual aspirations. Whether it shows that it is reasonable to believe that God exists, to suspect that God exists, i.e., to believe that a Creator’s existence is likely, or perhaps just reasonable to hope that a Creator exists is, for me at least, unclear. The value of this line of reasoning lies not in its conclusion as such, but in its rendering a particular explanation of historical events reasonable enough to believe. AN HISTORICAL ARGUMENT Like other exercises in natural theology, the abstract argument, even if we were to embrace its conclusion without reservation, brings us to a generic theism, not explicitly to the God of Abraham, Isaac, and Jacob, the God made flesh as Jesus of Nazareth. Conceivably, our universe was created not for the sake of a loving relationship with created persons, but for some other good reason. However, it coheres with the specifics of biblical faith. It concludes that there is a God who freely created this world for good reasons, while the Christian faith’s central claim is that God freely created this world for a good reason: for the sake of creatures who can share God’s Triune life. The historical argument turns to the historical specifics of the Christian faith and then at a critical juncture appeals to the abstract argument. By the late first and early second century C.E., many persons in the eastern Mediterranean region believed that Jesus of Nazareth had been resurrected. This belief is a matter of history for which we reasonably seek an explanation. Various possible explanations, historical, sociological, psychological are available. Some are clearly better than others, and perhaps there is a clearly best explanation. If there is a best explanation, then it is reasonable to believe it, but only if it meets some threshold of
Miracles
75
plausibility, only if it is a good enough explanation. 30 One very common and most often correct explanation of why people believe something is that it is true. For example, many people believe that John F. Kennedy was elected president in 1960 and, obviously enough, they believe this because it is true. The believing is historically connected, by testimony, to the event that makes it true. On the other hand, many people believe things that are false, and their beliefs call for some other explanation. For example, many people believe that Kennedy was killed by Lee Oswald, acting alone, but many people believe that he was killed by a conspiracy. Both beliefs cannot be true, so many people have a false belief. We can imagine future historians trying to explain why, decades after the assassination, many people believed that there had been—or had not been—a conspiracy, and investigating various hypotheses, some involving the truth of their belief, others not. Now suppose, for the sake of argument, that the best explanation of the widespread belief about Jesus being resurrected is that he was, in fact, resurrected. This is, of course, a very large claim that could be justified only by examination of the historical evidence. It can, I believe, be done, but any attempt to show this would take us well away from the task at hand, which is just to set out the kind of reasoning by which Christian faith could be justified. 31 Showing that this is so, that this is the best, or even the only reasonable, explanation, does not imply that Jesus was resurrected. The best explanation we can come up with might be false. Reasonable inferences to the best explanation—abductive arguments—do not guarantee that their conclusions are true, or even that they are probable. Sometimes even the best explanation is too unreasonable to believe in light of the rest of our beliefs. Imagine that we launch an investigation into crop circles. We consider a variety of explanations, for example that they are due to adolescent pranks, to fungal infestations, creative cows, meteorologically induced electromagnetic fields, attention craving farmers, and so on. We definitively eliminate every explanation we can think of except for one: crop circles are the work of visiting extraterrestrials. The only explanation not eliminated, it is by default the best explanation. However, this is a case in which the best explanation is not good enough. The reasonable course is to shrug one’s shoulders and say that there must be some other, better, explanation no one has thought of. The idea that space aliens travel to the Earth and do this is, in light of the rest of what we believe about the world, too implausible to believe, even if it is the best explanation. The fact that something is the best explanation of something we know occurred, such as the belief that Jesus was resurrected becoming widespread a half century or so after he was executed, is a good reason to accept it only if it is plausible in light of our background knowledge. Even if they agree that Jesus actually having been resurrected is the best explanation, many secular persons would respond to it just as we would
76
Chapter 3
respond to the (imagined) best explanation of crop circles: they would contend that there must be some better explanation we have yet to find; a crucified man being miraculously restored to life is just too implausible to accept. That would call for the existence of God, something they find too implausible to take seriously. For them, a reasonable view of the world leaves no room for such possibilities. At this point the abstract argument undermines and perhaps reverses this negative assessment of the plausibility of the best explanation. If the idea that there is a God is too implausible to take seriously then the miracle having actually occurred is too implausible an explanation to believe, even if it happens to be the only explanation we can think of that has not been refuted. However, there is some reason, even if it is not compelling on its own, to think that this world exists as the result of the free choice of a necessarily existent person—a Deity—who brought it into being for good reasons. This is the account that maximizes intelligibility consistent with the world’s contingency. Jesus, according to the testimony of those who knew him, conceived himself as being in some profound way the self-revelation of just such a God. They believed that he was miraculously resurrected by this God and their testimony convinced many others to believe it. If we had no more to go on than this alleged miracle, no matter how sincerely attested to and how widely believed in, we would not have sufficient reason to believe it actually happened. Even if this is the best explanation, it falls short of the threshold; it would be too implausible to believe. But with the very abstract considerations about contingency and intelligibility in view, it is no longer too implausible to believe the best explanation of these particular historical events. If Jesus was resurrected, then something occurred other than what would happen in the natural course of events, so there is a God. If Jesus was resurrected, then he was vindicated by God and we are entitled to identify the God who resurrected him as the God of Israel’s Scriptures and of the New Testament. The widespread idea that faith is belief which is for some reason laudable despite lacking rational justification makes Christianity irrational by definition. We need not endorse Richard Dawkins’ hyperbolic characterization of faith, construed as belief without sufficient evidence, as a form of mental illness, to see it as a departure from rationality to be avoided and as having no proper place in Christianity. 32 Faith, understood as trust in the God who resurrected Jesus, on this account is a matter of accepting the most reasonable explanation of the world at large and of human history. This is often the case with trust: I may not feel confident or find it easy to imagine the trusted person doing what I hope she will do, but I trust her because I have good reasons to do so. Faith and reason are often on one side, emotion and imagination on the other. Reasoning of the kind that appears in the abstract argument has relatively little power to capture the human imagination and compel assent. We are creatures for whom conviction most naturally arises not
Miracles
77
from abstract cogitation, but from causal connection to the world by way of the senses. The judgment that the best explanation of what we know happened in history is that Jesus actually was resurrected is one we can reasonably hope to justify on objective grounds, although doing this is difficult enough. The further judgment, that this explanation is not just best but also good enough, is categorically more difficult to justify, especially in our late modern culture, where belief in supernatural agents of some sort or other is no longer universally taken for granted. In this case, the issue of contingency and intelligibility is not all that is in play. An individual’s judgments of plausibility are subtly and intricately connected to many other things she believes, and beyond that to her values, hopes, personal history, and so on. This is not to retreat to the claim that these judgments are finally subjective, irrational, or a matter of choice. It is to acknowledge that the grounds for some of our most important beliefs can be complex, inchoate and largely implicit. These matters can, in principle, be treated objectively: our views on what is, or is not, plausible are proper matters of ongoing public, critical debate. In this regard they are of a piece with many of our most important beliefs, for example, various philosophical, ethical, political, and aesthetic judgments in which we regard ourselves as rationally justified, and our beliefs true, while acknowledging that other persons, viewing things from other perspectives, might reasonably and honestly disagree. In this book I develop an account of human beings as the material image of the transcendent, immaterial Creator who has characteristics that we at best only approximate, simulate, or possess analogously. God, in a sense, is the reality of which we are imperfect copies. The possibility cannot be summarily discounted that I have this wrong, and that such as we are, we are the reality and God a mere projection of our imaginations and hopes into the heavens. Faith is trust, and trust, leaving open the possibility of error, is consistent with confidence, but not with arrogant certitude. Belief in God is justifiable by means that subdue doubt, but they do not encourage us to dispense with humility. If belief in this God conflicts with what science tells us about ourselves and our world, that would be very good reason to reject the claim that Jesus was miraculously resurrected. The well-confirmed theories of science are, after all, much more securely established than our reasoning about why the world exists, the best explanations of remote historical events, or what explanations of them are good enough. But on no reasonable account is science all we know of the world, and on no reasonable account does science rule out knowledge of things beyond this world. Its reasonable constraint is that whatever else we claim to know must be consistent with what it tells us of the world and, in particular, what it tells us of the kind of creatures we are. It might, or might not, be reasonable to conclude that there is a deity, and it might, or might not, be reasonable to
78
Chapter 3
conclude that a particular event is divine intervention, but science claims no competence to evaluate such matters. The eminent American philosopher W. V. O. Quine once characterized naturalism as the commitment to “live within our means” epistemologically, acknowledging that we have no firmer ground for knowledge of the world than a natural science that generates empirically testable hypotheses. 33 This frugality does not demand that we deny that there is knowledge arrived at by other routes, but it recognizes that there are no ways of knowing that are more reliable than those of science. Yet we should recognize that the claim to know that this world, which we know best by way of science, is all there is, and thus to reject out of hand any inference from what is observed in this world to something beyond it is itself to live beyond one’s means. It is to place greater constraints on human knowledge than naturalism itself condones. An aspect of the intellectual humility appropriate to scientific naturalism is to acknowledge that there might be a great deal beyond our ken and that things might happen that are reasonably explained as miraculous. Christians are committed to a God who does miracles, making things happen that would not happen in the natural course of events. Among the events that most Christians regard as miraculous is the origin of humankind. Either overtly, as in the literally construed act of special creation described in Genesis, or subtly, by way of “guiding” the processes of nature, God brought the first humans into existence, causing something that would not have come about naturally. On this account, human origins lie beyond the reach of scientific explanation, and evolutionary biology’s claims to the contrary must be rejected. The task of the next chapter is to overturn this assumption and, by appeal to the revealed purposes for which God created, show that we should welcome the non-miraculous, scientific explanation of our origins. NOTES 1. John Macquarrie, Principles of Christian Theology, 2nd edition (New York: Charles Scribner’s Sons, 1977), 247. 2. The interpretation of the account, and whether we should believe the event actually happened, is not at issue here. The story’s symbolic theological significance might support the conclusion that it didn’t happen; on the other hand, one might suspect that Jesus actually did it for that reason, as an “enacted parable.” See N. T. Wright, Jesus and the Victory of God (Minneapolis: Fortress Press, 1996), 154–55. 3. David Hume, An Enquiry Concerning Human Understanding, Section X, Of Miracles, 2nd edition. Eric Steinberg, ed. (Indianapolis, IN: Hackett Publishing Company, 1993). 4. This example is, of course, contrived. Disconfirmations of well-confirmed theories are rare. The typical victims of disconfirmation are hypotheses that are not yet well-confirmed. The history of science is replete with causal or accidental observations launching scientifically controlled observation and experiment that led to the confirmation or disconfirmation of hypotheses, or more often simply new discoveries. How-
Miracles
79
ever, the causal observations are most often made by scientific researchers themselves, e.g., Alexander Fleming returned to his lab after a holiday, noticed that staphylococci in cultures contaminated with a fungus were destroyed, remarked, “That’s funny…” and proceeded with the investigation that led to the discovery of penicillin. See Frank Diggins, “The True History of the Discovery of Penicillin by Alexander Fleming,” Biomedical Science (March 2003), 246–49. Anyone might see something that makes it reasonable to subject a theory to testing that might result in disconfirmation, but chance favors the scientifically prepared mind. 5. This is how some of Jesus’ disciples are recorded as initially reasoning. Rather than concluding that God had resurrected Jesus, they thought that they were seeing a ghost. They had betrayed and deserted him, so this would not have been a welcome apparition. While as moderns we might see this as a supernatural explanation, they would have taken the post-mortem existence of an insubstantial residue of the dead, his shade or ghost, as a natural phenomenon. Jesus disabuses them of this conjecture by demonstrating his tangibility and ability to eat. Another natural explanation, episodically popular over the centuries despite its implausibility, is that they saw a Jesus who had not actually died. 6. Ilkkla Pyysiäinen’s description of the human mind’s innate propensity to believe in miracles, in Magic, Miracles, and Religion: A Scientist’s Perspective (Walnut Creek, CA: AltaMira Press, 2004), 81–89, is consistent with this normative constraint: unexplained events are regarded as miracles only if they can be satisfyingly explained as caused by the supernatural agents of one’s religion. 7. This applies to events anything like the miracles Christians actually believe in. There are conceivable events that would rationally compel anyone to believe that a miracle has occurred, e.g., the several hundred thousand stars in globular cluster M13 suddenly rearranged to spell out John 3:16 in koine Greek. 8. Arthur Peacocke, Theology for a Scientific Age: Being and Becoming—Natural, Human, and Divine (Minneapolis: Fortress Press, 1993), 139–40. 9. When direct divine action brings about some improbable event, it is true that it would not have happened naturally, although it could have. The interesting cases, which call for refinement of the simple counterfactual conception of miracles as what would not have happened naturally, involve probabilities that are not low. Suppose that some event has a .5 probability of occurring naturally, but God, wanting to ensure that it happens, acts directly to cause it. We are not entitled to say that it would not have occurred had God not acted. There might simply be no fact of the matter as to what would have happened had God not intervened. Suppose that some natural event has a .99 probability of occurring, but God, wanting to ensure that it happens, acts directly to cause it. Now are we entitled to say that it would have happened, even if God had not acted in this way, or just that it probably would have happened anyway? However we describe such cases, in any event God acts miraculously, intervening to ensure that a desired event that might not have happened naturally does in fact happen. 10. Some Christian thinkers have developed accounts on which God acts directly upon the world but does not intervene; see, e.g., Robert J. Russell, Cosmology from Alpha to Omega: The Creative Mutual Interaction of Theology and Science (Minneapolis: Fortress Press, 2008). This appears to rely on the assumption that interventions necessarily “violate” natural law. There is no obvious reason to make this assumption. If, as on standard interpretations of quantum mechanics, the laws of nature allow for causes that do not necessitate their effects, God could bring about an event that, so far as the laws go, could have happened, yet which would not have happened in the natural course of events. And to do that is, I believe, to intervene, no less than if God had done something physically impossible. Consider an analogy in positive law: President Coolidge could have intervened in the 1925 Scopes trial, and he could have done so either in a way that violated no laws, e.g., by filing an amicus brief, or illegally, by promising the judge an ambassadorship in exchange for a desired verdict.
80
Chapter 3
11. To avoid a kind of residual dualism, we have to keep in mind that what we have in view here are different levels of explanation of one reality, not different levels of reality. Ontologically speaking, the world is flat. Literally, levels of something are numerically distinct things (or parts of things), e.g., rock strata, layers of a cake, floors of a building. Talk of levels of reality easily blurs the distinction between thought (or language) and reality. There are multiple levels of description of the one reality but the structure of reality cannot be read off the structure of language. Complex things can be observed and described on a variety of scales, but to change the scale is not to look at a different thing. It is to look at the same thing in a different way, in more or les detail. The human brain, being extravagantly complex, can be described at a variety of scales, from that of its constituent elementary particles to its atoms, biochemicals, neurons, cortical columns, neural circuits and so on, up to its gross anatomy, and we can go on to apply our familiar folk-psychological idiom of belief, desire, intention, and so on. There are levels of description, explanation, organization, but not levels of being. So far as we know, there is just this very complex, multiply describable material creation and its transcendent Creator. 12. Configuring physical components this way is sufficient for the mental thing to exist, but, assuming that the mental is not reducible to the physical, it is not necessary for its existence. Mental things are multiply realizable. More on this in chapter 5. 13. Arthur Peacocke, Theology For A Scientific Age, 157. 14. Keith Ward, God, Chance and Necessity (Oxford, UK: Oneworld, 1996), 64–80. 15. Colin Gunton, The Triune Creator: A Historical and Systematic Study (Grand Rapids, MI: Eerdmans, 1998), 120. 16. How deep in the creation temporality resides, how close it comes to being fundamental in contrast to such features as color, consciousness, and even what we know as causality, which are superficial manifestations of the underlying physical reality, is an open question. For the view that time is a superficial feature of our universe see Sean Carroll, Something Deeply Hidden: Quantum Worlds and the Emergence of Spacetime (New York: Dutton, 2019) and for the opposing view see Lee Smolin, Time Reborn: The Crisis in Physics and the Future of the Universe (Boston: Houghton Mifflin Harcourt, 2013). In any event we should agree with Martin Heidegger, who suggests conceiving God as a “more primordial temporality” than that of this world, and with Ted Peters, who writes, “Rather than timelessness, I suggest that eternity be understood as everlastingness that takes up into itself the course of temporal history. See Heidegger Being and Time, John Macquarrie and Edward Robinson, trans. (New York: Harper and Row. 7th ed. 1962), 489 note xili, and Ted Peters, God−The World’s Future, (Minneapolis: Fortress Press, 1992), 109. 17. Quoted in Michael Ruse, Darwin and Design: Does Evolution Have A Purpose? (Cambridge, MA: Harvard University Press, 2003), 95. 18. Langdon Gilkey, “Cosmology, Ontology, and the Travail of Biblical Language,” Journal of Religion 41(1961), 194–205. 19. A paraphrase from the Book of Common Prayer, Holy Eucharist II, (New York: Seabury 1979). 20. Here I ignore the other the main contribution of the intelligent design movement, the alleged phenomenon of irreducible complexity in nature, i.e., complex structures which cannot function if a component is removed, and thus, the argument goes, could not have come about gradually by natural selection. There is at present no good reason to think there actually are examples of irreducible complexity, although, if there were, that would be a good reason to believe that an intelligent designer has been at work. On the other hand, instances of allegedly irreducibly complex structures explained in evolutionary terms undermine the argument. See the analysis in Sahotra Sarkar, Doubting Darwin? Creationist Designs on Evolution (Malden, MA: Blackwell Publishing, 2007), 93–116. 21. It is a decisive objection to forms of theistic evolution that blithely invoke biological evolution as the means God used to achieve specified aims, with no heed to the lack of fit of means to ends. Other species of theistic evolution avoid this by denying
Miracles
81
the explanatory adequacy of natural processes to produce the diversity, complexity, and adaptedness of biological life. They posit a sequence of unobtrusive divine interventions by means of which God guides evolution to its intended goals. The theistic evolution of The Material Image takes a different tack, severing design from creation. 22. St. Thomas Aquinas, Summa Theologiae, Volume 1: Prima Pars Q. 1–64. Dominican Brothers of the English Province, trans. (Scotts Valley, CA: 2008), I. Q. 1. A. 2. 23. If it were, then I could reasonably argue: Wacome believes that God exists, everything he believes is true; therefore, God exists. One can deny either premise without being an atheist and the second without being mistaken. 24. An older tradition disagrees, contending that the explanation that maximizes intelligibility is mind arranging things for the best. See Socrates in Plato’s Phaedo. David Gallop, trans. (Oxford, UK: Oxford University Press, 1979), 97b-99c. 25. If, as I will argue in the next chapter, a Creator who aims for the existence of distinct persons would employ natural processes which reliably achieve a general type of result yet leave specific outcomes to chance, the theistic view has the further virtue of explaining why, despite being fundamentally stochastic, the universe exhibits enduring regularities. 26. Robert J. Russell, “Cosmology, Creation, and Contingency” in Ted Peters, ed., Cosmos as Creation: Theology and Science in Consonance, (Nashville, TN: Abingdon Press, 1989), 177–209. While at face value the various scientific multiple world theories pose no problem for the Christian faith, modal realism, the philosophical theory that all possible worlds are equally real, removes the traditional distinction between the actual and the possible. This does conflict with Christian belief in creation. For its defense see David Lewis, On the Plurality of Worlds, (Oxford, UK: Basil Blackwell, 1986). 27. Late medieval theologians and philosophers realized that the laws of nature are contingent, the decrees of a radically free Creator. But how can anything be adequately explained without showing that it had to happen, and thus as not contingent? The solution was to conceive divine providence as grounding a surrogate necessity, what we now call natural or physical necessity in contrast to absolute (logical or metaphysical) necessity. The natural laws God freely authors can be known only by observation and experiment, not by a priori reasoning. Thus the requirement, inherited from Aristotle, that an explanation of natural phenomena must be a demonstration in which what happens is what has to happen in light of a law that has to be true, was abandoned and science in the modern sense finally became possible. 28. W. V. O. Quine, Word and Object (Cambridge, MA: MIT Press, 1960), 195–200. 29. The idea of a concrete individual which is a person yet exists necessarily conflicts with the popular supposition that matters of necessity are fundamentally straightforward, regular, bland, monolithic, predictable, uninteresting, etc., and that anything unique, individual or unexpected, anything with character, belongs to the realm of contingency. Yet this assumption does not survive an encounter with the domains of necessary truth we best understand, logic and mathematics. Here it soon becomes apparent that the realm of necessity is inhabited by things that have quite particular, unexpected, and often plain quirky natures. Consider the endlessly surprising discoveries in number theory, e.g., facts about the distribution of the primes, the weird space-filling geometrical objects generated with fractals, and the profoundly counterintuitive properties of cellular automata, the endless zoo of complex objects generated by a few simple rules from simple beginnings. These are areas where what is true is what must be true, yet the truth is often very far from anything we find simple or obvious. I don’t know how to put this sense of a connection between necessity and unique character clearly, let alone construct a natural theological argument from it, but without it I would find it harder to take seriously the idea of a necessarily existent free person. 30. See Peter Lipton, Inference to the Best Explanation, 2nd edition (London: Routledge, 2004), 59–62. 31. For a magisterial attempt to make this case see N. T. Wright, The Resurrection of the Son of God (Minneapolis: Fortress Press, 2003).
82
Chapter 3
32. As we will see in chapter 8, on religion, faith thus understood is normal, all too natural for human beings. 33. W. V. O. Quine, “Naturalism; Or, Living within One’s Means” Dialectica 49 (1995), 251–61.
FOUR Origins
BOTTLING THE UNIVERSAL ACID No aspect of scientific naturalism has been longer or more bitterly contested by Christians than the Darwinian explanation of human origins. Nowhere is resistance to the theory that the human species is the result of the “blind,” purposeless, mechanical processes of natural selection more intense than among fundamentalists and conservative Evangelicals, who not only vehemently reject it, but consider doing so a mark of authentic Christian commitment. Many other Christians, while not comfortable with an outright rejection of the scientific consensus, are uncomfortable with the implications of an unfettered Darwinism, and hope to blunt its impact on our conception of our origins and nature. Anti-Darwinism comes in varying degrees. Only the most conservative Christians reject evolutionary biology completely, but many fundamentalist and Evangelical Christians contend that all evolutionary change occurs within fixed species that God originally created. They accept microevolution but reject macroevolution, and with it the essential Darwinian explanation of the origin of species. Others grant the general competence of evolutionary explanations but make an exception for the human species, maintaining that although God did not supernaturally intervene to create other species, humans came into existence because of a miraculous act of special creation. Many other Christians acknowledge the competence of evolutionary biology to explain the diversity, complexity and adaptedness of organisms, including human bodies, but superimpose a dualist view of the human mind that keeps what is essentially human beyond the reach of scientific explanation. Others recognize the implausibility of mind-body dualism and agree that the human mind can to some extent be caught in the net of natural scientific explanation, 83
84
Chapter 4
but still hope to draw a line beyond which evolutionary explanation cannot proceed. In particular, proponents of this view reject the evolutionary explanations of human morality and religiosity we will examine in later chapters. Christians who adopt this position have many allies, for it attracts those who are wary of what Daniel Dennett calls the “universal acid” of Darwinism, the power of evolutionary explanation to dissolve the traditional human self-image. 1 In contrast to the various degrees of rejection, the thoroughgoing Darwinian maintains that the diversity and complexity of biological life, including human minds and, derivatively, the societies and cultures they create, can to a significant degree be understood as products of evolution. Human thought, feeling, and behavior can be explained as the functioning of a host of innate information processing mechanisms embodied in the brain’s neural circuitry. These mechanisms are adaptations to our ancestors’ natural and social environments. This perspective alone rejects the impulse to safeguard something uniquely human from the inroads of natural scientific explanation. It is the most starkly naturalistic view, the one with almost no Christian proponents. Whatever form it takes, Christian resistance to Darwinism depends on the assumption that it is antecedently improbable in light of the belief that God is the Creator of human beings, and that we are the image of the God who created us: if God is our Creator, this is not how we were made. This is not what we should expect the world to be like if it is God’s creation. This chapter challenges the judgment that Darwinism is improbable when viewed from the perspective of the Christian faith. On the contrary, contemporary science’s account of human origins has a high degree of prior probability in light of crucial Christian claims. Later chapters will extend this claim to the scientific project of explaining morality and religion in naturalistic, evolutionary terms. Surprisingly, a robust, thoroughgoing Darwinian point of view best fits with a Christian understanding of human origins. Christians have no need to try to bottle the universal acid of Darwinism. THE PLAUSIBILITY OF EVOLUTION Groundless claims about evolution being a theory in crisis notwithstanding, evidence in favor of natural selection being the principal explanation of the origin of species, including the human species, has steadily accrued for over one and a half centuries. Evolutionary theory, as a general proposition, is very well-confirmed and stands beyond reasonable doubt. Its defenders, impressed with the persistence of popular resistance to it, often proceed on the implicit assumption that it is a highly counterintuitive theory yet one that we are rationally compelled to accept by the weight of evidence. This strategy obscures the plausibility of evolution by natural
Origins
85
selection. 2 Given what we know about living things and their environments, evolutionary change is what we should expect. A simple story illustrates this. There’s a group of creatures, let’s call them murphs. About half are green, half white. They live in a place covered in green grass which makes the greens hard to see, so it is easier for predators to find the white murphs than the green ones. More white ones get eaten before they have a chance to have babies. When a green murph has babies, they are usually green too, though occasionally a green mother gives birth to a white baby. When a white murph has babies, they are usually white, though occasionally a white mother gives birth to a green baby. Once in a while, a white murph gets lucky and lives long enough to have babies, and sometimes predators get lucky and eat a green before it reproduces. But it is no surprise when, after many generations, there are proportionally fewer whites and more greens. Eventually, almost all of them are green. It is as though the environment, the green grass and the predators, have selected the greens for life and the whites for death. But no one intentionally selected anything; it was just nature: natural selection. Over many generations, these creatures have adapted to life in their green and grassy place. The murphs’ world does not stand still. At first, winter in the land of green grass was hardly noticeable, but the climate begins to change and most years the snow lasts longer than the year before. The greens are easy to see on the white snow but the white murphs, of which there are just a few, are now harder for predators to find. As generations go by, the greens’ chances of having babies are declining and the whites’ chances of reproducing improve. Eventually, snow covers the land all year, and almost all murphs are white. Just a few greens remain. The creatures have again adapted to their environment by natural selection. More time passes. The murphs have spread over a wide area. Things have gradually warmed up again; the snow is gone, and the land is green again. The melting snow causes the sea level to rise and a low-lying region floods, cutting off some of the murphs from the rest. Conditions on the island, and on the mainland, continue to change, though in slightly different ways. Both the islanders and the mainlanders adapt, so after a while they differ from one another. The predators on the island are sea birds that fly off with their prey. At first, just a lucky few of the murphs are too large to be suitable victims, but it is no surprise that, eventually, almost all the island murphs are too large for the birds to capture. They are now much larger than their mainland cousins. Things cool down again, the sea level drops, and the island becomes a peninsula. The two subspecies now mingle on the periphery of their territories. But now they barely recognize one another as possible mates. Island murphs tend to mate with similarly large island murphs, and the smaller mainland murphs tend to mate with smaller mainland murphs. The adventuresome few who ignore the difference in size find that mating attempts do
86
Chapter 4
not generally go well. More time goes by and the two groups continue to adapt in different ways, and interbreeding becomes impossible, not just unlikely. Temporary geographical isolation has resulted in permanent reproductive isolation. What started out as one species has become two. And so it goes. Biological organisms pass on many, but not all, their characteristics to their progeny. Some of those characteristics make it easier, and others harder, to reproduce. Those that make reproduction more likely tend to appear with greater frequency in later generations. Environments change, and inherited characteristics that facilitate reproduction in some affect it adversely in others. These simple facts do not, of course, prove that there is evolutionary change driven by natural selection, but they make it reasonable to suspect that there is and to seek confirming evidence. If it did not occur, we would need to explain why it does not. It could have turned out that the evidence was not forthcoming, that for any number of reasons this is not how things go in nature. Details of the actual mechanisms of reproduction might have made it impossible for natural selection to result in speciation, or even in significant change within a species. 3 However, the evidence is there in abundance, far more than can be easily summarized. Darwin died in 1882, leaving us with compelling evidence for “descent with modification” and the hypothesis that natural selection is its cause. That hypothesis was not generally accepted until the turn of the twentieth century, with the rediscovery of Mendelian genetics. As the century progressed fledgling evolutionary theory was augmented with the mathematics of population biology, making possible precise, empirically testable hypotheses. At mid-century, the identification of DNA as the bearer of hereditary information began the integration of evolutionary theory into the understanding of the causal mechanisms of biological life. In the 1960s, the shift in focus from the individual organism to the gene as the unit of selection expanded the scope of evolutionary explanation, making plausible its application to what is distinctively human. Today, many independent streams of scientific inquiry converge, making the Neo-Darwinian synthesis secure from reasonable doubt. A person of faith may accept all this, but hold out for an exception for the origin of the human species. He may judge that even though we cannot reasonably deny evolutionary theory as a general account of how species originate, the evidence for a natural origin of the human species is not yet so strong as to outweigh its antecedent improbability in light of the belief that human origins are miraculous. Christians reasonably believe that God can intervene in nature at will, bringing about events that would not naturally have occurred, so God could have created humankind in this way. But why should we believe this, and thus that there can be no adequate natural, scientific explanation? Conservative Christians
Origins
87
who call themselves creationists sometimes sound as though they do not really believe in creation. The question they often pose, “Did God create the human species, or did it come about as a result of evolution?” at face value presupposes, in contrast to the tradition of Christian theology, that if something comes about in the natural course of events, then God did not create it. This casts the natural world, with its law-governed processes, not as God’s creation, but as something alien into which God episodically intrudes. 4 But for the theist the legitimate question is not whether God created the human species, but how: miraculously or naturally? SECONDARY CAUSES? Traditionally, Christian theology has distinguished two ways God acts in the world. God brings about many things naturally, that is, indirectly, by way of secondary causes. God created the world in the beginning with its laws and initial conditions and in due course what God intends comes about. When something occurs in the natural course of events, this is something God does. Creation is continuing, not a one-time event completed in the past. However, God also acts directly upon the world, bringing about events that would not have happened in the natural course of events. As we saw in chapter 3, we have good reasons to expect that God acts directly, and thus miraculously, in “history,” that is, in interpersonal relation with created persons. However, there is no reason to suppose God miraculously intervenes in “nature,” that is, in the world prior to the appearance of created persons, or now in parts of the world remote from them. In this chapter, I claim that we have good reasons to suppose that the Creator would not act directly upon the creation until, in the natural course of events, persons came into existence. Christians have no reason to regard scientific theories that imply otherwise as improbable in light of their faith. On the contrary, given the purposes for which God creates, we can reasonably judge that an evolutionary explanation of the origin and nature of human beings is what we should expect. Christians confess faith in God as the maker of the heavens and the Earth, but it is longstanding Christian tradition that God is no less the Creator when divine ends are accomplished by secondary causes than when they result from intervention. Humans came into existence, and we rightly regard God as having created them. Given that God could have done this either directly or indirectly—neither is patently beyond the reach of divine power—why not be neutral as to how God actually did it until the scientific evidence decides the issue one way or another? Then, from this initial neutrality, why not go on to accept the strong consensus of the relevant scientific communities as decisive, and conclude that God brought human beings into existence by natural means? 5
88
Chapter 4
Despite the longstanding doctrine of divine action by means of secondary causes, there is a widespread tendency to denigrate the idea of God having created human beings by indirect, natural means. The idea that God created the human species by secondary causes that are described by evolutionary biology, rather than by means of miraculous intervention, is dismissed as tantamount to denying the theistic conception of creation. One writer contrasts evolutionary theory with “God having personally created the world and all its living creatures,” and asserts that Darwinian evolution implies that God is “disjoint from the world” and has no “active role in nature.” 6 The Christian philosopher Alvin Plantinga labels it, “a semideistic view of God and his workings.” 7 Law professor and antievolution crusader Phillip E. Johnson describes Christians who accept evolution as exchanging the Creator of the Bible for “the lifeless First Cause of deism.” 8 This is a judgment non-Christians share. John Dewey facetiously referred to creation by secondary causes as “design on the installment plan.” 9 Bertrand Russell, dismissing attempts to conceive God’s creative activity in naturalistic terms, writes, “Why the Creator should have preferred to reach His goal by a process, instead of going straight to it, these modern theologians do not say.” 10 In a similar vein, William Provine writes: Of course, it is still possible to believe in both modern evolutionary biology and a purposive force, even the Judeo-Christian God. One can suppose that God started the whole universe or works through the laws of nature (or both). There is no contradiction between this or similar views of God and natural selection. But this view of God is also worthless. Called Deism in the seventeenth and eighteenth centuries and considered equivalent to atheism then, it is no different now . . . religion is compatible with modern evolutionary biology (and indeed all of modern science) if the religion is effectively indistinguishable from atheism. 11
It is perhaps not surprising that secular critics of Christianity see the move from explaining things as due to divine intervention to explaining them as due to secondary causes as an unprincipled fallback position, an evasive stratagem to avoid the impact of science on traditional belief. But why should a Christian, fully convinced that there is a God who freely created the universe out of nothing and authored its causal laws, a God who works both miraculously and by way of secondary causes, agree? Why wouldn’t a Christian see what Dewey derides as “creation on the installment plan” as continuing creation, the Creator bringing new things into existence by way of the unfolding of the potentialities of the original creation?
Origins
89
DEFLATING THE DEISM CHARGE Let us dispose of the deism worry. The claim that it is deistic to think God created the human species by natural means is ill-conceived. Suppose we say that the planet Neptune came into existence, not because of divine intervention in the cosmos, but as a result of some long, complex but entirely natural process of planetary formation. No miracle was involved in getting from a universe without Neptune to one with it. Is this a “deistic” (or “semideistic”) view of the origins of Neptune? One can hold that God created Neptune by secondary causes and also say that God never intervenes in the creation, that Neptune’s creator is not the God of Christian faith but merely deism’s first cause. One can instead say that the God who created Neptune by secondary causes is the God who miraculously intervenes in the creation, indeed, that this is the God who became flesh as Jesus of Nazareth. Similarly, one can hold that God created the human species by secondary causes and also say that God never intervenes in the creation, that the human species’ creator is not the God of Christian faith but merely deism’s first cause. One can instead say that the God who created the human species by secondary causes is the God who miraculously intervenes in the creation, indeed, that this is the Creator who is incarnate as Jesus of Nazareth. The crucial contention between deism and Christian theism is not whether God has created some things by way of secondary causes; it is whether God ever acts directly upon the creation. In the context of judging the prior probability of God creating humans by evolutionary means charges of deism are spurious. We may disagree about how God chose to create humans, or anything else, but this is a question within theism, not between theism and deism. THE WRONG SECONDARY CAUSES? Those who find the evolutionary account of human origins theologically problematic might grant the possibility that humans came into existence by means of secondary causes, but contend that Darwinian evolution involves causes of the wrong kind. In the 1870’s the Princeton theologian Charles Hodge answered the question, “What is Darwinism?” succinctly: “It is atheism.” 12 Hodge did not object to Darwinism’s lack of accord with literalist readings of Scripture, nor to its claims about the antiquity of the Earth, the common descent of all living things, or even to human beings having come about by way of a natural evolutionary process. His rejection of Darwinism focused on its central idea: natural selection as the principal mechanism that accounts for life’s diversity and complexity in general, and specifically for the existence of the human species. God might have created species, including humans, by means of some sort of natural process, but for Charles Hodge, natural selection seemed obvi-
90
Chapter 4
ously not the kind of causal process the Creator would employ. Creating species by means of natural selection is, it seems, at odds with designing them, and critics like Hodge take for granted that God designs what God creates. DARWIN AND DESIGN The first humans came into existence as an effect of God’s initial act of creation; Darwinism does not challenge this. What it undermines is the idea that God designed the human species. Someone designs something if she intentionally causes it to exist and its specific characteristics are what they are because they are what she specifically intended. Darwinian evolutionary processes may not look like a reasonable means to realize specific creative intentions. We may with Richard Dawkins ask, “If natural selection and evolution is God’s way of designing life, why would he choose the one way which makes it look as though he doesn’t exist, which makes his own role completely superfluous?” 13 There is, first, the matter of sheer waste: Darwin’s Malthusian realization that evolution proceeds because the vast majority of individual organisms die before reproducing. Not only the premature demise of individuals, but the extinction of whole species, is an essential aspect of the evolutionary process. The “struggle for existence” was not simply something God allowed, but the intended means to the divine ends, if those ends are achieved by way of natural selection. Further—and this is an aspect that Darwin emphasized in his own objections to attempts to incorporate evolution by natural selection into a theistic worldview—there are the many structures whose current function is not that for which they were originally adapted. 14 Rather than exhibiting structures that look as though they appeared on schedule in accord with a rational plan, organisms look like Rube Goldberg contrivances, making use of whatever components happened to be available. Rather than the single most efficient mechanism for performing a function, the evolutionary process produces a variety of strategies, none optimal, most good enough to get by, but only until the competition intensifies. Close examination of form and function in nature presents not the appearance of rational engineering, manifesting wise foresight enacted by secondary causes, resulting in a perfect fit of organisms to environments, but something more like ad hoc tinkering. Preadaptations and exaptations, vestigial organs useful not to the creatures that bear them but to their ancestors, structures that appear in the embryo only to disappear later, homologous structures recycled for radically different functions, life taking profoundly new directions not in virtue of the gradual unfolding of inbuilt purpose but because of natural catastrophes inducing mass extinctions, species marooned at local optima, unable to change in ways that would ultimately be advantageous because
Origins
91
natural selection, oblivious to the future, cannot sacrifice immediate goods for greater future goods: all this tells a tale of a long, messy, chancy history, not a progressive march toward planned perfection, not a rational implementation of the creative intentions of an omniscient, omnipotent God. We would admire an engineer who designs a factory that, once built and set in motion, produces cars automatically, requiring no after the fact intervention. We might well regard its designer as superior to one who can devise only a factory that requires its maker periodically to intervene to get it to do what it cannot do on its own. 15 We would, however, have less admiration for that engineer’s skills when we discover that most of the factory’s output goes straight to the junk yard, that the operative vehicles rolling off the line are best-suited to road conditions that have not existed for 100,000 years, or that they had hubcaps for steering wheels, wiring for headlights crossing the windshield, an air conditioner made out of a modified manual transmission, or an anachronistic crank uselessly affixed to a car with an automatic starter. DETERMINISM TO THE RESCUE? None of this proves that the Darwinian evolutionary mechanisms are not God’s handiwork, nor does it prove that God did not design each species, or even each individual. If the basic laws of nature are deterministic, then no matter what evolutionary mechanisms science discovers, their results will be precisely what God intended from the first moment of creation. If God intends the initial state of the universe and intends its laws, and these laws guarantee specific effects, then God specifically intends everything that happens in creation, no matter how long after the beginning, no matter how complex the chain of causes and no matter how unpredictable—to us, but not to the Creator—the final effect. Rising sea levels reproductively isolate one group of creatures from another and speciation occurs. An asteroid formed early in the solar system eons later strikes the Earth and ends the ascendancy of the dinosaurs. A cosmic ray causes a particular mutation on a particular chromosome in a lemur three million years ago: all the necessary conditions for the eventual appearance of humans could have been in God's unfolding plan. The divine intention from the start could have called for disparate chains of causation to converge on preordained, inevitable effects. This way of doing things staggers the human imagination, but lies within the scope of the Creator’s omniscience and limitless power. God could have designed the human species in a way that hides divine design. The theistic evolutionist can insist that talk of chance and accident in Darwinism merely reflects human ignorance of the true causes of things, not inherent indeterminacy in the processes God used to bring humans into existence.
92
Chapter 4
Yet there can be a great distance between what we know is possible and what we can find plausible. It is no surprise that in the aftermath of Darwin so many rejected evolution by natural selection as utterly improbable from a Christian perspective, seeing it as for all practical purposes a denial of divine design. Why would God contrive a process for achieving a specific goal that looks as though its results are unplanned? Anyone convinced that God specifically intended the existence of humans may reasonably judge evolution by means of natural selection to have very low prior probability, and at least temporarily reasonably refuse to accept it, even in the presence of evidence that might otherwise be compelling. God, having the human species in mind, might have brought us into being by means of secondary causes; if so, we should expect to find evidence of natural processes elegantly, infallibly moving toward that goal. However, the evidence Darwin marshaled on behalf of natural selection does not point in this direction, nor does the vast amount of evidence that has accumulated in the ensuing century and a half. The philosopher Michael Levin succinctly summarizes where evolution by way of natural selection leaves us: The theory of evolution showed the appearance of design in nature to be misleading. That most features of living organisms display the goaldirectedness and efficiency of artifacts led mankind for many ages to think the natural world was created by a planner. As Darwin explained, however, design can be mimicked by randomly generated variations competing to reproduce under environmental constraints. The world may be an artifact in some ultimate theological sense, but so far as natural processes go, it merely simulates one. 16
INDETERMINISM Nonetheless, so long as the world in which natural selection operates is taken to be deterministic, it remains possible, no matter how implausible, to see it as the method a divine designer would choose. Abandon determinism, allow an element of real chance into nature, and the possibility vanishes. No rational Creator deploys indeterministic means to achieve specific ends. Indeterminism is the legacy of early twentieth-century physics. On standard interpretations, the fundamental physical laws of nature are indeterministic. 17 When an event of a particular type occurs, the basic laws of nature do not guarantee one effect. The laws fix the probabilities of the occurrence of various possible effects. Which effect occurs on any occasion is left to chance. Prior to quantum mechanics it was possible, even if difficult, to contend that God used natural processes, even those described by Darwinism, to realize specific creative intentions. No matter how weirdly messy and contingent the process appears,
Origins
93
it would produce precisely the results God intended from the beginning. After quantum mechanics, there seem to be no natural means for God to realize specific creative intentions, no way to reconcile divine design and any natural, scientific explanation of human origins. If nature’s fundamental laws merely assign probabilities to possible effects, then longterm natural processes lack precisely foreseeable outcomes, and nothing short of divine intervention can guarantee that things in nature proceed to the specific outcomes God intends. Therefore, those who see evolutionary theory as incompatible with God having created the human species are not necessarily being unreasonable. They are right to reject facile notions of “theistic evolution.” If God’s creation of the human species implies that we were designed, specifically intended, then Darwinian evolution in a world governed by indeterministic causal laws, is not a rational method. A process of that sort would not reliably lead to a specifically intended outcome. If the human species exists because God specifically intended it to exist—if God designed it—then the reasonable conclusion is that God miraculously intervened along the way to ensure the specifically desired outcome. On this view, God must have “guided” the evolutionary process to bring about specifically intended goals. Such divine guidance can only be miraculous intervention, God acting upon the world to bring about events that in the natural course of events would not have occurred. If divine intervention is required to explain the existence of the human species, then evolutionary biology cannot provide an adequate explanation of human origins, for the origin of this species, at least, is miraculous. The conviction that God specifically intended the human species lies deep in the Christian tradition. To abandon design requires that we revise what it means to acknowledge God as our Creator. We who believe in creation should abandon the tradition’s adherence to design only for good reasons. There are excellent scientific reasons to do so. Natural selection as the main motor of speciation and adaptation is a well-confirmed scientific theory, and the standard interpretation of quantum mechanics is indeterministic. However, complementary reasons lie within Christian theology itself, specifically in its account of the purposes for which God created us. When we give these considerations their due, it becomes clear that the traditional judgment about the antecedent improbability of evolutionary theory is dubious. Science portrays the human species as not having been designed. The Christian faith, confessing that God created us for the purposes revealed in Scripture and in Jesus Christ, concurs with the scientific view that we are not the product of design.
94
Chapter 4
THE GOD WHO CREATES To challenge the assumption that God specifically intended human beings to exist, and thereby reverse the familiar judgment that Christians must regard evolutionary biology’s explanation of our origins as improbable, requires focus on some large issues: what is the nature of God, and why did God create the world? The creation is contingent. It did not have to come into existence, and it did not have to contain human beings or any other persons. It exists only because God brought it into existence ex nihilo, out of nothing. God did not have to create anything. God would have been no less good had nothing been created. The world exists as a result of God's free choice, but that choice is not capricious or arbitrary. God acts freely, but with rational purpose: God created this universe for good reasons. Knowing why God creates constrains answers to questions about how God creates. The reasons for which God creates are grounded in God’s Triune nature, revealed in Jesus Christ. Christian theology is officially controlled by the biblical witness, but theological reflection on the nature of God often relies heavily on abstract philosophical considerations. This is epitomized in the perfect being theology that, in the tradition of Anselm, and traceable back to Christian NeoPlatonism, defines God as the greatest conceivable being and sets to work drawing out the implications of this concept. As valuable, and perhaps indispensable, as this strategy is in some contexts, it generates unease when we contrast God as portrayed in the Bible with the God of philosophical theology. We should be wary of moving too far from the God of Abraham, Isaac, and Jacob, who chooses to be revealed not in abstract metaphysics, but by way of a complicated, troubled, often tragic, sometimes joyous, narrative of a fraught relationship with human beings. The God we meet in Scripture is the Yahweh of Israel, not ultimately comprehensible and often unpredictable, but revealing a trustworthy character in the concrete history of a particular people. This is the God whose most salient attributes are faithfulness and loving kindness, a Deity whose nature comes into view in deepening engagement with human beings. The Scriptures speak of the righteousness of God, but this is not the abstract justice of the Greek philosophers; it is God’s faithful commitment to the covenant made between God and created persons, ultimately between God and all that is made. God is utterly good, but this goodness is God’s pursuit of what is good for beloved creatures. God possesses utmost wisdom and power, but this manifests itself in providential care for created persons and the world they inhabit, and in a commitment to do whatever it takes to save the needy creatures for shalom, blessed life in community with the Creator. In the world of the Bible, God’s defining characteristics have their meaning in relationship with created persons. The paradigm of human fellowship with God is not private religious
Origins
95
experience, but individuals’ participation in the life of the community of faith, sharing God’s work in this world and embodying the divine love for human beings. The idea that the one God is not a solitary being, but distinct persons in loving communion, is unique and essential to the Christian faith. As Peter van Inwagen points out, “Jews and Muslims believe that power and goodness and wisdom and glory are from everlasting to everlasting. But only Christians believe this of love, for the eternality of love is a fruit of the uniquely Christian doctrine of the Holy Trinity.” 18 The first Johannine epistle makes an exacting assertion about the nature of God: “God is love” (1 John 4:9). Trinitarianism is not an arcane and dispensable theological dogma; it is vital to Christian belief and practice. It underlies the central fact about God’s attitude toward us. The tri-personal God comes to us with a love that is as unconditional as it is demanding, rescuing us from self-inflicted, self-destructive alienation from ourselves, one another, the creation, and from our Creator, inviting us to share the everlasting life of the Triune God. The trinitarian confession decisively shapes what Christians can say about the creation. The God who creates is transcendent, ontologically distinct from the creation, but not a monolithic will to power over it. Instead, the Creator wills to exist in personal relationship with creatures. We must take care to avoid accepting views simply because they seem plausible from the perspective of a philosophical theology, or of our natural religious consciousness, and keep in view the God specific to Christian confession: the God who graciously opens the inner life of the Trinity to the creature, reaching out to those who are other than God. The New Testament proclaims that this God reaches out to human creatures in a concrete way, in the Incarnation. Christianity proclaims that creation exists by and for Jesus Christ. In him God’s ultimate purposes are revealed. The meaning of creation, the answer to the question why the world exists, lies in Jesus Christ, Immanuel, “God with us.” Jesus is God incarnate, God committed, for better or worse, as a creature among creatures. Jesus is God disclosed in human terms, as the Son of the Father. The world was brought to be from nothing to be the world in which God takes creaturely flesh and dwells with those God has made. The Christian tradition focuses on the Incarnation as the means devised by God to restore a fallen creation, to deliver us from sin and death. It is too often assumed that this was merely a divine response to the human condition, that God would not have become human if things had not gone wrong. I assume, on the contrary, that the world was always intended to be the setting in which God makes a life with created persons, and that the Incarnation would have occurred, though in a radically different context, even if there had been no Fall. 19 God created the cosmos as a home for personal creatures, always intending to be physically, empirically present to them. We exist because
96
Chapter 4
God wants us to exist and the universe exists for our sake. This audacious assertion resides at the heart of the Christian faith. 20 God’s aim in creation is the existence of persons other than God, persons who have an independent reality, a degree of autonomy, insofar as they may, but need not, accept the invitation to enter into that relationship and share in the everlasting life of the Trinity. God intended there to be persons distinct from God, capable of actions that are truly their actions, not God’s, so they can respond to God’s actions, and God to theirs, and so on. God does not encounter their actions merely as the remote but predictable consequences of the act of creating the universe. They may, but are not compelled to, see things the way God sees them. They may, but need not, want what God wants. They may, but are not forced to, do what God wants them to do. 21 The fact that this is why the world exists profoundly conditions our expectations about what sort of world it is. Any judgment about the prior probability of scientific hypotheses about human origins should be approached with this central confession of Christian theology in mind. DESIGNED PERSONS? Given that the Creator made this world for the purpose of mutual love and fellowship with creatures, how might we expect God to have gone about making them? Should we expect, as so many do, to find that God designed us? On reflection we should not find this at all likely. Our being designed would be at odds with the purpose for which God created us. Whatever our expectations, they might, conceivably, be overturned by evidence that God did things another way, but our estimation of antecedent plausibility properly influences how we weigh the evidence for competing hypotheses about human origins. God, a communion of loving persons, creates with the intention of existing in loving relation with persons other than their Creator. A principal consideration in wondering about how the Creator is most likely to have pursued this aim is the requirement that the created persons with whom God interacts have lives of their own, that they be independent persons, not mere extensions of their creator. As Langdon Gilkey once said, humans are an image, not a shadow, of God. 22 Our thought concerning creation needs to pay heed to what Colin Gunton refers to as the “proper substantiality of the creature.” 23 God does not annihilate or diminish the integrity and individuality of created things, but seeks to empower each created thing to become more fully itself. Creation’s purpose is that there be persons, not unrelated to or independent of God, but distinct from the one who created them and who calls them into relationship. Created persons lack God’s aseity and depend upon God for their exis-
Origins
97
tence, yet they are real, independent beings, for not even God can have a genuine relationship of mutuality with one who is not other than God. 24 A created thing is radically dependent upon its Creator. It is causally dependent upon God: it exists only because God acted to create it. God in some sense sustains it in existence; no creature exists at any moment except at God’s pleasure. 25 Yet God creates things that are other than God. A created thing cannot reasonably be thought of on analogy to an idea in the mind of God, for an idea is an occurrence in someone’s mind, not something distinct from the person whose idea it is. The relation between God and creatures cannot be compared to that between the author of a novel and its characters. The author can write a conversation between a character and author into the book, and in doing so she might feel that she interacts with someone other than herself, but in reality she doesn’t. She might find that her characters metaphorically take on lives of their own, but they remain fictional, not real, persons. Her specific intentions, even if known only in retrospect, are realized in what her character says no less than in what she says herself. In contrast, God creates, and communes with, persons who are literally other than God. It is not a paradox, but common human experience, that being related to another person in mutuality is possible only for those who have clear boundaries that delineate the integrity of a distinct self. Individuals who cannot properly distinguish themselves from others fall into pathological relations of co-dependency. Martha Nussbaum writes, “love requires not only the recognition of its object’s separateness, but also the wish that this separateness be protected.” 26 In one sense, the requirement that the things God creates be distinct, ontologically independent, entities is trivially satisfied. If God actually creates a thing, as opposed to merely contemplating its possibility, then something other than God comes into existence. However, the kind of independent existence that God’s purposes require for created persons is not guaranteed simply by their being creatures. In light of God’s omnipotence and omniscience, the question is whether there can be creatures that can relate to their Creator as distinct persons. When humans create, we always face the possibility of a gap between what we intend and what we actually cause to exist. We lack sufficient control over the processes of nature we employ to ensure that fully specified goals are realized in every detail. I build a gate for the cat enclosure, and I am satisfied so long as what I wind up with is close enough to what I intended for it to work. The product of my labors has features I did not intend, including some I intended it not to have, but if all goes well the thing works well enough and my intention is realized. The situation for God is dramatically different. When God creates, whether immediately or by way of secondary causes, nothing need prevent the result from being precisely what is intended. This leads to a crucial problem. Can a personal creature be truly distinct from her Crea-
98
Chapter 4
tor if her every characteristic is what it is because it is precisely what her Creator intended? Could fully designed persons have the independence that God’s aims require? Many Christians, especially those most ready to describe their faith in terms of a personal relationship with God, and wary of anything that makes God seem remote, not personally involved in the world, are adamant about finding evidence of design in creation, particularly evidence that human beings were designed. This is ironic, for our being designed is incompatible with our having a personal relationship with God. Given what we know of God’s aims in creating, what we should have expected to find in the creation is evidence of God having taken pains to create us without designing us. We are in general entitled to regard our practical knowledge of human relationships as at least as good a guide to God's nature and character as the philosophical considerations on which theology has so often relied. 27 One of our best images of the creative Trinity is that of a human couple, deeply in love and satisfied in one another, who choose to have a child, not out of unfulfilled need, but simply because it is the nature of their love that they desire to share it. They want to bring another into their shared life, not because their love is incomplete without him, but because they want to share its completeness. Their creative act is free in whatever way human actions are free, not a matter of necessity or obligation; yet they have excellent reasons for it. When this couple chooses to conceive a child they reasonably want some general knowledge of what they are bringing into existence. If they were going to conceive a hamster or an armadillo instead of a human infant they would want to know this before proceeding. This is about all many prospective parents would want to know. They are happy to start the process and let nature take its course to an outcome of which they possess only this general foreknowledge and over which they possess just this minimal control. The intention they act on is general, not specific. Most prospective parents will do whatever they can to make sure that the child will be normal and healthy, but go no further. Some would, if given the option, decide what sex their baby will be. And some might like to select many more, perhaps even all, of their child's characteristics. For them, having a baby would be like building a new house or buying a new car. As far as possible they want the child’s specific features to depend on their specific choices. They want to leave nothing to chance. They do not want to be merely causally responsible for their child's existence. They want to design it, not merely to set in motion a somewhat chancy natural process that leads to one of many possible children coming into actual existence. They want to select it, with all its characteristics, from the possible babies they could produce. We rightly find this desire inappropriate, in conflict with the intention to bring into existence someone to relate to as a person distinct from oneself. To love is to risk, to take one’s chances on how the beloved will
Origins
99
respond. We would not regard such parents as better or more caring or loving in virtue of making their child's characteristics depend as much as possible on their decisions. We would suspect their choices as springing from vanity, from the desire to have a child whose characteristics and achievements reflect well on them. Discussing the moral issue of genetic engineering intended to enhance one’s progeny, Michael J. Sandel writes that it “disfigures the relation between parent and child” and that it is at odds with “the norm of unconditional love,” revealing a lack of “openness to the unbidden.” 28 Parents ought not to design their children. Making a baby is not like making a house, for to make a baby is to make a person distinct from oneself. When building a house it is reasonable to try to ensure that it has precisely the characteristics one wants it to have, but it is unreasonable to try to relate to another person as something one has designed. It undermines the “psychological distance” that authentic interpersonal relations require. I am not capable of interacting with my parents on a personal basis if my characteristic ways of thinking, feeling, and acting are the guaranteed result of their intentions. One ought not to be able to regard another person as an extension or expression of oneself in the way things one has designed are extensions or expressions of oneself, reflecting one's desires and choices to whatever extent allowed by one's power over the natural material. There is a science fiction scenario in which a man, frustrated with the independence of human females, designs a perfect mate, choosing not just her physical appearance, but her values, preferences, behavioral dispositions, in sum, deciding what her character and personality will be. The moral of such tales is that even if he is delighted with what he has made and lives with her happily ever after he is sadly deluded, imagining that he enjoys a wonderful relationship with another, when in reality he is relating to a projection of his own fantasies, not crossing the boundary of himself to connect to a genuine other, with all the uncertainties and difficulties this involves. When she acts, and he responds to her, he is not responding to something other than himself, but to what he himself has specified. There is no real reciprocity; too much of what transpires between them, in too much detail, flows from his desires and intentions. He does not encounter the unpredictable, not fully knowable, mysterious other in her. Nor could she, made aware of her origin, see herself as an independent person, capable of freely entering into, or not entering into, a relationship with this man. He might have designed her to find him the ideal man, exactly the sort of person she most desires, but insofar as she knows this, she will be alienated from those desires and preferences, able neither to escape them nor to embrace them as truly her own. God could have created in such a way that what a creature wants she wants simply because that is what God wanted her to want, that she thinks the way she thinks because that is how God intended her to think, and ultimately that she does what she does because that is what God
100
Chapter 4
chose for her to do, but this would fatally undermine her capacity to see herself as God’s companion in the give and take of a genuine interpersonal relationship. She would be God’s creature, but not God’s friend. 29 There are good reasons for human makers of human persons not to design them. There are cogent ethical objections to positive genetic engineering. Human beings plainly lack the goodness, knowledge, intelligence, and imagination to assume this sort of responsibility without bad consequences, but none of these considerations apply to God. God would do nothing wrong by designing persons. Nonetheless, it would be incongruous for God to have brought us into existence in this way, for it would be at odds with the purpose for which we are created. We would not have the independence of our Creator required by the relationship for which we were created. God is wise: God’s means are suited to the divinely chosen ends. Therefore, God created us by means that did not involve design. For centuries, no one had any idea of how God could have done this. Miraculous intervention seemed necessary. Scientific inquiry has now supplied us with a reasonable account of what God’s means were: evolution by means of natural selection in a world governed by indeterministic laws. FAREWELL TO DESIGN God created the human species by some means. All possible methods were available, but in light of the revealed purposes behind the creation they are not equally likely. God could have used deterministic secondary causes to create us, employing natural processes guaranteed to produce the intended result with no need for intervention. Alternatively, whether the creation’s laws are deterministic or indeterministic, they might not suffice to bring into being creatures of the kind God wants, so intervention is required: God acts directly to bring into existence things that would not have come about in the natural course of events. This miraculous intervention could have been subtle, imperceptible divine guidance of natural processes, gradually accomplishing God’s specific purposes in nature. Or it could have been overt and immediate, as those who read primeval Genesis literally envision the creation of Adam. Both methods share the implication that God designed the first humans: God did not simply act in a way that in due time led to their existence, but decided what their characteristics were going to be and acted to ensure that specified result. If we suppose, with most Christians through the centuries, that God must have specifically intended human beings, these are the only possibilities. They conflict with what science tells us, which is that the human species came about as a result of chancy natural processes.
Origins
101
The neglected possibility is that God chose to create our species by means of an indeterministic natural process, and did not direct it toward a specific result by intervening along the way. Rather than designing us, God left to chance what specific kinds of personal creatures eventually came into existence. This coheres with the scientific evidence for Darwinism, evidence that so many Christians, convinced that the Creator must have intended, foreknown, and thus designed humanity, feel must be misleading. Which possibility best coheres with God's revealed purposes in creating? Which would the God whose purpose is to create persons for communion with their Creator choose: precise foreknowledge or a degree of unpredictability, complete control or specifics left to chance? So far as we have any reason to suppose, God could select the third option. God could set in motion a process that has a future outcome knowable within general parameters, yet not knowable in detail ahead of time. God could make a world whose fundamental laws fix the probabilities of future events but do not determine them. On standard interpretations of basic physical theory, God has done just this in devising this world which, while a good simulation of a deterministic creation, is at bottom quantum mechanical and thus, apparently, indeterministic. There are those who feel sure God could not really have acted in this way, but the burden of proof falls to anyone who claims it lies beyond God’s capabilities. The objection that God could not do it because God must completely control the future subjugates God to our conception of the divine attributes, as though this is simply God’s job, irrespective of what means best fit God’s freely chosen ends. 30 God is sovereign in all things, including creation but, as John Sanders says, “it is not up to human beings to dictate what sort of sovereignty God has freely chosen to practice.” 31 The same is true of how we should understand God’s omniscience. The future of a creation governed by indeterministic causal laws is knowable in general ways, but highly specified knowledge of its future is impossible. The proposal that God pursues the aim of there being persons distinct from their Creator by creating an indeterministic world precludes exhaustive divine foreknowledge. This does not impugn the idea of divine omniscience. God knows whatever it is possible to know, but God can do whatever it is possible to do, and one possible divine action is creating a world with a future that it is not possible exhaustively to foreknow. Someone might argue that God could not create such a world, but he would owe an account of why this lies beyond the power and wisdom of an omnipotent God. He might instead argue that God would not make such a world, that God would not constrain foreknowledge in this way, but this, I contend, is exactly what a Creator with the goal of making persons capable of fellowship with their Creator would do. It is at the very least plausible that the God we know in Christian faith could, and would, choose to create undesigned persons. Divine design
102
Chapter 4
appears at the roots of creation. God, intending to create persons, created the world, selecting basic laws and initial conditions to guarantee this general outcome. Setting in motion an intrinsically indeterministic process, the Creator left the specific outcomes—just what sort of personal creatures were going to exist—to chance. Which is to say that in making us, God acted with a concern for us we see reflected in loving human parents. The infinite qualitative difference between God and creatures notwithstanding, the analogy is evident: in both cases a person desires there to be another person, distinct from her maker, who will live her own life and be able freely to enter into relationship with the one who brought her into being. In both cases the purpose is relationship with another who is separate yet not distant. We may wonder whether indeterminism really makes this difference, rendering us not specifically intended. The quantum domain’s indeterminacy does not typically “bubble up” to the observable world. When vast numbers of particles are involved, low-level events cancel one another out, and we find ourselves in a good simulation of a deterministic world. Normally, for particular undetermined events to have an observable effect requires technological interventions, such as Geiger counters. However, at critical junctures, quantum indeterminacy matters. In the very early universe, which galaxies, stars, and eventually planets to which life would adapt, were to come into existence depended on the precise configuration of elementary particles. Once biological life came about, which mutations occur depends, to a degree, on what goes on in the quantum mechanical microworld. When a human couple sets out to produce a child, which of many possible human individuals will be conceived depends on which of millions of spermatozoa reach the ovum first, and this is influenced by the buffeting of Brownian motion, and thus, on standard interpretations of quantum mechanics, due to genuine chance events. Shortly before conception, who will be conceived is not yet decided. Further, the genome that comes into existence at conception contains too little information fully to specify the fine structure of the brain. At a fine level of detail, the brain circuitry on which our being who we are to a significant degree depends is also a matter of chance. 32 Someone might contend that these processes really are causally determined, and that which possible human person actually comes into existence is not left to chance. Or one might insist that God miraculously intervenes in these otherwise indeterministic processes, bringing about specifically intended results. If our theology predisposes us to think God must have had precise intentions as to who eventually inhabits creation, we would have reason to discount the evidence of randomness in the evolutionary process, sexual reproduction, and brain development, and seek evidence of design. If, instead, we think God would not have designed personal creatures, we can take all this apparent indeterminacy at face value, as just the sort of thing we should have expected, given God’s ultimate ends.
Origins
103
God acts so as to leave much about us as individuals up to no one, letting at least the characteristics with which we begin our lives be “decided” by impersonal natural processes, and thus making it possible for us to assume a measure of responsibility for ourselves and to relate to our Creator as independent persons. It is equally plausible that God let which species of personal creatures came into existence be “selected” in the natural way Darwinism describes. Judgments about prior probabilities can be difficult to substantiate objectively. I do not pretend to prove that God created us by means that did not involve design, but I believe I have shown that God’s choosing to create us in this manner accords with essential Christian claims about God's revealed character and purposes, and it is what science invites us to believe. Insofar as a Christian theology of creation grants these claims the central place they deserve, it assigns a high prior probability to evolutionary biology’s naturalistic claims about our origins. Those of us who believe that God created this world with a view to interacting with created persons should judge the theory of evolution as antecedently probable, as the way we might well have expected things to be in light of our theological convictions. Rather than demanding more and better evidence for Darwinism, Christians can reasonably accept it on the basis of less, and less good, evidence than others require! An evocative idea that originates in the mystical Jewish tradition of Kabbalah is that when God, infinite and everything, chose to create, this was an act of self-limitation. God contracted, withdrawing to make space for the creation, for what is other than God. God does not, of course, occupy and then vacate space to make room for creatures, so it is not obvious what literal sense can be made of the idea that God is self-limiting in creating. But the idea that God creates without designing, foregoing complete control and foreknowledge for the sake of there being persons truly other than their God, creatures who can act upon their Creator, literally links divine self-limitation to creation. 33 PURPOSE FOR NATURE Ever since Darwin, many Christians (and others) have hoped to sweeten what seemed the bitter pill of evolution with the idea that it is progressive, that it naturally aims at its results, but this is at odds with natural selection being the principal engine of biological complexity, diversity, and adaptation. 34 A hasty theistic evolution points to a possibility—”this is how God could have done it”—but fails to come to grips with the incongruity of God choosing to achieve specific ends by means of the mechanisms Darwin discovered, introducing a teleology that has no place in Darwinism. The account of human origins proposed here avoids this. It does not suppose that the laws of nature are contrived to achieve
104
Chapter 4
fully specified goals. On the contrary, the proposal is that they are contrived so as not to achieve fully predictable results, and that this, not design, coheres with God’s creative purposes. The “watchmaker,” as Richard Dawkins famously described natural selection, may well be blind, but its maker has good reasons for making him so. If this is correct, then Christians assiduously seeking evidence of design in nature are like government inspectors, suspiciously examining the roulette wheels in a casino, convinced that they must be rigged to turn up certain numbers. If the casino builder is honest, their assumption is mistaken, for guaranteed specific outcomes is precisely what he would not have arranged. The view on offer here rejects the notion of specific goals built into the natural processes that led to personal creatures, but it requires that it be a practical certainty that at some time, at some place, the creation gives rise to persons of some kind. God’s purpose for nature precludes there being a specific purpose in nature, but there is a general purpose for nature: the giving of life to created persons. 35 The Creator, bringing into being a universe for the sake of fellowship with personal creatures, can, and would, choose to leave the specifics to chance, but the basic laws and initial conditions of the creation must make the eventual existence of persons of some sort or other a certainty. There must be a secure connection between God’s general intention that there be persons and what actually happens as a result of evolutionary history. How reasonable is it to think that natural selection, despite its ultimate lack of direction, could be relied upon to produce persons? Defining a person as one who is rational, conscious, and thus self-conscious, we need only regard the arrival of rationality and consciousness somewhere in the universe as analogous to the arrival of vision, or flight, or photosynthesis, characteristics that have obvious adaptive value, and thus likely candidates for convergent evolution. Rationality, while obviously not the only evolutionary route to success—bacteria and beetles have flourished without it—does have profound adaptive value. A creature capable of reliably forming accurate internal representations of the world, and of transforming them in ways that preserve accuracy, and then regulating its behavior by them, would in many situations be at a significant competitive advantage. 36 Potential prey should prefer not to be tracked by a predator whose behavior is guided by expectations current perceptions generate about what it does not perceive, for instance, one that forms beliefs about the whereabouts of its unseen prey by inspecting its spoor. 37 Members of a social species will do better in competition for reproductive opportunities if they can internally model, and in so doing anticipate, the behavior of their conspecifics. Homo sapiens dominates this planet because of, not despite, its intelligence. 38 A problem arises from the fact that humans are the only persons on this planet. Many other creatures are, presumably, conscious and many
Origins
105
others have impressive cognitive powers, amounting to a kind of intelligence, but so far as we know these have led to self-awareness, and thus personhood, in humans alone. Unlike vision or flight the rational selfconsciousness that makes its bearer a personal being came about just once on this planet, or, if more than once, in closely related species, of which only we remain. This contrasts with phenotypes that have evolved many times independently. Some theorists conclude from this that the existence of persons is probably a one-time occurrence without parallel elsewhere. Even if we share the universe with other forms of highly evolved life, the only persons in the universe might be humans, here on Earth. For if rational, conscious organisms evolved just once here, where we know conditions are right, how probable is it that natural selection reliably produces persons, and that the existence of persons in the universe is not an unpredictable fluke or the product of divine intervention? An important consideration here is that persons might be exceedingly rare in the universe but still exist in great abundance. The age and size of the universe is so great that things that are in relative terms very rare nonetheless can exist in very large numbers. Isaac Asimov once asked, Do you know how rare astatine-215 is? If you inspected all of North and South America to a depth of ten miles, atom by atom, do you know how many atoms of astatine-215 you would find. . . . Practically none. Only a trillion. 39 Persons in the universe might be extremely rare, coming about only when a large number of independent conditions are satisfied. 40 Maybe there is just one species of person per every 100 billion stars. That seems very rare indeed, but it amounts to at least one per galaxy, and there are at least 100 billion galaxies in the known universe. 100 billion different species of persons would make them relatively extremely rare yet, absolutely speaking, extremely plentiful. The rarity could be increased by orders of magnitude and there would still be a very large number of kinds of personal creatures. Depending on point of view, the universe could be at once almost devoid of personal life and rich with multifarious versions of it. 41 However, a significant question remains: the so-called Fermi Paradox, which is not a paradox but Enrico Fermi’s observation, made casually over lunch in 1950, that given reasonable assumptions, if there were other intelligent species out there we should have heard from them by now. The universe could have produced intelligent species millions, even billions, of years ago, which gives them plenty of time to have signaled us, or for they themselves, or self-reproducing automata of their making, to have reached us. Yet so far no contact. A variety of attempts have been made to explain why, if they exist, we have not heard from them, but none are compelling. 42 The apparent absence of aliens is, I believe, a real problem. Indeed, it seems to me the best reason Christians could have to reject the evolution-
106
Chapter 4
ary explanation of human origins. While it is quite plausible that God would choose indeterministic means to create persons, it is not at all plausible that God would make the universe in such a way as to produce exactly one species of person. One is too close to zero. We can suppose that God would create a world in which indeterministic processes somewhere, at some time, bring persons into being, but not that God would create a universe where things might easily have gone the other way, so that no one is home in God’s creation. Personhood in the universe might be relatively very rare, but for Christians the reasonable expectation is that there are, in absolute terms, many instances of it. 43 The crucial question is whether personhood in the cosmos could be rare enough to account for our current solitude, yet common enough to be the predictable result of indeterministic natural processes set in motion at the beginning of creation. Too many imponderables are involved here for a definitive answer; for example, would non-human persons, with psychologies significantly different than ours, have the motive and means for intergalactic signaling or voyaging? Finally, the question whether there is anyone else in the universe is empirical. While the only near-term answer we can hope for is the positive one of a call or a visit, in the long term absence of evidence could accumulate to compel the conclusion that we are alone. This negative result would falsify the account of our origins put forward here in The Material Image. No doubt, an encounter with non-human persons could generate interesting challenges for Christian faith, but the persistent lack of an encounter might create a more serious challenge. 44 THE APE AND THE IMAGE Evolution by natural selection in an indeterministic world rules out the design of created persons and thus challenges a traditional Christian conception of human origins. For some, other aspects of the evolutionary account are at least as disturbing. How could a modified monkey be the image of God? Charles Woodruff Shields, who in 1865 became Princeton University’s first professor of “the harmony of religion and science,” contended that Darwinism means, “Man could see in himself only a developed animal; his highest and purest culture would be accepted as but the gradual outcome of savage bestiality; and the image of God be lost in the image of an ape.” 45 Evolutionary theory of course does not imply that humans descended from monkeys; our lineage diverged from that of chimpanzees at least five—and in the case of monkeys, at least thirty— million years ago, so present day monkeys are our cousins, not out ancestors. This will not mollify those who find the idea that we have nonhuman antecedents preposterous and wicked. For them, the likelihood of
Origins
107
any theory of origins that implies the imago Dei resides in creatures made from animals is vanishingly small. Despite the small number of times it occurs in the Bible, the assertion that human beings are the image of God occupies a central place in Christian reflection on human nature. It is reasonable for Christians to evaluate scientific hypotheses that have implications about human nature in light of this doctrine, and to assign low prior probability to those that conflict with it. However, while Christians all agree that we are made in God’s image, there is less than unanimity on its meaning. 46 A traditional, and still popular, way to understand what it means for humans to bear the divine image is to fix on characteristics of humans in virtue of which we resemble God. “Man,” Calvin wrote, “is called God’s image because he is like God.” 47 The resemblance most often pointed to lies in the fact that human beings, like the Creator, are rational, and possess moral agency. Other characteristics have been emphasized or added, as ways of conceiving God and ourselves have changed over the years. Today freedom, creativity, and a capacity for emotion and for interpersonal relations, are frequently cited as what make us like God. 48 The characteristics the similarity interpretation of imago Dei focus on are essentially those persons possess. We may reasonably take this traditional view as implying that to be made in God’s image is simply to be a person and thus that personhood is both necessary and sufficient for being made in the image of God. 49 The similarity interpretation is commonly contrasted with one that locates the imago Dei in God’s special relation to humans, rather than in our intrinsic properties. It is not entailed simply by our personhood, but is a matter of God’s choice. To be designated God’s image is to be called to the task of representing the Creator in the creation. God’s image is the creature freely chosen to serve as God’s steward, to exercise stewardship over creation on God's behalf. This appears closer to the intentions of the author(s) of the texts, which biblical scholars read as alluding to the practice, common in the ancient Near East, of a high king setting up his statue—his image and likeness—to signify his rule over a territory. 50 This vocational, relational understanding coheres with other aspects of biblical theology. It points to the continuity of God’s initial creative acts with God’s subsequent acts to redeem and reconcile creation. God calls Israel as the elect people not because of any special qualities it possesses, nor because of any supposed superiority among the nations, but quite the reverse; its special standing is entirely a matter of God’s gracious reaching out to the lowly (Deuteronomy 7:7). It extends further, to Christian teaching about salvation: God comes to us in the flesh not in response to some real or imagined merit in human beings, but in virtue of a loving commitment to needy creatures. Christians conceive their faith in God not as a sign that they are better or wiser than others—that is plainly false—but of God’s electing grace, in virtue of which we are privi-
108
Chapter 4
leged to know and bear witness to the God who in Christ acts to save the world. With this in view, we rightly hesitate to assume that our being the image of God is a matter of our superiority to other creatures. Given what we know of God’s proclivity to reach out to the least and lowly, we ought not to be surprised to discover that the universe contains all manner of persons superior to humans. Karl Barth wrote that the incarnate God, in freely chosen humility, descends to the lowest depths of creaturely life. Herein lies the glory of God. 51 The doctrine of the imago Dei testifies to the greatness of God’s grace, but it does not imply much about the greatness of human beings. 52 Ultimately, a relational understanding of God’s image resonates with the purpose for which God created us. We are God's image because, as persons, our vocation is a relation of mutual love and trust with the uncreated person who made us. Insofar as being a created person in the fullest sense is to exist in interpersonal relation, and thus to mirror the perichoresis—the reciprocal indwelling and mutual constitution of the persons of the Trinity—the two accounts converge. 53 Even if, with most theologians and biblical scholars, we see the relational account as superior, personhood remains a necessary condition for being made in God’s image. For nothing but a person could assume the responsibilities and take on the vocation to which God calls humankind. Given their lack of cognitive capacity, lack of freedom, and other ontological shortcomings, God could not have called tuna or turnips to bear the divine image, to be persons tasked with God’s work in the world. No less than the similarity interpretation, the relational interpretation construes rationality, consciousness, and whatever else follows from being a person as necessarily presupposed by being made in God’s image, even if it does not constitute it. Christians (and others) often find it inconceivable that something that has non-human ancestors could be a person, and thus satisfy this necessary condition for being God’s image. Sometimes, this objection rests on little more than a sense of propriety: it would be beneath the Creator’s dignity for the divine image to reside in a creature with such lowly progenitors. God’s image must be a superior sort of being, not something derived from lowly animals; humans are worthy of this exalted role only if they have no such embarrassing antecedents. Reminiscent of the God of Platonic philosophy, who cannot be tainted by the material world, the deity here envisioned is incapable of stooping to a creature too closely related to chattering monkeys or to hairy, smelly apes. Objections of this type are alien to the Christian faith, which teaches that God’s relation to creatures is grounded ultimately not in the creature’s merits, but in grace. Even if we resemble our Creator because we too are persons, we cannot parlay this into a denial of our obvious similarities with less impressive beings, even ones with tails. Being God’s image ought not to lead us to forget we are God’s image, not divine beings. On examination, our simi-
Origins
109
larities to God are not unequivocal; in crucial cases we can at best see ourselves as material analogues or simulations of the transcendent, immaterial God we image. We share certain important properties with our Creator, but not in a straightforward way. The point, ultimately, of our being made in the image of God is that we are God’s, not gods. The Christian confession is that humankind serves as the image of God not because we have a claim on this status, but because God, despite our unworthiness, condescends to us. Indeed, this kind of objection to Darwinism seems motivated by a desire to defend not God’s dignity but our own. 54 To the extent that it is, we risk an inflated view of what we are, one that ignores the prevailing biblical view of humankind as weak, vulnerable, and in constant need to trust in God’s care. 55 A more credible objection appeals not to impropriety, but to alleged impossibility: nothing evolved from non-human animals by Darwinian mechanisms could possibly possess the rationality, freedom, moral nature, and so on that a creature must have in order to be a person and thus be made in God’s image. This objection might be grounded, not in the nature of natural selection, but in the idea of God creating us by any sort of natural process. James McCosh, after admitting that Darwinism has limited explanatory power, asserts: Mr. Darwin does not attempt to show, and all attempts by others have failed to prove, that the law of selection, or any other, can account for the origin of life, the origin of consciousness, or of knowledge generally, and the origin of man with his high psychical qualities, such as his power of using speech, his intuitive reason, his appreciation of beauty, his conscious freedom and responsibility, his ideas of moral good, and his immortality. It is as true as ever, that we know no law of nature operating at present which is capable of producing these phenomena. It may be safely asserted that, if the origin of these powers be ever accounted for, it will be by far higher agencies than those contemplated by Mr. Darwin or Mr. Huxley. 56
This critical stance rejects the idea of human persons being brought into existence by any form of secondary causation. For McCosh and others, then and even now, this is a matter of philosophical incredulity; they cannot imagine any natural process resulting in such amazing properties as rationality, consciousness, freedom, and moral agency. In this book’s remaining chapters, I seek to show that whatever justification this incredulity had in the century after Darwin, it is no longer justifiable. Increasingly, human nature is being caught in the net of scientific explanation in general and evolutionary explanation in particular. We now have excellent reasons to believe that God could, and in fact did, bring about persons by way of natural causes, indeed those of the kind Darwin and his successors describe. We are now in a position to see ourselves as the evolved image of the Creator.
110
Chapter 4
In 1995 the National Association of Biology Teachers issued a statement on the teaching of evolution: “The diversity of life on earth is the outcome of evolution: an unsupervised, impersonal, unpredictable natural process that is affected by natural selection, chance, historical contingencies and changing environments.” 57 This elicited protests from those who saw it as going beyond what is scientifically warranted to endorse a secular vision of human origins, and under pressure the organization revised it. 58 On my account, this is ironic, for the original statement depicts the kind of process we might have expected from the God of Christian faith. Karl Giberson speaks for many Christians when he expresses the hope that there is an “infusion of divine creativity under the scientific radar,” and that somehow we are the product of design. 59 Despite the persistent desire to secure a place for design in nature, particularly when it comes to human origins, the creation as a whole appears to have been designed precisely to achieve God’s purposes without design. COVERT INTERVENTION Some contemporary thinkers accept that nature’s basic laws are indeterministic but suggest that at critical junctures God directly produces events at the quantum mechanical level that lead to divinely desired mutations. By this means God maintains the control design requires, guiding the course of evolutionary history to a specified outcome. For instance, Robert John Russell has developed the idea that God determines the outcome of quantum mechanical events and in so doing causes point mutations to guide the course of evolution in a desired direction. 60 This strategy retains divine design in a domain where to all appearances there is none, and it does so while honoring the letter—if not the substance—of Darwinism. If we had good reason to try to reconcile divine design and science, this might be our best option. But this expedient is not necessary. From the perspective of Christian faith, we have no reason not to take the apparent lack of design at face value. IS EVOLUTIONARY NATURALISM SELF-DEFEATING? The distinguished Christian philosopher Alvin Plantinga has advanced an argument against the human mind being the product of unguided evolution that calls for special consideration. 61 He asks about the antecedent probability of our cognitive faculties being reliable, given the naturalistic view that they are not the result of divine design, but result from the mindless, purposeless mechanisms of biological evolution. The answer, in Plantinga’s view, is that the probability is not high; it is either low or inscrutable. Therefore, the naturalist’s belief that naturalism is true should undermine his confidence in the cognitive faculties that give
Origins
111
rise to all his beliefs, including his belief in evolution. We have no reason to regard as reliable cognitive capacities that are not the product of design. If our beliefs are the product of mindlessly evolved mechanisms we have no reason to suppose they are true, so we have no reason to think evolutionary theory is true. 62 The heart of Plantinga’s objection is the claim that we have no reason to suppose that natural selection (in contrast to “special creation” by immediate divine intervention or by divinely guided evolution) would favor creatures that have true beliefs over those that have false beliefs. Initially, the notion that false beliefs might be no less adaptive than true beliefs sounds implausible: all things being equal, creatures genetically disposed to believe that predators are harmless are less likely to pass their genes on to the next generation than creatures genetically disposed to believe that they are dangerous; for most creatures, being eaten seriously lowers the chances of producing offspring. However, Plantinga knows that beliefs on their own have no effect on behavior. It is only in combination with goals, motivations, or desires that beliefs lead an organism to act one way rather than another. The belief that a predator is likely to eat you leads to aversive behavior because you want to avoid being eaten. If you wanted to be eaten, this belief would lead you to seek out, rather than flee, predators. There are, Plantinga points out, any number of possible pairings of desires and false beliefs that would produce adaptive behavior and thus be favored by natural selection. Plantinga offers as an example Paul, a prehistoric hominid approached by a tiger. Paul wants to become the tiger’s dinner, but he mistakenly believes that the best way to attract it is by waving a firebrand in its face. He does so, the tiger runs away, and thanks to this adaptive behavior Paul lives to transmit his genes to the next generation. For all we know, we are descendants of pre-historic Pauls and, like them, have systematically false beliefs. Natural selection is indifferent to the truth or falsity of beliefs as such. All that matters in evolutionary terms is the behavior they produce, and only one thing about that behavior matters: is the organism likely to have offspring that have offspring that have offspring, and so on? Natural selection will favor true beliefs if having them enhances inclusive fitness, but it will favor false beliefs if having them does so, and, if either true or false beliefs will do equally well, then we have no reason to suppose it will favor organisms whose genes produce cognitive mechanisms that produce true beliefs. If there are many possible belief-desire combinations that are adaptive, but relatively few involve true beliefs, it is not likely that our evolved cognitive mechanisms reliably produce true beliefs. How reasonable is it to think that true or false beliefs could do equally well? Plantinga’s argument assumes this, but the naturalistic response is that this is not even remotely plausible, and that the scenario he wants us
112
Chapter 4
to imagine is not scientifically credible. There could be creatures whose beliefs are systematically false yet adaptive, but the chances of Darwinian evolution producing such creatures is negligible. Consider, first, a simple organism that engages in behavior that enhances its chances of reproducing, where this behavior is caused by some internal state that represents some feature of the environment. A representation is caused by the presence in the vicinity of the organism of a predator of a particular kind, and the representation in turn causes the organism to try to get away. It is playing this causal role that makes it a representation of that kind of predator. Nothing intrinsic to it makes it mean this or anything else. Suppose that someone suggests that it might not be an accurate representation, that it really represents something else. This is not a sensible proposal. It is a representation of the predator just because of its causal role: it is characteristically caused by the presence of the predator and it characteristically causes avoidance. There is no further fact beyond this about what it represents that the scientific theory could be getting wrong. Plantinga would, I assume, accept this but insist, “none of this has anything to do with belief.” 63 In his view, there is a dramatic discontinuity between representations in such simple creatures, which are guaranteed to be generally accurate—though in what he regards as a trivial sense—and beliefs, which can be true or false. Obviously, there are very large differences between a simple creature that lacks language or even full-fledged concepts and the human mind. Yet evolutionary naturalism does not accept the radical discontinuity Plantinga assumes, for it is committed to there being an explanation of human minds as coming into existence in virtue of the operation of natural selection upon just such relatively simple non-human minds. What is not plausible is an evolutionary route from creatures whose mental representations—in some form more primitive than genuine beliefs—are systematically accurate, to creatures whose mental representations—in the form of beliefs—are systematically false. We have no reason to think the unplanned, piecemeal operation of natural selection upon an ancestral population of accurate representers would give rise to a descendent population whose false beliefs just happen to mesh with its desires so as to cause adaptive behavior. Mutations might cause an individual to represent certain features of its world inaccurately, for example, to represent tigers as harmless pussycats; on its own this would be maladaptive. And mutations might result in unusual desires, for example, to want to make friends with tigers, but on its own this too would be maladaptive. Natural selection would not systematically pair false beliefs with idiosyncratic desires. The evolutionary origin of our belief formation mechanisms casts no reasonable doubt on their reliability for our daily lives. Naturalists justifiably ignore Plantinga’s scientifically unrealistic scenario. It faces a dilemma. Either prehistoric Pauls were disposed to notice when their actions
Origins
113
failed to achieve their aims, in which case they found better means to attract tigers, got eaten, and they did not become our ancestors. Or they were not disposed to notice when their actions failed to achieve their aims, failed to achieve them, lacked fitness, and thus did not become our ancestors. Evolutionary theory implies our beliefs could be systematically false but imply they probably are not. SPECIAL CREATION A Christian might have a robust sense of what God is capable of doing by means of secondary causes but, because of a certain reading of early Genesis, remain committed to the special creation of the first humans, construing this as creation by means of miraculous intervention. In contrast to its description of the creation of other living things, which is on its face consistent with secondary causation, e.g., “And God said, Let the earth bring forth living creatures” (Genesis 1:24a), the biblical text portrays the first humans coming into existence by God’s direct action: “Then the Lord God formed man of dust from the ground” (Genesis 2:7a). The term “special creation” generally expresses the conviction that the first human beings were created by a direct, miraculous act, not by natural means. The naturalistic view of human origins is clearly incompatible with this. However, it is not incompatible with, and in fact embraces, much of what the term connotes: human persons have a unique and honored place in the divine economy. What is special about human beings is not how they were created, but why. For some, God creating humans directly signifies an intimate relation to them, caring for them in a way God does not care for impersonal creation. If what I have said in this chapter is correct, this is a mistake: God, making creatures for personal relationship with their Creator, would choose to create them by secondary causes, indeed by indeterministic ones. 64 Special creation can be recast in terms of the purposes for which God creates. This reconception of special creation in terms of God’s purposes, rather than God’s methods, is in keeping with the purpose of the ancient creation narratives, which is not to explain how things were created, but to portray the world, and humans, as God’s creation. In chapter 3, with miracles in view, I argued that a consideration of the purposes for which God created motivates a principled distinction between divine intervention in history and in nature, and leads to the conclusion that we should expect God to intervene in human affairs. Here, the same considerations lead us to expect that God refrains from intervention in nature, and in particular to conclude that God employed no miracles to bring the human species into existence. We should assume that after the initial act of creation, and prior to the forces of evolution bringing the first created persons into being, God did no miracles. This
114
Chapter 4
does not imply that God was inactive during that 13.7 billion year span, that for those eons the Creator was, as deists imagine, absent. With Genesis (2:3) in view, perhaps we should say that in this period God was at rest, having set in motion the natural processes that led to persons, bringing about generally intended and foreseen ends by way of secondary causes. For example, the sun came into existence something like four and one half billion years ago; one of the many things God did between the creation of the universe and the creation of the human species was to create this particular star. This was not a miracle; it was God’s action by means of natural, secondary causes. What transpired during this era was the Creator’s doing, but this is no reason to doubt the competence of science to explain what happened then, to explain how atoms, stars, planets, biological life, animals, and the human species came into existence. What happens in the natural course of events in God’s creation is precisely what science aims to explain. No miracles are called for until beings like us, the material image of God, appear, ready to meet our Creator. Finally, less an objection to the evolutionary account of human origins than an attempt to confine it and render it innocuous, is the idea that it pertains only to physical things, and that our capacities for reasoning, and free choice, and whatever else makes us persons, belong not to the body but to an immaterial mind or soul. This grants that the human body is the product of an evolutionary history traceable back through ape-like progenitors, but insists that what makes us persons has no share in that history; the soul is a supernatural addition. This, not anything produced by natural selection, is what enables us to be the image of our immaterial Creator. This view is at least implicit in McCosh’s talk of the need to invoke a “higher agency” to account for the “high psychical qualities” of human beings, and more recently it has been offered as a Christian compromise with evolutionary biology by Billy Graham and Pope John Paul II, among others. 65 More recently, Joan Roughgarden, an evolutionary biologist and devout Christian, assured readers, “we inherit our bodies from ancestors who were also the ancestors of other species. We do not inherit our souls from ancestors.” 66 On this account, the natural processes of evolution formed the first human bodies, to which God affixed an immaterial mind. What truly matters in the human self-image, our capacity for rational consciousness, free choice, moral responsibility, and our other mental properties, belong to us in virtue of our being more than our bodies. What makes us persons is not a product of the unguided course of evolution, for it is not part of nature. To render our evolutionary origins palatable by denying that the human mind is material drives a wedge into scientific naturalism, which contends that human beings are material things, products of Darwinian evolution. In the next chapter, I take up the naturalistic account of the mind and point to the compelling evidence for a materialist view of hu-
Origins
115
man beings. Many Christians find this just as unlikely as evolution by natural selection, but I make the case that it is the materialist account, not the mind-body dualism still dear to many persons of faith, that is consonant with Christianity. Christians should, in fact, welcome materialism as heartily as evolution. Even if they accept that the causal laws God used to create us are indeterministic, some may suspect that because our Creator remains the first cause of all that we are and all that we do, leaving us lacking the distinctness I have ascribed to us in this chapter. It might seem that we can be truly distinct from our Creator only if we transcend physical creation and possess freedom of a radical kind, freedom from natural causation. Later, in chapter 6, I describe the freedom and responsibility material creatures can have, and argue that it suffices for the purposes for which we were created. NOTES 1. Daniel Dennett, Darwin’s Dangerous Idea: Evolution and the Meanings of Life (New York: Simon and Schuster, 1995), 63. 2. Natural selection is not the only process involved in biological evolution. There is also sexual selection, genetic drift, linkage, and recombination, but natural selection deserves pride of place in virtue of being the only natural mechanism that clearly produces complex adaptations. There is some controversy as to what degree natural selection, in contrast to the other processes, should be relied on to explain the course of evolution. Beyond this, there is controversy as to what role should be accorded the constraints imposed on biological forms by basic physical laws, and by mathematical and geometrical principles. Some theorists, e.g., Stuart Kaufmann, maintain that the latter merit a salient place in the evolutionary explanation; see At Home in the Universe: The Search for the Laws of Self-Organization and Complexity (New York: Oxford University Press, 1996). However, most evolutionary theorists are content to take whatever constraints basic natural and mathematical laws impose simply as the background against which evolutionary explanation proceeds. 3. The murph story begins, as it should, in media res, providing no explanation of how there came to be living things capable of self-replication in the first place. The origin of biological life, in contrast to its diversification, complexity, and adaptation falls outside the scope of current evolutionary theory. Although there is a large and growing body of promising research on the origins of life, there is as yet no wellconfirmed explanatory theory; see, e.g., J. William Schopf, editor, Life’s Origin: The Beginning of Biological Evolution (Berkeley: University of California Press, 2002). The reasonable expectation is that such an explanation will be forthcoming. Such explanations often invoke phenomena akin to natural selection in the development of complex chemistry on the way to the first simple living things. Arguably, natural selection is not a strictly biological phenomenon, but a fundamental law of nature: anything, even certain molecules, capable of imperfect self-replication in an environment in which copies’ probability of self-replicating is a function of inherited characteristics will evolve over time, adapting to the environment. In that case, there could be an evolutionary explanation of life’s origin. Some Christians invoke God to fill current gaps in natural scientific explanations of the origins of life, e.g., Michael Behe, The Edge of Evolution: The Search for the Limits of Darwinism (Free Press, 2008), but this has no apparent motivation other than the judgment that the Christian faith renders a natural explanation of life’s origins improbable.
116
Chapter 4
4. I suspect that the denigration of divine action by natural, indirect means, as opposed to direct intervention, manifests an illicit projection of human limitations onto God. We can be alienated from what we bring about indirectly. We can act, intending to cause some future effect, but forget about it by the time it occurs, or change, so that by the time it finally happens we no longer care about it, or even regret causing it. When it occurs we can fail to notice, being preoccupied with something else. Further, our understanding of the situation in which we initially act is incomplete and sometimes faulty, as is our understand of the laws governing natural causation, so what we cause is not necessarily what we intended. None of this applies to God. The vast reaches of time that lie between creation’s beginning and now make no difference: God acts and what is specifically intended, or if natural laws are indeterministic, generally intended, occurs. 5. There is no shortage of Christians who contend that we need not wait to hear from science, since God has spoken in the Genesis creation narrative. The issue, of course, is whether this is rightly interpreted in some literal, historical, quasi-scientific, manner. This reading of the biblical text has become familiar to many, and we cannot reasonably expect them to abandon it, despite its implausibility, unless they become aware of a superior approach, one that more adequately integrates the creation story into the full biblical account of God’s action in creation and human history. That account, which is on offer from biblical scholarship and has the virtue of not being at odds with what science tells us about human origins, I will discuss and endorse in chapter 8. 6. Cornelius G. Hunter, Darwin’s God: Evolution and the Problem of Evil (Grand Rapids, MI: Brazos Books, 2001), 9, 165. 7. Alvin Plantinga, “When Faith and Reason Clash,” in Robert T. Pennock, ed. Intelligent Design Creationism and Its Critics: Philosophical, Theological, and Scientific Perspectives (Cambridge, MA: MIT Press, 2001), 130. 8. Phillip E. Johnson. An Easy-To-Understand Guide for Defeating Darwinism by Opening Minds (Downers Grove, IL: InterVarsity Press, 1997), 17. 9. John Dewey, “The Influence of Darwinism on Philosophy,” in The Influence of Darwin on Philosophy and Other Essays in Contemporary Thought (New York: Henry Holt and Company, 1910), 12. 10. Bertrand Russell, Religion and Science (London: Oxford University Press, 1935), 80. 11. William Provine, review of Edward J. Larson, Trial and Error: The American Controversy over Creation and Evolution (New York: Oxford University Press, 1985), in Academe (January/February 1987), 51–52. 12. Charles Hodge, What Is Darwinism? (New York: Scribner Armstrong and Company, 1874), 61. 13. Speaking at the Sunday Times Oxford Literary Festival, March 23, 2007. Video file at www.youtube.com/watch?v=02jYAr3wfG8. Accessed 11 November 2019. 14. See Michael T. Ghiselin, The Triumph of the Darwinian Method (Berkeley: University of California Press, 1969), 153–59. 15. This was noted by some of Darwin’s contemporaries, e.g., Charles Kingsley, who in his 1863 tale The Water Babies has a character allude to Darwinism, saying, “anyone can make things, if they will take time and trouble enough, but it is not everyone who, like me, can make things make themselves,” The Water Babies: A Fairy Tale for a Land-Baby (London: Macmillan, 1885), 307. 16. Michael Levin, “How Philosophical Errors Impede Freedom,” Journal of Libertarian Studies 14(1998–99), 133. 17. In general physical systems change through time deterministically, in accord with Schrödinger’s equation, but when a measurement is made the system changes in accord with Born’s rule, which supplies only probabilities for possible observations. The Copenhagen interpretation holds that what a measurement reveals—how the wave function collapses—is ultimately not determined and that nature is, at bottom, stochastic. Indeterminism lies in objective reality; it is not a function of human ignor-
Origins
117
ance. There are various alternative interpretations, but none have displaced the prevailing view. This is not to say that this interpretation, in contrast to quantum mechanics itself, is a well-confirmed scientific theory, but it is reasonable to believe it. For that reason it, and its implication that the universe is governed by indeterministic laws, is this book’s assumption. But not only for this reason: indeterminism is also what we should have expected, given the belief that this world was created by the God of Christian faith. Further, a major puzzle arises from the fact that we know measurements occur, since human beings make observations, e.g., we deploy an instrument to detect the emission of a particle, yet it is not clear what makes something a measurement. Do they occur just when humans make observations in laboratories, or are they pervasive in nature, e.g., whenever microscopic things interact with large objects in certain ways? Here, in part because of the implausibility of the idea that human cognitive activity as such plays so central a role in the physics of nature, and in part because it is likely in light of Christian theology, my assumption is that measurement pervades the universe. 18. Peter van Inwagen, “And Yet They Are Not Three Gods But One God,” in Philosophy and the Christian Faith. Thomas V. Morris, ed. (Notre Dame, IN: University of Notre Dame Press, 1988), 241–78. 19. This minority view is often associated with Duns Scotus. See Richard Cross, Duns Scotus, (New York: Oxford University Press, 1999), 127–29. 20. But not so audacious as to imply that God had no other reasons for creating, unknown to or even incomprehensible to us. Nor does it imply that God is not interested in, or concerned for, the impersonal aspects of the creation. Although, as Colin Gunton notes, the New Testament contains “very little suggestion that the created world is of interest apart from us, its most problematic inhabitants” Christ and Creation (Grand Rapids, MI: Eerdmans, 1992), 33, the larger biblical witness does portray God delighting in the non-human creation (e.g., Psalm 104, Proverbs 30:19, Job 39). As Arthur Peacocke points out, we cannot regard everything else in creation as a mere byproduct of the process by which we were created, for these things have value to God in their own right; “The Challenge and Stimulus of the Epic of Evolution to Theology” in Steven J. Dick, ed., Many Worlds: The New Universe, Extraterrestrial Life and the Theological Implications (Philadelphia: Templeton Foundation Press, 2000), 95. Still, the principal theme of the biblical witness is the travail of God relating to created persons, not to impersonal creation. 21. I do not want to suggest that God will not, in the fullness of time, persuade all created persons freely to accept the vocation of everlasting life with their Creator. I believe that God’s great work of creation and salvation will not fail. 22. Langdon Gilkey, Maker of Heaven and Earth: The Christian Doctrine of Creation in the Light of Modern Knowledge (New York: Doubleday, 1959), 64. 23. Colin Gunton, The Triune Creator: A Historic and Systematic Study (Grand Rapids, MI: Eerdmans, 1998), 101. 24. This reflects reality’s deepest fact: in the one God, the Father, the Son, and the Spirit are distinct persons, bound in a mutuality of love. 25. God conserves creatures in existence; their existence from one moment to the next depends on God’s willing that they exist. But I believe the mainstream of the theistic tradition goes too far when it understands this as God having to act at each moment to keep created things from ceasing to exist. I assume that God, like us, has the power to create something that goes on existing until something destroys it. A created entity exists at any moment only because God wills its existence, insofar as God refrains from willing its demise. The traditional view, with its metaphors of images in mirrors and ideas in minds, tends toward rendering created things as not truly other than God. It is reminiscent of a Platonic idea of creation: if the divine Forms were withdrawn, the things of this world would immediately disappear, collapsing back into the nothingness of formless matter. In contrast, I want to uphold the doctrine that God creates real things distinct from their Creator.
118
Chapter 4
26. Martha Nussbaum, Upheavals of Thought: The Intelligence of Emotions (Cambridge, UK: Cambridge University Press, 2001), 225. 27. In the essay “Chance, Potentiality and God” in his Evolution: The Disguised Friend of Faith? Selected Essays (Philadelphia: Templeton Foundation Press, 2004), 62, Arthur Peacocke advised, “The Christian understanding is that the meaning of the cosmic process revealed by science is ultimately to be expressed in personal terms in the sense that the language of human personality is the least misleading for describing the direction in which the process moves.” 28. Michael J. Sandel, The Case Against Perfection: Ethics in the Age of Genetic Engineering (Cambridge, MA: Harvard University Press, 2007), 47–49. 29. If human beings were “agent causes,” i.e., the ultimate originators of their decisions and actions, this would mitigate the loss of distinctness divine design implies. Nothing we do would be an effect of how we were designed. However, in Chapter Six I will contend that no human being can be an agent cause. 30. I once heard a theologian argue against the idea of God possessing less than complete control of the future because, if we were to accept this, we would have no guarantee that things in the end will turn out well for us and the rest of creation. This seems to me out of tune with a Christian conception of things: our only guarantee that “all shall be well, all shall be well, all manner of things shall be well” lies in the resurrection of Christ, in which the ultimate victory of God is proleptically present. The beloved quotation is from Julian of Norwich, Revelations of Divine Love, trans. Grace Warrack (Grand Rapids, MI: Christian Classics Ethereal Library, n.d.), 57. 31. The God Who Risks: A Theology of Providence (Downers Grove, IL: InterVarsity Press, 1998), 11. 32. As Stephen J. Gould said, if the “tape of the universe” were rewound and replayed, we would not be here; Wonderful Life: The Burgess Shale and the Nature of History (New York: W.W. Norton and Company, 1989), 283 33. Christian theology extends the idea, connecting the divine self-limitation in creation to kenosis: becoming incarnate, the Second Person is emptied of deity to become human (Philippians 2:7). See Jürgen Moltmann, God In Creation: A New Theology of Creation and the Spirit of God (Minneapolis: Fortress Press, 1993), 87–93. 34. There is no general progress, no goal-directedness, in the evolutionary process, but this calls for two qualifications. First, we can expect that over time as a species occupies an ecological niche it becomes better adapted to it, and eventually approaches optimal adaptation to this local environment. This does not, however, imply that individual organisms become better on any other standard. Becoming better adapted might involve becoming inferior by our lights. A well-known case is that of cave fish who, living in permanent darkness, are sightless, although they are descended from sighted fish. For this lineage, “progress” involved becoming sightless. Second, we plausibly suppose that there is a general trend among living things toward greater size and complexity, simply because there are more ways to be large and complex than to be small and simple. As species become specialized for particular environments and differentiate, it is no surprise that individuals grow in size and complexity. Further, as organisms increase in size, it is no surprise that nervous systems develop, in response to the need to coordinate the activities of a large body’s disparate parts. It’s not much of a leap from this is to thinking that natural selection would hit on using those nervous systems to represent the external environment and thereby regulate the behavior of organisms. So, while there is no guarantee that any given lineage will tend toward greater size and complexity, let alone intelligence, it is not unreasonable to suppose that after a few billion years of evolution, creatures with minds might appear. And, given world enough and time, one could even expect there to be persons of some sort, somewhere. 35. As Ted Peters and Martinez Hewlett put it, we should look for the divine purpose where it belongs, in God, rather than within nature. Evolution from Creation to New Creation: Conflict, Conversation, and Convergence (Nashville, TN: Abingdon Press, 2003), 159. Unfortunately, this crucial distinction is not always so clearly made in
Origins
119
discussions of science and faith, where the ambiguous question is whether the universe is purposeful (or teleological). One might, like Aristotle, believe that there is purpose in nature—final causation, natural teleology—yet not believe that the universe was created for a purpose. Others, like Thomas Aquinas, believe that there is both purpose in and purpose for nature. Other theists believe that God created the universe for a purpose, but that the universe contains only efficient causation, no natural teleology, i.e., God created a “mechanical” universe. However, Christian thinkers often disparage mechanism, in part because they see it as precluding human freedom. Further, mechanism is often described as scientifically outdated. We should note, however, that even a casual perusal of the literature across the scientific disciplines suggests that the quest to find what is referred to as “causal mechanisms” is flourishing. However, quantum entanglement does suggest that physical reality at the most fundamental level is no more mechanical than it is material; mechanism, and with it causation and even temporality, as significant as they are in human experience, might be superficial features of the creation. See Wesley C. Salmon, Scientific Explanation and the Causal Structure of the World (Princeton, NJ: Princeton University Press, 1984), 239–79. 36. See Kim Sterelny, Thought in a Hostile World: The Evolution of Human Cognition (Oxford: Blackwell, 2003), 3–50, for a plausible scenario. Whether there is need for an explanation of consciousness, distinct from the mind’s cognitive capacity to represent the world and itself in efficient ways, is a controversial issue. My supposition is that there is no such need, and consciousness should be understood in terms of mental representation. For a defense of this view see Michael Tye, Consciousness, Color, and Content (Cambridge, MA: MIT Press, 2000). 37. In fact, I suspect it is difficult to underestimate the role of this, and other prehistoric forms of reasoning, in the development of the human intellectual capacities we today apply in scientific reasoning, even though this was not an activity of the Pleistocene. See, e.g., Louis Liebenberg, The Art of Tracking: The Origin of Science (Claremont, South Africa: David Philip, 1990), Alison Gopnik and Andrew N. Meltzoff, Words, Thoughts and Theories (Cambridge: MIT Press, 1997), and Edwin Hutchins, Cognition in the Wild (Cambridge, MA: MIT Press, 1995). 38. This is true even if rationality is as much a potential mixed blessing as any other product of selection. Pessimistic scenarios, futures in which a rational species uses its intelligence to destroy itself by way of war or industrial pollution, are sobering, but do not change the fact that current human hegemony is due to our intelligence. 39. Isaac Asimov, Only A Trillion (New York: Grosset and Dunlap, 1976), 8–9. 40. See Peter Ward and Don Brownlee, Rare Earth: Why Complex Life is Uncommon in the Universe (New York: Springer, 2003). 41. Attempts to estimate the odds (e.g., the Drake equation) are often applied to the Milky Way galaxy, not to the entire universe, a practice that can make extraterrestrial intelligence appear much more rare than it might be. 42. See, e.g., Stephen Webb, If the Universe Is Teeming with Aliens . . . Where Is Everybody? Fifty Solutions to the Fermi Paradox and the Problem of Extraterrestrial Life (New York: Copernicus Books, 2002). See also Paul Davies’ discussion in The Eerie Silence: Renewing Our Search for Alien Intelligence (Boston: Houghton Mifflin Harcourt, 2010), 116–39. 43. Were our universe just one of many, or just one component of a vastly greater multiverse, and it were possible for personal life to come about in a significant number of universes other than ours, this would render the apparent dearth of aliens less problematic. The answer to Fermi’s question might be that they are not here—visiting our planet—because they are not here—in our universe—but in other universes. Possibly, if things in this universe had gone slightly differently, it would contain no persons, but the odds might remain overwhelming in favor of its appearance in some universe. Once we realize that in choosing how to create, God would choose to be distanced from the specific results of an initial act (or acts) of creation, theories that posit a multiplicity of universes may seem plausible. They make unavailable design
120
Chapter 4
arguments based upon the “fine-tuning” of the cosmological constants of our universe, but theists may have other, perhaps better, reasons to find them probable. Some Christian thinkers, drawing on the idea that God’s creation is the best possible, but that no one universe contains all possible varieties of goodness, believe that God would have created many, perhaps infinitely many, universes. See, e.g., Timothy O’Connor, Theism and Ultimate Explanation: The Necessary Shape of Contingency (Malden, MA: Blackwell, 2008), 111–29. Proponents of the multiverse, asserting that it has always existed, take this as implying that God did not create it. See, e.g., Victor Stenger, God and the Multiverse: Humanity’s Expanding View of the Cosmos (Amherst, NY: Prometheus Books, 2014). However, so long as it exists contingently we can reasonably ask why it has always existed and advance the answer that it has always existed as an effect of God’s everlastingly free choice for it to exist. 44. Several of the essays in Steven J. Dick, ed., Many Worlds: The New Universe, Extraterrestrial Life and the Theological Implications (Philadelphia: Templeton Foundation Press, 2000) explore the problems for faith alien contact might engender. 45. Quoted by David N. Livingstone in Darwin’s Forgotten Defenders: The Encounter between Evangelical Theology and Evolutionary Thought (Grand Rapids, MI: Eerdmans, 1997), 111–12. 46. For a survey of modern theological reflection on the imago see Gunnlaugur A. Jónsson, The Image of God: Genesis 1:26–28 in a Century of Old Testament Research, Lorraine Svendsen, trans. (Stockholm: Almquist and Wiksell International, 1988). 47. Ford Lewis Battles, John T. McNeill, trans. and ed. Institutes of the Christian Religion (Philadelphia: Westminster Press, 1960), I. XV. 3. 48. One might suspect that the move away from regarding rationality as definitive of human uniqueness is motivated by the fact that machines now begin to approximate rationality. 49. Colin E. Gunton, The Triune Creator: A Historical and Systematic Survey (Grand Rapids, MI: Eerdmans, 1998), 208. 50. See, e.g., Gerhard von Rad, Genesis: A Commentary, John H. Marks, trans. (Philadelphia: Westminster Press, rev. ed. 1973), 59–60, and Hans Walter Wolff, Anthropology of the Old Testament, Margaret Kohl, trans. (Mifflintown, PA: Sigler Press, 1996), 160–61. 51. Karl Barth, Church Dogmatics, II.1. T. H. Parker, W. D. Johnston, Harold Knight, J. L. M. Haired trans. (New York: Continuum), 516–17. 52. For a defense of the idea that a perfectly good God might not create the best creatures possible, see Robert M. Adams, “Must God Create the Best?” in his The Virtue of Faith and Other Essays in Philosophical Theology (Oxford, UK: Oxford University Press, 1987), 51–64. 53. See especially John D. Zizioulas, Being as Communion: Studies in Personhood and the Church (Crestwood, NY: St. Vladimir’s Seminary Press, 1993). 54. Genesis portrays Adam as made from clay or dust, which is no more personal than an ape, but some find the most well-known of our simian relatives, monkeys and chimpanzees, particularly undignified examples of impersonal nature. They seem not to notice that these creatures’ risibility lies precisely in their similarity to humans. And, while invoking God’s dignity, they seem to forget that this is the God who accepted the abject humiliation of a Roman crucifixion. 55. See Wolff, Anthropology of the Old Testament, 10–31, 149–55. 56. James McCosh, The Method of Divine Government: Physical and Moral (New York: Robert Carter Brothers, 8th edition, 1890), 157. 57. Reports of the National Center for Science Education 17(1995), 31–32. 58. For an account of this episode see Karl W. Giberson, Saving Darwin: How to be a Christian and Believe in Evolution (New York: Harper Collins, 2008), 166–69. 59. Karl Giberson, Saving Darwin, 220. 60. See, e.g., “Special Providence and Genetic Mutation: A New Defense of Theistic Evolution,” in Robert John Russell, Cosmology from Alpha to Omega: The Creative Mutual Interaction of Theology and Science (Minneapolis: Fortress Press, 2008), 212–25. For Rus-
Origins
121
sell, this elegantly allows for the possibility of divine design by means that do not involve what he calls intervention, i.e., God bringing about naturally impossible events. On my account in Chapter Three, we should heed neither constraint: God does not guide evolution’s course but may well bring about events that could not have naturally occurred. So far as I know, the basic idea was first expressed by William G. Pollard in his Chance and Providence: God’s Action in a World Governed by Scientific Law (New York: Charles Scribner’s Son, 1958). 61. Alvin Plantinga, “Introduction: The Evolutionary Argument Against Naturalism,” in James Beilby, ed. Naturalism Defeated? Essays on Plantinga’s Evolutionary Argument against Naturalism (Ithaca, NY: Cornell University Press, 2002), 1–15. 62. This argument is analogous to, and in some respects the contemporary descendant of, a persistently popular argument against materialism put forward by C. S. Lewis more than a half century ago in Miracles: A Preliminary Study (New York: Macmillan, 1947). I consider that argument in chapter 5. 63. See his “Reply to Beilby’s Cohorts,” in James Beilby, ed. Naturalism Defeated? Essays on Plantinga’s Evolutionary Argument against Naturalism (Ithaca: Cornell University Press, 2002), 258–60, 263–65. 64. Philosopher Harry G. Frankfurt interprets the early Genesis account of the creation of Adam as revealing that this creature is unique precisely because his nature is not simply a function of God’s creative will: this is the creature that uniquely requires God’s ongoing intervening action. See “On God’s Creation” in his Necessity, Volition, and Love (Cambridge, UK: Cambridge University Press, 1999), 117–28. 65. For Billy Graham, see David Frost and Fred Bauer, Billy Graham: Personal Thoughts of a Public Man (Colorado Springs, CO: David C. Cook: 1997), 72–4. For Pope John Paul II see “Message to the Pontifical Academy of Sciences: On Evolution” (22 October 1996.) 66. Joan Roughgarden, Evolution and Christian Faith: Reflections of an Evolutionary Biologist, (Washington, DC: Island Press, 2006), 17–18.
FIVE Mind
TWO QUESTIONS Human beings are physical objects and nothing more. Daniel Dennett expresses this central tenet of scientific naturalism when he writes, “we human beings are a part of nature—supremely complicated but unprivileged portions of the biosphere.” 1 Dennett is right, at least if being “unprivileged” is a matter of how we are made, for we are made of the same things other living creatures are made of: quarks and electrons, atoms and molecules, cells and organs, and these parts are assembled in the same way: by mindless natural selection. Christian faith confesses other, more significant, ways in which, as God’s image, we do occupy a privileged position in the creation, but it is abundantly clear that we do not transcend the natural world. We are the material image of the transcendent Creator. There is no immaterial soul, no human mind but the brain. Our thoughts, feelings, and choices are events in the physical universe, governed by natural law. Sensing, emoting, reasoning, believing, deliberating, and all our other mental activities are activities of the physically and culturally embodied brain. No aspect of scientific naturalism more directly challenges traditional ways of understanding ourselves than the materialist (physicalist) account of what we are. Many, both secular and religious, consider the idea that human beings possess a non-physical, “spiritual,” component—an immaterial soul—as definitive of religious belief. 2 Many Christians in particular see materialism as starkly opposed to their faith. This chapter begins the task of showing that a materialist conception of human nature, rather than being inimical to Christianity, is plausible in light of it. As with the question of how God created the human species, the view with the higher prior probability in light of Christian theology is 123
124
Chapter 5
the one decisively favored by the scientific evidence, not the one long associated with the Christian faith. The evaluation of scientific naturalism’s materialist account of human nature encounters a false dichotomy akin to one surveyed in Chapter One. There we noted the popular opposition of naturalism, conjoined with atheism, to an anti-naturalist view conjoined with Christian faith, as though these exhaust the possibilities. This “either-or” obscures the possibility of a Christian naturalism and condemns Christian faith to endless conflict with science, which unequivocally supports naturalism. Similarly, the narrower issue of whether humans are material things is often portrayed as a contest between “materialism,” understood not as the idea that human beings are material things, but as the idea that only material things exist. So defined, materialism implies that there is no God. This erases the distinction between two questions: Does God, presumed to be an immaterial being, exist? Are human beings, at least in part, immaterial beings?
These questions might be related, but it is unreasonable to presuppose that we must answer no to both or yes to both. Labeling “materialism” the view that answers no to both silences two other voices: that of the atheist who believes human beings are non-physical, and of the theist who believes that human beings are completely physical. The assumption—either humans are immaterial or nothing is—needs only to be explicitly stated for it to be obvious that we are not entitled simply to assume it. 3 In this chapter “materialism” will be used only for the theory that human beings are totally material beings, not for a grander claim about reality in general. As an account of human beings, materialism is the theory that mental properties are properties of the brain; the mind in some sense just is the properly functioning brain in its embodied, social context. The contrasting dualist view is that the mind is not the brain, for mental properties are properties of a non-physical entity. Thus, a complete account of a human person involves two distinct substances, the body and an immaterial mind or soul. The question of what Christians can reasonably make of what contemporary science says about the human mind is closely related to the issue of human origins addressed in chapter 4. Once we realize that God could have brought persons into existence either in a way that involves design, or in a way that realizes only a general intention to bring persons of some sort into existence, we must go on to ask which, if either, is antecedently more probable given what we know of the reasons for which God creates persons. Despite prevailing opinion in favor of divine design, it is more plausible in light of central convictions of Christian theology that God would choose not to design us but to employ indeterministic natural processes, such as those known to evolutionary biology, to bring us into existence. What natural science tells us of human origins is what we
Mind
125
should expect if Christianity is true. This chapter shifts the focus from human origins to human nature, but the underlying issue remains. God could have created creatures with immaterial minds or souls, and God could have created purely material persons. All things considered, which of these two possibilities is our Creator more likely to have selected? I begin by canvassing the reasons, not only in science but in ordinary experience, that unequivocally support materialism. Then I examine, and reject, explicitly biblical and theological considerations that convince many Christians that their faith clashes with materialism. Having shown that Christian theology as such does not make materialism improbable, I move on to some philosophical arguments for dualism and against materialism. These, while not explicitly theological, are often put forward by Christian thinkers who assume their faith requires dualism. I suggest that these arguments presuppose a problematic conception of what it means for human beings to be created in God’s image. A better understanding of the imago Dei portrays human creatures as the material image of the Creator. MATERIALISM A dualistic conception of human nature is so deeply entrenched in Christian tradition that it is worthwhile to consider the serious challenges it faces. The principal reason to believe that materialism is true and that dualism is false is that while everything we observe is consistent with materialism, there are observed facts that are very unlikely from the dualist point of view. To focus on dualism’s conflict with what we know about the world, let us first consider one clear case. Mike, a devotee of single malt scotch, begins drinking his favorite beverage in earnest. The alcohol is absorbed through his large intestine into his bloodstream, which carries it to his brain. There, the ethanol molecules penetrate the cell membranes of neurons, potentiating the activity of the inhibitory neurotransmitter (GABA) and impairing the excitatory receptor NMDA. Two overt effects are soon apparent. Mike’s motor control begins to deteriorate; he cannot, for instance, safely operate heavy machinery or drive a car, or even walk a straight line. His speech slurs as he begins to lose control of his vocal tract. Also, his sensory experience begins to be negatively affected; his vision becomes blurred and his kinesthetic sense malfunctions as he gets dizzy. Dualists should not find these initial effects surprising. They believe that the brain functions as a kind of transducer, sending sensory information to the immaterial mind, and receiving instructions from it for control of the body’s movements. From the dualist point of view it is predictable that the introduction of alcohol into the brain will lead to a loss of coordination. The immaterial mind sends the proper instructions, but the im-
126
Chapter 5
paired brain fails properly to send them on to the nerves and muscles. Nor is it surprising that a malfunctioning brain sends defective sensory information to the mind, which then experiences blurred vision and dizziness. These early effects are no more surprising for the materialist. According to materialism, the brain controls the movements of the rest of the body and the derangement of its neural machinery by alcohol renders it less capable of exercising control. Similarly, since sensory experiences are events in the brain itself, anything that interferes with its operations can impair sensory experience. If there were no further effects of imbibing alcohol in quantity, materialism and dualism would be on an equal footing, since what we observe in the first stages of the “experiment” agrees with the predictions of both theories. However, as more alcohol goes to work on Mike’s brain, what we observe continues to be what materialism predicts, but for the dualist it becomes anomalous. As he consumes more scotch Mike’s ability to reason is adversely affected. Rather than exhibiting his normal mental acuity, now he cannot think clearly. Also, his emotional state changes. He becomes euphoric. There are appropriate emotional responses to the loss of bodily coordination, and to distorted sensory experience, but this is not one of them. Further, practical judgment about behavior appropriate for the circumstances worsens and inhibitions dissipate. When, for instance, the Highway Patrol officer pulls him over and asks him to submit to a sobriety test, rather than being upset, Mike finds the situation hilariously funny, feeling sure that it is a good idea to remove the policeman’s hat, put it on his own head and impersonate Smokey the Bear. Why, if dualism is true, should alcohol in the brain have these effects on the immaterial entity that emotes, reasons, and makes decisions? There is no alcohol in Mike’s mind; it is only in his brain. Ethanol in the brain should degrade the quality of information the reasoning mind receives from its brain, but it should not derange the mind itself. It should impair the brain’s ability properly to implement the mind’s decisions, but it should not cause the immaterial mind to make foolhardy choices. For dualism, inebriation should affect the mind in certain ways, but not in these ways. 4 There is nothing special about the familiar effects of alcohol upon the mind. Many changes to the brain result in mental changes that make sense from the materialist perspective but which are unexpected on dualist assumptions. If dualism were true, it would not be surprising that a severe blow to the head renders the brain incapable of transmitting sensory information to the mind, or incapable of receiving instructions about bodily movement. What is unexpected is what happens: not a conscious person suffering from sensory deprivation and paralysis, but unconsciousness. This is not surprising if consciousness itself is a property of a properly functioning brain. The fact that physical damage to the brain
Mind
127
drastically impairs reasoning ability is not unexpected if it is the brain that reasons; nor is it surprising that various chemical imbalances among neurotransmitters have profound effects on a person’s ability to reason normally, or upon her emotional states, or upon her personality. Perhaps most troubling for the dualist, who regards the immaterial soul as the seat of moral judgment, is when frontal lobe damage, as in the famous case of Phineas Gage, leads to dramatic changes in the victim’s moral character. 5 This is difficult to understand if a moral agent is an immaterial soul, but it is predictable if human moral judgment is realized in neural circuits in that part of the brain. 6 Scientific investigation agrees with and extends the conclusion of ordinary empirical experience. What occurs in the human mind correlates with what transpires in the neural machinery of the brain in precisely the fine-grained way physicalism calls for. Yet, no matter what is observed with respect to the correlation of mental events and physical events in the brain, the dualist can insist that it is mere correlation, a systematic correspondence between the states of two radically distinct realities. However, in the absence of good reasons to believe otherwise, the one plausible explanation of this correlation is the simplest: mental things are the same things as material things. The standard dualist view is that the immaterial mind or soul causally interacts with the body, by way of the brain. Whenever the mind produces an effect in the brain, a physical event will to all appearances “just happen,” since it has no observable, physical cause. When, for instance, someone voluntarily moves her left index finger, there should be a corresponding event in the brain’s motor cortex, one that has no observed cause. At face value, this implies a violation of the conservation of energy. Those convinced this principle cannot be violated must reject interactionist dualism, but there is no compelling reason to believe the conservation law must be unlimited in scope. There is no need to deny that, if there were non-physical human minds, they would be able to act upon the physical world. 7 However, it is one thing to grant that it would be possible for immaterial human minds—supposing they exist—to act upon their bodies, but quite another to have any reason to believe there are physical events caused by them. Decades of neurophysiological investigation of the brain at the cellular level offer no evidence for forces at work other than ordinary physical ones. 8 Dualists can conjecture about how the mind might unobtrusively influence the brain, but nothing we observe invites an explanation of what goes on in the brain by appeal to an unobservable, non-physical cause. The absence of evidence for uncaused events in the brain should be sobering for the dualist. He would be in an entirely different position if there were anomalous events in human brains for the physicalist to contend with, but as things stand, while the empirical evidence is uniformly consistent with materialism, the most that can be said on behalf of
128
Chapter 5
dualism is that it does not render belief in a non-physical soul impossible. In this case, absence of evidence is not proof of absence, but it is good evidence for it. DUALISM If the materialist conclusion can be avoided, there must be some decisive reason to reject it and embrace dualism, despite the one-sided empirical evidence. The most common defense of dualism that makes no appeal to theology focuses on what science has so far failed to explain. Science has not explained how it is possible for a material thing to be conscious, in the sense of having subjective experience. Jerry Fodor, the philosopher of cognitive science, wrote that the materialist faces three main questions: 1. How could anything material be conscious? 2. How could anything material be about anything? 3. How could anything material be rational? 9 Fodor unabashedly delivers what he regards as the bad news about the first question: “Nobody has the slightest idea how anything material could be conscious. Nobody even knows what it would be like to have the slightest idea how anything material could be conscious.” 10 There are a variety of unsolved problems in the scientific study of the mind, some of great complexity and difficulty, but it is at least generally clear what methods a solution requires, and what a solution would look like. In contrast, the problem of consciousness—what some refer to simply as the “hard problem”—seems to many less a scientific problem than a mystery. The dualist is happy to announce that the problem of consciousness is so difficult for the physicalist because it is insoluble: consciousness is not a property of material things, but of immaterial minds. However, the dualist can draw no real encouragement from the mystery of consciousness. It is a hard problem, but not hard enough to make dualism plausible. For it is also true that no one has the slightest idea how an immaterial substance could be conscious. 11 The Christian dualist points out that we already know that at least one non-physical mind exists, and that God’s existence shows that there can be an immaterial mind. The Christian materialist agrees, but counters with the fact that we also know that there are material minds: those of cats and dogs, chimpanzees, pigs, dolphins, and so on. These creatures lack the self-consciousness that makes for personhood, but they presumably are conscious; they are subjects of experience. In the phrase popularized by Thomas Nagel, there is “something it is like” to be a bat and various other non-human creatures. 12 Those who claim a material thing could not possibly have subjectivity face a dilemma: they must grant either that, for example, cats
Mind
129
are unconscious automata that do not have sensations, or that cats do have sensations, but that feline sensory experiences are not neural events in cats’ brains but events in immaterial feline souls. Neither is plausible; a cat is a material thing and nothing more, but it feels pain when someone trods on its tail. There are both non-physical and physical things that are conscious. The question is whether the rational, conscious human beings—the persons— that God has created are material or immaterial. Materialism may have little illuminating to say about consciousness, but this is not the case with respect to Jerry Fodor’s other two questions. Indeed, dualism’s current focus on consciousness reveals how decisively the intellectual tide has turned against it. In centuries past, it was not consciousness, but rationality, that posed the central challenge to materialism. What seemed obvious was that our capacity for meaningful thought governed by norms of reason was what set us apart, not the capacity for sensation, emotion, or other kinds of conscious experience that we share with sentient non-human creatures. However, as Fodor goes on to say after expressing his pessimism about the problem of consciousness, we now have some notion of how a material thing can be capable of meaningful thought and we have one very good idea of how a physical thing can be rational: the computational theory of the mind. Much—though so far not all—of human cognition can be understood as information processing occurring in the mind-brain’s neural circuitry. 13 It is reasonable to be confident that in the future more of what our minds do will be explained in this way. It is the reigning paradigm of the cognitive sciences. The dualist focus on science’s present inability to explain how a physical thing can harbor subjectivity draws attention away from the fact that what was long regarded as the crucial problem for materialist theories of the mind has been solved. BIBLICAL ANTI-MATERIALISM Much opposition to materialism on the contemporary scene originates with Christians. Conservative Christians often contend that we should accept dualism because it is taught in the Bible. Christians have a range of views about the Bible, about how it should be interpreted and what weight we ought to acknowledge its teachings to have. Here, in keeping with my essentially traditional stance, I assume a high view of Scripture, one that regards the Bible as the word of God, the uniquely authoritative witness to God’s character, purposes, and actions in human history. If the Bible appears to teach dualism, then it is reasonable for Christians to evaluate the evidence for materialism with a high degree of skepticism, as antecedently improbable. Serious discussion of hermeneutical and exegetical issues would take us well outside the scope of this book; here I briefly raise some general issues and consider a few salient instances. 14
130
Chapter 5
Biblical writers employ terms that are translated into English with terms that a reader who takes dualism for granted but who is unfamiliar with the ancient Hebrew or Greek hears as presupposing dualism. On closer examination it often becomes clear that the ancient authors were not talking about anything like the immaterial mind posited by dualism. The most well-known case of this is the Hebrew term nephes, which, while often translated “soul,” most often bears the meaning “living thing” and is also applied to non-human creatures and on occasion to God. Far from showing that the Hebrew Scriptures embrace dualism, human transcendence of physical reality, the term and its cognates often focus attention on the vulnerable and needy physicality of the human person. 15 Sometimes the language employed in Scripture really is dualistic, and when it is we need to ask whether it bears “ontological commitment.” Human language has built into it various conceptual distinctions that are implicitly dualistic, but we often use it without thereby committing ourselves to a dualist ontology. 16 If, for instance, Daniel says of his friend Ryan, who has been spending more time in the gym than the library, that he has been developing his body rather than his mind, this belies no commitment on Daniel’s part to mind-body dualism. We find it natural to speak this way irrespective of our philosophical beliefs about human nature. It is no more reasonable—to take a familiar biblical example—to infer from Jesus’ words in Matthew 10:28: “Do not fear those who kill the body but cannot kill the soul; rather fear him who can destroy both soul and body in hell,” that humans are composed of two different entities. Jesus is referring to different ways in which a human being can be damaged or destroyed, e.g., to the difference between breaking a leg and being morally corrupted. Similarly, when St. Paul writes to the church at Corinth that he is with them in spirit, though not bodily present, we cannot plausibly suppose he was acknowledging the possibility of paying them an “out-of-body” pastoral visit (1 Corinthians 5:3). More interesting cases are those in which a biblical writer appears to believe in a non-physical mind or soul. For example, St. Paul in 2 Corinthians 12:3 describes an experience he had either “in the body” or “out of it,” he does not know which. If Paul was a dualist, is this just an ancient background assumption we can legitimately “filter out,” as we extract the essential message from the text? Whoever wrote, in Genesis 1:6–7, that God created the “dome” of the heavens, probably believed that the overarching sky is made of metal. No matter our confidence in Holy Scripture, we do not believe that there is a metal dome overhead. The fact that this text’s author took its existence for granted, and refers to it in his description of God’s creation, does not tempt us to believe he was right on the matter. We automatically abstract the theological content of the text—that what we see above is God’s
Mind
131
handiwork—from the way that the author—along with the rest of the ancient world—conceptualized it. It is not always so obvious what it is reasonable—and faithful—to regard simply as a background belief of a biblical author that has no claim upon us, and what we should hear as the word of God, one that can challenge our assumptions about the nature of reality. It seems obvious for us—although it was not obvious for earlier generations of Bible readers—that we should not accept the biblical authors’ assumptions about the Earth being flat and immobile, nor about the legitimacy of slavery, about the treatment of women as chattel, or lending money at interest. But where are lines to be drawn? We should at least not draw them arbitrarily, but Christians sometimes seem to do so. They readily accept that biblical authors’ assumptions about cosmology, geology, and biology are inaccurate, and they acknowledge that receiving the text as God’s word does not involve our trying to believe what we know is false on these matters. Yet, when they move from these ancient “scientific” conceptions of the natural world to the biblical authors’ “philosophical” conceptions of human nature, they hesitate to admit the need to abstract the theological content from the framework of ancient cultural assumptions. This difference in hermeneutical approach has no obvious justification. For the Christian, the biblical text should be read from the perspective of the good news about the God who is with us and for us in Jesus Christ. This does not, of course, resolve all our hermeneutical issues, but it frees the reader from any inclination to regard the Bible as setting forth a theory of the universe, whether cosmological, biological, anthropological or philosophical. Scripture teaches about the character of God, God’s aims in creating the world, and the actions in history by which a disordered creation is reconciled and restored. As noted in chapter 2, these revealed facts have implications about the world and its inhabitants: they have empirical content. What God reveals in the Bible might render a particular theory of the world or of human nature less likely—or more likely: this book’s central contention is that the latter is the case with respect to the theories that comprise scientific naturalism—but the fact that the human writers by whom God is revealed held to a particular theory is no reason for us to suppose that God wants us to accept it. Their beliefs can be regarded as the “packaging” for the good news of a gracious God who in Christ acts to deliver humankind. In terms of the Christian gospel, is a biblical author’s allegiance to mind-body dualism essentially different than his belief in a flat Earth under a bronze dome? The Bible’s purpose is no more to convey philosophical theories of human nature than it is to teach cosmological or biological theories. 17 Even if, as presumably is the case in some instances, a biblical writer presupposes a dualism of soul and body, this in itself is not a reason to enlist the authority of divine revelation behind this theory. 18 There is no reason to be
132
Chapter 5
conflicted because the biblical authors believe and say various things about the natural world, or about human nature, that we find unbelievable. We should not suppose that God, self-revealed through the limited minds of human writers, was limited to those who had only accurate beliefs about inessential matters. 19 A biblical text of particular interest in this context is Genesis 2:7. A traditional reading sees it as recounting the special creation of the first human beings, understood as God miraculously intervening into the natural world to create a human body and then affixing an immaterial human soul to it: “Then the Lord God formed man from the dust of the ground, and breathed into his nostrils the breath of life; and the man became a living being.” First, taken simply in its own terms as part of a literary text, there seems little to recommend a literal reading of this passage. Leaving aside any issue of its relation to scientific knowledge, there are good reasons internal to the Bible to understand this passage, like the rest of primeval Genesis, not as an attempt at a quasi-historical, quasi-scientific account of origins, but as semi-poetic, polemical theological meditation on God’s purposes in creation. From this hermeneutical perspective, there is no actual, pre-historical event to be accounted for, and we can assume that God created the human species by means of the same natural processes as the other species. As I suggested in chapter 4, this does not mean we must abandon special creation. This theological idea is best understood not in terms of a special etiology of our species, but in terms of the human vocation, our special relation to the Creator, and the meaning of human existence in relation to God’s creative purposes. The mythological form of early Genesis conveys the important literal truth that God made the world for the sake of the persons who inhabit it and bear the divine image. Perhaps this is, on reflection, closer to the intent of the original story: God stoops down, scoops up a handful of dirt, shapes something from it, holds it close, and breathes into it, sharing God’s own life with it and thereby making it a living thing given a special relation to the Creator. However, even if one were to insist on a literal reading of this account, it is not dualism that it supports. Instead, what it would support is the vitalism common to the ancient world, and indeed taken for granted by almost everyone prior to modern biology’s discovery of the microscopic causal mechanisms that underlie biological life. God forms from the soil a lifeless body and then imparts biological life to it. God’s breath, God’ spirit, animates the initially lifeless, human-shaped being. It becomes a nephes, a living thing. Nothing in this text suggests the immaterial rational soul later envisioned by dualism.
Mind
133
A SPIRITUAL BEING For many Christians, the fact that human persons are spiritual beings implies that dualism is true. “Spiritual” is commonly used by Christians and others as a synonym for “immaterial” or “non-physical,” so it seems that materialism denies our spiritual nature. Yet there is little warrant in Scripture or in Christian theology for making this its principal meaning. The biblical use of Hebrew and Greek terms translated as “spirit” or “spiritual” originates in ancient vitalist ideas about wind, air, and breath as the animating source of life that has its source in God, the giver and sustainer of life. Humans are spiritual beings because they, among the living creatures who partake in the life that comes from God, can do so knowingly and willingly. We are spiritual not in virtue of being nonphysical, but in virtue of being properly related to God as whole persons, irrespective of what ontological account of human nature—dualist or physicalist—is correct. How little sense it makes to construe “spiritual” as equivalent to “immaterial” is evident in St. Paul’s grappling with the question of how resurrection is possible. He writes that when we are resurrected we will have “spiritual bodies” (1 Corinthians 15:35–44). Paul’s meaning is not transparent, but he did not mean that we will possess non-physical bodies. At face value, an immaterial body is a contradiction in terms. Indeed, the absence of dualist language here is striking, since it is the one instance where the apostle, who most likely was some sort of dualist, makes an extended, theoretical attempt to deal with the issue without invoking an immaterial soul. Similarly, biblical admonitions to be spiritual (as opposed to “carnal,” i.e., “of the flesh,” e.g., 1 Corinthians 3:1–4) are not admonitions to be immaterial things, nor, on the model of Greek philosophy, to turn away from the material world, the body, and the senses toward supposedly higher things, but to put our trust in God and the salvation offered in Christ, rather than in the attitudes and values of a world that rejects God. THE IMMATERIAL IMAGE? Human beings are made in God’s image. It is often assumed that this means that we are similar to God. Simple, but plainly fallacious, reasoning concludes from this that human beings are also immaterial beings. The fact that God has a characteristic, and humans in some way resemble God, obviously does not imply that humans share that characteristic. If it did, we could reason from the fact that God is infinite, perfect, omnipotent and uncreated to the absurd conclusion that we have these attributes. We resemble God in some, but not in all, ways. Being immaterial might, or might not, be one of the ways in which we are similar to our Creator.
134
Chapter 5
In any case, as indicated in chapter 4, the imago Dei is best conceived not as similarity to God, but as our being called to a relationship of love and trust with God and as receiving from God a particular vocation. However, our being made imago Dei presupposes similarities between human beings and their Creator even if they do not constitute the image. Plausibly, the characteristic of human beings our being made in God’s image presupposes is personhood. Only a personal creature, one endowed with consciousness, rationality, and the capacity for free decision and a responsible response to God’s call, can be invited to share in the life of God, and to serve as God’s image in the creation. So, like God, we are persons. This leaves open the question of whether, like God, we are immaterial persons, or whether we are physical persons, the material image of the God who is an immaterial person. THE DUALIST IMAGINATION Neither the scientific evidence, nor the current limitations of scientific explanation, nor Scripture, affords dualists genuine comfort, yet they persevere, advancing a variety of explicitly philosophical arguments against physicalism. Often, these rely on appeals to modal intuitions, our a priori judgments about what our concepts of mind and matter reveal about what is possible or impossible. A contemporary argument along these lines invokes the fact that if it is true that the mind is identical to the brain, this is a necessary truth, one we could not conceive of being false, and from this concludes that the mind is not identical to the brain, because we can conceive of a fully functioning human body (and brain) totally lacking consciousness, devoid of subjective experience yet in its appearance and behavior indistinguishable from a normal human. 20 These “zombie” arguments generate a variety of logical, metaphysical, and epistemological questions; dealing with them individually would take us well beyond this book’s purview. What is worth noting here is the methodological issue. They rely on the assumption that our reasoning about our concepts affords us reliable knowledge of the nature of reality that can compete with science, specifically insight that the mind is not, and in fact cannot possibly be, identical to anything physical. The naturalist is much less sanguine about this capacity. From the naturalistic perspective, while we cannot deny that a priori reflection on concepts can give tell us something about reality—e.g., contemplation of my concept of the number 17 suffices for me to know that it is prime but not composite, not green, not a mammal—we should not suppose that a material mind crafted by natural selection is optimally suited for reliable intuitive judgment about how things can and cannot be. Science has a record of overturning intuitively obvious ideas about what is, and is not, a genuine possibility, revealing the inadequacy of
Mind
135
even some of our most seemingly adequate concepts. For example, for centuries philosophers offered as an example of a necessary truth that the shortest distance between two points is a straight line, but thanks to Einstein and general relativity we now know this is not only not a necessary truth, but typically false. For the naturalist, the fact—if it is a fact— that we can conceive of something that has all the physical properties that according to materialism suffice for a conscious mind yet which lacks consciousness reveals not the non-identity of mind and brain, but the limitations of our concepts of mind and matter. After all, the mind in which these concepts reside did not develop them because they contribute to a perspicacious metaphysical theory of reality, but because they, or at least the capacity to acquire them, enhanced the reproductive fitness of our ancestors. With this in view, materialists responding to modal arguments for dualism typically see their task as explaining why our concepts generate misleading conclusions about the identity of mind and body, by attempting to diagnose, and explain, the “conceptual illusions” at work, and thus to show why things that seem impossible really are not. 21 IS MATERIALISM SELF-REFUTING? Dualism’s encounter with the evidence of ordinary experience and of science is unpromising, and it is difficult effectively to enlist the Bible in its support. So, it is not surprising that Christians committed to dualism have sought to dispose of materialism with abstract arguments that make no appeal to specific empirical or biblical data. Among these is the argument that materialism is self-refuting: if materialism were true we would have no reason to believe it. I have no reason to think a belief is true if it can be fully explained as a result of physical causes in my brain. Whatever causes the belief would cause it whether it is true or false. So, if materialism were true, then all my beliefs, including my belief in materialism, would be ones I have no reason to have. This argument achieved wide circulation in C. S. Lewis’s 1947 book Miracles. Borrowing J. B. S. Haldane’s words, Lewis writes, “If my mental processes are determined wholly by the motions of atoms in my brain, I have no reason to suppose that my beliefs are true . . . and hence no reason to suppose my brain to be composed of atoms.” 22 Although unsound, the argument is superficially plausible. 23 Some ways of physically causing beliefs undercut their rationality. If we know that someone’s belief that there are bats in the attic was caused by her eating a spoiled burrito, then we have no reason to suspect that her belief is true. Not all causes of a belief are good reasons for it; but this does not support the conclusion that all ways of physically causing beliefs undermines their rationality. 24 Speaking of causes and reasons, Lewis asserts the “two systems are wholly distinct.” Beliefs that can be adequately
136
Chapter 5
explained as the effect of natural causes ipso facto are not rationally grounded, for, in his view, the only mechanism by means of which some beliefs could cause others is association, which does not transmit truth. If Daniel’s belief that the car he saw advertised during the Super Bowl will make him popular with beautiful women is merely the result of the images of the beckoning models featured in the ad, his belief is not rationally justified. However, Lewis’ view of ways in which beliefs can be caused is too narrow. The norms of rational thought can be embodied in the causal relations of a physical mechanism. This possibility has been in view since the nineteenth century, when it became apparent that the laws of deductive logic can be realized in electrical circuits. Switches wired in series correspond to conjunction and those wired in parallel correspond to disjunction. Therefore, causal connections within a device could, for example, be so devised that they mirror the rule known to students of elementary logic as modus ponens: imagine, for instance, that one of its states represents Puff is a cat, and another represents If Puff is a cat, then Puff has fleas, and imagine further that its being in these states simultaneously causes it also to be in the state that represents Puff has fleas. If the two input beliefs are true, then so is the output belief; the causal connection preserves truth from input to output. The “two systems”—reasons and causes—can coincide. In fact, since the work of Alan Turing in the 1930s, it has been known that any process of rational thought that can be clearly described can be embodied in a physical mechanism. 25 Turing’s crucial insight, the origin of the modern computer, lies at the heart of cognitive science and of the computational theory of the mind. MIND, MECHANISM, AND GÖDEL The suspicion that the materialist account of the human mind is selfrefuting reflects underlying concerns about human nature that depend upon a flattering, but untenable, conception of what we are. This comes into view when we turn to another well-known argument against materialism, one that shares this allegiance to an inflated, and dispensable, idea of what we are. Kurt Gödel’s famous incompleteness results imply that there are ineluctable limitations on the powers of formal reasoning, i.e., reasoning that can be mechanically embodied in a computer. Were it to turn out that human minds are not subject to these limits, and thus have cognitive powers that computers cannot have, this would be a serious blow to materialism, since, as we saw above, the only good idea anyone has ever had about how a material thing can reason is that it is a matter of computation. 26
Mind
137
Envision a powerful computer programmed with a set of axioms and with rules for logically deriving mathematical theorems from them; the machine mechanically embodies a formal system. Left to run forever, the machine would prove and print out infinitely many theorems of mathematics. Gödel’s surprising discovery implies that no matter how long the machine runs, no matter how many theorems it proves, there are mathematical truths it will not produce. There are truths of mathematics that are not theorems, not provable. The Christian philosopher J. R. Lucas seized on this, pointing out that Gödel, and anyone else who follows his reasoning, can grasp that certain mathematical statements are true precisely because they cannot be mechanically proved. This, Lucas claimed, implies that there are things that humans can know that cannot be known by means of the formal, mechanized, reasoning computers are capable of, and therefore that human reasoning cannot be a matter of mere computation. 27 If human thought is not computation, then in all likelihood the mind is non-physical. Lucas offers a gratifying contrast: the machine stupidly, inflexibly, mechanically grinding away, endlessly trying to prove what flexible, supple, human reason sees cannot be proved. No matter how complex the formal rules with which the computer is programmed, all it can do is follow those rules. It cannot transcend them, reflecting upon its operation as a whole from “outside” its program, and from that perspective grasping what cannot be known by mere mechanical means. A machine can be programmed with the capacity to monitor itself, and to represent itself representing itself, and so on. It might be designed to modify its program over time as a result of what it learns as it monitors its own performance. It might, for instance, be designed to modify its set of axioms after trying and failing for a million years to prove a particular theorem. Nonetheless, at any moment the device—no matter how complex, no matter how many levels of mechanized self-reflection it embodies—comes up against the limits of its current program and simply follows the rules, doing what someone not bound by those rules, viewing the system of rules from the outside, might see as stupid, mindless—mechanical. Lucas’s conviction is that we are not like this. Human reasoning differs radically from machine reasoning, from rule-governed, physically embodied computation. Human minds, he believes, are infinitely selfreflective, not merely potentially, as a computer is in virtue of the possibility of adding new levels of self-monitoring, but immediately. The human—as opposed to the mechanical—mind knows itself knowing itself knowing itself . . . endlessly, but not because while it is doing something, part of it monitors itself doing it, and another part monitors the monitoring of it, and so on, in a sequence of subsystems monitoring other subsystems—the way, for instance, my laptop comes up with its “system resources are severely strained” messages—but in an altogether different way. Unlike any computational device, in which recursion is purchased
138
Chapter 5
at the cost of complexity, the mind is metaphysically simple; it is not composed of parts. It can, because of this, at any point freely take a further self-reflective step and in some holistic way, know itself knowing itself. 28 No material thing, no mere object, can do this. Physical things change from one state to the next as the laws of nature require. The human mind transcends the material world of mere objects that do what they do in accord with the laws that govern them. Unlike physical objects, we are subjects through and through, our minds are never stupidly trapped in the causal matrix of nature, but always able to step beyond our current thinking, evaluate it, and correct it. Lucas’s Gödelian argument, and C. S. Lewis’s self-refutation argument, reject the possibility that the human mind is a complicated object, a mere thing that, like everything else in the world, behaves as the laws of nature require, even when doing its best as a self-reflective, self-critical thinker. What Lewis saw as implicit in the description of the mind as a material, causal entity, John Lucas saw as implied by the logical constraints on computation. If our neural circuitry is wired in a certain fashion then we will reason a certain way, irrespective of how irrational or mistaken it happens to be, and there is nothing we can do about it. We could not even be aware of it. If human thought is just what the evolved neural machinery does, we can never step beyond it to validate our thinking. We must blindly reason—for better or worse—in accord with it. An essentially similar concern animates the self-refutation argument Alvin Plantinga directs at evolutionary naturalism, which we examined in Chapter Four. If the mind is not fabricated by a divine designer intent on endowing it with the power of reason, but is the result of thoughtless natural selection, then what we think of as our subtle, flexible rationality, might—for all we know—be a risible parody of the real thing. If mindlessly evolved minds come up with an evolutionary account of their own origins, there is no reason to believe that their reasoning tracks truth, and thus no reason to believe their evolutionary conclusions. Even if, as I believe, Lewis and Lucas are mistaken in believing that a material mind could not embody rational thought, and Plantinga is wrong to think we have no reason to believe that natural selection would reliably give rise to rationality, the underlying conception of the human mind they share warrants examination, for it has found a home in the Christian tradition. This shared conviction is that the scientific portrayal of human reasoning as computation implemented in neural circuitry crafted by mindless natural selection is profoundly inadequate to the reality of what we are, for our minds are essentially different than anything that might transpire in computers, evolved brains, or anywhere else in nature, at least in nature left to its own devices and not configured, naturally or miraculously, by divine design. Genuine reasoning—as opposed to some sort of mechanical simulation of it—is conceived as some-
Mind
139
thing only a transcendent, immaterial being could do, not a being fully implicated in the causes and effects of the natural world. Why should those of us who believe we are the image of the God who transcends the natural world believe that we—at least our minds—are likewise transcendent? Why should we see such a pleasing account of human nature as likely from the perspective of Christian faith, rather than as at odds with it? TOO LOW OR TOO HIGH? Recall from chapter 1 that from the standpoint of Christian faith, we can identify two ways to go wrong in understanding ourselves. We might have an image of what a human being is that is too low, one that pays insufficient attention to the ways we, as persons called into a relation of mutuality, love and trust with God, differ from creatures not so privileged. We might pay too little heed to our uniqueness among creatures, to our honored role as the image of the Creator in creation. We might fail to take account of the distance that separates rational self-consciousness from mere animal sentience, being a person from being something else. Enthralled by the explanatory successes of naturalistic science, we might see ourselves as less distinctive than we actually are. As science draws more and more of what is characteristically human into its explanatory net, we might be tempted to see ourselves, as B. F. Skinner infamously put it, “beyond freedom and dignity.” 29 Since the beginnings of modern science, persons of Christian faith intent on avoiding this error, have resisted the naturalistic implications of science and contended for a high view of human nature, one that emphasizes our dissimilarities to other created beings and our similarities to God. The conviction that human minds must be more than highly evolved, physically embodied computation, that they must somehow stand beyond the material realm, or at least elude scientific explanation, exemplifies commitment to this exalted self-image. The modern era has seen relatively little concern with the other possible error, that of too high a view of human nature. In light of the longer Christian tradition, the relatively recent investment in combating tendencies toward too low a view of human nature, and the neglect of opposing tendencies to aggrandize the human self-image, is anomalous. The biblical tradition, while confessing that humankind is made in the image and likeness of God, has long seen the greater danger not in a tendency to regard ourselves as too much like our fellow creatures, but in the temptation to imagine ourselves as more than creatures, as gods. The Genesis account of humankind’s fallen condition portrays the root of our selfdestructive alienation from God, from one another, and from the rest of creation, as originating in the desire to be like God—or like gods—knowing
140
Chapter 5
good and evil (Genesis 3:5). Whatever the precise intention of that seminal text’s author, we cannot avoid concluding that rebellion against being a mere creature is deeply engrained in us and implicated in our fallen condition. Human pride, manifest in discomfort with being less than fully autonomous, with being made by someone else, an object assembled of ordinary impersonal parts, and in that way all too similar to creatures that are not called to be God’s image, lies at the heart of what Christianity calls sin. When we find ourselves saying things like, “The glory of Man is that he is the image and likeness of God . . . he is not an animal, not a thing,” we should wonder whether we are implicitly rejecting our status as God’s creatures. 30 The dark side of the human story—from the Edenic temptation to the execution of God incarnate for sedition and blasphemy—is that we do not want to be God’s creatures but instead want to take God’s place. We readily embrace inflated conceptions of what we are, just as we are all too ready to justify ourselves, to regard ourselves as in the right, vindicated in light of norms we endorse, relying on our own propriety, rather than on the grace of God. None of this implies that any particular conception of human nature is opposed to the Christian faith, but it is cautionary for creatures that have inveterate tendencies to self-congratulation and resistance to being mere creatures. Rather than encouraging natural propensities to pay ourselves metaphysical compliments, receptivity to evidence deflates an inflated selfimage. When scientific inquiry undermines pretensions to being godlike, and reveals human minds as part and parcel of the natural world, not ideally rational intellects contingently tied to bodies, but material minds shaped by the contingencies of a particular evolutionary history, Christians have no call to resist. On the contrary, we have good reason to welcome the therapeutic incursion of science into our self-conception. When we dispense with the once, but no longer, reasonable philosophical assumption that God could not have created material persons, the ancient affinity for the dualist conception of human nature becomes groundless and the scientific account of what we are looks increasingly likely from the perspective of Christian faith, even if it affronts an image of ourselves long held dear. THE MIND NATURALIZED We are not ideally rational Cartesian souls, constrained only by creaturely finitude, always able to avoid error, if not ignorance, always capable of critical self-reflection. Scientific inquiry has begun to show us what it means for there to be no transcendent “I” that stands beyond, operating the neural machinery. We are that machinery, and that machinery was assembled not for truth and reason as such, but for disseminating the
Mind
141
genes that construct it and the body that houses it. We manifest not only the limitations necessary for all physically realized cognition—those to which J. R. Lucas drew attention—but all manner of contingent, idiosyncratic constraints due to the particularities of our evolutionary history and biological constitution. 31 Ways of thinking, feeling and acting that seem so natural to us as to require no explanation are the products of an evolutionary trajectory that could have gone differently. Human reasoning is shaped by unconscious factors in ways not obvious to casual observation, nor accessible to introspection, but which scientific scrutiny brings to light. Our most sophisticated reasoning, be it theological, philosophical, or scientific, is a matter of finding new uses for tools evolved for other uses. Much of our reasoning is the operation of specialized information processing modules, crafted by natural selection, not for the godlike contemplation of eternal truths, but for the “fast and frugal” thinking that made it possible to avoid predators, find food, and impress prospective mates on the African savanna in the upper Pleistocene. Minds so made are not ideally equipped to abstract the truth of the world from the exiguous, ambiguous input of sensory data. They bring to the world a host of innate assumptions about how things are, “hard-wired” into our neural circuitry. When these assumptions are not fulfilled, reasoning fails. Rather than embodying perfect finite rationality, brains so constructed often employ “satisficing” heuristic strategies; we are evolved to arrive at “good enough” solutions quickly. Because of this we are vulnerable to various “cognitive illusions;” we confidently think in ways that from an objective, i.e., scientifically constrained, perspective are less than rational. Equipped with Stone Age minds—evolutionary change lags far behind social and cultural innovation—we are prone to behave in ways that stopped making sense when our ancestors left the hunter-gathering life for agriculture, 10,000 years ago. 32 Further, the cognitive architecture of the mind-brain exhibits a significant degree of encapsulation: information available to some information processing modules is not available to others. The idea of a fully unified self, capable of complete self-knowledge, is an illusion. The self is real, but more likely a kind of spokesperson generated by the parts of the brain that are linguistically articulate and socially interact with other selves. Coherence of cognition and character is the result of effort, reliant on cultural scaffolding and always incomplete. Not having complete access to everything that goes on in the mind, the self will engage in spectacular feats of confabulation as it seeks to make sense of what it experiences. We have less self-awareness than we suppose; we are certain we have introspective knowledge of matters about which we are simply mistaken. 33 This does not imply that we are not really rational. It reveals an imperfect rationality possible for a certain kind of creature, an animal to which Darwinian evolution has bequeathed cognitive powers that are in ways
142
Chapter 5
that matter to us far superior even to those of our primate relatives. Human beings are rational animals, but this does not mean we are hybrids of non-rational matter and rational souls. To understand our minds we need not imagine ourselves transcending the mundane material world. We have synaptic interconnections that embody, in rough and ready ways, the norms of rational thought, and thus capable of reasoning and being moved to act with rational purpose. Our thinking and feeling is the operation of neural circuitry extremely similar to that of our primate relatives; small differences in organization, most likely the result of a relatively small number of regulatory genes that slow fetal development, result in the differences between a chimpanzee and a human being, an animal that is, like its Creator, a person. But a person of a profoundly different kind. What we know of ourselves should not tempt us to conclude that our minds are very much like the mind of God. We are material, mechanical images of a mind of a radically different order of reality. Our fellowship with God, our role as the divine image, depends not on our being much like God, but on God’s choice to reach down to us, no matter how great the distance. Psychologist Gary Marcus, in his book Kludge: The Haphazard Construction of the Human Mind, asserts, “If humankind were the product of some intelligent, compassionate designer, our thoughts would be rational, our logic impeccable. Our memory would be robust, our recollections reliable. Our sentences would be crisp, our words precise, our languages systematic and regular.” 34 For Marcus, as for many others, the reality of our material, evolved minds indicates that we are not creatures of a designing God. This, we should agree, is true, but not because we are not God’s beloved creatures. What we find when we examine the human mind is what we should expect if, as the Christian faith confesses, we were created by a God who calls us to share in the Triune life and thus eschews design. THE GHOST OF THE GHOST IN THE MACHINE Gilbert Ryle, in his influential The Concept of Mind, derided mind-body dualism as belief in the “ghost in the machine.” Although the analytical behaviorism he espoused is long since abandoned, Ryle’s 1949 book marks the beginning of the end for dualism in the academic world. Today, despite the continuing commitment to dualist ideas by most persons of faith, relatively few scientists, philosophers, or even theologians will readily admit to believing in an immaterial mind or soul. However, many are uncomfortable with a robust naturalism, and seek to mitigate the impact of materialism on our traditional conception of a mind that transcends the physical world. The hope is that we can accept what science tells us, and not deny its materialist implications, but find loopholes
Mind
143
that make it possible to avoid the stark conclusion that we are “nothing but” an arrangement of quarks and electrons. One manifestation of this on the contemporary scene is the popularity of non-reductive physicalism. A generation ago, many materialists envisioned the reducibility of the mental to the physical, in the sense of an identification of mental categories with neurophysiological categories. Pain, for example, would be identified as a certain type of brain state. As it turned out, complex reality did not cooperate with the simple theory, and this, along with other reductionist projects, was a failure. Instead, the main current of physicalism has sought to identify mental states more abstractly, as states that can be realized in a variety of physical ways. States of mind are multiply realizable: Jay believes that cats are adorable in virtue of his having such and such neural properties, ones that only a human brain can have, but a Martian, whose brain is constructed on quite different principles, might have the same mental property, and thus share Jay’s conviction, although he cannot have the same neural properties. The prime candidates for such multiply realizable properties are defined in terms of function. Take, for instance, the property of being a mousetrap: this cannot be identified with any particular set of physical properties. My mousetrap is made of wood and steel configured in a certain way, but someone else might bring the world to her door by building a better one made of plastic and string. What makes something a mousetrap is what it does—catching mice—not what it is made of or how its parts are configured. There is no plausible definition of mousetrap in physical terms. There are indefinitely many ways of assembling physical things to make a mousetrap. Many materialists subscribe to functionalism, contending that what makes something a belief, or a hope, or a fear, or a pain is not its intrinsic nature but the role it plays in the overall economy of a mind. These roles are generally taken to be causal, e.g., a pain is a state characteristically caused by bodily damage and causing avoidance behavior; the belief that there is a cat nearby is characteristically caused by the presence of a cat and it characteristically causes dispositions to do things like say, “Here kitty, kitty!” Functionalism allows for the possibility that physical things that are otherwise extremely different, made of different kinds of stuff and assembled in different ways, are nonetheless minds, in virtue of realizing the same causal roles. Paradigmatically, computation is a matter of functional roles embodied in causal connections among physical states, with indifference to the physical nature of those states: the same software can run on a machine made of transistors, vacuum tubes, water running through pipes and valves, or teams of trained ferrets. The identification of mental states and physical, as opposed to functional, states is a problem for theists, since we ascribe mental states to God. Christians reasonably reject old-style reductionist theories of mind in favor of functionalist theories. 35 What is not altogether clear is why
144
Chapter 5
many Christian thinkers see the advent of non-reductive physicalism as making a vital difference to our conception of human nature. They suppose that while materialism as such does not conflict with the Christian faith, reductionist versions of materialism do. Reductionism excludes the properties something must have in order to be a genuine person and thus threaten the status of human beings as the image of God. In a classic statement of this view, Arthur Peacocke, in the course of defending a “Christian materialism,” writes, “Christian believers, affirming as they do the reality, dignity, and value of the human mind, are opposed to reductionism.” 36 The rejection of “reductionism” is more understandable if it is understood not as the idea that mental kinds are identical to physical kinds, but as eliminationism. Eliminative materialism is a response to the impossibility of reducing the mental to the physical. It holds that mental kinds are not identical to physical kinds, that pain, for example, cannot be a type of neural event and thus, at least for theoretically serious purposes, should be dispensed with, as not mapping onto what is really there and therefore unreal. Ironically, it is not the reducibility of some mental things to the brain, but their irreducibility that prompts some physicalists to deny their reality. 37 If a neurophysiological state exists, and a mental state is identical to it, then that mental state exists. Something cannot fail to exist in virtue of being something that does exist! Dualists, convinced that being immaterial is an essential feature of mind are entitled to regard the reduction of the mental to the physical as tantamount to denying the existence of the mental, but no physicalist, irrespective of her views on reduction, can believe this, and thus she cannot reasonably reject reductive physicalism on the ground that it denies the mind’s reality. Sometimes, the impetus to celebrate the demise of reductionism appears to arise not from concern for the existence of mental things, but for their causal powers. If reductionism is taken as equivalent to epiphenomenalism, then it is reasonable to reject it, if one's concern is to preserve the value and dignity of the mind. 38 For, while epiphenomenalism does not deny the existence of conscious experience, it accords it a secondary status in the economy of the mind. It denies its causal efficacy, asserting that the real causes of the body’s motions are neurophysiological events, but not those to which our beliefs, desires and choices are identical. Here too we see the rejection of something called reductionism motivated not by the implications of the reduction of the mental to the physical, but by the apparent implications of denying reducibility. If a type of neurophysiological state is causally efficacious, and a type of mental state is identical to it, then that mental state is causally efficacious. Something cannot lack causal efficacy in virtue of being the same thing as something that is causally efficacious. 39
Mind
145
What non-reductive materialism does imply is that mental categories do not map smoothly onto the categories referred to in the causal laws of the natural sciences. We cannot identify sensations of pain with a type of neural state. The mental language we employ to speak of sensations, desires, beliefs, and so on resists translation into the theoretical vocabulary of neuroscience. The impossibility of such reductions for some inspires hope that the human mind, although part of the physical world, is an ontologically quite special thing, one that can, in virtue of being irreducible, somehow be extricated from the causal order of nature and not caught in the net of natural scientific explanation. However, irreducibility is far too common to fulfill these hopes. Consider the mousetrap referred to earlier: because there are indefinitely many ways to build a mouse trap, the property of being a mousetrap cannot be reduced to the physical. The same goes for mountains, oceans, moons, genes, hearts, keys, pizzas, and anything else, natural or artificial, that is conceived in essentially functional terms and therefore realizable in indefinitely many ways. The human mind does not reduce to the physical, but this hardly makes it unique. Mental kinds are not identical to physical kinds, but this leaves open the possibility that a particular mind is identical to a particular physical thing, a particular mental state or event to a particular neural state or event. We cannot define mousetrap in physical terms, but the mousetrap under my bed is identical to a particular physical object. Likewise, even though, for example, belief cannot be defined as a particular neural state of a human brain, Jay’s believing at this moment that the cat is on the roof just is his brain being in a particular functional state. The failure of the classical reductionist project offers no comfort to those hoping that the human mind is special in some deep, metaphysically satisfying sense. Metaphysically, it reveals the mind’s superficiality, not its uniqueness. EMERGENCE “Non-reductive physicalism” is sometimes used to express the more ambitious idea that the mind is emergent in some sense that means that it is not fully explicable in terms of natural law. “Emergence” can mean a variety of things, some obviously true, but trivial, others interesting but, if not obviously false, at least without warrant. The challenge is to say something significant, i.e., to make emergence mean something beyond mere irreducibility, something that might mitigate the implications of materialism, and which we also have some reason to think is true. Some properties are emergent in the sense that they belong to wholes but not to their constituents. No molecule in a mug of beer has the property of liquidity, but the volume of beer as a whole has this property. Individual neurons are not conscious or rational, but these properties
146
Chapter 5
emerge from a large number of them properly interconnected inside a human body. Further, configuring large numbers of simple things in certain ways gives rise to properties that are novel; nothing in the world had them prior to these parts being assembled in this way. Once, for example, there were no living things in the world, but at some point a combination of non-living things was a living organism. In the past, the natural world contained no minds, although it contained all the elementary components that biological evolution eventually assembled into minds. The combination of parts into complex wholes gives rise to properties no one familiar only with the properties the parts exhibit in isolation, or in simpler configurations, would have predicted. A part of a complex whole will often behave in ways that can be made intelligible only by paying heed to the fact that it is part of that whole, interacting with other parts. The behavior of a neuron, for example, can be partially explained by the chemical gradients in its synaptic clefts, and the molecules impinging upon its semi-permeable membranes, etc., but an adequate explanation of what it is doing may require knowing, say, that it is part of the memory circuit in the hippocampus of a depressed human being. Things being emergent in these ways does not conflict with the most robust scientific naturalism, not even with classical reductionism’s identification of physical and mental kinds. Yet there are ideas of emergence that have implications that are not obvious and that are not plausible from the perspective of naturalism. Consider consciousness, a prime candidate for an emergent property: We suppose it is, with respect to the physical world, irreducible. However, we also know that configuring physical things in a certain way guarantees that the complex whole they comprise will be conscious. This physical state is sufficient for consciousness although it is not necessary for it. Quite different kinds of physical things, and—says the theistic physicalist—even a non-physical thing, can realize this property. Once we know that a certain kind of configuration has the emergent property, we will be in a position reliably to predict that this sort of configuration will be conscious. Even if initially our knowledge of the properties of the constituents would not have allowed us to make this prediction, once we know that assembling 100 billion densely interconnected neurons in the way they are assembled in a human brain brings a mind into existence, the emergence of consciousness becomes predictable. The crucial question is whether we can move beyond ex post facto predictability toward intelligibility. Can we explain why assembling parts in a particular way results in consciousness? Can we explain why configuring neurons the way they are configured in the brain produces a conscious mind? Can we come to understand this emergent property as a necessary manifestation of the underlying physical structure?
Mind
147
What can be labeled strong emergence—to distinguish it from uncontroversial versions—is the idea that no such explanation is possible. Assembling the non-conscious parts this way does, as a matter of fact, produce a conscious mind, but that it does so is a basic fact that admits of no further natural explanation. In contrast, the naturalistic assumption is that it is possible to explain why configuring physical things this way is one of the ways to make a mind, and that we can reasonably hope that scientific inquiry will eventually deliver that explanation. The scientific naturalist thinks that consciousness, as inscrutable as it may be now, will ultimately be explained in a way analogous to the way biological life was explained in terms of biochemistry, despite the fact that the properties necessary for life—homeostasis, metabolism, self-replication—cannot be reduced to any physical or chemical properties. These properties emerge in complex configurations of physical things, but that they do so is explicable by appeal to the forces, fields and particles known to physics and chemistry. What difference would strong emergence make to our conception of ourselves as persons? I suggest that the principal attraction is that it offers some of the consolations of dualism without requiring that we accept a theory of human nature so clearly at odds with the evidence. It blocks the scientific explanation of the human mind without casting the mind outside nature. Even if it is established that such and such a mental state emerges from such and such a brain state, there will be no explanation of why this is the case and thus no scientific understanding of the mind as part of the physical world. The mental properties in virtue of which brains are not just brains but minds are arbitrarily superimposed on the physical reality. This is illustrated by the fact that if consciousness were a strongly emergent property, there could be an exact microphysical duplicate of this world, indistinguishable from it except for its lack of consciousness. The bodies, brains, and behavior of its inhabitants would be exactly like ours, but they would belong to creatures completely devoid of consciousness. Rather than appealing to these “zombies” in support of dualism, as in the modal argument we saw earlier, their alleged possibility can be enlisted on behalf of strong emergence. If our world includes minds and that other world lacks them, but the explanation has nothing to do with the physics of either world, then minds are in that sense independent of anything physical. Since the physical facts have no explanatory role to play with respect to the emergent mental properties, the naturalistic program of integrating our understanding of the mind into the physical world is guaranteed to be frustrated. The objectification of the human person that science aims at, and which many fear, will be forever stymied. Science will never succeed in portraying us as one kind of thing among others in the natural world. The crucial facts that make us what we are are opaque to science: they are added only after science has had its say. In a way, the idea of the strong emergence of the mind is a
148
Chapter 5
replacement for old-fashioned, discredited dualism, a ghost of the ghost in the machine. Many persons of faith find this an agreeable implication. If there is no natural explanation of why mental properties being to some complex physical entities, there must be a supernatural explanation: God directly endows human brains with consciousness and thus personhood. Strong emergence provides a reason to conceive ourselves as in a way transcendent, forever eluding capture by scientific explanation, not mere things in the natural world but something more, something more like God. Christians, affirming that humans are the image of God, and thus personal beings, are entitled to be suspicious of theories that suggest we are less than persons, but we are no less obligated to doubt theories that manifest discomfort with our being mere creatures. It might turn out that there is such a phenomenon as strong emergence, but it is not at all obvious that it is something Christians have reason to expect. ON BEING ONTOLOGICALLY SUPERFICIAL For some, it might seem that I have too quickly separated the question of human nature from that of God’s existence, for they see the materialist account of human nature as having immediate implications about the fundamental nature of reality inconsistent with belief in God. There are two traditional conceptions of what is ultimately real, one theistic, the other atheistic. The mind-first view is that the ultimate, selfexistent reality on which all other things depend for their existence is mind. This in essence is the theistic view: the underlying reality is God, a personal being, a self-existent mind. All material things exist because God created them:
Mind does not come from matter; both created minds and mindless material realities owe their existence to an immaterial, mental reality. Since mind is utterly distinct from matter, a more complete picture por-
Mind
149
trays created mind and matter as created independently by the underlying divine mind: 40
In contrast, the matter-first view is that the ultimate, independently existing reality is matter, and minds exist only because matter has been arranged in certain ways; there is no uncreated mind. On this, the atheistic view, reality at bottom is mindless and impersonal:
However, once again, we are not forced to choose between two simple possibilities. Christians of course hold that the ultimate, self-existent reality on which all other things depend for their existence is God, an uncreated, immaterial mind. Reality, Christians confess, is ultimately personal. But this does not commit us to concluding that human minds are directly created by God. Instead, we can propose that God directly created matter which, when suitably arranged, gives rise to human minds:
150
Chapter 5
This incorporates the matter-first view that a suitable configuration of matter suffices for the existence of human minds, but it also embraces the fundamental mind-first claim that all material things, and all human minds, depend ultimately on God’s creative action. This account—mindfirst, then matter, then human minds—coheres with the scientific evidence and with the Christian conception of human persons as created by God. The secular scientific naturalist is committed to explaining the human mind in terms of the non-mental, i.e., in terms of the physical. The Christian naturalist endorses this enterprise, but in a qualified way, seeing the quest for naturalistic explanation of mind in the natural world as aiming for penultimate explanation. The ultimate explanation of why human minds exist is that God created them. This supernaturalized naturalism denies us the right to think of ourselves as among the basic constituents of God’s creation. We are not, as dualism maintains, among the things made directly by God. Instead, we are made of the things God made indirectly, assembled—by an indeterministic evolutionary process—of simpler, more basic, components. This is no reason to doubt our existence, even if we could do so. But it calls for ontological humility. We are things that exist only because other things have been brought together in certain patterns; we are here only because, and so long as, those patterns persist. Ironically, given the enthusiasm it generates, the failure of reductionist accounts of mind does not mitigate this. On the contrary, it highlights
Mind
151
our superficiality. The mental, the personal, cannot even be systematically identified with configurations of the world’s basic material constituents. There is a perspective from which, had we never existed, the catalogue of real things would have been no different; the catalog of the particles, fields and forces of physical science would be unchanged. What would be missing would only be some complex, improbable configurations of the basic constituents. We easily could have been absent: had the world’s highly contingent history differed in any number of ways, it would not have included us. The traditional human self-image, with its commitment to our minds transcending the material world, is innocent of this unsettling realization. This is, perhaps, most pronounced among Christians intent on contrasting the truth that we are persons, free moral beings, made in the image of God, with what they take to be the error of thinking that we are no more than atoms and molecules, assembled by chance and evolution. However, the confession that we are persons, called to be the image of the Creator, need not be a stratagem for denying our place in the shallows of being. Instead, it points to something entirely different in its account of what most matters about human beings: we are loved by the one who made the worlds. SHALLOW MEANING Possibly, nothing more vividly illustrates human ontological superficiality than the Chinese Room argument. John Searle’s famous argument is a frontal assault on the computational theory of mind. It is directed at intentionality, the property of mental states in virtue of which they are about something; they have content. I have an idea of Polly the cat, and this is an intentional state; there is something it is about: Polly. One cannot simply have an idea; it must be an idea of something. Most materialists believe that the principal route to an explanation of human cognition, our capacity to have meaningful thoughts, lies in seeing it as computation. Brains are minds in virtue of their computational powers. They are biologically evolved information processing organs. A brain receives input from the sense organs, forms mental representations of features of its natural and social environment, manipulates those representations in systematic ways so as to produce output, i.e., efferent nervous signals to the muscles that control the motion of the body in ways that, if all goes well, enable it successfully to navigate the world it inhabits. Mental states represent—they mean something—in virtue of the roles they play in this evolved causal network. Mental states have intentionality—they mean something—not in virtue of any intrinsic meaningfulness, but in virtue of their place in a pattern of cause and effect. Searle’s argument challenges the computational theory. He contends that computation cannot possibly give rise to intentionality. An imprecise
152
Chapter 5
but useful characterization of computation is that it is the manipulation of symbols in accord with formal rules so as to realize an input-output function. Searle’s Chinese Room thought experiment is designed to show that computation cannot amount to genuine understanding. 41 If our minds were, at bottom, computers, we could not have thoughts that are actually about anything. Searle envisions a room with two slots, one for input, one for output. Questions written in Chinese ideograms are put in. From the other slot reasonable answers appear, also written in Chinese. It looks as though someone inside understands Chinese, but in Searle’s story, this is an illusion. For what’s inside is you, and you don’t understand a word of Chinese. The room contains a rule book, written in English, and it instructs you to match incoming symbols with other symbols which you then send through the output slot. You have no idea what the incoming or outgoing ideograms mean. To follow the rules, you need pay heed only to what they look like, to their form, their shape. What goes on inside the room is entirely syntactic, not semantic; it is about the form of the symbols, not their content, not what they mean. The extraordinarily complex and clever rules, and your swift but careful execution of them, generate the illusion that someone inside understands Chinese. Having secured acknowledgement that no matter how adept at following the rules the person in the Chinese Room is, she does not understand Chinese, Searle points out that nothing essential changes when the rule-following human being is replaced with a computer. As it executes its program, one physical state, say a pattern of high and low voltage states in a particular bank of transistors, causes another state to occur, a different pattern of electronic activations. These transitions occur in accord with the program, formal rules embodied in the hardware. This cause and effect relation depends entirely on the physical properties of these electronic “symbols,” irrespective of what any human being uses them to mean. In themselves, they mean nothing. They are symbols only because someone who already understands uses them to mean something. Their intentionality is derived, not original. Making the computer, or its program, more complex cannot endow the manipulated symbols with meaning. Nor does it matter what the computer is made of: meaning cannot arise from the rule-governed manipulation of symbols whether they are patterns realized in transistors, neural interconnections, or anything else. We cannot, Searle concludes, get from syntax to semantics. Meaningful thought cannot be computation. We can endow inherently meaningless things, like the states of a computer, with derived meaning, using them to represent things, but we must first possess the real thing, original intentionality. Since Searle launched his argument in 1980, computationalists have responded in various ways. Perhaps the best response is that Searle’s claim, so far as it goes, is true, but that it has no bearing on the computational theory of mind. A physical state of a computer, whether artificial or
Mind
153
biological, means nothing . . . so long as it lacks the requisite causal connections to the world. Using the computer on which I am writing this, I type the word “cat,” creating a symbol, a particular array of high and low voltage states in transistors, but that symbol possesses only derived, not original intentionality. That physical thing means cat only insofar as I—with the help of several intervening levels of software—use it to refer to cats. Running some other program, the same pattern might play a completely different role; inherently, it means nothing. However, give a computer control of a cat-hunting robot and things change. The robot has a camera, connected to the computer by way of a transducer, so the presence of a cat in the vicinity reliably causes a particular pattern of low and high voltages in some transistors. In turn, these interact with other internal states of the computer and are ultimately connected by various actuators to other parts of the robot—its wheels and weapons. Thus, it causes the robot to advance toward the cat it has detected, take aim, and so on. The activation pattern, embodied in the onboard computer, acquires the causal role of representing the cat. If we sought to explain the behavior of such a device, it would be natural and reasonable to identify that internal computational state as a representation of the cat. This is not to say that the machine understands; presumably, it does not. But it does not understand because it does not possess sufficient intelligence, or consciousness, not because there is nothing going on in it but computation. 42 This response is, I believe, correct, but it amounts to little more than a reiteration of the fundamental computational assumption Searle rejects. Computationalists acknowledge that there is nothing inherent in the physical things that computation involves, whether they are slips of paper in the Chinese Room, the electronic states of a digital computer, or the neural states of a human brain, that makes them mean something. Their semantic properties are derivative. In other contexts, playing different causal roles, they might mean something else, or mean nothing. If thought is computation, then what is obvious in the case of language—the word “cat” refers to cats not in virtue of any intrinsic characteristic of this physical item, but because we use it to represent cats—is no less true of thought—my idea of cats is an idea, and it is the idea of cats, not because of what it is, but because of its complex causal relations to other things going on in my brain, to other speakers of English, and to the rest of the world, particularly to the cats it contains. Whatever goes on in my brain that represents cats, it is just a thing used to mean something. But of course it is not me who endows it with its derived intentionality. That would launch an infinite regress. This bit of the natural world has been recruited to mean something not by anyone at all, but by the mechanisms of natural selection, as it crafted creatures whose complex nervous systems realized sensory input-behavioral output functions that led to inclusive fitness. Not all meaning can be derived; some must be original,
154
Chapter 5
but there is no plausible place in the natural world for intrinsic meaning. Nothing in our heads necessarily, in virtue of its inherent “aboutness,” connects to the world. All such connections derive from the mental state’s role in an evolved pattern of causal relations. Searle’s story does not show that understanding is not computation, because there is no route from mere syntax to semantics; what it shows is, as Daniel Dennett put it, “brains are syntactic engines that can mimic the competence of semantic engines.” 43 DEEPER INTO SUPERFICIALITY Roger Scruton reminds us “we belong to the surface of the world, and apply to it the classifications which inform and permit our actions”; for human beings, “the meaning of the world is enshrined in conceptions that science does not endorse . . . conceptions that grow in the thin topsoil of human discourse” 44 He cautions, “our perspective on the world is not sovereign, but a by-product of the evolutionary process which created us. Its authority stands always to be usurped by the imperial ambitions of scientific theory.” 45 As constituents of the world, we are superficial, neither fundamental constituents nor reducible to them. The same is true of the world as we represent it. Human beings evolved to have true beliefs, mental representations that accurately represent the natural and social environment, to whatever extent that mattered to reproductive fitness. Whatever does not bear upon fitness is up for grabs. The mind’s representational scheme can simply ignore what is not relevant to fitness and, when it represents what is relevant, what natural selection devises need not closely correspond to what is really there. Our representational scheme is not a window through which we clearly see the world’s true nature. The world as we naturally represent it, the world of human experience, is superficial. Its kinds and categories do not map smoothly onto those with which nature’s laws can be formulated. Our minds represent things as having characteristics they do not, or even cannot, have. What seems to belong to reality itself may in fact belong to us, projected onto the world, not found in it. The “manifest image” of the world in various respects amounts to a useful, but superficial simplifying gloss on the objective reality science painstakingly and only partially reveals. 46 The theory of color vision is a salient example. When I look out my window on a June day, I see lots of green: leaf-laden trees surrounding a parking lot and, beyond, grassy playing fields. It looks as though green is out there, adhering to the surfaces of objects, to the leaves and the blades of grass, but we know instead that physical objects are, in one obvious and literal sense, colorless. Objects reflect electromagnetic radiation at various frequencies in virtue of the invisible microscopic structure of
Mind
155
their surfaces, and the reflected photons impinge on retinal cells, setting in motion the neural computation that culminates in color vision. Neither their constituent atoms and molecules, nor the packets of photons that bounce off them, are green in the way the leaves and grass appear to be green. Early modern philosophers distinguished primary qualities—properties of the objects themselves, such as their mass and shape—from secondary qualities—the properties they cause in us when they act upon our sense organs. Objects are green in the sense that they cause minds like mine to have a particular kind of sensation. But this does not locate green in the mind, rather than in the world. For it cannot be there either. When I look out the window I have an experience, a sensation of green things, but clearly that is not green. My sensations are events in my brain, but inspection of it would reveal nothing green. (Or, if it did, this would be a sign of something gone wrong, not of my having a normal visual sensation of grass and leaves.) My brain is as much part of the natural world as the leaves and grass and it cannot be green in any way that they cannot be green. Were my materialism mistaken, and the mind that senses those things an immaterial substance, this would not save the green. Whatever properties an immaterial mind might possess, being green is not one of them. The inescapable implication is that reality has no place for green. 47 The mind represents things as being green, but nothing strictly has this property. Nothing could have it. It looks as though a physically heterogeneous array of objects all share a simple intrinsic property which I readily detect when, in fact, they do not. A green car is parked under the trees. The chemical composition of its paint and of the trees’ leaves have little in common. It just so happens that in the circumstances the light they reflect interacts with my eyes, and thereby initiates multiple levels of perceptual information processing in my brain, to cause the same kind of mental representation: I represent both the car and the leaves as having that property; both look green. If, as we assume, color vision is an adaptation, natural selection tracked some feature of objective reality and equipped us to detect it. There must have been something to which differential response made for a difference in fitness. However, that feature need not be and, in the case of color vision, is not, overt, not something as obvious as objects causing us to represent them as green if, and only if, they reflect light at a given frequency. The physical commonalities creatures equipped with color vision detect when they represent things as being green are subtle, complex, and relational. 48 But natural selection economized and invented, constructing minds that represent them as possessing an obvious, simple, and intrinsic property: we represent them as green, no matter that this is not a property things could really have, except in the sense that they can cause us to represent them as having that impossible property. 49 In this latter sense it is true that the grass, the leaves, and the car are green. Yet these truths are framed with concepts that belong not to the world independent of us, but to our representation
156
Chapter 5
of it, and it is inevitably inaccurate representation. Teaching us that a feature of our experienced world as salient and ubiquitous as color is a kind of illusion, science disillusions us of pretensions about our place in the natural order. 50 Whether color sensations, meaning, knowledge, or rationality, the traditional human self-image ascribes to us properties material creatures, particularly creatures produced by biological evolution, cannot unambiguously possess. We know, but our knowledge is always the product of fallible mental mechanisms; it never amounts to unmitigated certitude. We reason, but our reasoning is the operation of natural computational mechanisms we cannot ultimately transcend: we are not infinitely selfreflective. We think meaningful thoughts, but their meaningfulness is of the kind of which an evolved material mechanism is capable: our minds harbor no intrinsic intentionality. Later, in chapter 6, we will see that the same is true of our conception of ourselves as free and responsible. The recurrent lesson is that we possess these characteristics only in an analogical way, or as an approximation to, or perhaps as a kind of simulation of the thing we hazily envision but which is not what we are and, to all appearances, not a physical possibility at all. For some, this result is destructive, amounting to the conclusion that human beings do not really know, cannot really reason, and cannot really think meaningful thoughts. 51 However, we can more reasonably conclude not that we lack these capacities, but that our initial conceptions of them were idealizations from which the reality we are diverges. Taking up the naturalizing project of contemporary science we seek to ascertain the ways such humanly significant things fit into the objective world known to science. Those of us who confess that humankind is made imago Dei can go further, and see in this ways to flesh out what it might mean to be the material image of the immaterial God. WHAT THE SKIN HORSE SAID In Margery Williams’s classic story for children, The Velveteen Rabbit, the rabbit, a stuffed toy, approaches the skin horse, the oldest and wisest of the nursery’s inhabitants, to ask “What is REAL?” “Does it mean having things that buzz inside you and a stick-out handle?” The skin horse explains, “Real isn’t how you’re made. . . . It’s a thing that happens to you. When a child loves you for a long, long time, not just to play with, but REALLY loves you, then you become real.” 52 Like the rabbit in the sentimental story, we easily imagine that being “real,” being a person, is a matter of being made a certain way, being a certain kind of thing. We fear that if we are nothing but these fragile, constrained bodies, not mysteriously different than other things in the material world, then we might not be special enough to matter to God or even to ourselves. We feel the
Mind
157
threat of a science that erodes ancient, cherished ideas of how, being persons, we must be made. That venerable image is increasingly precarious. The illusion that if we are not made that way, then we cannot be privileged among this world’s creatures, convinces many—not only persons of faith—that we can defend the reality and worth of human personhood only by rescuing humanity from the clutches of scientific explanation and preserving the now implausible idea that we are something other than material beings. Failing that, we seek surrogates in irreducibility or emergence. Christian faith has no stake in this project, for it has no legitimate affinity for the desire to make ourselves mysterious, to see ourselves as transcending the world of mere objects. Faith can warmly welcome the scientific portrait of human beings as physical persons, trustfully acknowledging our calling as the material image of the transcendent Creator. Today, despite the preponderance of reasons, including theological reasons, to reject dualism and acknowledge that human persons are material things, many Christian thinkers remain committed to it. For them, dualism remains an indispensable presupposition of the essential Christian confession that God’s love for us is not defeated by death, and that God will resurrect human beings, as he resurrected Jesus. Materialism implies that a human person is a functioning human body. At death, the body becomes permanently incapable of higher mental functioning and the person no longer exists. Sooner or later after death that body will also cease to exist; its constituent atoms and molecules going their separate ways. How, in the absence of an immaterial mind or soul that continues to exist after the body’s demise, is the Christian claim that there is life after death even possible? This, the most significant objection to materialism, I defer to chapter 9, where I take it up along with other eschatological matters. There I will contend that the naturalistic account of human nature supports a conception of the identity of persons through time consistent with the possibility of resurrection. Before that, the intervening chapters turn to the freedom, morality, and religiosity of the human person conceived of as the material image of God. NOTES 1. Daniel Dennett. “Foreword” to Ruth Garrett Millikan, Language, Thought, and Other Biological Categories: New Foundations for Realism (Cambridge, MA: MIT Press, 1984), ix. 2. See, e.g., Owen Flanagan, The Problem of the Soul: Two Visions of Mind and How to Reconcile Them (New York: Basic Books, 2002). 3. A further analogy is the widespread view of divine intervention cited in chapter 4: either God intervened to create the human species, or he does not intervene at all. 4. I believe that this kind of example originates with Peter van Inwagen, Metaphysics (Westview Press, 3rd ed., 2008), 247–48.
158
Chapter 5
5. For an account of this railway worker whose brain was damaged when an explosion sent an iron rod through his head, see Michael S. Gazzaniga, Human: The Science Behind What Makes Us Unique (New York: Harper Collins eBooks, 2009), 119–20. 6. For an examination of the brain’s role in moral behavior, see Patricia Churchland, Braintrust: What Neuroscience Tells Us About Morality (Princeton, NJ: Princeton University Press, 2011) and her Conscience: The Origins of Moral Intuition (New York: W. W. Norton and Company, 2019). 7. Christian theists, dualists or not, believing that God acts miraculously, intervening in the natural world, certainly have no reason to accept it. 8. See David Papineau, Thinking About Consciousness (Oxford, UK: Oxford University Press, 2002), 253–55. 9. Jerry Fodor, “The Big Idea: Can There Be A Science of Mind?” Times Literary Supplement, 3 July 1992, 5. 10. Fodor, “The Big Idea,” 5. 11. Imagine disembodied minds, having no idea how to explain consciousness, concluding that it must be the property of some material thing, and then responding to the objection that there is no evidence for the existence of such a thing by pointing out that absence of evidence is not evidence of absence. 12. Thomas Nagel, “What Is It Like to be a Bat?” in his Mortal Questions (Cambridge: Cambridge University Press, 1979), 165–80. 13. For accounts sympathetic to this theory of mind while sensitive to its possible limitations see Tim Crane, The Mechanical Mind: A Philosophical Introduction to Minds, Machines, and Mental Representations (London: Routledge, 2nd ed. 2003), and Andy Clark, Mindware: An Introduction to the Philosophy of Cognitive Science (Oxford, UK: Oxford University Press, 2nd ed. 2013). 14. For an in-depth treatment of the issue see Joel B. Green, Body, Soul, and Human Life: The Nature of Humanity in the Bible (Grand Rapids, MI: Baker Academic, 2008). Warren S. Brown and Brad D. Strawn insightfully examine the problematic practical effects of mind (soul)-body dualism on Christianity: The Physical Nature of Christian Life: Neuroscience, Psychology and the Church (Cambridge, UK: Cambridge University Press, 2012). 15. Hans Wolter Wolff, Anthropology of the Old Testament, Margaret Kohl, trans. (Mifflintown, PA: Sigler Press, 1996), 7–25. 16. It appears that dualism permeates thought and language in virtue of a quirk of the brain’s cognitive architecture that makes it hard for it to believe that it is a brain, and not immaterial. On the innate plausibility of dualism explained materialistically see Paul Bloom, Descartes’ Baby: How the Science of Child Development Explains What Makes Us Human (New York: Basic Books, 2004), 209–27. 17. The same is true for the facts of history and ethics. 18. No more than the fact that ancient Hebrew writers often ascribe mental properties to bodily organs, e.g., the heart, justifies claiming that the Bible teaches materialism. See G. C. Berkouwer, Man: The Image of God, translated by Derk W. Jellema (Grand Rapids, MI: Eerdmans, 1962), 200–201. 19. It would be presumptuous to assume that if the revelation was perfectly accurate about these matters it would be something we could understand. 20. The historical antecedent is an argument Descartes advanced in his 1641 Meditations on First Philosophy, Donald Cress, trans. (Indianapolis, IN. 3rd edition, 1993) to the effect that since he could not doubt his own existence, but could doubt the existence of his body, he could not be his body: he could not both have, and not have, the property of being dubitable. The modern revival of this general type of argument, but not subject to the obviously fatal objections Descartes’s argument faces, is due to Saul Kripke, Naming and Necessity (Princeton, NJ: Princeton University Press, 1980). 21. See, e.g., David Papineau, Thinking About Consciousness (Oxford, UK: Oxford University Press, 2002) and John Perry, Knowledge, Possibility, and Consciousness (Cambridge, MA: MIT Press, 2001).
Mind
159
22. Miracles: A Preliminary Study (New York: Macmillan, 1960), 22. Lewis takes the Haldane quotation from Possible Worlds and Other Papers (New York: Harpers, 1928), 209. The argument in the later, 1960, version of Miracles differs from the 1947 version in ways that suggest Lewis saw the force of objections made by Elizabeth Anscombe at a meeting of the Oxford Socratic Club in 1948; see Humphrey Carpenter, The Inklings: C. S. Lewis, J. R. R. Tolkien, Charles Williams, and their Friends (Boston: Houghton Mifflin, 1979), 216–17. It might be that this episode helped turn Lewis’s focus away from philosophical theology and to his splendid fiction writing; if so, we owe Anscombe our gratitude. We can see Alvin Plantinga’s argument that belief in our evolutionary origins is self-refuting, discussed in the preceding chapter, as a descendent of Lewis’s argument. This kind of argument is attractive not only to theists. A version appears, e.g., in Thomas Nagel’s Mind and Cosmos (Oxford, UK: Oxford University Press, 2012), 26–28. 23. For an instance of its ongoing popularity see Victor Reppert, C. S. Lewis’s Dangerous Idea: In Defense of the Reason Argument (Downers Grove, IL: InterVarsity Press, 2003). 24. In chapter 8, we examine the cognitive theory of religion, which gives rise to a “debunking argument” that needs to be taken seriously. 25. See Charles Petzold, The Annotated Turing: A Guided Tour Through Alan Turing's Historic Paper on Computability and the Turing Machine (Indianapolis, IN: John Wiley & Sons, 2008). 26. What matters here is not what computers that currently exist cannot do; there are many things they obviously cannot now do. What matters would be cognitive tasks provably impossible for any computer, including products of any technology, no matter how advanced. 27. At the heart of Gödel’s reasoning is a very clever scheme for coding mathematical statements so they can refer to themselves and say things equivalent to “I am not provable.” When we contemplate such a statement, we realize that it must be true. If it were false then mathematical reasoning could produce a contradiction, since that would mean that it is provable. Mathematics is logically consistent, so there are mathematical truths that are not mathematically provable, which is to say that mathematics is incomplete. The proof is available in Kurt Gödel, On Formally Undecidable Propositions of Principia Mathematica and Related Systems, B. Meltzer, trans. New York: Dover Publications, 1992). 28. J. R. Lucas, “Minds, Machines, and Gödel” Philosophy 36 (1961) 112–27. 29. B. F. Skinner, Beyond Freedom and Dignity (New York: Knopf, 1971). A contemporary argument that this is what science reveals is Alex Rosenberg’s The Atheist’s Guide to Reality: Enjoying Life without Illusions (New York: W. W. Norton and Company, 2012). 30. Harry R. Boer, An Ember Still Glowing: Humankind as the Image of God (Grand Rapids: Eerdmans, 1990), 1. 31. See, especially, Steven Pinker, How the Mind Works (New York: W. W. Norton and Company, 1997) and The Blank Slate: The Modern Denial of Human Nature (New York: Viking, 2002), Jerome H. Barkow, Leda Cosmides, and John Tooby, eds. The Adapted Mind: Evolutionary Psychology and the Generation of Culture (Oxford, UK: Oxford University Press, 1992), and David Buss, Evolutionary Psychology: The New Science of the Mind. (New York: Pearson Education, 4th rev. ed. 2013). 32. See, e.g., Daniel Kahneman, Thinking, Fast and Slow (New York: Farrar, Straus and Giroux, 2010), Daniel Kahneman, Paul Slovic, and Amos Tversky, eds, Judgment under Uncertainty: Heuristics and Risks (Cambridge, UK: Cambridge University Press, 1982), and Gerd Gigerenzer, Rationality for Mortals: How People Cope with Uncertainty (Oxford, UK: Oxford University Press, 2010). 33. See, e.g., Dan Ariely, Predictably Irrational: The Hidden Forces that Shape Our Decisions (New York: Harper, rev. ed. 2010), Timothy D. Wilson, Strangers to Ourselves: Discovering the Adaptive Unconscious (Cambridge, MA: Harvard University Press, 2004), Daniel M. Wegner, The Illusion of Conscious Will (Cambridge, MA: MIT Press,
160
Chapter 5
2002), George Ainslie, Breakdown of Will (Cambridge, UK: Cambridge University Press, 2001), and Nick Chater, The Mind is Flat: The Remarkable Shallowness of the Improvising Brain (New Haven, CT: Yale University Press, 2018). For the difficulty of extricating genuine reasoning from rationalization see The Enigma of Reason, by Hugo Mercier and Dan Sperber (Cambridge, MA: Harvard University Press, 2017). 34. Gary Marcus, Kludge: The Haphazard Construction of the Human Mind (Boston: Houghton Mifflin, 2008), 1. 35. The divine mind is not a human brain, so when we ascribe mental properties to God, e.g., “God loves us,” we do not refer to some state of a brain. Further, theists who adhere to functionalism do not suppose that God’s mental states relate to one another in ways governed by natural, causal laws. 36. Arthur Peacocke, “A Christian ‘Materialism?’” in How We Know, Michael Shafto, ed. (San Francisco: Harper and Row, 1985), 149. For the view that the abandonment of reductionist theories of mind is something that should be of particular good news for persons of faith see the essays in Warren S. Brown, Nancey Murphy, and H. Newton Malony, eds., Whatever Happened to the Soul? Scientific and Theological Portraits of Human Nature (Minneapolis: Fortress Press, 1998). 37. For the irrelevance of non-reductive physicalism in this regard, see Donald Wacome “Reductionism’s Demise: Cold Comfort,” Zygon 39(2004), 321–37. 38. A thoroughgoing epiphenomenalism about consciousness is infeasible: if there are conscious states and we directly know them, they must have some effect on us. The reasonable concern is that conscious acts of choice do not actually cause action; see Benjamin Libet, Mind Time: The Temporal Factor of Consciousness (Cambridge, MA: Harvard University Press, 2004). 39. Having dispensed with the reduction of mental types to physical types, there remains the question whether token-identity, i.e., the identification of any given mental event with a particular physical event, allows for the causal efficacy of the mental. The causal exclusion argument challenges this. On plausible assumptions, if mental types are not identical to neural types, then the mental as such causes nothing. See Jaegwon Kim, Physicalism, Or Something Near Enough (Princeton, NJ: Princeton University Press, 2004), 15–22. This kind of epiphenomenalism is, I believe, innocuous: the fact that a neural state instantiates a mental property has nothing directly to do with its causal powers. Nonetheless, the physical system of which that neural state is a component exists only because nature has selected for an organism that has such mental states. Here we see not that the mind is epiphenomenal, but that it is a superficial aspect of physical reality. But such superficialities play significant roles in causal explanation. For argument along these lines see Karen Neander, The Mark of the Mental: In Defense of Informational Teleosemantics (Cambridge, MA: MIT Press, 2017), 59–60. 40. The classic articulation of this is found in John Locke’s An Essay Concerning Human Understanding, revised edition, edited by Peter H. Nidditch (Oxford, UK: Clarendon, 1979), Book IV, Ch. X, Sec. 10. 41. John Searle, Minds, Brains and Science (Cambridge, MA: Harvard University Press, 1983), 31–34. 42. Someone might contend that the intentionality of the cat seeking machine is derived, because it is just being used, by whoever designed and built it, to represent cats. But we know that such devices are not always products of design, e.g., a dog, a product of natural (along with some artificial) selection, mentally represents the cats it pursues. Mental canine states mean feline present. Robust human meaning is no more intrinsic than primitive mental representation in animals or current machines. It arises out of social interaction, the arena of commitments and entitlements, and thus of rules and norms, which speaking creatures share. This inferentialist account of meaning focuses on the “space of reasons,” not on natural causation, but thought and language are not self-contained. The normative realm of meaningful thought and talk is tied to perception and action, and thus finds its explanation in the causal, evolutionary structures natural science delineates. See Joseph Rouse, “Naturecultural Inferentialism” in
Mind
161
From Rules to Meanings: New Essays on Inferentialism, Ondřej Beran and, Vojtěch Kolman, Ladislav Koreň, eds. (New York: Routledge, 2018), 239–48. 43. Daniel Dennett, “Self-Portrait,” in Daniel Dennett, Brainchildren: Essays on Designing Minds (Cambridge, MA: MIT Press, 1998), 357. 44. Roger Scruton, Modern Philosophy: An Introduction and Survey (New York: Penguin Books, 1994), 240–41. 45. Scruton, Modern Philosophy, 208. 46. The term “manifest image,” contrasted with “scientific image,” is due to Wilfrid Sellars, “Philosophy and the Scientific Image of Man,” in Wilfrid Sellars, Science, Perception and Reality (London: Routledge and Kegan Paul, 1963), 1–40. 47. This is hard to believe about a simple property like being green, but we know that we can have visual experiences of complex properties that nothing could possibly have, e.g., looking at some of M. C. Escher’s drawings, one has a visual representation of physical objects that could not possibly exist. 48. See, e.g., Paul Churchland, “On the Reality (and Diversity) of Objective Colors: How Color-Qualia Space is a Map of Reflectance-Profile Space,” in Paul Churchland, Neurophilosophy at Work (Cambridge: Cambridge University Press, 2007), 198–231. 49. See Frank Jackson, “Mind and Illusion,” in Anthony O’Hear, ed. Minds and Persons (Cambridge, UK: Cambridge University Press, 2003), 261–63. 50. With this in view, the hard problem of consciousness might not be so hard after all. Natural selection has constructed minds that represent things that have properties they possess only in the sense that we project them upon them. Perhaps what we understand as consciousness, a property we spontaneously ascribe to ourselves and to other things, is mysterious just because it is a property no material being could literally have. For a theory of this type, see Michael Graziano, Consciousness and the Social Brain (Oxford, UK: Oxford University Press, 2013) and Rethinking Consciousness: A Scientific Theory of Subjective Experience (New York: W. W. Norton and Company, 2019), Daniel Dennett, From Bacteria to Bach and Back: The Evolution of Minds, (New York: W. W. Norton and Company, 2017), 335–70, Daniel M. Wegner and Kurt Gray, The Mind Club: Who Thinks, What Feels, and Why It Matters (New York: Viking Press, 2016), and the essays in Keith Frankish, ed. Illusionism as a Theory of Consciousness (Exeter, UK: Academic Imprint, 2017). 51. A recent example of this view is Alex Rosenberg’s The Atheist’s Guide to Reality: Enjoying Life without Illusions (New York: W. W. Norton and Company, 2007) and How History Gets Things Wrong: The Neuroscience of Our Addiction to Stories (Cambridge, MA: MIT Press, 2018). 52. Margery Williams, The Velveteen Rabbit: Or How Toys Become Real (New York: Avon Books, 1975), 16–17.
SIX Freedom
A TWO-FOLD CHALLENGE Belief in human freedom and responsibility, like belief in miracles, needs to be defended on two fronts: is it consistent with scientific naturalism, and is it consistent with God as our Creator? Human beings are the material image of the immaterial God. Like the God we image, we are persons, and thus free, morally responsible agents. Unlike the God who made us, we do not transcend the physical universe. Human thought, including all deliberation and choice, occurs in the neural machinery of our brains and is governed by nature’s causal laws. Any decision a human makes, trivial or momentous, reasonable or foolish, good or wicked, is a physical event and has physical causes. Scientific naturalism’s portrayal of human agents as inhabiting a world governed by causal law challenges our idea of ourselves as free, morally responsible beings. If our actions are effects of events that occurred long before we were born, then they do not ultimately originate in us and, in some sense, we are not ultimately responsible for them. We lack libertarian free will. 1 Whatever our freedom and responsibility amount to, it must be of a sort possible for creatures who are completely enmeshed in the causal order of nature. For many, who assume we either possess libertarian freedom or nothing, this means that our image of ourselves as free, morally responsible persons is illusory, a residue of a pre-scientific conception of human nature. Others, agreeing that this is what materialism implies, try to avoid it by embracing dualism, insisting that the human person who deliberates and chooses is not merely part of the physical universe. Yet, as we saw in the preceding chapter, the evidence that we are material beings is compelling. Whatever transpires in the human mind occurs in nature. This prompts a quest for a loophole in the naturalistic picture, a place for free 163
164
Chapter 6
agency in the physical universe, thanks to the indeterministic nature of the quantum domain or the mind’s possession of causal powers that transcend the physical. In this chapter, I will show that neither strategy plausibly delivers human choice from the web of physical causation in a way that matters for freedom and responsibility. The attempt to integrate this naturalistic conception of human action into a theistic view of the world presents a second challenge. The Creator freely chooses the initial conditions of the creation and the laws that govern it. Whatever natural events cause proximately, God, the first cause, causes remotely. If each human action is a link in a chain of causes and effects that originates in God’s initial act of creation, can we truly be the authors of our choices, or is God the only real actor, in virtue of being the most remote cause of all human actions? If God causes us to do what we do, intends it, and foreknows it, can we be free and responsible in relation to our Creator in the way personal relationship requires? Some forms of theism, even some varieties of the Christian faith, celebrate a God who exercises complete control over creation, specifically intending everything that happens, and readily accept the denial of human freedom and responsibility that this entails. This cannot be reconciled with the central Christian affirmation that the God who is love creates for the sake of fellowship with persons distinct from their Creator. Christians worship the God who creates not as a demonstration of limitless power, but to share the joy of the Triune life with truly distinct created persons. This requires that as creatures we possess a kind of autonomy, the ability to act on our own in relation to our Creator. Our actions must be our actions, not God’s. With this twofold challenge in view, this chapter delineates a naturalistic conception of freedom and responsibility, one appropriate to material persons created for relationship with God. COMPATIBILISM Some scientific naturalists abandon the idea that human beings are free and morally responsible. They take it for granted that if our choices are the effects of natural causes, then we are not free and responsible agents. This denial of moral agency is not, however, an inevitable feature of a naturalistic outlook. Many naturalists believe that there is a place for freedom and responsibility in the world contemporary science discloses. We have the best of reasons to believe that the human mind is the functioning human brain, part of the material world, and explicable in causal terms. We also have the experience of contemplating what appears to be an open future, one we shape by choosing among equally available alternatives. And we spontaneously judge that when we make these choices we are morally responsible and thus worthy of praise or blame, punishment or reward. Why not simply accept both these facts about ourselves,
Freedom
165
and grant that our choices can be both caused and free? Why not assume that freedom and causation are compatible? Many philosophers find this compatibilist option plausible, but scientists, theologians and almost everyone else generally take for granted the incompatibility of freedom and causation. Often, the question is framed: “Are we free or determined?” Thus, incompatibilism is implicitly assumed, the compatibilist option rendered invisible. Even when compatibilism is in view it is often rejected out of hand. This is puzzling: consider a set of three, mutually inconsistent, assertions, one grounded in our ordinary experience, one due to science, and one a philosophical theory: 1. Our choices are free. 2. Our choices are caused. 3. Our choices cannot be both free and caused. One of these statements must be false, and almost everyone assumes it must be either the first or the second. But why hold fixed the philosophical claim that freedom and causation are incompatible, so we must reject either the testimony of longstanding human experience or of scientific inquiry? Scientific inquiry can, and sometimes does, do away with established components of the human self-image, but we should hesitate to accept such revisions on the basis of a controversial philosophical idea, particularly if we do so merely on the basis of intuition. Of course, there are cases where finding that a choice had a certain sort of cause reveals that it was not free, but the idea that free choices cannot have a causal history of any kind is the sort of general philosophical assertion of which we should be wary, particularly in the absence of very powerful arguments for it. Why not take what science tells us about ourselves as revealing that freedom and causation are not mutually exclusive after all? REASONS AS CAUSES Typically, those who believe that our choices can be both free and the effect of causes assert not just that causation is compatible with free choice, but make the stronger claim that to be free a choice must be the effect of causes. An uncaused human choice—if such a thing were possible—would not be free. Nor is one that has the wrong sort of causes. Free choice and action is the effect of causes of the right kind. Generally, the right causes are the desires of the person who makes the choice, together with her beliefs about how to attain what she desires. These desires and beliefs are the reasons for which the person chooses and acts. Compatibilists describe a choice as free, one for which a person is morally responsible, just if it is caused by her own, reflectively considered, beliefs and desires.
166
Chapter 6
We make sense of what persons do by appeal to the choices they make, decisions that in turn we account for in terms of their beliefs and desires, the reasons for which they choose and act. 2 There are things we want and we have beliefs about what actions are likely to achieve them. Generally, choices are intelligible in light of these reasons. Karen chooses to go to the Szechuan Inn, because, all things considered, what she most desires now is some Chinese food, and she believes that, all things considered, going to that restaurant is the best way to get some. This at face value is a free choice. We employ this belief-desire psychology to explain and predict human action; without it, human behavior would be largely inexplicable. 3 Were it not generally possible to point to beliefs and desires as the reasons for which someone acts she would not be rational, and we could not regard her as a free agent. To act rationally is to act reasonably, on the basis of good reasons. Someone whose behavior is systematically unreasonable is not free; she is ill or defective. Being responsive to reasons, in the sense that her choices are, in the main, explainable by appeal to them, is necessary for being a morally responsible actor. Someone incapable of regulating her behavior by considering her reasons is not morally responsible for what she does. If there is something someone would choose to do no matter what reasons she has not to do it, her behavior is compulsive, obsessive, or otherwise falls short of behavior for which she is morally responsible. Along with practical rationality, the compatibilist sees freedom and responsibility as involving autonomy, in the sense that the person acts on reasons she can reflectively endorse as her own, as constitutive of her identity. This explains why victims of coercion are often not to blame for what they do to avoid a threatened harm. When David kidnaps Zoe the cat and threatens to do away with her unless Mike, her owner, throws a rock through the window of the dean’s office, we reasonably blame David, not Mike, for the broken window. Mike acts on a desire that is in him, one that is efficacious in producing the decision to break the window, but it is not Mike’s very own desire. It has been placed in him by David, who, knowing Mike’s devotion to Zoe, also knows that threatening him will create the desire to vandalize the dean’s office. So far as moral evaluation of the incident is concerned, the person who really wanted the window broken and got what he wanted is David. He is the one to blame. Mike was forced, not free; he chose, but not freely. The idea that free, morally responsible human action is the exercise of rational autonomy is widely accepted. What is specifically compatibilist is the idea that our reasons cause our free choices. Why believe that our reasons cause our choices? One consideration, important to the naturalist, is that this makes it possible to locate free, rational agency in the physical world. Ultimately, by subsuming rational human action under the causal laws of nature, it promises the integration of the human sciences into the natural sciences. Karen’s wanting some Chinese food is a particular state
Freedom
167
of her embodied, socially located mind/brain, as is her believing that the best way to acquire some now is by paying a visit to the Szechuan Inn. Likewise, the mental process of deliberating about what to do, the episode of practical reasoning that results in the choice to go to that particular restaurant, is neurally embodied computation. Reasons and choices being related as causes and effects makes clear why we invoke a person’s beliefs and desires to explain her choices. It integrates the explanation of rational human action into the ubiquitous practice of explaining events as effects of causes: things happen because of earlier events that happened. Those who deny that reasons cause choices owe us an alternative account of why we explain choices by invoking beliefs and desires. Why do we say that she made the choice because of her reasons if this is not a matter of cause and effect? There must be something more to acting for a reason than simply having that reason to do it and then doing it, for it is possible to have a reason to do something, and to do it, but not do it for that reason. Jack wants to shoot Lee, so as to collect the money he has been promised for doing so. As it happens, Lee has a twin brother Harvey, and Jack has a reason to want him dead too, one independent of his contract to do in Harvey. Jack sees Harvey, but mistakes him for Lee. He shoots Harvey, thinking he is shooting Lee. Jack had a reason to kill Harvey, and he killed Harvey, but he did not kill Harvey for that reason. A reason to do something might be present in a mind but not in a way that explains the act. 4 Conceivably, when we say that someone did something for a reason this does not imply that this reason caused her to do it, but if not then the relation of our reasons to our choices and actions is mysterious. Whether a particular mental event, such as someone’s choosing to perform an action, is caused or not is an all or nothing matter, but whether it is caused by reasons one has rationally examined and identified as his own is a matter of degree. Compatibilists therefore regard freedom and moral responsibility as a matter of degree. This coheres with ordinary moral judgment where, for example, we hold that a twelve year old is less culpable for a harmful act than an adult, but more blameworthy than a six year old. The capacity to reflect upon and critically assess one’s desires and beliefs before acting on them is a skill human beings develop over time, so we see freedom and responsibility as increasing as cognitive powers develop. Similarly, we see freedom and responsibility as diminished for persons who are cognitively impaired. When we ask whether someone is morally responsible for what she has done, the practical question is not whether her choice to do it was caused, but whether she did it for her very own, reflectively considered reasons.
168
Chapter 6
NEITHER FATED NOR FORCED Despite the consistency of compatibilism with our practices of praise and blame, many find the idea of a choice being both caused and free absurd, and contend that if our choices are caused by anything, even our very own, reflectively examined and embraced beliefs and desires, then we are not free and responsible. A common compatibilist response is to suggest that the incompatibilist confuses the idea that all our choices are caused with fatalism. Fatalism is incompatible with free choice, but causation is not fate. If something is fated to occur, it will happen no matter what. If Oedipus is fated to kill his father and marry his mother, he will do so irrespective of what he believes and wants. Had patricide and incest struck him as fine ideas, he would have killed his father and married his mother, though the story would have been rather different. What he wants and believes make no difference to the outcome. If I were fated to choose to order pepperoni on my pizza, then I would do so no matter what reasons I have and no matter what else I do or choose to do. I would select it even if I hated pepperoni, even if I think it is poison. In contrast, compatibilism presupposes that if our desires or beliefs had been different in various ways, then we would have made different choices. My actual beliefs and desires caused me to order pepperoni, but if I had remembered that I promised my wife not to eat any more pepperoni this week, then of course I would not have ordered it. My reasons cause my choices, and this means that they make a difference to what happens. Fatalism implies that they make no difference. If I am fated to act a certain way, no matter what I want or believe, then I lack the sensitivity to reasons moral responsibility requires. If, when tempted with pepperoni pizza, I cannot help myself, and no matter what reasons I have against it, I choose it anyway, then I am not a responsible agent, at least in this context. In contrast, my choices being caused by my reflectively considered beliefs and desires means I have the responsiveness to reasons that, according to the compatibilist, moral responsibility requires. Incompatibilists sometimes speak as though causation were coercion. If something causes me to make a particular a choice then I am forced to make it and thus not free. For compatibilism, when someone is forced to do something, then his action has causes of the wrong kind for freedom. A victim of coercion can say that whoever threatened him forced him to choose a certain way. The victim of a compulsive disorder can say that the desires that cause his choices force him to make them. These are cases where freedom is absent. Yet it is hard to take seriously the idea that when someone carefully considers various courses of action, examines his desires and his beliefs about how best to fulfill them, and this deliberative process eventuates in a judgment about the best thing to do in the circumstances, and this in turn causes him to choose to do it, he is being forced to choose it. Normally, one is forced to do something only if it is
Freedom
169
contrary to one’s best judgment, at odds with what he wants to do. Were David, apprehended after killing Mike’s cat, to try to exculpate himself by saying, “After closely examining my most heartfelt desires and critically considering my beliefs about how best to fulfill them, I judged that the thing to do was to kill that cat, and this made me do it!” he would be laughed out of court. We are forced to do things we see ourselves as having good reasons not to do. If someone sees himself as having, all things considered, good reasons to do something, including desires he acknowledges as his own, he cannot seriously claim that he was forced to do it. Suppose that someone lacked the power to choose contrary to his best judgment: could he reasonably want to acquire it, so as to become truly free and responsible? The compatibilist asks what more a human being could aspire to: would I be more free–‒or really free–‒if there were no reliable connection between my practical reasoning and my choices? In reality, of course, this is not a capacity human beings lack. We sometimes choose to act in ways we know we should not act, contrary to the beliefs and desires we endorse as our own. This weakness of will exemplifies not freedom, but a failure of practical rationality, a defective capacity for rational action. Though not infallible, the causal connections between our reasons and our choices do not make us victims of our wants and beliefs; they make free, responsible rational agency possible. The more secure the connection between considered reasons and action, the more free we are. An incompatibilist might agree that causation differs essentially from fatalism, and that our own well-considered reasons do not force us to act, but still insist that causation precludes freedom and responsibility. For if Jack’s choices are caused by anything, even his own, reflectively examined reasons, there is a sense in which he could not have made any other choice. A cause, we reasonably assume, suffices for its effect; once it happens, the effect must follow. 5 It seems obvious that a person is responsible for what he does only if he could have not done it, only if he could have done something else instead. We reasonably blame someone only for what he could have avoided doing. Jack commits a crime, deciding to do so in way that is, according to compatibilism, free. His choice was caused by his reflectively examined and embraced beliefs and desires. Jack’s criminal associates have offered him a large sum of money to do away with Lee. His desire for the money, and his belief that he would get it in exchange for the crime, together with his belief that he had a good chance to avoid capture and punishment, sufficed in the circumstances to cause Jack’s decision to do the deed. Jack is responsible for choosing to kill Lee only if he could have chosen not to do it. Could Jack have made a different choice? The compatibilist answer is: Yes, Jack could have made a different choice, for he would have made a different choice if his beliefs or desires had been different in various respects. If, for instance, he had believed
170
Chapter 6
that he would be caught, convicted and executed, rather than having the mistaken belief that he would get away with the crime and live to enjoy the proceeds, he would not have accepted the contract. Similarly, if Jack’s desires had been different, for example, if he had a less intense yearning for wealth, greater empathy for other human beings, or simply a more efficacious desire not to do wrong, then he would not have chosen to kill Lee. After all, Jack was not fated or coerced to make the murderous choice. He could have chosen otherwise in the sense that the choice was in response to his own reasons. Patricia Churchland writes, “If you are intending your action, knowing what you are doing, and are of sound mind, and if the decision is not coerced (no gun is pointed at your head), then you are exhibiting free will. This is about as good as it gets.” 6 This is not good enough for the incompatibilist. Jack is culpable only if he could have chosen not to kill Lee while having exactly the same reasons. Freedom is being able to make a different decision even while one’s beliefs and desires are no different. Nothing about the world prior to Jack’s choice, not even what went on in his own mind, guaranteed the choice to commit the crime. Jack’s choice was free and blameworthy only if the future at that moment was open. The decision not to commit the murder, no less than the decision to do it, could have followed the beliefs and desires Jack actually had. The incompatibilist claims that we are free only if we could have chosen differently without qualification. There are reasons to think the incompatibilist insistence that we act freely only if we could have made a different choice, no matter what, is mistaken. There are conceivable, if implausible, scenarios in which an individual could not have chosen differently, but in which he was nonetheless free and responsible for choosing as he did. For example, a highly skilled neurosurgeon, in the pay of Curley, Beantown’s notoriously corrupt mayor, covertly monitors Randy’s brain and, if he detects an inclination to choose to vote for Smiley, intervenes, causing him to choose to vote for Curley instead. But, if he detects an inclination to vote for Curley, he does nothing, and permits Randy to choose Curley. 7 Suppose Randy votes for Curley: he could not have made any other choice; the neurosurgeon would not have permitted it. Yet it seems obvious that Randy is responsible for voting for Curley. The fact that if things had gone differently, he would have been manipulated to decide for Curley anyway, does not make his choice less than free, nor mitigate his responsibility for voting for the corrupt incumbent. It appears that the incompatibilist objection to free choices being caused must rely on something other than a commitment to an unqualified “could have chosen otherwise” requirement. 8
Freedom
171
THE LONGER VIEW Much can be said on behalf of compatibilism. However, a strong objection comes into view when we step back from the issue of whether free, morally responsible choices might be caused by our beliefs and desires, and consider the full causal history of our choices. On the compatibilist account, the reasons that cause choices are themselves the effects of earlier causes, and these have causes that have causes, and so on, back to a time before that individual existed, back to the beginning of the universe. This implies that our choices are the effects of causes over which we have no control, since we have no control over the effect of causes that are not in our control. If Zoe eats sardines, she gets sick. Julie has no control over what Zoe eats, and there is nothing she can do to keep her from getting sick if she eats sardines. So, when Zoe eats sardines and gets sick, it is a mistake to blame Julie for letting her get sick. It is wrong to blame or punish Julie for something she could do nothing about. The chain of cause and effect “transfers” Julie’s lack of control over what the cat eats to the effects of eating it. Incompatibilists contend that if we are not responsible for the causes of our choices, then we are not responsible for them. Suppose that Jack’s murdering Harvey exemplifies the rational autonomy that, according to compatibilism, suffices for freedom and moral responsibility. His choice was caused by his very own, reflectively examined beliefs and desires. Yet those beliefs and desires were the effects of earlier causes, events that in turn had earlier causes, and so on, back to a time when there was no Jack to have a say over anything. Jack cannot, after all, now choose what to want or what to believe. No doubt, some of his current beliefs and desires have his earlier choices in their causal histories, but those choices in turn have causal histories tracking back to Jack’s infancy, before he was capable of making choices, to his genetic endowment and the natural and social environment in which it was expressed, and ultimately back to a time prior to his existence. The fundamental incompatibilist argument is that if Jack’s choice is simply one link in a chain of causes and effects disappearing into the dim past, then, no matter what its immediate causes, it happened as a result of events that happened long ago, when Jack was not around, and therefore we cannot sensibly regard him as morally responsible for his choices. They do not originate with him. This challenges the compatibilist to explain why, if all our choices were “in the cards” before we were born, the idea of moral responsibility, and the practices of praise and blame, reward and punishment it underwrites, are not founded on an illusion. Why, if human choices are the effects of factors over which we have no say, is it any more reasonable to hold us responsible for them than it is to reward someone for having blue eyes or punishing her for disliking the taste of sardines?
172
Chapter 6
For incompatibilism, pointing out that Jack could have chosen differently if other things had been different, where ultimately Jack had no control over those other things, is myopic. The integration of human practical reasoning and the choices it causes into a system of universal causation implies that there is a sense in which we can make no choices other than those we actually make. The compatibilist conception of freedom and responsibility appears to depend on a narrow focus on the proximate causes of choice, and to ignore the crucial fact that the reasons for which we act ultimately are effects of primordial causes beyond our control. To the incompatibilist, it seems unreasonable and wrong to hold Jack responsible simply because he rationally pursues his ends without coercion, manipulation or defect, while ignoring the fact that his ends, either directly or indirectly, are given to him, not of his choosing or making. If Jack’s decision was caused by the state of the world billions of years ago, can we seriously treat him as morally responsible for deciding as he did, no matter that some process of practical reasoning in his brain was among its proximate causes? Here the crucial incompatibilist challenge comes to the fore. How can we see a human being’s choice as originating with her, rather than as something that happens in her as a result of things that happened long ago? And if we cannot see it as ultimately originating in her, can she be responsible for it? THE IRRELEVANCE OF INDETERMINISM One response appeals to the fact that even though Jack’s decision to kill Lee was the effect of causes, it was not necessitated by the way things were eons ago. Standard interpretations of physical theory portray the basic laws of nature as indeterministic. The laws that govern the microphysical components of all macroscopic entities do not guarantee that a particular event will occur, but only fix the probability of its occurrence. Yet, at the scale of ordinary human experience this indeterministic world behaves as though it were deterministic. In objects constituted by trillions of elementary particles, the law of large numbers implies that things will average out, so while the behavior of any individual particle is in certain ways unpredictable, the behavior of the whole is virtually as predictable as if it were part of a deterministic universe. When applied to large numbers of microscopic entities, the stochastic laws of quantum mechanics imply that the probability of certain things happening is so high, and of other things happening so low, as to be for practical purposes indistinguishable from certainty. It is physically possible for the cat that I put in the cellar last night to appear this morning in the kitchen, thanks to all its constituent particles quantum tunneling through the floor, but the probability of this trick is so close to zero that it would be irrational to take it seriously
Freedom
173
in an attempt to explain how she escaped confinement. When he takes aim, Jack can ignore the fact that the trajectory of individual particles in the bullet he fires at Lee is inherently unpredictable, and proceed as though the bullet’s path is deterministic. Human brains, including, so far as anyone knows, the neural components causally relevant to decision making, are among the things so large as normally to render quantum mechanical considerations negligible. Still, a macroscopic material object, such as Jack’s brain, could be structured so as to be sensitive to goings on at the quantum level. Artificial devices can be devised to channel the effects of particular quantum mechanical events up to the observable realm; a Geiger counter, for example, detects the decay of a nucleus of a radioactive element, and the emission of a beta particle, and responds with an audible signal. Conceivably, Jack’s brain is similar, a naturally occurring chaotic system, in the sense that very small initial differences in physical state can cause large differences later. Suppose there is a particular quantum mechanical event Q, with a probability of .5 of occurring in a particular neuron in Jack’s prefrontal cortex between 10:45 and 10:46, such that if Jack has a certain set of beliefs and desires at 10:45 and Q occurs in the next minute he will choose to kill Harvey, while if it does not occur, then he will choose not to. If this is the way Jack is constructed, his choices are not causally determined by the state of the world years before he existed, or even when he got up this morning. Nor are they determined by his reasons. As of 10:45, things could have gone either way. However, as critics of the attempt to enlist quantum mechanics in the cause of free will have long noted, this channeling of microlevel indeterminacy up to the level of a brain’s psychological properties—even if it is physiologically realistic, which is dubious—does not make possible what incompatibilists regard as free choice. In this scenario it is true that however Jack chooses, he could have chosen differently, even if all his desires and beliefs had been exactly the same. His reasons do not necessitate his deciding to commit the crime. A nanosecond before the wave function collapsed, the future was open and things could have gone either way. How he chooses depends on the outcome of the “measurement” that occurs in his brain, and that is governed by a law that assigns only probabilities to the possible outcomes. However, this does not afford Jack the avoidability that, according to incompatibilism, moral responsibility requires. Which undetermined event takes place is not in Jack’s control. Jack’s choosing to kill is still necessitated, though by a very recent undetermined event, rather than by a deterministic causal chain originating at the beginning of the universe. Whether Q happens, and what it causes, is no more up to Jack than whether the Big Bang, and its effects, happen. If the remote cause renders the choice not free, so does the proximate quantum mechanical event. 9
174
Chapter 6
As far as human freedom and responsibility in a world governed by causal law is concerned, determinism might as well be true. So far as we can tell, the operations of the human mind/brain approximate to being deterministic no less than those of a cat, a car, or a computer. Quantum indeterminacy is of no use to the incompatibilist seeking to make the physical universe safe for free choice. In general, if causation conflicts with freedom, it conflicts with it whether it is deterministic or indeterministic. However, later we will see that indeterminism does make a difference when God causes human choices. DOWNWARD CAUSATION We saw in chapter 5 that the human mind is not reducible to the brain, and that the mind in some sense emerges from an underlying physical reality. Yet this does not mitigate the implications of materialism: it sustains no vision of human beings transcending nature. Reduction is rare and emergence is commonplace. Nonetheless, some thinkers appeal to these facts in the hope of securing human freedom and responsibility despite our being material things. The phenomenon invoked in this context is downward causation. We sometimes opt for causal explanations in which what we cite as the cause, and what we cite as the effect, are on “different levels.” When the cause is an event on the “higher” level, there is an instance of “downward causation.” We might, for example, explain a neural event as the effect of a psychological event. What occurs in someone’s brain might be the effect of what she is thinking or feeling. Thus, the mind acts upon the brain; the mental acts upon the physical. We can say, as Nancey Murphy does, that downward causation, “allows agents to exert downward control on neural processes and behavior.” 10 This language is not objectionable, unless it gives rise to a covertly dualistic picture in which the mind is one thing that acts upon another thing, the brain. When it does, the discussion is haunted by the ghost of the ghost in the machine that appeared in chapter 5. Unless dualism is true, and there is an immaterial mental substance distinct from the physical world, reality in the case of human beings is flat. There are no levels of reality, just levels of description of the one reality. When we say, for instance, that Ralph chooses to pick up the book, we refer to a particular event in his brain that could, in theory, be referred to with the terminology of neuroscience. Ralph’s brain being in a certain neuroscientifically describable state is, at this moment, in this context, sufficient for Ralph to be choosing to pick up the book. We might go on to say that Ralph’s choosing to pick up the book activated a group of neurons in his motor cortex. If, as materialists contend, particular mental events are particular brain events, and some brain events cause other brain events, it is not a surprise that some mental events figure in the
Freedom
175
causal explanation of some brain events. But this is just to say that some physical events which can also be described as mental events have physical events as their effects. This is what we should expect if every human choice and action is the effect of physical causes. Indeed, even if some very robust form of emergence were true, so that it is an inexplicable, basic fact of the world that whenever neural event x occurs, the mental event y occurs, this would not change the fact that whatever causes x also causes y. Downward causation offers no reason to suspect that we have any freedom or responsibility beyond what compatibilism ascribes to us. It cannot free our choices and actions from the regime of natural causation. Downward causation, like quantum mechanical indeterminacy, fails to supply what incompatibilism demands for freedom: the human individual as the ultimate originator of her actions. For libertarian free choice the incompatibilist must resort to something radical: agent causation. AGENT CAUSATION For incompatibilists, there can be no real freedom, no genuine moral responsibility, no rational or moral justification of blame and punishment, unless, in choosing, we initiate an absolutely new chain of causes and effects in the world. Freedom and responsibility demand that the person be the ultimate origin of his choices and actions, and that he be more than a conduit through which chains of causes and effects pass. Jack decides to do away with Lee and this sets in motion a sequence of events resulting in Harvey’s death, but Jack is responsible for the choice only if it was not the effect of prior events in Jack, since those events are the effects of earlier causes over which he had no control. Yet, if the only alternative to Jack’s choice being an effect of earlier events is its having no cause, there is no hope for responsibility of the sort the incompatibilist envisions, for Jack cannot be responsible for an event that has no cause, one that “just happens.” Some incompatibilists propose to escape this dilemma by radically revising our understanding of causality. The standard modern understanding of causation is as a relation between events: something happens and then something else happens: Rein throws a rock at the window and the window breaks. Asked what caused the window to break, we might report that Rein caused it to break, but we would more precisely mean that one event—the window changing from not being broken to being broken—was the effect of another event—the change from Rein not throwing the rock to throwing it. The idea presented to make the world safe for libertarian freedom and responsibility is agent causation: the person herself, not some event in her, not her having particular beliefs and desires, not her deliberating, but simply the person—the agent—causes the choice of a particular course of
176
Chapter 6
action. If Jack’s wanting to get the money offered in exchange for killing Lee, together with his beliefs about killing Lee being a good way to get it, caused him to choose to kill Lee, the choice does not ultimately originate with him, for that desire and that belief are effects of causes outside him. Jack does not initiate the choice; it happens to him. The sequence of cause and effect passes through him; it does not begin in him. The defender of agent causation contends that Jack is responsible for his choice and deserves indignation, blame, and punishment because he, not something that happened in him that was caused by something else, is the ultimate source of the wicked deed. Events in people don’t kill people; people kill people. Nothing short of this terminates the regress of causes that precludes our having what the incompatibilist regards as authentic responsibility, rather than some compatibilist imitation. We cannot be morally responsible unless we are causally responsible in this fundamental way. The proponent of agent causation faces, first, the challenge of saying precisely what agent causation could be. He needs to explain how, if choices are not the effects of changes within the agent, why they occur at some times rather than at others. Jack exists on Monday, and he does not choose to attack Lee, but on Tuesday he does. The mere fact that he exists obviously does not suffice for his choosing to commit the crime. It is easy to suspect that he acquires reasons to do it, and this change in him—this event—explains the choice. Thus, we are back to asking how reasons explain choices yet do not cause them. Beyond providing an account that makes agent causation intelligible in the abstract, its advocates owe us an account of how a human being, a material thing inhabiting a universe governed by causal law, could be an agent cause. Perhaps a radically different mode of causation calls for a profoundly different mode of being. Dualism invites the image of the freely choosing human person controlling things from outside the realm of natural cause and effect. But it is highly implausible that the mind/brain, a complex physical entity whose states from moment to moment are governed by causal laws, could have this mysterious capacity. The fact that agent-causation is mysterious does not entail that it does not exist, or even that all things considered it is unreasonable to believe in it. After all, Christians and other theists believe that God is an immaterial person whose actions are radically free, necessitated by nothing, yet done for good reasons. We reasonably believe that God is an agent cause. In chapter 3, I argued that our best reasons to believe that God exists involve explaining this contingent world’s existence as due to the creative act of an agent cause. We are not agent causes, but we exist, having the creaturely freedom we have, because of agent causation. However, even if we could believe that a material being could be an agent cause, this would not be enough for us to be endowed with the responsibility that the incompatibilist ascribes to us. 11
Freedom
177
IMPOSSIBLE FREEDOM Freedom, responsibility, and rationality are inextricably interconnected. Whatever free, responsible action is, it involves having, and choosing to act on, reasons, i.e., on one’s desires and beliefs about actions that might fulfill them. Even if human beings somehow were agent causes of their choices, they would still choose in accord with their reasons, and these are desires and beliefs for which the individual cannot be ultimately responsible. Perhaps there could be creatures that uniformly behave in ways that resist rational explanation, but they would neither choose freely nor be responsible for what they do. Human beings do, of course, episodically act in ways that cannot be adequately explained by their reasons, but an excess of this warrants a diagnosis of incompetence and loss of status as a responsible actor. In general, we act freely and responsibly as we act in ways that make sense in light of what we want and what we believe about how to get it. Denying the causal link between reasons and choices does not change this. We cannot be ultimately responsible for our choices unless we are, first, ultimately responsible for the reasons for which we choose, and no creature has that responsibility. I can, for example, choose now to act in ways that I think might result eventually in my liking beer and coffee less and vegetables and tofu more. Our ability to bring about changes like this is limited, but we have it to some degree, and the degree to which we have it is itself probably sensitive to past choices and actions. What I cannot do is desire at will. I cannot simply choose what to want. Recall, as we saw in Chapter Two, that the will is no less impotent when it comes to belief. Whatever my earliest choices were, I made them in virtue of beliefs and desires that were not themselves the result of choices I made. They were, for me, givens, due to my genes and my early environment, not matters over which I had a say. Galen Strawson succinctly argues that when we act, we do what we do because of the way we are, and we cannot ultimately be responsible for the way we are, so we cannot be ultimately morally responsible for what we do. 12 The serious problem for the incompatibilist conception of freedom is not, finally, human choices being the effects of causes. The deeper problem is that it requires the impossible: that we freely choose to bring ourselves into existence, with whatever reasons, and ways of reasoning, that explain the choices we make. Of course, nothing brings itself into existence. The only person the incompatibilist might be entitled to regard as free, as ultimately responsible, is a deity that has always existed, always freely choosing to be a certain way, everlastingly choosing to have the reasons that account for divine choices. Creatures are not what they are as a result of their own free choices, and thus cannot possess ultimate responsibility. Creatures, whether brought into existence directly, or products of secondary causation, act on the basis of
178
Chapter 6
reasons that are given, not freely chosen. What the incompatibilist conception of freedom and responsibility finally collides with is not human beings being material creatures, but with our being creatures. To be an agent cause is to be godlike. When philosopher Roderick Chisholm introduced the idea of agent causation onto the modern scene in 1964, he acknowledged that it implies “we have a prerogative which some would attribute only to God: each of us, when we act, is a prime mover unmoved. In doing what we do, we cause certain things to happen and nothing—or no one—causes us to cause those things to happen.” 13 Chisholm revived an idea initially worked out by Medieval Christian thinkers. Holding that God acts, causing things to happen in the world, but also that God is immutable, they needed to explain how there can be events (changes) in the world that are the effects of God’s action when there are no corresponding events (changes) in God to be their causes. The solution was what we now call agent causation: God, not events in God, is the cause of effects in the world. 14 Even if, as I believe, they were mistaken in portraying God as immutable, they knew that God, the Creator of nature and author of its laws, transcends the realm of cause and effect material creatures inhabit. However, we are not everlasting, transcendent beings and should reject the metaphysically inflated selfimage that assumes we are. No creature can be ultimately responsible for her choices and actions. The conviction that she must be manifests the misplaced fear that modern science portrays humans in too low a fashion, as lacking what our being made in God’s image involves. On reflection, commitment to ultimate responsibility, agent causation, and incompatibilist free choice for creatures should be rejected as too high, as portraying human beings not as God’s image, but as tantamount to gods themselves, acting on the natural world from beyond. Whatever our freedom amounts to, it is that of a material creature. Whatever our responsibility is, it does not require that we be prime movers unmoved. In previous chapters, we saw that we are prone to imagining that we share certain impressive properties with God, but these properties in fact have no place in the physical world to which we belong. This is the case with our freedom and responsibility as construed by incompatibilists. In contrast to our transcendent, everlasting Creator, we are not the absolute uncreated originators of our actions. The freedom and responsibility available to creatures must be sought along compatibilist lines. RESPONSIBLE CREATURES The essential Christian confession is that the Triune God who is an everlasting community of loving persons freely seeks fellowship with persons other than God. For this reason the universe, and the persons who find
Freedom
179
themselves in it, exist. This does not require that we possess the kind of freedom that only God has, but it requires that we have freedom and responsibility of the kind required for relationship with other persons, human or divine. To exist as a person in relation to other persons presupposes responsibility understood not as ultimate origination, but as responsiveness to reasons of one’s own. Suppose Jack’s decision to carry out a killing for hire was caused by his having the particular desires and beliefs he had at the time, computationally processed by the neural circuitry in his brain, and that these events were the most recent in a series of causes and effects stretching back to a time before he was born: would we be wrong to punish Jack for killing Harvey? Can we reasonably blame him for what that desire caused—the choice to commit the murder—if we cannot blame him for having that desire? Would blaming him and putting him in prison be the moral equivalent of, for instance, blaming and punishing Jill, a victim of Tourette’s syndrome, when she shouts obscenities in the courtroom? In either case, the behavior is caused by an event in a human brain that is a remote effect of events that took place eons ago. Neither has a causal history that originates in the agent’s conscious act of will. Nonetheless, there are differences between Jill and Jack, and they justify holding Jack, but not Jill, responsible. The relevant difference is that Jack’s behavior, in contrast to Jill’s, is regulated by reasons. What Jill wants or believes has no bearing on whether she disrupts the courtroom. She might be horrified at the thought of shouting obscenities at the judge, but she does it anyway. Her beliefs are equally impotent; even if she had believed that this judge orders the summary execution of anyone who disrupts proceedings in his courtroom, she would have done it. It makes no sense to demand from Jill an accounting of her behavior in the sense of justifying it, showing that she had good reasons for acting as she did. Her reasons play no explanatory role. Uncovering the causes of her behavior, we seek to explain it, not to justify it. Jack, in contrast, did what he did for reasons; he would have acted differently had his beliefs and desires differed in various ways. We ask him, “Why did you do it?” and expect an explanation in terms of his reasons, which we might—but in all probability will not— agree are good reasons for what he did. He is accountable for what he did when he killed Harvey in the sense that others can reasonably demand an accounting; they seek and weigh his reasons. For compatibilism, being responsible is being responsive to reasons. Reasons are also the causes of a person’s choices and actions, but this does not excuse him from being a participant—even if an unwilling one—in the forums in which choices and actions are evaluated and judged as justified or unjustified. He is fully implicated in the order of cause and effect, but this does not keep him from being a person among persons, one rational agent interacting with others. Moral responsibility
180
Chapter 6
is a normative social status. To be a morally responsible person is to have standing in the tribunals, whether formal or informal, public or private, judicial or otherwise, where actions, choices, and the reasons for them are appraised. A responsible person is someone it is reasonable to try to reason with, to try to persuade to perform, or to refrain from, actions. In doing this we enter into his point of view, acknowledging that, like us, he acts for reasons, but that he might have reasons we do not share. Treating him as a responsible actor we recognize him as similar to, yet distinct from, ourselves. We acknowledge that he has a point of view from which things are valued in various ways, and things appear in various ways, and these ways need not be ours. This, not an impossible absolute origination of action, is what belonging to a community of free persons presupposes. 15 A responsible individual goes through life doing things because of her reasons. She differs from creatures whose behavior is not explainable in terms of reasons, whose doings we can make sense of only by finding their non-rational causes. Responsibility is being subject to the demand for reasons; to be morally responsible is to be subject to the demand for moral reasons. Normal, adult human beings are in this way responsible, while very young children, the insane, the coerced, and dogs and cats are not. This is true although we are not the ultimate originators of what we do; we have not created ourselves. This is a species of responsibility, and correlatively a kind of freedom, possible for creatures. We can imagine a more radical freedom, but it belongs to the Creator, not the creature. IMAGING AGENT CAUSATION We are the material image of our Creator, the God who absolutely originates actions. Explanations of what God does always terminate with God, with the divine character everlastingly freely chosen. No human being is an uncreated agent cause, but it is open to us to one degree or another to image the agent cause who freely created us. Our freedom is an image of God’s freedom. We do not choose what to want. Yet we have as an ideal being someone all of whose desires are the products of, or at least vetted by, critical self-reflection, so that one’s desires are consciously and reflectively endorsed as one's own. In this way, what one wants can be construed as a matter of free choice, though only in a compatibilist sense. The cigar smoker who, after serious self-reflection and consideration of the pros and cons, wants to want to keep smoking, is responsible for that desire in a way he would not be responsible if this were a desire he regretted, disapproved of, or resisted even as he acted on it, or if he simply took it as an unquestioned given. Obviously, this is an ideal to which we at best only
Freedom
181
approximate. To whatever degree we approximate it, what we approach is being a material, created simulation of an agent cause. The case with belief is less straightforward. No rational being would choose what to believe, even if she could. What we can reasonably aspire to is that the processes of acquiring beliefs, our procedures of inquiry and modes of reasoning, approximate as much as possible to being the results of one's rational, informed choices. We can choose to act in ways that make it more likely that we will believe things if, and only if, they are true and worth knowing. Someone who unreflectively and uncritically believes what he has been told, or what he wants to be true, is not responsible for his beliefs in the way he would be if they were products of sustained reflection. In this sense the agent can be said to endorse his beliefs, and take responsibility for them, though not, typically, responsible for their being true. In these ways, the reasons that cause the material agent's choices and actions, her believing and desiring as she does, move toward, but cannot conceivably reach, the condition of being fully and freely chosen by her. For her actually to be in such a state would mean that she created herself from nothing or that, like God, she has always existed, neither of which is true of human beings. Ultimately, we believe, want, and choose as we do because of how we are made, and because of what befalls us. But penultimately, to the extent that she lives a life of critical self-reflection, the informative explanations of why she has the reasons she has, and thus why she chooses and acts as she does, appeal not to causes outside her but to her own choices, choices in turn explained by beliefs and desires she reflectively endorses as her own. In reality every choice she makes remains an effect of earlier causes, eventually ones that precede her existence. Yet to the degree that she succeeds in simulating agent causation those causes are of decreasing value in making sense of why she believes and wants as she does now, and of why she is making the choices and performing the actions that she is making and performing now. This accords with our common-sense moral psychology: no human agent's actions are ever decoupled from her genotype and the behavioral dispositions it programmed for, nor from her past environment and experiences. When she is young these non-rational causes predominate in explaining why she does what she does. As she matures, such explanations recede into the background. What comes to the fore are facts about her previous critically reasoned decisions. They figure significantly in the story of how she came to be the person she is today. Had she not made them, she would be significantly different. In this limited sense she is a self-made woman. We are entitled to see her, and she is entitled to see herself, as the penultimate origin of her actions. She becomes more free and more responsible for herself in virtue of her increasing responsiveness to reasons.
182
Chapter 6
There is much to be said on behalf of the compatibilist account of freedom and responsibility. It portrays the kind of free, morally responsible agency possible for persons who inextricably belong to the realm of natural cause and effect. This deflationary account of human agency will not satisfy those convinced that genuine freedom and responsibility lie in being the ultimate initiator of one’s choices and actions. Nonetheless, and crucially, compatibilism casts the human agent as a person among persons, a member of a moral community, interacting with similar agents in relations of mutual accountability. COMPATIBILIST FREEDOM FOR CREATURES A familiar argument purports to show that divine omniscience is incompatible with human freedom. If God now knows that tomorrow Sam will choose to go to the zoo, then Sam must choose to go to the zoo and thus does not freely choose to go. The flaw in the reasoning is readily detected: the “must” it relies on pertains to logical implication, not causation. If God (or anyone else) knows that something will happen, then it will happen, simply because whatever anyone knows is true. I know that the moon will rise tonight at 12:30 a.m. This implies that the moon will rise tonight at 12:30 a.m., yet my knowing this has no effect on the moon. If mere knowing were a threat to freedom, it would not be limited to foreknowledge. If God (or anyone else) knows that yesterday Sam chose to go to the circus, this implies that yesterday Sam chose to go to the circus, so he had to go. If God (or anyone else) at this moment knows that Sam is choosing to stay home, this implies that at this moment Sam is choosing to stay home, so he must be making that choice. The absurd conclusion: no known choice—past, present, or future—could be free. Unfortunately, the challenge that divine foreknowledge poses to the freedom of creatures cannot be so easily deflected. This becomes clear as soon as we ask how God knows creatures’ future choices. There is just one good answer: God knows what will happen in virtue of knowing the present and knowing the deterministic laws that govern the universe. 16 Having perfect knowledge of the creation at any moment, perfect knowledge of the causal laws, and infallible knowledge of what together they entail, God has exhaustive knowledge of the future. God knows future human choices, but only because they are causally determined. If free choice is incompatible with determinism, then divine foreknowledge rules out human freedom, not because it is foreknowledge, but because it presupposes that human choices are causally determined. Here, we might be tempted to say, “So much the worse for incompatibilism!” We can dispense with it and conclude that the compatibilist understanding of freedom and responsibility coheres with a Christian vision of ourselves as the material image of God. In itself, the fact that our
Freedom
183
choices are the effects of what happened long ago and far away does not imply they are not free, or that we are not responsible for them in ways that matter for human life and, crucially, for God’s aims in creating us. What matters, the compatibilist insists, are not the remote causes of choices, but their proximate causes, whether they are the person’s very own, reflectively considered and endorsed beliefs and desires. For this, not being the ultimate originator of choices and actions, is what participation in a community of persons requires. Our freedom is categorically different than God’s, but this is what we reasonably expect: we are not gods, but God’s created image. However, this happy compatibilist conclusion runs afoul of a tacit assumption about the remote causes of our choices: they are not the actions of some other person who controls or manipulates us. The fact that Jack’s choosing to shoot Lee was the effect of events that occurred in the remote past, long before Jack or any other human being existed, does not excuse him from his peers’ demand for an accounting, but things would be very different if some other person had control of Jack’s mind. Suppose that the reasons for which Jack acts are beliefs and desires surreptitiously implanted in his mind by Rolf, an evil genius neuroscientist. He wants Lee dead and deploys his highly reliable techniques of mind control, using Jack to achieve his goal. Prior to choosing to do in Lee, Jack reflects upon these desires and beliefs and concludes that, all things considered, he ought to kill Lee—it is what he has best reasons to do—and this deliberation causes him to choose to do so. In Jack’s subjective perspective the conditions for compatibilist freedom and responsibility appear to have been satisfied, but this is not the case. Despite how things seem to him, the reasons for which he chooses to kill are not fully his own. Rolf’s role in the causal history of his action seriously undermines his freedom and responsibility. The more Jack’s beliefs, desires, choices, and actions are what they are because that is what Rolf intended them to be, and the more reliable his methods of realizing his aims through Jack’s practical reasoning, the less we can regard Jack as acting on reasons that are truly his own. Discovering the causes of Jack’s reasons, we see him not as free and responsible but as Rolf’s puppet. For Christians and other theists, the concern is that if our choices and actions have God’s initial creative act as their remote cause, then we are in a situation essentially similar to Jack’s. Even when we seem to satisfy the requirements for freedom and responsibility as compatibilism conceives it, in reality we do not act freely and responsibly, for our reasons are never authentically our own, but always precisely what God caused and intended. Even if compatibilism as a general view of the relation of freedom and causation is true it seems that it cannot be true in a world created by God, a world in which our actions are intended, caused, and foreknown by someone else. Theism threatens to take back what compatibilism offers: a kind of freedom and responsibility appropriate to materi-
184
Chapter 6
al creatures. God created a world that contains creatures that make choices, and these created persons are not agent causes, but material persons. This implies that God intentionally causes them to have the reasons they have, to make the choices they make, and to do what they do. In light of this, can we believe that we are free, responsible persons? This challenge is urgent in light of the singular Christian confession that God created, not for the purpose of having a universe to control, but for relationship with created persons. A relation of love and trust where everything one person does is the intended and foreknown effect of what the other does is not conceivable. Suppose, to extend a science fiction scenario from chapter 4, that a dissatisfied wife invents a spousal compliance device, a machine that unobtrusively acts on her husband’s brain, causing him to have precisely the beliefs and desires that she knows will cause him to choose, and thus to act, in accord with her wishes. Even if (implausibly) she is so good and wise as to use the device always to control her spouse in ways that are best for him, the project is inimical to a genuine interpersonal relationship. If she does it without him knowing, she deceives him, and the relationship is ipso facto defective. If he knows she is doing it, his sense of himself as a person in his own right, distinct from his wife, erodes. He will be in the untenable position of never being able unequivocally to regard his desires as his own: “I want this, but do I want this just because she caused me to want it? Is this what I really want? Would I want it if she had not taken control of my mind?” Never being able to act on reasons he unequivocally identifies as his own, he is deprived of the possibility of compatibilist free agency and the relationship is fatally impaired. Current technology does not give human beings the capacity systematically to control one another, and we may well hope it remains impossible. But we have no reason to doubt that God could have created a world inhabited by persons whose actions are precisely what God foreknew and intended, but we have good reasons to reject the idea that God did so. Knowing the purposes for which God created, we can be confident that God would not choose control. No less than design, the control determinism implies is at odds with God’s creative aims. The considerations in view here figured in chapter 4’s examination of human origins, but now they arise in a more pronounced form. We noted there that reasonable parents, desiring to bring a child into existence, would decline the opportunity to design their offspring. They would not make the child’s specific characteristics depend on their intentions, not simply because of the obvious limitations on their goodness and wisdom, but because no matter how good and wise they otherwise might be, to do so is at odds with the aim of genuine relationship with the child as a person distinct from themselves. This fundamental fact makes it reasonable to think that the God revealed in the Bible, the God whose aim is to share the divine life with created persons, would create them by means
Freedom
185
that do not involve design. God causes the human species to exist, but did not design it, a conclusion consistent with what contemporary science tells us. It is even less plausible that this God would make the choices and actions of individual created persons depend on their Creator’s intentions. Therefore, from the perspective of Christian faith, we have strong reasons to doubt that this world is governed by deterministic laws. Even though freedom is compatible with causal determinism in a godless universe, it is not compatible with it in a universe God creates. THE RELEVANCE OF INDETERMINISM Simply causing someone to perform an action does not in itself raise a problem for freedom. Marvin, while having lunch with friends in a restaurant, relates an anecdote about being attacked by a cat as a child. Jackie, sitting nearby, overhears, and this causes her to recall with bitterness the sad demise of her cat at Harvey’s hands. Dwelling on this, her hatred for Harvey grows and gives birth to a desire for revenge. Later, when the opportunity presents itself, this desire, together with her beliefs, becomes the reason she chooses to kill Harvey. If Marvin had not told the story, then Jackie would not have killed Harvey, but although his story telling had a necessary role in the causation of her action this obviously does not imply that Marvin intended her to do it. Rolf the neuroscientist does, but Marvin the raconteur does not, undermine freedom and diminish responsibility. Rolf intends to cause Jack to kill but Marvin has no such intention with respect to Jackie killing Harvey. Acting on this intention, Rolf manipulates and controls Jack. Marvin, having no such intention, merely causes Jackie to do what she does, and this has no bearing on whether she acts freely. Unlike Rolf, Marvin does not undermine the compatibilist conditions for free choice. Suppose, though, that someone persuades someone to do something. Marvin, wanting Sally to go to Pizza Palace and knowing her devotion to pineapple pizza, informs her that it is on sale there. She believes him and this, in conjunction with her desire to save money, causes her to choose to go there, where she procures the pizza. We may suppose that she would not have done so had he not told her about the discounted pineapple pizza. This is a scenario in which Marvin causes Sally to do what he intended her to do. How can Sally be free, despite Marvin’s role in the causal history of her choice, while if this is a deterministic universe in which she chooses as she does because God intentionally caused her to do so she would not be free? The crucial difference is that whatever reason Marvin provides Sally for making a particular choice, there are other, competing reasons he does not know of, not provided by him, over which he has no control, and in her deliberations, she might, for all he knows, have acted on them,
186
Chapter 6
rather than on her love of pineapple pizza and her desire to be thrifty. When we say that Marvin caused her to do something by persuasion, we are pointing to the fact that his action was a salient event in the causal sequence that eventuated in her making the choice he intended. This one act did not suffice to cause her choice. We do not mean that Marvin had the knowledge or the power to ensure that his action would cause the intended choice. The Creator of a deterministic universe acts on that knowledge, exercising that power. Marvin may be a good persuader, but he does not control her. A human being giving Sally a good reason to do something may make it likely that she will do it, but there is no guarantee and no control. If her choice is bad, the fact that someone gave her the reason to do it does not exculpate her. Coercion and manipulation deprive her of freedom; persuasion does not. 17 In contrast, if God were to create a deterministic universe, then every belief and desire, every reason, Sally has is intended by, caused by, and foreknown by her Creator. Whatever reasons cause her to choose as she does she has because God intended, caused, and foreknew them. And God, the free author of the deterministic laws, intended, caused, and foreknew she would make that particular choice. God would be Sally’s controller, not her persuader. She would lack responsibility. She would not be free. Human cognitive limitations often make for causation without intention. Lacking exhaustive knowledge of the laws of nature and of the conditions in which we act, as well as being unable reliably to draw all the implications of what we do know, we regularly act in ways that have consequences we do not foresee or intend. In contrast, free of all cognitive limitations, God intends all the causally determined consequences of the initial act of creation. If human choices and actions are precisely what God intended them to be from the beginning of creation, then compatibilism cannot rescue freedom and responsibility. If the creation is deterministic, then God not only causes all our actions, but intends them. God freely brings the universe in its initial state into existence, and at that moment, knowing precisely what the effects will be billions of years later, knows that Jack will kill Lee. This event, like everything every creature does, would be one of the many inevitable effects of that original act of creation, all of which are known and intended by God. Indeed, it might be hard not to conclude that if the creation is deterministic, then it is God alone who truly acts, controlling a world of puppets. God freely acts, bringing the universe with its laws into existence, knowing the guaranteed effect will be that Jack kills Lee. Add to this the fact that God, unlike creatures, freely chooses the causal laws, and we wonder how we can avoid seeing God, rather than Jack, as culpable for the crime. Jack is simply the means by which God kills Lee. In contrast, indeterminism severs specific intention from causation. If the laws of nature are indeterministic, so that when God initially created
Freedom
187
there were no specific inevitable consequences, then although God causes creatures’ choices and actions, they were not intended. 18 This requires that creatures’ choices and actions be very improbable. Mere indeterminism does not suffice for someone to cause an event without intending it. Jack, let’s suppose, is not a particularly skilled marksman. When he takes aim at Lee, the odds that the bullet he fires will find its target is only one out of five. Unfortunately, this time his aim is true; the bullet strikes Lee and kills him. Apprehended, Jack cannot plausibly appeal to the fact that he is not a very good shot to make the case that he did not intend to kill Lee. In contrast, even if Marvin, before telling the story about the cat attack, had fleetingly considered the possibility that his doing so might set in motion a sequence of events eventuating in someone choosing to exact murderous revenge on behalf of a cat, he would not have intended for Jackie to murder Harvey. One does not intend the possible but highly improbable effects of one’s actions, even when one knows what they are. This does not imply—absurdly—that the universe is an unintended consequence of God’s act of creation. The Creator’s intentions are fully realized in this indeterministic creation, but the divine intentions in creating are general, not specific. Human action is encumbered with significant ignorance of the possible effects of our actions, especially over long periods of time, and with comparable ignorance of the probability of the possible effects we can conceive. In contrast, when God called into being a universe and chose for it to be governed by indeterministic laws, every possible effect of this creative act was fully known—all the possible futures for the creation—and God knew precisely how probable each of them was. On reasonable assumptions, general features of the universe were virtually guaranteed as it developed over time: it would give rise to particles, atoms of hydrogen, nuclear fusion and stars, supernovae and the heavy elements, later generation stars with planets, complex carbon chemistry on planetary surfaces, and eventually, biological evolution and somewhere at some time, persons of some sort. God caused, and intended, all these things, but all this is a matter of general intentions, goals that can be realized in many different ways. The specific ways the general creative aims were achieved was left to chance. No specific realization was at all probable. God intended a universe containing stars, planets, living things, and persons, but it is thanks to indeterministic natural processes that the star we call the Sun, the planet we call the Earth, and dogs, cats, the human species, and you and I actually came into existence. As disconcerting as this may be if we are in thrall to the idea that God must have designed all that was created, it is, as argued in chapter 4, entirely reasonable if God’s aim was persons truly distinct from their Creator. It is no surprise that the God of Christian faith caused us but did not intend us. The universe as it has developed answers to our Creator’s general intention to create persons. To return to the parental analogy, when a human couple gives birth to their child, the fact that she exists
188
Chapter 6
does not realize a specific intention, for they had no such intention. Their intention was appropriately general. They intended to bring a human person into existence and left just which of the many possible persons it would be to indeterministic natural processes. For the project of causing persons to exist, to have only general intentions, to refrain from design, and to rely on indeterministic means are the same thing. As we saw above, indeterminism does not give us incompatibilist free will, but it does preserve compatibilist free will in a universe created by God. RESPONSIBILITY SHARED Human beings enjoy freedom of a kind possible for creatures, a freedom that involves no escape from the matrix of natural cause and effect. This compatibilist conception of free agency implies that we are not the absolutely original initiators of our actions. That involves an impossibly high view of human nature, one that ascribes to creatures the agent causation properly ascribed only to the everlasting God who transcends nature and its causal laws. If there is a candidate for being the ultimate originator of human choice and action, it is God, since the initial act of creation is the first cause of all that we do. However, the Christian faith leads us to the view that creation’s future lies open in its particulars. The specifics of its development over the eons, including who will inhabit the world and what they will do in it, is not determined, and thus not intended by its Creator. God is the first cause, the most remote cause, of all that transpires in the creation, but not the ultimate originator of our choices and actions in the sense agent causation calls for, since they are not intended or foreknown. When, for example, Sam chooses to adopt a ferret, and this choice is the effect of a deliberative process in which he carefully examines and embraces various beliefs and desires as his own, he intends to adopt a ferret. When, at the inception of the universe, God set in motion the sequence of causes and effects that in due time cause this event, there was no such intention, nor does God intend for Sam to exist, ferrets to exist, or for Sam to adopt one. God’s intention was to bring into being a world in which Sam, the ferret, and Sam’s choice to give it a home was among a vast array of wildly improbable possibilities. Sam does choose to adopt the ferret, and in an obvious sense this choice does originate in him: his very own, reflectively considered reasons are its proximate cause. This is origination, though not the ultimate origination hypothesized by advocates of agent causation and libertarian free will. Earlier events caused Sam to have the beliefs and desires that caused him to choose to adopt the ferret, and those events, in turn, had earlier causes, the causal sequence regressing finally to God’s choice to bring the world into being. A full explanation of why Sam adopted the animal refers proximately to his very own reasons, and ultimately to the reasons for
Freedom
189
which God created the world, and to the immensely complex intervening history of natural causes. God and Sam have explanatory roles, but neither is the ultimate originator of Sam’s actions in the sense agent causation requires. No one is the ultimate originator of Sam’s actions in the sense agent causation demands. The incompatibilist will object that this account implies that human agents, not being the ultimate originators of their actions, are not ultimately responsible for them. Rather than contest this, we should agree. There is only one First Cause, but the Creator chooses to create an indeterministic world in which not all that happens is divinely intended. The Creator, acting for love’s sake, steps aside. Pursuing fellowship with created persons, God devises a world in which some things are left to chance, the future stands open, and personal creatures act on reasons that are uniquely theirs, not their fellow creatures’ and not their Creator’s. They are persons, members of a community of persons each with his or her own point of view and reasons for action. Rational agents, responsive to reasons and always subject to being called upon to explain and justify their deeds, they engage in the discourse of moral evaluation. They identify themselves and others as morally, as well as causally, responsible for what they do. Still, the free actions for which humans are morally responsible are always the effect of reasons—beliefs and desires—that they did not choose. No matter a created person’s success at embodying rational self-reflection, and thus becoming the material image of agent causation, on the final accounting, no creature can be solely and absolutely responsible for who she is or what she does. 19 The only absolute originator of action is God. We humans possess the compatibilist freedom and responsibility available to created persons. We are not the absolute initiators of our actions; the causal chain does not begin with us but runs through us, back to God. We originate our actions in the way possible for material beings. To say who is responsible for human action, we must refer both to human persons and to the God who created them. Created persons and their Creator share responsibility for what we do and what we are. This conclusion is entirely at home in the Christian faith. God, we confess, is faithful. Even when things go very badly, God never denies responsibility for creation. God responds in sorrow and anger to our destructive behavior, but never abandons us. God intervenes, taking action on our behalf to reconcile, redeem, and restore. God does not give up on the project of becoming incarnate, making a home with creatures, but persists, becoming incarnate as their savior. The great mystery of the Christian faith is, of course, the death of Christ. The full meaning of this event transcends our attempts at theological explanation, but in it we see God taking responsibility for a creation gone wrong. 20 An unholy alliance of religious and political authorities puts Jesus of Nazareth to death for blasphemy and sedition. Yet the innocent victim on the cross is the
190
Chapter 6
crucified God, whose death atones for the sin of the world. God comes among us in humility, as a human individual, giving us our creaturely freedom, giving us space to do our worst. We do it, but God overcomes even that, saving us from ourselves and setting the stage for the final triumph of God’s creative endeavor: creatures and Creator in everlasting communion. NOTES 1. The Material Image shares various themes with, but on this major issue disagrees with, what is known as “open theism.” 2. The traditional paradigm of human action in which the person consciously deliberates, considers alternatives, and then chooses what to do, is an idealization. In most cases we act much more spontaneously; the preferred action is obvious and we simply do it. Further, the human individual’s capacity rationally to control her behavior, and thus function as a free, morally responsible agent, involves a great deal of neural computation, primarily in the frontal cortex. The causation involved here is exceedingly complex, most likely exhibiting “top-down” causation in a self-organizing, non-linear dynamic system; see, e.g., Warren S. Brown, “Nonreductive Human Uniqueness: Immaterial, Biological, or Psychosocial,” in Nancey C. Murphy and Christopher C. Knight, eds. Human Identity at the Intersection of Science, Technology and Religion (Burlington, VT: Ashgate, 2010), 81–93. Such complexity, though not extricating human action from the realm of natural causation, does render much of human behavior for practical purposes unpredictable when described in neurophysiological terms. It is the simplification that the application of intentional concepts, such as desire and belief, provides that makes behavior often predictable and intelligible, irrespective of our ignorance of the underlying causal structures. 3. Many cognitive scientists and philosophers believe that an evolved mental module embodies this belief-desire psychology, our innate theory of mind, our “folkpsychology;” see, e.g., Stephen Pinker, How the Mind Works (New York: W. W. Norton and Company, 1997), 24–5, 60–64, and Bertram F. Malle’s comprehensive analysis in How the Mind Explains Behavior: Folk Explanation, Meaning, and Social Interaction (Cambridge, MA: MIT Press). 4. This type of case appears in Donald Davidson’s seminal 1963 essay “Actions, Reasons, and Causes” in his Essays on Actions and Events (Oxford, UK: Oxford University Press, 1980), 3–20. 5. Later we will ask whether construing causation as fundamentally indeterministic, so that a cause just fixes the probability of its effect, makes a difference here. 6. Patricia Churchland, Touching A Nerve: The Self As Brain (New York: W. W. Norton and Company 2013), 180. 7. This thought experiment depends on the intelligibility of there being a detectable neural precursor to a choice. A well-known series of experiments indicates that this not only intelligible, but actual: Benjamin Libet, Mind Time: The Temporal Factor in Consciousness (Cambridge, MA: Harvard University Press, 2004). 8. This type of example is due to Harry Frankfurt, “Alternate Possibilities and Moral Responsibility” in Harry Frankfurt, The Importance of What We Care About: Philosophical Essays (Cambridge, UK: Cambridge University Press, 1988), 1–10. This has generated a large literature, much of it carefully analyzed in Derk Pereboom, Free Will, Agency, and Meaning in Life (Oxford, UK: Oxford University Press, 2014), 9–29. 9. For a sophisticated, but I believe unsuccessful, attempt to ground free choice in the brain’s amplification of undetermined microscopic events see Robert Kane, The Significance of Free Will (Oxford, UK: Oxford University Press, 1998).
Freedom
191
10. Nancey Murphy, “Introduction and Overview” in Downward Causation and the Neurobiology of Free Will, Nancey Murphy, George Ellis, and Timothy O’Connor, eds. (Berlin: Springer, Berlin, 2009), 1–28. Murphy and others go on to make the worthy point that the conceptual resources of neuroscience, let alone physics, do not suffice for an adequate explanation of human behavior, and thus that there is no possibility of neuroscience producing a deterministic account of what we do; see Nancey Murphy and Warren S. Brown, Did My Neurons Make Me Do It? Philosophical and Neurobiological Perspectives on Moral Responsibility and Free Will (Oxford, UK: Oxford University Press, 2007). This does not, however, have any bearing on the physicalist view that the physical state of the universe at one moment fixes the probabilities of its state at the next moment, i.e., it does not extricate human action from natural causation. The human mind/brain, like all material things, is governed by the laws of nature, but as is the case with most material things, there are no physical laws that refer to it as a mind/ brain. 11. If humans were agent causes, this would mitigate the implications that if we are designed we are not distinct from our Creator. Even if we are designed and inhabit a deterministic world, our actual choices could not be intended or foreknown by God. That we almost certainly are not agent causes reinforces the claims of chapter 4. 12. “The Impossibility of Moral Responsibility” Philosophical Studies 75 (August 1994), 5–24. 13. “Human Freedom and the Self” in his On Metaphysics (Minneapolis: University of Minnesota Press, 1989), 12 14. See Peter van Inwagen, An Essay on Free Will (Oxford, UK: Oxford University Press, 1983), 135–36. 15. To become a person is to inhabit the “space of reasons,” leaving our fellow creatures entirely in the “space of causes.” This is not to deny that the reasons on which we act are effects and causes. But reasons are subject to normative evaluation while other causes are not. A significant achievement of modern philosophy is the displacement of the distinction between the physical and the non-physical with that between the space of causes and the space of reasons as crucial to understanding ourselves. Prefigured in Kant, the classic source is Wilfrid Sellars, Empiricism and the Philosophy of Mind (Cambridge, MA: Harvard University Press, 1997), Sec. 36. 16. An alternative, suggested long ago by Boethius in The Consolation of Philosophy, c. 524 C.E., is to describe God as timeless. This implies that God has no foreknowledge since, for God, there is no future. God timelessly knows everything about what we call past, present, and future. God timelessly—eternally—knows that on July 4, 2021, Sam chooses to go to the circus. Whatever merit there might be in the idea that God is atemporal—timeless, rather than everlasting in time—it does nothing for the freedom of creatures; on the contrary, it imperils it. A timeless deity is immutable, and nothing that happens in the creation affects this creator in any way. This Creator knows what happens in creation only insofar as it is included in the one timeless creative intention. The Creator knows that on July 4, 202,1 Sam chooses to go to the circus, but only because this is an effect of the one timeless act of creation guaranteed to realize that specific creative intention. This God knows what Sam will do in virtue of knowing what Sam must do, and it is what Sam must do because it is what God intends for him to do. God’s knowledge of Sam’s choice is not a case of Sam acting on God, but of God’s self-knowledge. If anything could occur in creation other than what the Creator timelessly intends, that Creator could not know about it. 17. However, if Sally systematically does whatever Marvin gives her reasons to do, reasonable doubt arises as to whether she really satisfies the conditions for compatibilist free action. We will suspect that she is choosing and acting not on the basis of rational deliberation, but that she slavishly does whatever he wants her to do, perhaps out of an unreasonable desire to please him. We will think he controls her. 18. And thus they might, or might not, be what God wants them to be. 19. It seems that compatibilism leads us to conceive freedom and responsibility not as given in nature, but as a social status. Our choices, free or not, are caused, and their
192
Chapter 6
causal histories are complex and not fully known. The concept of a choice being caused by one’s own desires and beliefs, and the concept of it being self-reflectively endorsed, often provide us with no firm grip on the complex reality that leads to choices. Often, when someone does something, no one, including her, knows whether she could have done otherwise. We cannot always clearly distinguish a free and responsible action from the unavoidable effect of genetic endowment, upbringing, and environment. The implicit social consensus is that in the absence of good reasons to think she could not have done otherwise—she appears at least minimally responsive to reasons—we regard her as responsible and expect her to assume responsibility for her actions. That responsibility, and with it freedom, emerge because we are adapted for the matrix of interpersonal relationships, that it is a forensic, not a deep, objective, feature of human beings, will to some seem an unacceptably shallow account of such an important aspect of human life. Trinitarian Christians, for whom reality’s fundamental fact is the three-person God, should have no such problem. 20. Appealing once more to the parent-child analogy, we might think of the incarnate God as like a conscientious human parent, not to blame for the misdeeds of her foolish and wicked child, but putting herself in harm’s way to rescue him because she bears causal, and assumes moral, responsibility for him, although she never intended or foreknew the trouble he would get into.
SEVEN Morality
EXPLAINING MORALITY Contemporary science leaves no room for serious doubt that humans are material beings, fully implicated in the causal order of nature. The freedom and responsibility that can be ours involve no transcendence of the physical universe. Yet for some the hope remains that the encroachment of science into what is essentially human can finally be halted at our moral nature, our capacity to identify, and respond to, actions and situations as right or wrong, fair or unfair, just or unjust, and so on. Perhaps, as moral agents, we escape the reach of scientific explanation and as moral beings connect to something that transcends physical reality. Often, Christians and other persons of faith feel there is more at stake in preserving morality from scientific inquiry, for in their view it is as moral beings that we relate to God, the voice of morality is the voice of God, and our relation to God depends on responding to a divine call to moral virtue. A serious challenge to this arises from evolutionary psychology. It describes the moral nature of human beings as a product of natural selection, no more than an evolutionary adaptation to social life. Our capacity for moral sentiment, moral cognition, and our disposition to behave morally are features of our evolved biological nature. In this chapter I outline this naturalistic account of morality and then proceed to the widespread objection that it is implausible because it debunks morality. The fear is that this, and probably any, theory that naturalizes morality means that it is unreasonable to let moral considerations trump self-interest. We will see that these concerns are largely unfounded, and that in laying bare the origins of our moral sensibilities science does not do away with reasons to behave morally. However, while science does not seriously 193
194
Chapter 7
threaten to debunk morality, it does dethrone it, rendering unlikely a divine or transcendent origin of the moral truth. Indeed, it imposes limits on the sense in which morality can be objective, rather than dependent upon human beings. The remainder of the chapter challenges the assumption that Christians should resist this extension of the naturalizing project and oppose its demotion of morality. Christians have no more reason to reject the evolutionary explanation of morality than to hold out for the dualist view of human nature, the miraculous origin of the human species, or freedom free of natural causation. On the contrary, Christians have good reason to welcome what contemporary science tells us about the mundane origins of our moral capacities. The scientific project undermines claims for the ultimacy of morality and “puts it in its place” as a natural, merely human reality. This, I argue, coheres better with the Christian faith than traditional views of morality. Admittedly, this calls for reconsideration of some familiar views about God’s connection to morality, and thus morality’s relation to the Christian faith. In particular, two traditional theological doctrines, the Fall and original sin, which have long been shaped by traditional assumptions about morality, appear in a new light once morality has been properly located and its pretensions chastened. As in other contexts, the theological revisions that make sense in light of what science now tells us about ourselves are ones we have good reasons to accept on grounds independent of science. The idea that morality can be explained as a product of the evolutionary history of the human species must be distinguished from other ideas. It is helpful to distinguish different types of question about morality: 1. Normative questions are our familiar questions about what’s right or wrong, good and bad, fair and unfair, and so on. “Was it wrong to lie to Sally about what Marvin said about her?” “Should Jack be executed for murdering Harvey?” 2. Metaethical questions, which pertain to the nature of moral truth. Here we ask not what is right or wrong, but what makes the moral truth what it is: why is what is right, right and what is wrong, wrong? What is its ground? Here we ask whether the moral truth is objective, independent of human beings, or relative, somehow dependent on us. 3. Questions of moral psychology, which are about the nature and origin of human moral emotions, concepts, judgments, reasoning and behavior. Of these, only the third lies in the domain of scientific inquiry. For science to “explain morality” is for it to explain how the human species acquired its capacities for moral sentiment, cognition, and action. This project is distinct from attempts to derive the normative truths of morality from
Morality
195
science. Scientific inquiry might provide knowledge that, when conjoined with moral principles, enables us to reach normative moral conclusions, e.g., we take account of what neuroscience tells us about the functioning of the brain when faced with moral decisions about the care of a comatose patient. But scientific knowledge on its own never answers our moral questions. Science is the enterprise of explaining what we see in the physical world by finding natural causes. It makes no judgments about what we ought to do. 1 The scientific program this chapter examines purports to explain how there came to be creatures that know, and care about, the moral truth. It does not purport to instruct us in the moral truth. An evolutionary account of the human capacity for moral cognition and behavior is no more a source of moral knowledge than an evolutionary explanation of our mathematical capacities will tell us the square root of 117 or whether the Riemann Conjecture is true. Answers to mathematical questions come from mathematical reasoning, not from evolutionary biology. Similarly, answers to moral questions come from moral, not scientific, reasoning. Science can tell us how human beings came to think about, and care about, whether we should lie to save Mike’s feelings, encourage Bob and Ray to marry, or euthanize Uncle Bruce, but it cannot tell us what we ought to do. Evolutionary theory might explain why some human beings have innate desires to engage in, or avoid, certain kinds of behavior, but this implies nothing about whether that behavior is morally right or wrong. Nor can science answer metaethical questions. However, although metaethical questions about the nature of the moral truth lie outside scientific inquiry, the results of scientific inquiry can constrain answers to them, rendering some implausible. It may make some conceptions of the nature of moral truth unreasonable. As we will see, the discovery that human moral psychology is a product of natural selection bears on the issue of whether, and if so in what ways, the moral truth is objective, part of the fabric of reality, independent of us, rather than in some manner dependent on us. From the constraints scientific explanation imposes on metaethics arise the worry that we do not really have the powerful reasons to do what is right that we seem to have. I will contend that this concern is unjustified. EVOLUTIONARY MORAL PSYCHOLOGY Human beings have the ability, and often the inclination, to do things for the sake of other persons. We behave altruistically. At its heart morality is a body of norms that require us to act altruistically, constraining the pursuit of our own interests and directing us to pay heed to those of others. To do what is morally right, we must sometimes refrain from
196
Chapter 7
doing what we want to do, because doing it would harm someone else, and on occasion we must do things that are costly for us in order to benefit others. There may be more to morality than altruism but the idea that we must forego benefits to ourselves and incur costs to avoid harm to others lies at its core. Explaining altruism is the principal task scientific naturalism faces when it tries to find a place for morality in the natural world. There can be a natural scientific explanation of the moral nature of human beings if, but only if, science can explain why human beings have both the desire to act altruistically and the conviction that they sometimes have decisive reasons to do so—that this is what they ought to do—even when they have competing desires not to do so. Although they are strengthened or attenuated and generally shaped by patterns of moral education that differ across cultures, the altruistic tendencies that lie at morality’s heart appear to be innate in human beings. Empathetic feelings, and expectations that others will be similarly moved by one’s own needs, manifest themselves in early childhood, they occur spontaneously, without benefit of conscious calculation, and they occur across all human cultures. Simpler analogues of human moral emotion and behavior are observed in the non-human primates. 2 We reasonably seek an evolutionary explanation of such characteristics. However, altruism, even of the most primitive sort, seems made to order for resisting evolutionary explanation. Because of fortuitous mutations an organism might bear genes that dispose it to behave altruistically, e.g., it utters a warning cry at the sight of an approaching predator. This rudimentary altruism does not depend on self-awareness or calculation. We need not suppose that the individual that utters the warning cry intends to warn others, or even that it is aware of doing so. Its behavior is altruistic because it alerts nearby members of its species to flee and save themselves although it thereby draws the predator’s attention to itself. 3 As a result, it, and the genes for this altruistic behavior, are eaten and copies of its genes do not reach the next generation. Other members of its species, having escaped the predator thanks to its self-sacrifice, live to send their genes for non-altruistic behavior to future generations. It appears that natural selection will blindly consign such genes to oblivion. As Nevin Sesardic pointed out, “inveterately altruistic creatures have a pathetic tendency to die before reproducing their kind.” 4 At face value, there is no way for altruism to gain a foothold. The species as a whole might be better off if each of its members bears a copy of a gene for altruism, but natural selection lacks foresight. It is not obvious how it could possibly produce altruistic creatures. For a century after The Origin of Species, any sort of altruism, and certainly the full-fledged morality unique to humans, seemed safely beyond the reach of evolutionary explanation. A conceptual breakthrough in evolutionary thought in the 1960s profoundly changed the prospects for applying evolutionary explanation to
Morality
197
the human mind and culture, and to our moral nature in particular. A shift of the focus of explanation from the organism that makes copies of itself by means of its genes, to the genes that make copies of themselves by means of making organisms, dramatically enlarges the explanatory scope of natural selection. Natural selection can take the form of kin selection. Genes that dispose individuals to altruism that benefits close relatives are likely to propagate to later generations. Suppose that an individual carries a gene disposing her to sacrifice herself for the sake of her siblings, and that one day she does so. She dies, but as a result her siblings survive and reproduce. In a sexually reproducing species, the odds are at least one out of two that a sibling also carries that gene. A copy of the gene perishes but the probability that other copies will survive and be transmitted to the next generation increases. 5 The altruistic individual has greater inclusive fitness, even though the genes that increase her fitness in this respect might prompt her to sacrifice herself for her kin and reduce her own chances of survival to zero. Genes that cause altruistic behavior could go to fixation in a population comprised mostly of closely related individuals. 6 This is an evolutionary opening for altruism, but what is still needed is a strategy to explain trans-kin altruism, i.e., altruistic behavior benefiting those to whom one is not closely related. Cooperation is the route to trans-kin altruism Cooperative behavior is altruistic, since it calls for giving up something for the sake of someone else, even if in exchange for something one values. Mary, who has more apples than she needs, sees that Matt is hungry and wants one of her apples. She sees an opportunity to make herself better off. She admires his hat and offers an apple in exchange for it. Matt hands the hat over to Mary who, now in possession of both the apples and the hat, realizes that while she would be better off with the hat than with the extra apple, she is even better off with both and runs away, leaving him hungry and hatless. The presence in a population of many actors like Mary, unwilling to give up something for someone else even reciprocally, when they can cheat instead, makes cooperation impossible. Apples can be traded for hats only among reciprocal altruists, individuals willing to sacrifice something in exchange for something from someone else, i.e., willing to cooperate. A lone reciprocal altruist surrounded by cheaters will be no better off than an individual who gives what he has to strangers out of sheer generosity; his genes will be driven to extinction. Initially, the odds against reciprocal altruism taking root in a population seem dim. This is a coordination problem, a situation in which everyone involved would be better off if everyone cooperates, but no one cooperates because everyone knows everyone else has a motive to let others cooperate, but to defect herself, and thus receive the benefit without contributing. The mathematical theory of games investigates situations that have this structure. The simplest, and most famous, scenario is Prisoner’s Dilemma. Rein and Dietrik, partners in crime, have been ar-
198
Chapter 7
rested and are interrogated separately. Without a confession, the police do not have enough evidence to convict either of a major crime, but they have enough to put both away on a lesser charge. The police try to play one prisoner against the other to get a confession and thereby impose a long sentence on a least one of them. The possibilities: • If neither Dietrik nor Rein confesses, each serves a one-year sentence. • If Dietrik confesses but Rein does not, Dietrik goes free and Rein gets five years. • If Rein confesses but Dietrik does not, Rein goes free while Dietrik gets five years. • If both confess, both get a two-year sentence. Prior to being captured, Rein and Dietrik, knowing police tactics, agree not to confess if caught. Now each on his own must decide whether to do as he promised, to cooperate and remain silent, or to defect from the agreement, to cheat by confessing. Unfortunately for the culprits there is only one rational option: to confess. The risk of being the one who keeps silent while the other confesses is so great that it makes it irrational to stay mum, even though both know that they would both be better off if they both keep silent, and both know that the other knows this, and so on. If they behave as rational, self-interested actors, both confess and go to prison for two years, serving twice as long than if neither confessed. Individual rationality makes cooperation impossible. Natural selection is mindless, and involves no rational calculation, but the structure of the evolutionary dilemma is similar to prisoner’s dilemma. 7 All members of a species would be better off if each carried genes for reciprocal altruism, but from the standpoint of any individual, the rational course, i.e., what natural selection favors, is not to cooperate, but to be a “free rider,” feigning willingness to cooperate. Whether it is individuals rationally calculating their self-interest or the blind operation of natural selection on a population, the implication seems dismal: there is no route to a cooperative regime in which all concerned are better off. However, a hopeful prospect emerges for iterated games. When individuals interact not just once, but encounter one another many times, and how one acts in later encounters depends on how the other has behaved in previous encounters, reciprocal altruism—cooperation—can take hold and come to dominate a population. Individuals can proceed through a sequence of encounters with others guided by various strategies, e.g., always cooperate, always defect, and the payoffs of the various strategies can be calculated. Computer simulations demonstrate that the winning strategy, over a wide range of situations, the one that brings the best payoff in the long run, is tit-for-tat, i.e., begin by cooperating and on the next encounter with that player do whatever he did on the previous round. 8 In their initial encounter, Matt tried to cooperate with Mary, who cheated
Morality
199
him, defecting from the agreed-upon exchange, making off with both her apple and his hat. Employing the tit-for-tat strategy, when Matt encounters Mary again, he cheats her if he gets the chance. But if Mary has a change of heart, and tries to cooperate, on a third encounter Matt will return to cooperation and continue to do so, as long as Mary does not defect. If at some point she cheats again, Matt retaliates on the next round, but his strategy holds open the possibility of a return to mutually beneficial behavior in the future. For creatures with the requisite cognitive capacity to keep score and to generalize, the expectation generated by altruistic acts becomes open ended. Gort, an ancestral individual living on the African savanna, has a good day hunting. He returns to camp and shares his extra rabbits with Klatu, who has had no luck. When a week later Klatu finds some especially nice rocks suitable for shaping into hand tools, Gort expects him to share and is indignant when he refuses. If he won’t share, Gort most likely will refuse to share the next time he has extra rabbits. The fact that Klatu is stingy with rocks, not rabbits, does not mitigate Gort’s sense of having been cheated, and it will not keep him from refusing to share with Klatu in the future. There is no need to suppose that Gort engages in self-interested calculation, that he initially gives the rabbits to Klatu because he wants to create an obligation in him. Animals incapable of engaging in this kind of abstract thought can be reciprocal altruists. Gort’s mind/brain need only be so constructed that he wants to help others and later, when those he has helped do not reciprocate, he wants to respond negatively—he feels he has been wronged—and he wants to punish that individual in subsequent encounters. Gort is rationally self-interested in that the strategy implicit in his behavior is the winning tit-for-tat, but he need not be thinking in these terms. The strategy, or indeed any calculation of his interests, need not be explicitly represented in his mind. He may be moved by pure fellow feeling in the first instance, and in the second by indignation at having been mistreated. For him, it is simply obvious that people should help one another in these ways. So far, this is direct reciprocal altruism: individuals keeping track of their interactions with specific individuals. A further step leads to indirect reciprocal altruism: individuals act for the benefit of others not because doing so is likely to lead to reciprocity from the specific individuals on whose behalf they have acted, but because doing so builds a good reputation. Gort, because he habitually shares his catch, is generally regarded as the sort of helpful, reliable fellow with whom others want to cooperate and, when the need arises, to help. His reputation as someone who pays attention to others’ needs, even when doing so is not in his immediate interests, would be among his valuable assets, one he would be inclined to protect and enhance. For ancestral humans, survival, and thus reproductive fitness, was crucially dependent upon the good will, help, and
200
Chapter 7
cooperation of the small band with whom one spent one’s entire life, and thus there were strong selection pressures in the direction of indiscriminate altruism toward members of one’s group. Recognizing opportunities for cooperation, and detecting and punishing free riders who benefit from others cooperating without doing so themselves, skills requisite to navigating the Stone Age social environment, were no less valuable than the skills deployed in navigating the natural environment. Again, we need not assume Gort has, as his reason for acting altruistically, the desire to enhance his reputation. He might have that desire, but the theory does not require it. Indeed, what might most enhance his reputation is others believing that he is not calculating his interests, but simply wants to help, and the most effective means to getting them to believe this is for it to be true. Nor does the theory require that when he desires a good reputation, this is the reason he acts for the sake of others; it might be that he would act this way even if he did not give a fig for his reputation. The theory requires only that Gort, because of his genetic endowment and the neural circuitry it produces, wants to help the people around him and expects them to do likewise. He cares about others, wants to cooperate with them, and will sometimes put their needs over his own. If he reflects upon why he does this, he finds it self-evident that people being in need is a good reason to help, and that he has a good reason to abide by the agreements he has entered into, simply because cheating is bad. Gort identifies actions, individuals, and social situations as having moral characteristics. Klatu is stingy. Barada is generous to his friends but treats his mate badly. Nikto is a liar and not to be trusted. Yesterday’s parceling out of the fish was fair, but Moog got more bananas than she deserved, etc. The detection of these features of his social world arouses various emotions and prompts him to act in certain ways. It constrains his own behavior; he feels guilty when he realizes he has behaved unfairly. Others’ misbehavior moves him to anger and contempt. When he knows others are behaving badly, he is angry and moved to try to influence how they act, blaming and imposing punishments. Further, he will be disposed to impose second-order sanctions, punishing those who fail to impose sanctions on wrongdoers. Gort has become a moral being. 9 Natural selection adapted human beings for life with other human beings, providing us with a complex set of dispositions to categorize human actions, and the social situations they lead to, in certain ways, and to respond emotionally and behaviorally to them thus conceptualized. When, upon reflection, we take note of the moral features of the world and the responses that seem appropriate, we engage in moral reasoning, considering the moral reasons for or against courses of action, evaluating our own behavior and that of others in moral terms. To be a creature endowed with these moral dispositions living among similar creatures is to inhabit a realm of moral meanings. It is to be party to a scheme of
Morality
201
sanctions, to be subject to the demand for justification, criticism, praise and blame, rewards and punishments, and to have the status of a morally accountable person, and thus, in the naturalized sense of the preceding chapter, to be a morally responsible free agent, a being responsive to reasons. This responsiveness to moral reasons is a function of information processing structures (modules) in his mind/brain by means of which natural selection has adapted him for life with others. Gort has not become a moral being on his own, but as a member of a community. 10 How people ought to act toward one another, issues such as whether we should endorse or reject the way Barada treats his mate, how we should divide the fish, what we should do to people who start fights, and so on, and on, are not just matters of perennial human interest, they are matters on which consensus is necessary for harmonious life in a small group. We can plausibly imagine our ancestors debating these issues over the prehistoric millennia. What comprises the innate core of morality, and what is lore honed, made consensual and passed down through the generations, is not always clear, but both are indispensable. The assertion that evolutionary psychology explains human morality is tempered by the fact that significant parts of the explanation lie not in biology, but in anthropology and history. 11 But the innate altruistic dispositions, on which evolutionary explanation focuses, underlie all the rest. Gort experiences, and is moved to action by, a complex array of interrelated evaluative emotions and judgments. Some of Gort’s far future descendants will be ethicists, striving to delineate clear boundaries between moral concepts, judgments, emotions, and behavioral dispositions and those that are not matters of morality. This may be challenging when they see that clear lines are often not easily drawn and, when they are drawn, the boundaries that distinguish morality from other norms differ from one culture to another. What in one place seems a morally irrelevant prohibition on eating animals of a particular kind, in another is regarded as akin to the moral prohibition of theft or murder. They find the claims of morality tangled with those of the local religion, as gods, spirits, or ancestors are invoked to monitor compliance and punish infractions. The requirements of morality are no less tangled with the various purity codes in terms of which groups form an identity which distinguishes them from others. Gort’s philosophical heirs eventually discern the outline of a common, somewhat indeterminate core human morality, and attempt to refine and articulate it in a coherent system of principles, but unsurprisingly they meet with only limited success. No one way of making explicit the dispositions to categorize, emote, judge, and act with which natural selection endowed human beings as they were adapted to social life is likely to be decisively superior to all others. Evolution has endowed human beings with a robust, yet rough and ready and somewhat inchoate, package of emotive, cognitive, and behavioral disposi-
202
Chapter 7
tions, not a logically perspicacious set of moral principles, but a complex skill needed for living with other human beings. 12 This naturally acquired skill is necessary if human beings are to live together peacefully in a world in which the Creator is not unambiguously, empirically present. GENUINE ALTRUISM? The strategy of The Material Image is to examine the most patently naturalistic plausible explanations available and ask whether, despite widespread expectations to the contrary, they cohere with the Christian faith, or are perhaps even antecedently probable in light of it. However, a widespread view is that evolutionary psychology’s account of morality does not deserve to be taken seriously because rather than explaining morality, it debunks it, and that to believe the theory is tantamount to denying that there are real moral motives, that there are true moral judgments, or that we have good reasons to be moral. We will consider these objections before asking how the theory appears from the perspective of Christian faith. An initial concern, one to set aside before turning to more substantial objections, is that the evolutionary account of morality portrays human beings as lacking genuine altruism and as acting always on the basis of self-interest, rather than as true moral agents. Evolutionary psychology advances explanations for contemporary human behavior that invoke competition among genes in the ancestral human, and pre-human, population. Some genes did better than others at enabling the individuals that carried them to navigate their social environments, and those individuals were more likely to attract mates, successfully raise children and thus transmit their genes to the future. Among the genes that increase inclusive fitness were those that endowed individuals with the array of affective, cognitive, and behavioral dispositions that constitute our innate moral nature. The purely mechanical processes of selection have resulted in a particular distribution of genes in the human species, and thus the behavioral phenotypes that characterize us today, including those that comprise our innate moral psychology. Such is the mundane, and perhaps uninspiring, origin of human morality. Popularizing the shift in evolutionary thought that began in the 1960s, Richard Dawkins introduced the now infamous term selfish gene: genes, acting with no regard for anything (as befits mindless molecules) act as though they were concerned exclusively with their representation in future generations. As illuminating as Dawkins’s metaphor was when he introduced it in the 1970s to highlight the fact that whatever happens in evolution is ultimately a matter of some genes being more successful than others at getting themselves copied, it has been a source of confusion. 13 The human mind/brain in general, and specifically the affective and cog-
Morality
203
nitive mechanisms that constitute our moral psychology, are products of genes that have always been out for themselves. Can unselfish persons result from selfish genes? Isn’t the so-called altruism that allegedly grounds human morality not altruism at all, but really self-interested behavior, mere prudence, not disinterested morality? 14 When Anna buys Randy lunch, does she act not out of generosity, but out of a calculation that in so doing she will obligate him to help her in the future, or a calculation that she is thereby strengthening her reputation as a helpful person, or even calculating that this loan of a few dollars somehow will incrementally improve her reproductive fitness? Invoking those ancient selfish genes to explain our behavior has no such implications. It is far from uncommon for human beings to do good for others, and to avoid doing them harm, without any thought of good or bad consequences for themselves. The presence within Anna of the emotional, cognitive, and behavioral dispositions that were activated by her realization that Randy was broke and hungry was due to the evolutionary history of the species to which she belongs, but this does not make her generosity less than genuine, a facade for self-interest. 15 To see how profoundly this objection is misconceived, consider an innate human disposition that is plainly a product of our evolutionary history: a mother’s love for her child. A human mother typically has intense spontaneous feelings of love for her infant, an intense desire to care for and protect it, and a tendency to act in a self-sacrificing way on its behalf. The adaptive value of this phenotype is obvious: the woman’s mind/brain was constructed by genes that are copies of those that were better than others at constructing organisms that were successful at reproduction, i.e., selfish genes. Shall we say in light of this that she does not really love the child? What occurs in her subjective experience are various feelings for, and beliefs about, her child; she loves it and believes it is precious. She spontaneously sees the infant being in need or danger as a compelling reason for her to act on its behalf. These mental states have causes, proximately in the neural structure and biochemistry of her brain, and remotely in the differential effect of natural selection on mothers who cared more, rather than less, about the fate of their offspring. The fact that her distant ancestors’ genes figure in the causal history of her feelings has no bearing on their genuineness. She really is ready to sacrifice herself for the child. To suppose otherwise is, in Steven Pinker’s words, to “confuse the real motives of the person with the metaphorical motives of the genes.” 16 Or consider a parallel case, where the behavior is not admirable: Jared is highly sensitive to insults, whether real or imagined, especially from other young men, and is prone to become enraged and violently retaliate. When he kills Sam in a bar fight, should the fact that there is a compelling evolutionary account of why young males are inclined to this kind of violence lead us to conclude that his rage was not authentic? Evolution
204
Chapter 7
explains why humans have certain innate emotional responses that, together with innate ways of conceptually organizing experience, often motivate them to behave in certain ways, some good, some bad, but these explanations do not deny the reality of what they explain. A related objection shifts from denying the reality of altruism because of its origins, to claiming that evolutionary moral psychology implies that the individual who appears to be acting altruistically really acts out of ulterior, selfish motives. Consider, first, an example remote from moral concerns. I like the taste of hamburgers and this desire is due in part to the neural circuitry that results from my genetic endowment. Evolutionary theory might propose an explanation of why my genes incline me to like the taste of hamburgers: in the ancestral population, people who liked the taste of meat tried harder to obtain it and thus ate more of it than those less enamored of it. Those who ate more meat got more protein and as a result were healthier and thus had greater inclusive fitness. Genes that built brains that desire meat were passed to future generations, including mine. That’s why—on this simplistic explanation—I like the taste of hamburgers. My current taste for hamburgers is an effect of my ancestors’ selfish genes. Part of their metaphorical strategy for getting themselves copied was causing people to like the taste of meat. If someone thinks that this implies that I eat hamburgers because I want to imbibe protein and be healthy, he fails to grasp the selfish gene explanation. It could be true that I eat hamburgers because I want to be healthy and I know that they contain protein and believe that eating it makes one healthy. 17 But that would not explain why I like the taste of hamburgers. Someone might eat hamburgers to be healthy despite disliking their taste. I liked the taste when I was a child, oblivious to the existence of protein and insouciant about my future health. The same is true of Vader who, being a dog, loves to snatch hamburgers off the grill, but not because she thinks this is a way to get protein and be healthy. Were someone to persist and say that the evolutionary explanation implies that I eat hamburgers in order to send more of my genes into future generations, this would raise the misunderstanding to an absurd level. Someone might want to eat hamburgers as a means to reproductive success, e.g., an obstetrician advises a pregnant woman that she needs more protein, so she orders a hamburger for lunch, but again, even if it were true, it would explain only why she decides to eat hamburgers, not why they taste good to her. Perhaps she dutifully eats the hamburger even though she would prefer a kale and tofu salad, but in most cases human beings eat hamburgers not for health, but for the taste, even when doing so is a threat to good health. Perhaps no one so drastically misunderstands evolutionary psychology’s explanations as to believe that human beings have actual reasons that mirror those metaphorically attributed to the ancient genes. Howev-
Morality
205
er, an exception seems to arise when what is explained is sexual behavior, and the connection between the behavioral adaptation and reproductive success is less remote. One psychologist, criticizing evolutionary explanations of the differences between the mating behavior of women and men, portrays them as telling us that the male’s “true purpose in life is to get his genes copied as often as possible,” and that females, in contrast, want to limit the number of offspring they have so as to ensure that the maximum number can actually be brought to maturity. 18 Human males, in comparison to females, are indiscriminate in their selection of sexual partners and more likely to be promiscuous, because they want to produce as many children as possible and calculate that this is the best strategy for doing so. This is so plainly false that this critic draws back and tells us that what the theory implies is that promiscuous males unconsciously want to send their genes to future generations and unconsciously calculate that promiscuity is the best means to achieve this end. Thus, the desires, calculations, and strategies metaphorically ascribed to the ancestral genes appear in the minds of their descendants today, albeit unconsciously. This is no more plausible than thinking that I like the taste of hamburgers because I have an unconscious desire for protein, health, fitness and the propagation of my genes into the future. Promiscuous men have sex with many women because they want sex, and typically do so despite the conscious desire not to cause pregnancy, or at least while not caring whether they impregnate their partners. Whatever exactly unconscious desires and strategies might be, someone might devise a theory that ascribes them to promiscuous men and thereby explain their bad behavior, but it would not be the kind of explanation on offer from evolutionary psychology. It seeks the ancient causes of our actual, conscious desires. Deep in the past, there were genes that metaphorically wanted people really to want sex, hamburgers . . . and to be moral. The genes’ metaphorical calculation was that placing these real desires in people was an effective way to get copied. They did get themselves copied, and we now have brains recent copies of those genes helped construct; as a result we really want sex, hamburgers, and to be moral. Evolutionary psychology casts no reasonable doubt on the reality of these innate desires. A familiar feature of human life is that people on occasion pretend to act for the benefit of others, but have ulterior motives and really act just to benefit themselves. If the evolutionary explanation amounted to an attempt systematically to expose such motives for seemingly altruistic, moral behavior, we would be entitled to find it implausible, as advising that in reality human morality does not exist, since we are incapable of real altruism. But the theory has no such implication. It describes human minds as being furnished with innate dispositions, e.g., upon seeing someone hurt, to feel empathy, to want to alleviate the suffering, and to be angry at whoever inflicted it. Human beings are born with such inclinations in virtue of genes that they inherit from their remote ancestors,
206
Chapter 7
and we inherit these genes because they enhanced fitness in the ancestral population. Then, individuals who had these moral sentiments did, as a matter of fact, have enhanced inclusive fitness. As a result we share their capacity for the moral emotions they incite and nothing in the scientific account of their causal history calls their genuineness into doubt. THE SUPERFICIALITY OF MORAL REALITY Any serious doubt evolutionary theory casts upon the morality it aims to explain arises not from what it says about the origin or nature of our moral capacities, but from what it implies about the place of moral properties in the world and the nature of moral truth. The metaethical implication is that moral reality is superficial: the moral properties, the goodness and badness, the rightness and wrongness, that we ascribe to people and their actions belong not to reality as it is in itself, but to the world as represented in human minds. We saw in earlier chapters that science reveals, and the naturalistic point of view embraces, the ontological superficiality of human persons. Human persons are not among the world’s fundamental constituents. Creation’s basic elements appear to be the particles, forces, and fields of physics, if not hypothesized strings or branes. Personal creatures’ privileged place in creation has to do with our place in God’s plans and purposes, not with being ontologically foundational. We are complex entities that exist only because contingent historical processes have arranged some of the world’s fundamental constituents in improbable ways. The world as human beings experience it is in many respects likewise superficial, and this includes the realm of moral values and meanings. The scientific explanation makes superfluous any thought of a distinctive type of moral fact, present in the world prior to and independent of humans being adapted to social life. It portrays the moral facts as parochial, depending on a specifically human perspective on objective reality. Crucially, there is no need to presuppose moral truths in order to explain how evolution brought forth creatures capable of making moral judgments. Setting out with a world containing no moral properties, evolutionary theory explains how there came to be creatures that ascribe moral properties to what they experience, categorizing actions as right and wrong, fair and unfair, kind and cruel, and so on. So far as the explanation is concerned, moral truth enters the world only as human beings represent it. It is not there in reality waiting to be discovered. The metaethical implication of the evolutionary explanation of human moral psychology is a kind of relativism. The moral truth depends upon—is relative to—human beings, but relativism appears in various forms, and it is vital to distinguish the ways in which truth can be relative. The truth about some things is relative to the individual (subjective).
Morality
207
Elizabeth believes that coffee tastes good, but Sally thinks it tastes horrible. Here there is only difference, no real disagreement, because the question as to who is right, and the objective truth about whether coffee tastes good or bad, makes no sense. For other matters, the truth is relative not to individuals, but to society or culture. Individual relativism is colloquially signaled by “true for:” “Coffee is good” is true for Elizabeth but false for Sally. Elizabeth cannot be wrong in believing that coffee tastes good, because here what is true or false depends on her alone. In contrast, what she is wearing might seem fashionable to her, but she can be mistaken. The truth about fashion depends on human beings, but it does not depend on the individual. The truth about what it is fashionable to wear is relative to culture; it depends in some complicated way on the attitudes and behavior of many interacting individuals. However, although any given individual can be wrong about what is fashionable, the culture as a whole cannot be. People today might find what people wore in the 1970s weirdly unattractive, but they cannot seriously believe that they have finally found the truth about these matters where their predecessors erred. There are no objective truths of fashion; the truth depends on human beings. The evolutionary account of our moral capabilities makes plausible a third kind of relativism: morality is relative, but not to the individual, nor in its core, to society or culture, but to the human species. The moral truth is species relative. We are justifiably wary of claims about the relativity of moral truth. If anything in this domain is obvious, it is that the moral truth does not depend on the individual, that while Randy believing that beer tastes good is all it takes to make it true that beer tastes good, his believing that it is morally good for him to suffocate his roommate to make the snoring stop does not make it morally right for him to do so. We can properly say that “Beer tastes good” is true for Randy, but if we say, “Killing someone to make him stop snoring” is true for Randy, now all we can plausibly mean is that this is something he mistakenly believes. Social or cultural moral relativism is almost as unreasonable as individual relativism. Most persons in some society might believe that they are morally justified in enslaving some minority group, but this implies that a false moral belief is widespread there, not that this is a society in which slavery is morally right. We might report, “For them, slavery is morally permissible,” but all we mean is that this is a society where many people have the false belief that it is morally right. In practice, caution is called for when it comes to judging that a society’s moral beliefs are false. What common human morality calls for in a given situation can be highly sensitive to local non-moral facts, and practices that would be morally wrong under conditions prevailing in our cultural setting might not be wrong, or not obviously wrong, in a different setting. Nonetheless, in general, we properly reject the idea that what is morally right or wrong
208
Chapter 7
depends on what humans, either individually or socially, happen to believe. Both individual relativism, and social or cultural relativism, would, if true, seriously undermine morality. It is not clear what remains in our practices of moral reasoning and evaluation if we can no longer intelligibly criticize individuals, or groups of individuals, even whole societies, as having false moral beliefs. Morality’s point is to constrain what human beings do to and for one another, to move them to act in altruistic ways. These versions of relativism leave it powerless to motivate altruism. If Mike wants to take Sally’s coffee, and his believing that it is right to do so makes it right for him to do so, then he has no moral reason not to do it. If the majority in a country feels entitled to subjugate a minority, and their believing this makes it true, then they have no moral reason not to do so. If the moral truth cannot be insulated from what individuals and societies want, then it cannot give us reasons to do things we do not want to do. Evolutionary psychology’s explanation of human morality offers no support for either individual or cultural relativism. Portraying our moral psychology as a product of our evolutionary history, it characterizes it as a human universal, and in so doing undermines these familiar versions of moral relativism. An innate core of affective, cognitive, and behavioral dispositions that constitute human morality underlies divergence across cultures. The relativism it entails leaves intact the possibility of what morality calls for being at odds with what any individual wants to do, or with what any society approves of. The moral truth does not depend on the individual, or on society, but it does depend on something we can call human nature. 19 The truth about moral matters, like the truth about various other things, came about, in the course of evolutionary history, contemporaneously with the human species. A salient case that is in some ways analogous is color. Humans inhabit a world of colored objects, yet this pervasive feature of the world as we experience it is paradigmatically superficial and species relative. Color is a feature of objects as human beings represent them, not of the objects in themselves. We spontaneously categorize objects by representing them as one color or another. In complex ways the evolved representational scheme tracks ways in which the surfaces of objects reflect the various frequencies of electromagnetic radiation our eyes detect. These differences correspond to further differences in the physical constitution of the objects. Thus, for example, our ancestors acquired the useful ability to detect at a distance whether a piece of fruit hanging on a branch is ripe: not yet when it is green, but ready to eat when it is red. Evolving color vision, natural selection equipped us to detect subtle patterns involving physically heterogeneous things. Two red pieces of fruit might have nothing in common beyond the fact that it was adaptive to categorize them together. The same piece of fruit might cause a different color sen-
Morality
209
sation, or no color sensation, in species not adapted to detect the differences that matter to our, or at least to our ancestors,’ fitness. If the minds of hypothetical Martians represent ripe bananas as having a property very different from what we call yellow, it makes no sense to ask who is right, to wonder what color bananas really are, not from the human, or the Martian, point of view, but in themselves. The truth about color is superficial, relative to species. Colors and color vision came into existence together; colors did not precede our capacity to detect them. Color, despite being a prominent aspect of our experienced world, is ontologically superficial. Our color categories do not smoothly map onto the categories of nature left to its own devices; their application is restricted to reality as represented in the mind/brain of a particular species. This does not make the facts about color unimportant when it comes to practice. Superficiality does not translate into insignificance. When Fred shops for a car, the fact that a vehicle is painted pink might be of much greater moment than the fact that it has a mass of 2000 kg. He reasonably ignores someone who advises him that he ought not to care what color it is since, after all, color is superficial, unlike the car’s mass, not deeply rooted in objective reality but relative to the perceptual apparatus with which evolution has provided the human species. To turn to an example which brings us closer to morality, consider the fact that some things are disgusting. 20 I catch a fish, but rather than cleaning and cooking it right away, I leave it on the back porch. Returning to it after a few days anticipating a delicious meal, I am disappointed. The fish smells very bad; it disgusts me. I correctly ascribe to it the property of being disgusting. Of course, this would be true relative to a human perspective. The hungry cats milling about and trying to get at it find it very appealing. Nothing is disgusting in and of itself, but only relative to a point of view, one that is to some degree a result of one’s evolutionary heritage. 21 Presumably, in the past human beings who found rotting fish palatable were less likely to become our ancestors than those who gagged on it. Cats, products of a different evolutionary lineage, have no such evolved disposition to being disgusted by old fish and to the avoidance it motivates. That disgustingness is an objective property of the decaying fish, and not essentially tied to my reaction to it, is a compelling illusion, yet on sober reflection I know that nothing is disgusting in itself, and that there is nothing here that I get right that the cats get wrong. As with Fred and the pink car, the realization that the property in question is superficial does not make it not matter. I belong to a species adapted to have unpleasant sensations in the proximity of such things, sensations I want to avoid having and which I can avoid having in greater intensity only if I refrain from eating it. I might have further reasons not to eat it. My ancestors certainly did: those who ate decayed fish tended to get sick. Disgust in this instance is a subjective stand-in for the presence of disease causing microorganisms.
210
Chapter 7
Behavior that was adaptive for our remote ancestors might, but need not be, good for us now. My visceral aversion to the rotting fish might be like my liking for red meat: once adaptive but now maladaptive. Even if— implausibly—my only reason to eschew decaying fish are these unpleasant sensations, my experience of it as disgusting remains a good reason not to eat it. Knowing the evolutionary origins of one’s disgust reaction, and thus recognizing that it involves no necessary perception of inherent danger in the object, does little to weaken one’s resolve to avoid what disgusts us. 22 Natural selection has endowed us with mental machinery that draws us into the illusion that things in themselves are pink or disgusting, but learning this does not extinguish, or even seriously weaken, our reasons to avoid pink cars and rotting fish. The negative reaction itself is a good reason to avoid it. Scientific revelations of ontological superficiality may have little or no practical import. In many matters they leave our reasons to act in response to species-relative properties of things intact. Humans, in virtue of being adapted to social life, have the capacity reliably to identify certain acts as, say, morally wicked. The property of being morally wicked is superficial. The category is an artifact of our conceptual organizing of what we perceive. This does not mean that it is trivial or arbitrary, like, say, the category of things that go bump in the night, or the category of things within ten meters of Marvin on February 29, 2021. The human mind is built to structure the experienced world in light of human interests, or at least the metaphorical interests of ancestral genes. Scientific inquiry, the attempt to “cut nature at its joints” and so discern natural laws and the natural kinds that figure in them is one of our interests among others. Framing and deploying concepts which are superficial and relevant only from a distinctly human perspective on the world is not capricious. Natural selection has produced minds wired to use concepts that empower us to better cope with the real natural and social world. Being supplied with color concepts enables us better to deal with the physical world, to find the ripe fruit and avoid the unripe. Telling right from wrong, like telling green from red, is an adaptation that made our ancestors better at reproducing, despite the fact that the qualities themselves are not independent of us. Superficiality and species relativity do not imply a complete lack of grounding in reality. Concepts for which there has been selection must be reliably grounded in the objective, natural world, however indirectly. Evolution leads to a species detecting patterns in the world only if they are really there, and only if discerning them enhances inclusive fitness. When I perceive a piece of fruit as red, a piece of fish as disgusting, or an action as morally wrong, the properties I ascribe belong to the superficial realm of the world as represented by a particular species, and have no place in a hypothetical “view from nowhere,” yet it is precisely by means of schemes of repre-
Morality
211
sentation replete with such idiosyncratic concepts that our species successfully makes its way in the world. The fact that the moral properties we detect in human social life are superficial does nothing to debunk morality. If Marvin condemns some act of Fred’s as morally wicked, expresses moral indignation at it, and proceeds to impose sanctions on him, he cannot reasonably tell Marvin that his cognitive, emotional, and behavioral response is a mistake in light of the fact that moral wickedness is a category not of things as they are in themselves, but a parochial way of organizing things contrived by human evolutionary history. This would be no more plausible than if on these grounds we were to advise someone to go ahead and eat the disgusting fish or the green fruit. If the evolutionary explanation threatens to undermine morality, the problem lies deeper than in the exposure of its superficiality and relativity to the human species. MORALITY DEBUNKED? Rummaging in his closet, Marvin discovers clothes that he wore in the 1970s. Delighted that they still fit, he dons them for a night out with friends. Unfortunately, while this outfit was the height of fashion in the days of disco, it is now drastically out of fashion. We helpfully point out Marvin’s fashion faux pas, but once he makes it clear that he is indifferent to such considerations, there’s no more to say. He has a reason not to wear these clothes only if he cares about dressing fashionably or about the consequences of failing to do so. Absent some connection with Marvin’s interests—with what he wants or needs—the facts of fashion are, for him, volitionally inert. The demands of morality are not so easily dismissed. Marvin announces that he is fed up with Sally’s howling cat and intends to poison it. We admonish him, pointing out that killing his neighbor’s beloved pet would be wrong, but he replies that he does not care about morality. This hardly terminates discussion. We say that when it comes to morality, in contrast to fashion, his lack of desire to abide by its prescriptions is irrelevant; he still ought to refrain from killing the cat. We insist that he has a reason to do what is right irrespective of his desires and interests. Morality’s prescriptions are, as Kant famously said, categorical, not hypothetical. The reasons of fashion do not survive the lack of desire to be fashionable, but moral reasons appear impervious to one’s disregard for them. Evolutionary theory explains how human beings came to have moral beliefs and it does so without assuming that they are true. With this the possibility of a debunking explanation arises: we can explain why someone believes something without having to suppose that his belief is true. One day, Gort believes that he sees a goat. That night, he believes that he sees a ghost. It would be interesting scientifically to explain both epi-
212
Chapter 7
sodes, and to discern the evolved mental capacities involved in them. We have different assumptions about the two explanations. We expect the scientific explanation of Gort’s goat detection capabilities to start with the assumption that the world contains goats that causally interact with human sense organs and brains. A successful explanation casts no doubt on goats; it presupposes them. In contrast, we expect the scientific explanation of Gort’s belief in ghosts to show that under certain conditions he believes that there is a ghost in the vicinity, whether or not there actually is one about. The explanation works just as well even if there are no ghosts for Gort to detect. This does not imply that there are no ghosts. It does not prove that Gort’s belief that he sees a ghost is false, but it undermines it. Explanations that explain beliefs without having to suppose that they are true tend to debunk them, to show that there is no good reason to have them. Rationally acquired beliefs are those we tend to have if, but only if, they are probably true. Beliefs that one would have whether or not they are true are not rational. Nicto catches a fish, and he wants to eat the whole thing rather than share it with Gort. Thanks to natural selection, Nicto is wired to believe that he ought to share—that he has a good reason to share—despite his desire not to do so. Maybe this prompts him to give away half his fish, even though he knows Gort will never be able to reciprocate, and that no one will know that he has done this good deed. Despite this, he believes that it is what he ought to do. The explanation succeeds whether or not, as a matter of fact, Nicto actually has an objective reason to share. The adaptive altruistic behavior comes about when human beings believe that they have reasons to behave altruistically not grounded in their needs and desires, and the belief does the job whether it is true or false. Evolutionary moral psychology is indifferent as to whether there actually are any categorical imperatives. It requires only that human beings believe that there are and act accordingly. Crucially, from a naturalistic point of view, these beliefs appear to be false. Rational agents, and their reasons, emerge naturally from a world that contains no agents, and no reasons—only causes. There are no practical reasons simply there in reality for us to discover. Reasons reside in the mind’s subjectivity, not in objective reality. 23 There is, as J. L. Mackie put it, no objective prescriptivity. The world does not supply prescriptions for action. They arise out of rational agency in pursuit of perceived needs and desires. 24 In light of this, the evolutionary account of morality as an adaptation to social life is an error theory. It explains how, and why, this cognitive illusion was foisted upon us. How to get someone to act contrary to his own best interests? By deceiving him. If we imagine natural selection trying to solve the problem of how to socialize rational, self-interested creatures, the solution which presents itself is to cause them to feel that they have powerful, objective, reasons that can trump their own interests. This belief could not possibly be true, but this does not matter; all that
Morality
213
matters is that those who have it act in the altruistic ways that enhance inclusive fitness. Human individuals are metaphorically tricked into adopting the metaphorical reasons of their genes as their own. Our selfish genes “want” us to sacrifice ourselves to safeguard our siblings, and to cooperate with strangers even when cheating is a viable option, so they construct a brain disposed to believe that this is what we have reasons, sometimes compelling reasons, to do. Given that the raison d’etre of morality is to move us to act even when doing so is at odds with our needs and desires, whatever they happen to be and however strong they happen to be, the strategy of devising the appearance of a reason of a special kind, not tethered to anything we want or need, but objectively outside us, is not surprising. What follows from this scientific unveiling of the moral machinery? Does exposing the cunning of evolution reveal that we have no reason to be moral unless doing so happens to be in our interests? The power of science to erode morality’s practical force is, I believe, exaggerated. This becomes clear when we pay heed to some significant features of human moral psychology. DEBUNKING DEFEATED If natural selection has tricked us into being moral, the trickery has two components. There is the cognitive component: belief that the mere fact that an act is morally right is a reason to do it, and if it is morally wrong not to do it, irrespective of our interests. There is also the affective component: our emotional responses to behavior we count as morally relevant. When we judge that Marvin did something cruel when he poisoned Sally’s cat we probably do not do so dispassionately. His wicked deed makes us angry, and this moral indignation motivates action. We want to keep people from doing this kind of thing, to stop doing it if they are already doing it and, if they have already done it, to inflict punishment. Further, even though Marvin felt justified when, annoyed by the cat’s howling, he committed the crime, he may well in retrospect share our emotive response, although in him it takes the form of feelings of guilt that move him to make amends. Our moral judgments are often accompanied by strong emotional responses that motivate action. These desires—our moral sentiments—are in themselves reasons to do those things. The mere fact that one has a desire to do something at face value is a reason to do it. Therefore, evolutionary moral psychology indicates not that we lack reasons to do what is morally right, but that the reasons that we have are grounded in our evolved moral emotions, not in objective reality, despite the powerful innate disposition to believe otherwise. Should we conclude from this that we have no good reasons to be moral unless it happens to be in our interests?
214
Chapter 7
Scientific explanation debunks the belief that there are objective prescriptions. It thereby undermines the objective component of our moral psychology, and leaves us with reasons for being moral grounded in nothing deeper than our moral sentiments. The evolutionary account of our moral psychology dismisses the notion of objective prescriptivity and pushes us to acknowledge that our reason for being moral is that we want to be moral. Does morality survive this? Only if it implies that these are not good reasons, and thus that it is not reasonable to be moral. There seems no reason to believe that it implies this. Earlier, we noted that the evolutionary account of a mother’s altruistic impulses toward her infant casts no doubt on their genuineness. Returning to that example, we may suppose that she is inclined to believe that there is something ineffably wonderful about her infant, that it is absolutely and uniquely worthy of her love and self-sacrifice. She has compelling reasons to lavish care and affection on this marvelous creature and this, she initially imagines, has nothing to do with the fact that it happens to be her child. A normal human mother acts as though she believes that there is a categorical imperative, an objective prescription to act on behalf of the infant, even when doing so is at odds with what she wants and needs. However, on sober, scientifically informed reflection she recognizes that natural selection explains her disposition to these beliefs, as well as her emotional response to the infant. She sees that they are cognitive and affective sides of the same evolutionary coin. Mothers in the past who believed and emoted in these ways were more likely to tend to their child’s needs and thus had greater inclusive fitness. Suppose we suggest to the doting mother that her genes are exploiting her, making her believe that she has a good reason to care for the child, when she really doesn’t? Nothing but her feelings for the child stand behind her altruistic behavior, so she should pay heed to her own interests and let the child fend for itself. The proposal will strike her as preposterous. She wants the child to flourish, and this constitutes her reason to care for it. Prominent among the things she wants is the child’s well-being. However, human beings sometimes have desires that rational self-reflection leads us to realize that we are better off not acting on; sometimes, getting what we want is not in our interests. The conclusion that the mother should cease her altruistic ways depends on identifying her desire for the child’s well-being as one she is better off without. The fact that a person wants to do something is at face value a reason to do it. Further reflection might reveal it as one on which she cannot reasonably act and that it is best to disown it. A smoker might crave nicotine yet have a desire not to smoke; she wants to smoke, but she wants to not want to smoke. She finds a desire within her that she does not fully endorse as her own. Perhaps coming to understand her suscep-
Morality
215
tibility to peer pressure and to advertising has helped her identify her desire to smoke as not authentically her own. She feels she has been conned into wanting to smoke. She realizes that acting on the desire makes the satisfaction of other, more important, desires less likely. She realizes that if her health is impaired, or her life shortened, some of her other, much more important, desires will be unsatisfied. She may still act on her desire, but when she smokes, she knows she is lured into acting contrary to her interests. However, the bare fact that one desire is at odds with others does not make it reasonable to disavow it and to regard it as opposed to one’s interests. Ryan hates getting up early; he has a strong desire to sleep late. Unfortunately, this conflicts with his desire to be employed, to get paid, to buy food, and so on. It would be an obvious mistake to say that this shows that his desire to have a job and the things his pay buys is opposed to his interests, and that he should be advised that while he has a reason to get out of bed, it is not a good reason. Conflict of desires on its own tells us nothing about which desires one has a good reason to act on, rather than to disown and resist. No behavioral disposition is more obviously subject to being explained as due to our selfish genes than mothers’ love for their offspring, but this gives them no reason to doubt that they have good reasons to take care of their babies, even when doing so is inconvenient, unpleasant, or costly. She wants to sleep, to clean her house, to advance her career, and her desire to meet the child’s needs might conflict with these desires, but she rejects as ridiculous the idea that she should disown her altruistic desire to care for her child and acknowledge that she has no good reason to care for it. On reflection, she experiences her love for the child as a gift, grounding reasons for action which she did not originate but which she now emphatically endorses as her own. She regards her love for the child, and the self-sacrificing behavior it calls for, as an important part of what makes her her. Acting in the child’s interests is in her own interests, if anything is. Similar considerations apply to our commonplace moral behavior. Elizabeth is hungry and for this reason buys a hot dog from a street vendor. Two blocks farther on, a stranger asks her where she got it. She directs him toward the cart where she bought it. She does this because he asked for help and she wanted to help him. This minor altruistic episode is motivated by her mild, generally not even conscious, benevolence to her fellow human beings. She has no desire rudely to ignore him, to tell the stranger to get lost, or to misdirect him, lying about the location the hot dog vendor. Had such possibilities come to mind, she would have immediately rejected them. She does not want to be that sort of person. She wants to be the sort of person who helps people. She dislikes people who cannot be bothered to help others, or who play nasty tricks on them, and she does not want to be one of them. Earlier, if she had fleetingly
216
Chapter 7
imagined grabbing the hot dog and running off without paying, and realized that she could get away with it, she would have instantly dismissed this course of action too. She disapproves of people who steal and has no desire to be one. Elizabeth wants to be a decent human being much more than she wants a free hot dog. She finds the proposition bizarre that she should think getting a free hot dog without paying is more important than being honest. The claim that as a rational, self-interested person Elizabeth should see her desire for the hot dog as truly her own, one she has a good reason to act on, but that she should resist the desires to be honest, cooperative, and helpful as not truly her own, as desires she has no good reason to indulge, is groundless. It is arbitrary to suppose that because the desire to help others has been instilled in us by natural selection we do not have a good reason to act on it, while assuming that the desire to eat, also instilled by natural selection, is one we do have good reasons to act on. The debunking claim that in the absence of external social and legal constraints it is not rational to be moral is not so much false as it is empty. What’s true of all our desires cannot disqualify some and not others. Conceivably, we might discover some good reason to rid ourselves of our desire to be moral, but the evolutionary account of our moral psychology does not supply it. We do well to reflect, to deliberate, to seek explanations of why we want what we want, and thereby better advance our genuine interests. We sometimes find that we have been caused to want to do things we are better off not doing. We reasonably seek to diminish the role of these desires in our lives. This is the best human beings can do. We have no prospect of finding desires we have not been caused, by one route or another, to have. We cannot sanely aspire to act only on desires we have rationally chosen. No human being chooses what to want. We can at most endorse desires after the fact. We can no more choose what to want than choose what to believe. What fundamentally moves us to act is given. To reject desires merely because we were caused to have them, whether by the processes of biological evolution, or by any other means, is to presuppose the impossible. It reflects the impulse to deny that we are creatures, not ultimately self-made, but made by someone else. MORALITY DEMOTED What reason do we have to be moral? In the absence of social and legal sanctions, only that we want to be moral. Moral emotions, and the reasons for action to which they give rise, are constitutive of being human. Human beings characteristically judge them to be important, integral components of what they are and want to be. Evolutionary moral psychology reveals the origin of these practical reasons, but in so doing
Morality
217
does not show that they are not good reasons. There is no categorical imperative, no objective reason to be moral, but there remain good reasons, even if humbly grounded in nothing more than our evolved sentiments. The debunking argument fails, but it does not leave morality as traditionally understood pristine. It dispels the illusion that the imperatives of morality speak to us from beyond the human condition. 25 This becomes clear in the case of the hypothetical amoralist. Marvin, criticized for behaving badly, announces that he does not care about morality’s demands. Perhaps he simply lacks the moral sentiments that move normal human beings to action. Or he has them, but does not acknowledge them as good reasons to act. For him the desire to avoid harming others when it is in his interests, like the smoker’s desire to smoke, is something he judges he is better off without. He knows that the deed counts as morally wrong and that for others this is a reason not to do it, but he insists that this gives him no reason to avoid it. The account of human moral psychology on offer from evolutionary theory leads us to admit that what he says could be true. Marvin lacks reasons to be moral other than those supplied externally. The proper response to the amoralist most often is to deploy the power of social condemnation and, in extreme cases, judicial penalties to ensure that he does have a reason to do what morality requires. Most likely, the amoralist is merely hypothetical. When Marvin claims that he could not care less about morality we do not believe him, and in face of his denials insist, “That was wrong and you ought not to have done it!”, hoping that he will admit he really does see that we all have good reasons to avoid such acts. Still, no practical reasons reside in objective reality, so it is conceivable that his claims are true. As an error theory evolutionary moral psychology explains how we acquired the false belief that we have practical reasons completely independent of what we want. Some error theorists make a stronger claim. They contend that it is morally wrong for Marvin to kill the cat only if he has a reason to not to kill the cat independent of anything he wants or needs. If there are no objective prescriptions, then nothing is morally right or wrong. There are no categorical imperatives, so our moral beliefs are systemically erroneous. It is false that it is morally wrong to kill the cat. It is false not because in this particular case we mistakenly ascribe the property of moral wrongness to the act. This leaves open the possibility that other acts have the property of being wrong. The problem lies not in applying the moral concept to this act, but in the concept of being morally wrong itself. It refers to a property no act could possibly have. Objective prescriptivity is an essential feature of moral beliefs. The identification of moral facts purports to disconnect reasons for action from human interests, but they cannot be disconnected, so there are no moral facts. There is no objective prescriptivity, so nothing is morally right or wrong.
218
Chapter 7
This analysis, if correct, is unsettling, since some sort of moral nihilism seems the obvious implication. Yet we do not find error theorists exclaiming, “Yippee! No morality! I can do whatever I want!” Instead, they take pains to find some way for us to persist in our moral lives. Moral beliefs are false, yet we have good reasons, grounded in our interests, to act as though we really had moral beliefs. They are fictions we have good reason to treat as true. 26 But nothing can really have the properties moral truth requires. No practical reasons have a footing in objective reality. Whatever the feasibility of this strategy, moral fictionalism is not the only course open to us. The analysis of the meaning of moral language that motivates it, i.e., the view that nothing can be morally wrong or right unless we have reasons to do it, or not to do it, independent of our interests is not obviously correct. What obviously is correct is that to admit that there are no objective practical reasons denies morality a kind of power over us which, even if not essential to it, is a significant aspect of it. To go on making moral assertions requires pervasive revision in the meaning of moral language. In the wake of abandoning belief in categorical imperatives, to continue ascribing moral properties to human behavior requires us to recognize that our traditional concepts contain error and inconsistency, and that their continued deployment is reasonable only as they are modified. This would be akin to the conceptual revision scientific naturalism invites in other crucial domains. In the preceding chapter, we saw that while someone might, with some plausibility, insist that the concept of free action involves choices that have no natural causes, and thus that we must abandon the belief that human beings are free and morally responsible, we can more reasonably conclude that our traditional concept of freedom, insofar as it involves the idea that we are the ultimate originators of our actions, is flawed and needs to change in light of what science tells us about ourselves. The pattern appears in other contexts: someone claims that the concept of knowledge involves a certitude that is not available to evolved material minds, and thus infers that science leads, paradoxically, to skepticism. Someone claims that creation involves design, and from that concludes that science, in ruling out divine design, precludes creation as well. And, in our final chapter, we will contend against those convinced that the very idea of life after death presupposes a disembodied existence that contemporary science portrays as impossible. In all these cases, the better course is to accept that the naturalistic picture of human beings that emerges from contemporary science requires us to revise concepts inherited from a pre-scientific age, not to abandon belief in freedom, knowledge, creation, and the life to come. Our evolutionary history has endowed us with the strong but dispensable illusion that the truths of morality, and the reasons for action they give us, are grounded beyond us in the very fabric of reality. Yet morality’s power to move us is largely outside our voluntary control.
Morality
219
When we see what Marvin did to Sally’s cat, we cannot help but categorize his act as wrong, and we cannot help but be moved to feelings of moral indignation. This is a surrogate for genuine objectivity. Given the force of our moral emotions, and the general improbability of finding good reasons to disown them, morality’s hold on us is not likely to weaken. Revealed as a natural, human phenomenon, it is dethroned but not debunked. It does not prescribe human action from a standpoint that transcends our desires. We have good reasons to do what’s right, but nothing in this world sustains morality other than us and our evolved desires. The consequences are mixed. We should not deny that if we lose the conviction that an act being morally wrong is in itself a reason to avoid it will not in some way weaken moral motivation. I suspect, though, that the practical effects on public and private morality of this are trifling in comparison to the effects of the loss of belief that there is an omniscient, omnipotent deity who infallibly punishes the wicked and rewards the virtuous in the next life, if not in this one. By and large human beings have been motivated to do what is right less by the false belief that there are objective prescriptions, than by false religious beliefs about punitive deities. On the other hand, putting morality in its place by bringing its evolutionary origins to light can have welcome effects. Knowing that I am disposed to certain moral responses by an evolutionary process that tracked, not an objective moral reality, but inclusive fitness in the ancestral human population, I have room to scrutinize those responses. There is no reason to suspect that critical self-refection will generally lead us to disavow and resist acting on our moral feelings, but there are cases where it might and we can be glad of it. For example, suppose someone finds homosexual acts repugnant; they seem to him inherently wrong, something anyone has a compelling reason to avoid. This reaction is a plausible candidate for explanation by natural selection: those who had, and acted on, the desire for sex with members of their own sex were, all things being equal, less likely to produce offspring, while those with a strong aversion would be more likely to reproduce. Realizing this, one might conclude that he has no reason to believe that his repugnance signals inherent wrongness, and that he has no reason to despise or persecute homosexuals, or to support legal measures against them. Realizing that this negative emotional response may motivate mistreatment of gay and lesbian people, he may decide to try to resist the response and try to rid himself of it. Another example: someone feels that there is something inherently unjust in unequal distributions of material goods. Even when the inequalities are due not to wrongdoing, but to voluntary exchanges, differing mental and physical natural endowments, different priorities in life, or plain luck, they may give rise to moral indignation and the conviction
220
Chapter 7
that property ought to be redistributed. Recognizing this aversion to inequality as a product of cognitive mechanisms natural selection created for the detection of those who cheat in cooperative endeavors, and realizing that market economies make the least well off better off in absolute terms than they would be in an egalitarian society, we may come to see it as worthy of being disowned. 27 However, nothing guarantees that rational reflection on our moral responses will attenuate or extinguish them, rather than strengthening them and broadening their scope. The altruistic dispositions with which we are endowed are often parochial. Seeing someone close to me, familiar to me, or similar to me in need, I have strong empathetic feelings and the conviction that it is my duty to help. Similar or greater need endured out of sight, by those unfamiliar to me or different from me, gives rise to at most mild feelings and little sense of being obligated to help. As critics of our common sense morality have pointed out, when we see a sick child suffering in front of us, we experience a powerful conviction that we ought to provide aid, yet the knowledge that in some far off locale large numbers of children are dying for lack of something as simple as mosquito netting generates no corresponding sense of obligation. On reflection, we may come to regard as arbitrary and unreasonable the judgment that while we are obligated to help the child who is close at hand, saving the distant children is merely supererogatory. Natural selection has endowed us with both moral sentiments and the capacity for rational self-reflection on those sentiments. Identifying with and embracing our empathetic response to suffering children, we may decide that we do not want to be the kind of persons who arbitrarily care about nearby suffering while ignoring those who are out of sight. We find the difference morally irrelevant. The exercise of that capacity for rational self-criticism might make us not less, but more, sensitive to the pull of our moral sentiments. GOD AND MORALITY The idea that morality’s ground lies within us and that our experience of its objective, more than merely human, authority depends on nothing more than our inability to avoid projecting our emotionally powerful and motivating moral responses upon the world, is not new. Evolutionary theory simply adds a convincing explanation of how we came by the disposition to do this. What seems to be the detection of features of objective reality, features that in themselves give us reasons to act, is revealed as the operation of a mind adapted to social life. Arguably, while the general case for evolutionary moral psychology is quite strong, it is not so strong as to make rejecting it patently unreasonable. It differs in this regard from the claim that humans are material things, the product of biological evolution. The fact that a species came
Morality
221
about as a result of natural selection does not guarantee that any particular feature of it, such as humanity’s moral psychology, has an evolutionary explanation. As noted in chapter 2, it is reasonable to demand more and better evidence for hypotheses that are improbable in light of other beliefs. If, as I assume, the evolutionary account makes it very likely that morality is relative to the human species, then it might be reasonable for Christians to reject it, maintaining that the evidence is not so good as decisively to confirm it. This depends, however, on the assumption that the Christian faith makes it probable that morality is fundamentally objective. Relativist demotions of morality have long been resisted by Christians and other persons of faith, for whom the defense of an objective moral order is axiomatic. When what is at issue is the rejection of individual, or of cultural, moral relativism, this is understandable: religious or not, we have good reasons to reject these forms of relativism. However, the evolutionary account offers no encouragement to either view; indeed, its claim that the core of moral response is a human universal is at odds with them. What reason, then, is there to see evolutionary moral psychology, and the implied species-relativity of morality, as unlikely from the perspective of Christian faith? The answer is not entirely clear, but it seems to lie in the perceived need for morality to possess divine authority. The voice of morality must be identifiable as the voice of God; its prescriptions must be understandable as God’s directives, either in some immediate way or by God somehow having built them into the creation. It is, for example, morally wrong to lie for no other reason than that God forbids it. Or, like the fact that e=mc 2, it is a fact inherent in God’s creation for humans to discover and respond to appropriately. Abiding by the objective prescriptions embodied in created reality, we obey our Creator. In contrast, relativism of the kind indicated by the evolutionary explanation makes the connection to God very indirect. In our world some things are sweet and others sour, some things red and others blue, some disgusting, interesting, beautiful, ugly, funny, fashionable, impolite, illegal, and so on. On reflection, we acknowledge that these characteristics are not simply there in reality, but projected upon it because of what we are, thanks to our biological evolution, culturally conditioned socialization, and personal experience. Had things gone differently, there might have been no creatures that experience the world in these ways. We cannot reasonably suppose that it mattered to God that the universe would eventually bring forth creatures capable of seeing the sky as blue, being disgusted by rotten fish, and dressing fashionably, that these outcomes were specifically intended. God left it to chance whether the universe would become a home to creatures who know of, care about, and respond to such things. God did not build blueness, or disgustingness, fashionableness, and so on into the world, independent of the chancy evolutionary trajectories that brought about creatures that represent the
222
Chapter 7
world as bearing such properties. A world in which creatures project such qualities upon what they experience was just one among many possibilities. Evolutionary moral psychology appears to place morality in the same category, making it a matter of no unique and inherent interest to God. How can it be something with which God is profoundly concerned if it is a product of our contingent evolutionary history? For some, this is, to say the least, implausible. How can God not supremely care about morality? It is not a complete caricature to say that in the popular religious imagination, being the enforcer of good morals is God’s job; it is what God is for. Sin is moral wrongdoing, salvation is the reward for doing what is morally right, and those who persist in morally bad behavior are posthumously dispatched to hell. The deity’s principal concern is that creatures be morally good. The sophisticated avoid such simplistic reductions of theology to morality, but the underlying assumption that the moral behavior of human beings is of first importance to God is almost undisputed. The scientific naturalizing and relativizing of the moral truth might be mitigated. The course of evolution that led to our moral psychology appears contingent, but if nature’s basic laws are deterministic, then our moral nature is precisely what God intended. The superficiality of the moral truth could be mere appearance; the moral facts that result from our projection of moral categories onto the world could correspond to facts more deeply rooted in reality. Antecedent to creation, it would have been true, for example, that it is wrong to lie, and in selecting deterministic laws and initial conditions for the universe, the Creator saw to it that the personal creatures it brought forth evolved to believe that it is wrong to lie and see this as a good reason to refrain from lying. Adaptation to social life was the means God employed to endow creatures with a nature conversant with objective moral reality. This account, whatever its value in preserving the ultimate objectivity of morality in face of the evolutionary explanation of human moral psychology, suffers from what is for Christian faith the decisive disadvantage of portraying creation as deterministic and creatures as designed. This is at odds with the aim of creating persons distinct from their Creator and as such something we can accept only if the evidence for it is compelling. Possibly, God could create with the aim of there being moral creatures without having to govern the universe with deterministic laws. In this, our apparently indeterministic world, there was a reasonable expectation that in due course natural processes would eventuate in persons of some description. Perhaps we can go further and assume that whatever persons came about would be social beings and, adapted by natural selection to life together, would possess a moral psychology akin to that of humans. The silicon based creatures of Kepler-62e, although biologically, psychologically and culturally very different than their human neigh-
Morality
223
bors—1200 light years distant in the Milky Way—might have evolved what is at some very general level of description an essentially similar body of moral concepts and practices. And perhaps something like this is true of persons throughout creation; convergent evolution reliably leads to morality, or something functionally equivalent, wherever it leads to persons. 28 In itself, this would not render the moral facts objective: universality is not objectivity. If it so happened that, for example, evolution by various routes converges on color vision throughout creation, this would not make color an objective feature of objects, rather than a way they are represented in the minds of creatures endowed with the capacity. And, if it did so, it would still be unlikely that disparate organisms uniformly represent things as having the same colors; the four-eyed inhabitants of Kepler-62e, who detect light in the deep ultraviolet, might have color experiences unimaginable to us. 29 The same, presumably, would be true of ubiquitous morality: its universality would be highly abstract, with much of its content varying in accord with the idiosyncrasies of the various species. Only the most abstract—thin as opposed to thick—moral judgments would survive interspecific translation. Nonetheless, the committed theistic defender of moral objectivity could contend that prior to all this stands the eternal moral truth, objectively grounded somehow in the will or nature of the God who set the universe going in a way sure to result in creatures with moral psychologies that in some abstract fashion correspond to it. 30 The scientific account of morality as an adaptation does not necessarily disconnect it from God’s creative aims nor compel us to deny its ultimate objectivity. However, we can connect morality to God’s aims in creation without committing ourselves to its universality or objectivity. There is a preferable account in which science puts morality in its properly human place. It embraces the surprising conclusion that morality per se is not of profound concern to God and enables us to reappraise its significance in light of essential Christian commitments. Contemplating the convergence of Christian and scientific explanations of human origins, in chapter 4 we focused on the essential confession that God is love—the Triune community of loving persons—and inferred that accounts of human origins should be constrained by God’s aim of sharing life with persons distinct from their Creator. Whatever interest in the moral behavior of human beings we ascribe to God is not a matter of independent concern, but subordinate to divine love. It is not, of course that God is indifferent to the moral quality of human behavior. Morality imposes constraints on how we treat the human persons that God loves. It prescribes some actions that benefit human persons and proscribes many acts that harm them. What morality calls for, and what God cares about, significantly overlap. Given morality’s natural function, which is to move human be-
224
Chapter 7
ings to altruistic behavior, God’s general endorsement of it, irrespective of its nature and origin, is perfectly intelligible. Yet the divine endorsement of morality is not unconditional. God’s actual aims do not completely coincide with the metaphorical aims of the genes that gave rise to the moral psychology of the human species. A human being has a powerful, innate inclination to believe that a type of behavior is wrong and ought to be avoided and prevented. He has it because in the Stone Age humans who had it had greater inclusive fitness: they were more likely to reproduce. This was no guarantee that acting on this inclination in any particular situation actually benefitted any given individual then, and even less so that acting on it is to anyone’s benefit now. In the ancestral population, actions motivated by this moral emotion often did benefit human beings, if not the individual acting on them, at least other members of the community. Inclusive fitness correlates positively with things that make an individual’s life better; one must be alive and minimally safe and healthy to produce viable offspring. Moving from the ancestral population to the present day, in which the genome is largely unchanged but conditions have changed drastically, there is greater scope for what is good for individuals and the promptings of our innate moral sense to diverge. From the perspective of Christian faith, we have no reason to believe that God approves of morality when its prescriptions are at odds with our Creator’s love for human persons. In a frequently cited passage in his Institutes, Calvin wrote that God condescends to human limitations, accommodating to our ways of seeing things in order to be revealed to us. 31 In contrast to Calvin’s well-known views on the matter, God has left the specifics of creation to chance. It is on an ad hoc basis that God uses what in the natural, indeterministic course of events becomes available in order to pursue the aim of calling the created persons to share in the trinitarian life, love, and work. When the universe eventually gives rise to a species of persons adapted to social life, and thus endowed with a complex of dispositions to altruistic behavior, God has something of use. God wants human beings to care for one another, to love one another, and thus at the very least not harm one another. Natural selection has implanted in the human species something that can be recruited for God’s ends, but our moral proclivities stand under divine judgment. God’s endorsement of our morality is critical. We may do our best to discern what is morally right and wrong, but God’s final word can be a rejection of our righteousness. And God’s righteousness is loving faithfulness to created persons. Despite the powerful inclination to make morality God’s principal concern in dealing with humankind, Christians have every reason to endorse the demotion of morality to an imperfect means to God’s ends. The heart of the Christian faith is the good news that God dismisses the demands of justice and acts out of unconditional, undeserved love for us.
Morality
225
Morality demands that we get what we deserve; those who do evil ought to be punished and those who do good ought to be rewarded. Christians confess that God sets justice aside, forgiving everyone indiscriminately, as wicked as they might be, and giving to all far more and better than we deserve. We inhabit a world in which the demands of justice are authoritative, the last word, but God steps in and overrides them. In sharp contrast to the reasonable calculations of human justice, God’s justice— the righteousness of Yahweh—is no less than God’s love for creatures with no regard for who ought to get what. God’s justice is faithfulness to God’s creative intentions. 32 When Luke (4:18) portrays Jesus announcing that he has come to set the prisoners free, there is no qualification that indicates he has in view only the unjustly incarcerated. From the moral point of view, the grace of God is a scandal. Someone might agree that human morality lies under divine judgment, but contend that this means that God corrects it, revealing the truth about morality in opposition to human error. God shows us what is really right and wrong. This objection has some plausibility, but only because of ambiguities of the term morality. In this chapter, I have used the word in a narrow sense, to denote a fairly well-defined set of norms that pertain to behavior that affects other persons. Others use the term much more broadly, to refer to whatever we have best reasons to do overall. Thus questions like, “What sort of person should I try to be?” and “What kind of life is worth living?” are thought of as moral questions. The Christian rightly sees herself as having the best possible reasons to do what God wants her to do. After all, God loves her, wants what is ultimately best for her, and has the ability to bring it about. If there is something that God wants her to do, she can be sure that doing it is in her ultimate best interests, and thus that she has a good reason to do it, indeed a reason that trumps any competing reason not to do it. Insofar as her convictions about what pleases God are the most important factor in the overall ordering of her life, she can reasonably say that she aspires to live by the norms of her “Christian morality,” the true, “higher” morality unqualifiedly endorsed by God. Confusion arises, however, when the narrow and broad uses converge. The Christian woman believes that she ought to pray, worship, be part of a local church, evangelize, and study the Bible, but she cannot reasonably contend that secular persons, and persons of other faiths, behave in a morally bad way in virtue of not engaging in these Christian practices. She cannot reasonably regard her neighbor’s failure to attend worship services as on a par with his embezzling from his employer or beating his children. Persons who are not Christians are not necessarily bad persons, or even morally inferior to Christians. Further, when Jesus counsels, “If anyone strikes you on the cheek, offer the other also,” we cannot reasonably understand him as making a claim about what is morally right. 33 We have no moral obligation to let others hit us as they
226
Chapter 7
please. In some circumstance, not just refraining from hitting back, but refraining from trying to get away, may be supererogatory, i.e., morally laudable although more than duty demands. But this kind of response to an attacker might often be merely morally wrong: if I let him hit me again he will probably knock me down and move on, unimpeded, to attack the children and elderly ladies down the street. Jesus is not criticizing or correcting our ordinary judgments from a moral perspective, but showing us that in the kingdom of God the demands of morality are left behind. He is not saying that it is really morally wrong to try to avoid getting hit again, or even to hit back. The point, to the contrary, is that morality does not have the last word. Our present faithful attitudes and actions can point to a future when God’s aims are fully realized, for example, I can ignore the fact that I have a moral right to defend myself or to retaliate. Nonetheless, what makes sense in light of the Christian gospel will often correspond to what is morally right. Sometimes, it will be a supererogatory course of action. Very often, it will be morally indifferent, like most things we do, having no particular moral quality. Sometimes, it will be contrary to what morality calls for, but those who claim that God calls them to do what is morally wrong have the burden of justifying this. Generally, doing what is morally wrong harms the human beings God loves, so we need exceptionally good reasons to conclude that we are in a situation where God wants us to ignore the requirements of morality. THE FALL AND ORIGINAL SIN Scientific explanation does not deprive us of good reasons to be moral. It does not debunk morality, but it does deny the lofty status often accorded it. Evolutionary theory portrays moral reality as an aspect of the way social creatures are adapted to represent the world, not of how things are in themselves. In so doing, it does not expose doing what is right as irrational, but it does put it in its place, revealing that those reasons have no ground deeper than the entrenched desires that make us human. We find in this world nothing deeper than ourselves to underwrite morality. So far as this world goes, we are its origin, its advocates and defenders. Some will find this too flimsy a foundation for something as awesome as the moral law. The further implication that the voice of the moral conscience is not the voice of God, and that while God endorses morality, it is only in a qualified and contingent way, is one many of the religious, devoted to a necessary connection of God and morality, will find appalling. On the contrary, with its most central claims about the nature of God’s purposes for creating clearly in view, Christianity can accept the demotion of morality at the hands of science as unproblematic. The scientific naturalizing of morality is a valuable
Morality
227
counterweight to the moralizing that has infected Christianity over the centuries. There may, however, be doubt that Christianity survives morality’s demotion, let alone agreement that a superior articulation of the Christian faith is available now that science lays bare morality’s humble natural origin and relative nature. I will conclude, then, by briefly and speculatively considering the implications for two interrelated theological concepts that have come to be understood in moralistic terms: the doctrine of the Fall and of original sin. The identification of sin with morally bad behavior is a given in our general culture, but it cannot stand against even casual scrutiny. A popular understanding of the Fall, sin’s origin, portrays it as an actual historical event. 34 Initially, human beings were morally innocent, but became morally bad as a result of disobeying God. This account faces insurmountable difficulties. There was no time in the past when human beings were not inclined to selfishness, cruelty, cheating, deception, etc. We have innate dispositions to morally bad—as well as to morally good— behavior, not as a result of choices made by early human beings, but because of the evolutionary processes that brought the species into existence. Dispositions to selfishness, deception and cruelty, as well as to empathy and a rudimentary sense of fair play, are observed in today’s non-human primates, and we reasonably assume that these were the raw materials out of which natural selection crafted genuine human moral psychology in the five or six million years since the our lineage separated from that of today’s chimpanzees. 35 The moralistic account of sin and how we came to be sinners coheres no better with Christian faith than with scientific knowledge of human origins. Despite the prevalence of the idea of sin as moral wrongdoing, it is profoundly at odds with Christian belief. Sin is the condition of creatures who reject their Creator. God created a world which in the natural course of events gave rise to persons distinct from their Creator, persons who can freely share in the life of the Trinity, becoming members of a communion of persons freely bound to one another and God in love. We are sinners insofar as we spurn this invitation, refuse to trust God, deny God’s role as source and sustainer of life, and ultimately will to be the self-sufficient, and self-justifying, “gods” of our own world. The sinner is the creature who absurdly refuses to acknowledge himself a mere creature. This misbegotten impulse to envision ourselves as more than creatures and as impossibly transcending the natural world, and the role of contemporary science in delivering us from it, has been an ongoing theme of The Material Image. The rebellion manifests itself in many ways, one of which is morally bad behavior. By and large, the moral sentiments which evolution has embedded in us move us to do good for other human beings, and at the least to avoid harming them. To acknowledge and accept our Creator is
228
Chapter 7
to care for what God cares for, and morality calls for minimal concern for the persons God loves. The rejection of God commonly manifests itself in selfish indifference to the well-being of others and thus to morality’s prescriptions. However, the fundamental refusal to let God be God can express itself in morally correct behavior no less than in moral wrongdoing. What the biblical story of the Fall does not say is noteworthy: Adam and Eve are tempted, not to do evil, but to become like God—or gods—knowing good and evil. At face value, the point of acquiring this knowledge is to be able to do good and avoid doing evil, a worthy aim. The story’s obviously mythological elements—the original human pair, the talking serpent, the magical fruit—do not obscure the theological point. Genesis does not proffer a simple equation of sin and morally bad behavior. It portrays human estrangement from God not as unwillingness to do what is right but, on the contrary, the desire to discern, and do, what is best with no need to trust God. Fallen humans are no more likely to be morally wicked than to be self-justifying moralists, confident that they are in the right and, like gods, lord it over their fellow human beings. 36 Throughout the biblical witness the righteousness of God is opposed to human righteousness. This has often been understood as a contrast of genuine moral goodness to defective or spurious human morality. Biblical writers, the prophets especially, are acutely aware that human beings are inclined to obtuseness, hypocrisy and evasiveness in face of the demands of morality. The rich might pride themselves on their moral propriety even as they exploit the poor. The God who condescends to our evolved sense of fairness, and to our evolved empathy for the needy, condemns the morally wicked exploitation of the people God loves. What cannot be theologically sustained is the ultimate identification of God’s justice with human justice. The justice of God is nothing less than relentless fidelity to the covenant with what is made, to the deeper aims of creation, to love for creatures. As we saw above, God enlists our moral sensibility, but in the end, God is an affront to it and overturns it. We naturally approve of a deity whose power underwrites our sense of who deserves what, but the God of biblical faith can be disreputably indifferent to moral matters. This God treats the morality we want to objectify and, in a way, deify, as the merely human thing it is, often but not always useful for divine purposes and as such always subject to divine judgment. Having excised the moralistic accretions, what can we make of the Genesis story of the Fall and its consequences for the human condition? In chapter 4 I suggested that we revise the traditional doctrine of special creation, interpreted as a scientifically and theologically implausible assertion about the means by which God created human beings, and understand it as bringing to the fore the special place that human beings (and any other personal creatures there might be) have in God’s creative aims.
Morality
229
Here, I suggest a similarly motivated revision of our reading of early Genesis. The biblical portrayal of the Fall as an historical event, as something that occurred in the remote past, expresses the contingency of our alienation from God. Human beings are sinners: one way or another, we choose to get by without God, ignoring the call to know, love, and trust our Creator. But this is not inevitable. Sinfulness is not simply a manifestation of human nature. We are not forced; we choose, exercising the freedom, and acquiring the responsibility, of which material creatures are capable. We should reject the view that the Fall of humankind is how things must be for finite beings, for creatures, for material beings, for humankind. Likewise, we should reject the view that the Fall was a good thing, that it was a necessary step out of innocence on the path to eventual maturity. With this we should also reject the idea that our sinning was intended by God, and that in disobeying our maker, we do what God really wants us to do, because our rebellion is a secret means to God’s ends. We trust that God finally achieves creation’s ends despite our rejection, but this does not mean that God intended us to turn away, or even knew that we would do so. The contrary assumption is that God knew the possibility and, if there is a fact of the matter, the probability of human beings refusing his call. For all we know, this probability was high. Perhaps, given the kind of creatures humans are, the odds were against us making reasonable choices in response to the encounter with God. What matters is that humans choose freely. As we well know, human beings can find it hard to make certain choices that it is possible and reasonable for them to make. Perhaps there are many other free creatures in God’s creation for whom a rational, grateful response to the call to share in God’s life is virtually certain. For such superior creatures the choice to trust God might be as patently obvious as the choice to eat food instead of dirt. It might be that in the divine self-revelation, God reached down as far as possible, taking greater risks than with other creatures. Our accepting the invitation to share God’s life and invite our Creator into our world was far from a sure thing. That things went badly was, perhaps, no surprise. If so, then it is no more surprising that God intervened to make reconciliation with this particular species, becoming part of the creation precisely here, as a member of this least promising, most wayward species. God, being God, reaches out in love to what is other than God, so we reasonably suspect that in choosing how to become part of creation, God reaches down as far as possible. Human beings are the image of God. This implies that we are graciously called to share in God’s life, but it is no guarantee that we are the most godlike of creatures. It may indicate the opposite. For all we know, we are the prodigal children, the lost sheep, the lepers of God’s creation, those in most dire need. This is, of course, speculative, but it has the minimal virtue of cohering with what we now know of the world thanks to science, as well as
230
Chapter 7
with the most central beliefs presupposed by faith in Christ. Further problems, arising out of the idea of original sin, invite further speculation. Traditional theology preserves the contingency of the Fall but at the cost of various implausibilities. There was no first human couple, a pair of human beings who did not have human parents. This is true not because there have been infinitely many human beings, but because no sharp line marks the beginning of the human, or any other, species. There were the earliest humans to whom God chose to be known, but their parents would have been no less human than they were. Whoever first received, but chose to reject, God’s call thereby became sinners. Their sinful condition in some mysterious, unspecified way has been passed to the rest of us. Thus the obscure doctrine of original sin. That we are born into our sinful condition increases the implausibility of equating being a sinner with being guilty of moral wrongdoing. No one can inherit moral guilt; to be guilty one must have done something. At most, one could be born with a disposition to moral wrongdoing and as a result sometimes freely choose to do evil. If sin is inherited, and moral guilt cannot be inherited, then whatever being a sinner is, it is not being morally guilty. Further, if everyone chooses to do something, we rightly suspect that it is irresistible and not something we can freely choose not to do. Sin is not moral wrongdoing, yet we cannot easily abandon the conviction that like moral wrongdoing, it is in some way a matter of free and responsible choice. We are sinners because we freely reject the call to trust God. So, while we can dismiss the idea of original sin as innate moral guilt, the problem that remains is how we can be born into a sinful state. Further problems arise from the idea that original sin means that human sinfulness is both innate and due to past choices made by our ancestors. God could respond to a primordial choice to rebel by modifying the human genome, thereby ensuring that these sinners’ descendants are all born with a disposition, perhaps irresistible, to follow their forebears and sin. But why would God do this? No plausible answer presents itself. The reasonable assumption is that any innate behavioral dispositions found in human beings are the result not of a singular event in the past, but the operation of natural selection over many generations. With all this in view, we must seek an understanding of original sin that construes our sinfulness as something into which we are born, yet not innate, and despite its universality, involving free, responsible choices. This is a major challenge, but I will conclude this chapter by gesturing in the direction of such an account. My speculation is that to be fallen is to be born into a world in which the existence and nature of God is not empirically obvious, i.e., a world in which God is not incarnate in a way that compels the belief that God exists and loves us and intends our good unconditionally. When God is not evident to our senses in this overt way, we might still have true beliefs about God, but we are so
Morality
231
constituted that this fails to reliably induce the reasonable response. An unfortunate, but indisputable, feature of human psychology is that mere theoretical knowledge, no matter how well grounded, can fail to move us to appropriate action. We know, e.g., that eating too many hamburgers is likely to have deleterious effects on our future selves, yet we drastically discount, or simply ignore, this fact and imprudently overeat. We know, e.g., that the plight of far off strangers morally obligates us to aid them, yet this knowledge fails to move us to the morally appropriate response. What is unseen, merely in virtue of being remote in space or time, often fails to engage emotion and imagination and fails to motivate us to act in accord with what we know. The same or worse might be the case with an unseen God. I conjecture that rational, properly functioning human beings, when they literally and unambiguously see the true God respond in the only reasonable way, which is to obey, trust, and love. When the same normal, properly functioning human beings are born into a world from which God has withdrawn, and is not empirically available, they are not moved to a positive response. Instead, they find it reasonable to proceed as though there were no God, and vainly to project themselves into that role, imagining themselves transcending the visible world, and trying to be like gods, determining what is best for themselves and one another. 37 In a way, this account invokes human nature to explain sinfulness. Human beings are sinners because of the inadequacies of our cognitive architecture, the incomplete integration of cognition with the imagination and emotion that moves to action. It is not a surprise that the creation of a good and wise God contains such creatures. We can imagine a universe home to many kinds of creatures that are persons. They are all descendants of creatures who were not persons, and all stand, at one remove or another, from their ancestors who were not rational, self-conscious beings. Some species bear more, and others less, resemblance to their nonpersonal ancestors. Perhaps some kinds of created persons are eminently reasonable, their various mental functions so thoroughly integrated that they invariably act in accord with what they know. When their Creator is revealed to them, by whatever means, they freely and joyfully choose communion with the one who created them. For them, turning away from the God they know wills their good and has limitless power to achieve it is possible−they are free−but it is simply “unthinkable.” 38 Not all created persons are so blessed. Some species lie closer to the threshold of personhood. They can, and often do, know what is good for them, what they have best reasons to do, yet choose to do something else. They are not reliably moved to appropriate action by what they know. The truth must be made vivid to them by way of their senses if they are not to go astray. That such lowly creatures exist does not reflect negatively on their Creator. God always intended to be palpable to creatures, dwelling with them as the incarnate Creator. God is fully present with
232
Chapter 7
them, all is well; when God is not in view, things go very badly. These creatures have some unsavory innate propensities that make it easy to choose to act in ways destructive to themselves and others. When their God is present to them in the way optimal for their limited capacities, this is not a problem. God’s convincing presence makes it patently absurd to act in these ways. But left to their own devices, all hell breaks loose. Why does God permit the situation to go bad by not being empirically evident to these creatures? Further inherent limitations of these creatures can be hypothesized to explain why God is not immediately manifest in a manner that elicits a positive response. They are persons, and thus free, responsible beings. They are so constituted that their choices can be caused by their very own, reflectively examined beliefs and desires. This sensitivity to reasons is part of what makes them persons, but for them the conditions for this compatibilist freedom are easily subverted. Their freedom is easily compromised; the overt presence of God easily overwhelms them. It threatens to render their response less than voluntary. God seeks to persuade, not compel. Recall from our discussion of human origins that the challenge of making creatures who are not mere extensions of the Creator, but distinct persons, capable of genuine personal relationship with their God, is not trivial. It explains why God chooses to create persons by means of eons of unguided natural selection in an indeterministic universe. It might also explain God’s slow and careful pursuit of fellowship with creatures, particularly those with low degrees of psychological integration and fragile selves. The dilemma is that if God is not manifest patently, to the senses, they act as though there is no God, but if God does so, they do not respond in genuine freedom. The solution is to proceed with care, taking time with them, becoming visible, but at first enigmatically, leaving room for those who hope there is no God, or who disapprove of God, to avoid, reject, even to evict the Creator from the world as they experience it. These creatures have room to say no, to seek their own way, as time and generations go by, building a cultural regime in which God seems simply absent. The free, responsible creature, born into this world, finds it all too reasonable to concur in the long-established rejection of God. 39 The situation worsens if, because of accidents of their cognitive architecture, these creatures, typically unmoved by the unseen, are endowed with a countervailing propensity to find obvious the existence of supernatural persons, but irresistibly to envision these beings, the various gods of their religions, as condemning and controlling, as projections of themselves. Culturally entrenched institutions supply emotionally engaging sensory experiences, along with explanations that motivate behavior, but this ritual responds to something antithetical to faith in the true God. 40 Incipient inclinations toward taking God seriously are subsumed into the local culture’s religiosity. Belief in the God who is love becomes implau-
Morality
233
sible. Imagined gods are recruited to police morals as locally conceived and to enforce group identity. To be fallen, to be the bearer of original sin, is to be born into a world in which God is empirically absent, but religion is empirically compelling. This is the world in which the true God, the relentless pursuer of beloved creatures, condescends to our natural religiosity just as to our natural morality. The Christian version of our story portrays God taking responsibility for creatures in a way that does not undermine their responsibility and distinctness as persons. God loves us too much not to let us be ourselves, but also loves us too much not to deliver us, even from ourselves. God lets creatures do their worst, rejecting the God who loves them for the sake of the empty promise of being like gods and going their own way. When at last God becomes one of them incognito, their foolish pride is provoked and it drives them to put the incarnate God to death for blasphemy and sedition, as the one who challenges their authority. But divine love is not finally defeated. The God who resurrects Jesus Christ demonstrates that nothing beloved but rebellious creatures can do defeats God’s commitment to them. They do their worst and this too is turned to their salvation. Now, between the times, that resurrection is the guarantee that in time’s fullness God will at last be fully present with us in a way that persuades each of us freely to welcome our Creator among us. Among the features of human beings God enlists in this project is our natural religiosity, which is the subject of chapter 8. NOTES 1. Sam Harris, in The Moral Landscape: How Science Can Determine Human Values (New York: Free Press, 2011) is representative of those who deny this, but he depends on the mistaken idea that science can identify a morally neutral conception of human well-being.. 2. See, e.g., Frans De Waal, Primates and Philosophers: How Morality Evolved (Princeton, NJ: Princeton University Press, 2006),13–67. 3. In the biological sense, in contrast to the ordinary sense, behavior is altruistic so long as it benefits others, whether or not it benefits oneself. Sally goes to the store to buy bananas. She could run out the door without paying for them. She knows she could get away with it, being young and fast, but she chooses not to, and cooperates, going to the checkout to exchange some money for them. Both she and the store are better off: she would prefer having the bananas over the money, while the store prefers her money to the bananas, so there is reciprocal benefit. Sally would be even better off if she got the bananas for free, but she has an innate disposition to find stealing objectionable. She could, of course, simply give her money away, having neither it nor the bananas, and that self-sacrificing behavior would also be altruistic. But it is Sally’s commonplace honesty that most needs to be explained, not rare cases of self-sacrifice. Social life is possible only if most people most of the time, like Sally, do not want to steal even when they can get away with it. 4. “Recent Work in Human Altruism and Evolution,” Ethics 106.1(1995), 128–57. 5. There is a story that the great evolutionary biologist J. B. S. Haldane, asked whether he would give up his life for his brother, did a quick calculation on a napkin and answered, “No, but for two brothers or eight cousins.” But it is important to note
234
Chapter 7
that this account supposes that no one actually calculates other than the biologist who employs the mathematics of population biology to ascertain which behavioral phenotypes promote inclusive fitness. 6. The limit case is the social insects, species in which large numbers of sterile individuals labor their entire lives for the benefit of the queen, their sister. The attempt to understand these altruistic insects played a major part in the discovery of the significance of kin selection and opened the way to sociobiology and later evolutionary psychology. See E. O. Wilson, Sociobiology: The New Synthesis (Cambridge, MA: Harvard University Press, 1975). 7. See Robert Trivers, “Reciprocal Altruism” in Natural Selection and Social Theory: Selected Papers of Robert Triver (Oxford, UK: Oxford University Press, 2002), 3–55. 8. See Robert Axelrod, The Evolution of Cooperation (New York: Basic Books, 1984). 9. For an evocative account of the central role of evolved propensities to cooperation in the making of the human species see Matt Ridley, The Rational Optimist: How Prosperity Evolves (New York: Harper, 2010). Leda Cosmides and John Tooby make the case that humans have an innate mechanism to detect cheaters in social interactions in “Cognitive Adaptations for Social Exchange,” in Jerome H. Barkow, Leda Cosmides, and John Tooby, eds., The Adapted Mind: Evolutionary Psychology and the Generation of Culture (Oxford, UK: Oxford University Press, 1995),163–228. 10. Originally, a person (as in persona) was a role occupied by a human being, not the individual whose role it is. With this in mind, Alva Noë, in Strange Tools: Art and Human Nature (New York: Hill and Wang. 2015), 178–79, writes, “personhood is defined, crucially, in relation to praise, blame and evaluation. . . . We are persons insofar as we are subject, always and implicitly, to the standards of our community.” I have used person in its contemporary sense, to denote the rational, self-conscious individual, an organism of a particular kind, but this older sense illuminates what it is to be such a thing, at least one of the human type. 11. Philip Kitcher, The Ethical Project (Cambridge, MA: Harvard University Press, 2014). 12. For morality as a skill, see Paul Churchland, “Rules, know-How, and the Future of Moral Cognition” in his Neurophilosophy at Work (Cambridge, UK: Cambridge University Press, 2007), 61–79. In contrast to what I claim about the limits of ethical theory Derek Parfit devoted his monumental On What Matters to defending the full objectivity of moral truth and to this end argued for the underlying unity of the principal ethical theories: On What Matters Volume 1 (Oxford, UK: Oxford University Press, 2013). 13. Richard Dawkins, The Selfish Gene (New York: Oxford University Press, 1976). It is worth noting that if we press the metaphor further, we must acknowledge that the genes are in their way altruistic: an individual gene helps build a brain which acts in ways that can lead to that gene being destroyed, sacrificed for the sake of exactly similar but numerically distinct genes in other organisms, so that other, not yet existing, exactly similar but numerically distinct descendant genes will exist in the future. These entities are metaphorically indifferent to the fate of the organisms they construct and inhabit, but they are altruistic when it comes to genes just like them. Or, to indulge in a perhaps more accurate metaphor, they are chauvinistic. 14. Here I pose the objection as it is usually made, ignoring the implicit mistaken assumption that self-interested behavior is necessarily morally bad, selfish behavior. All behavior is self-interested, since all choices are caused by one’s own reasons. This includes what we do for the sake of others. We often identify the interests of others as our own, as with the mother’s reason to act on behalf of her child’s needs. In contrast, selfish behavior is self-interested behavior that pays insufficient heed to the rights and needs of other persons. 15. The contrary view is widespread, e.g., the moral theorist John Hare bases a critique of the explanatory power of evolutionary moral psychology on the assumption that the behavior it identifies as altruistic is really self-interested; “Is There an Evolutionary Foundation for Human Morality?” in Philip Clayton and Jeffrey Schloss,
Morality
235
eds., Evolution and Ethics: Human Morality in Biological and Religious Perspective (Grand Rapids: Eerdmans, 2004), 187–203. 16. Steven Pinker, The Blank Slate: The Modern Denial of Human Nature (New York, Penguin: 2003), 192. 17. For me, like most people not living in the Stone Age, when one had to burn many calories to obtain meat, but in a culture where meat is readily available in quantity, the belief that eating more hamburgers makes me healthy is, alas, false. 18. Mary Stewart Van Leeuween, “Of Hoggamus and Hogwash: Evolutionary Psychology and Gender Relations” in Journal of Psychology and Theology 30 (2001), 101–111. 19. Grounding the moral truth in human nature, scientific naturalism’s account of morality corresponds to the ethical naturalism that dominated thought about morality from Plato and Aristotle until early modern times, with the crucial difference that while on the traditional account facts about human nature directly imply the moral facts, the scientific account only explains why humans impose moral concepts upon the world. 20. We sometimes describe morally bad behavior as disgusting. There is evidence that disgust can be closely linked to moral judgment; see, e.g., David Pizarro, Yoel Inbar, and Chelsea Helion, “On Disgust and Moral Judgment” in Emotion Review 3.3 (2011), 267–68. 21. Evolution is only a partial explanation of the complex phenomenon of disgust in human beings. For an exploration of the cultural side of the subject see William Ian Miller’s The Anatomy of Disgust (Cambridge, MA: Harvard University Press, 1997). 22. It would be a mistake to say they can do nothing. If I know that the fish that would have sickened and killed members of the ancestral human population but will not harm, and instead nourish, me, and I am starving with nothing else to eat, then my reason to avoid it is overcome and, if I am reasonable, I will force myself to eat the fish, despite the disgust it occasions. 23. This view of practical rationality, associated with David Hume, who described reason as the slave of the passions (Treatise of Human Nature, Book Two, Part III, Sec. 3) does not overtly follow from scientific naturalism as I have defined it. I will, however, assume the Humean account, in keeping with the overall strategy of examining for coherence with the Christian faith the most radical version of naturalism that is not plainly implausible. For a defense of this account see Bernard Williams, “Internal and External Reasons” in his Moral Luck (Cambridge: Cambridge University Press, 1981), 101–113 and for a case against it in response to Williams, but still within naturalism as The Material Image conceives it, see John McDowell,”Two Sorts of Naturalism” in his Mind, Value and Reality (Cambridge, MA: Harvard University Press, 1998), 176–97. 24. J. L. Mackie, Ethics: Inventing Right and Wrong (New York: Penguin, 1977), 15–49. 25. What’s true of morality is true of human values as such. We at first suppose them to have a grounding in objective reality, but learn that they do not: we care about certain things in virtue of our evolutionary history. Valuing these things increased inclusive fitness in our ancestors. Knowing this, we critically reflect on what we value, determining whether we ought to continue doing so. But when we decide we should, there is the further realization that the processes of reasoning that reached this conclusion are themselves the product of those same evolutionary processes. The attempt to ground our values is recursive. It stops only when we accept that we cannot succeed, and that like the mother who loves her child, we can only take a stand in their favor or abandon them. It is not surprising that those who see this as world without a God who takes a stand with us and for us can feel forlorn, and hope to find some ground for human values outside us. However, while the human mind’s tether to its biological origin can be long and slack, it cannot be severed. In this regard see Keith E. Stanovich, The Robot’s Rebellion: Finding Meaning in the Age of Darwin (Chicago: University of Chicago Press, 2005). 26. For a defense of this approach see Richard Joyce, The Myth of Morality (Cambridge University Press, 2007).
236
Chapter 7
27. See, e.g., Paul H. Rubin, Darwinian Politics: The Evolutionary Origin of Freedom (Piscataway, NJ: Rutgers University Press, 2002), 87–112. 28. For speculations see Michael Ruse, “Is Rape Wrong on Andromeda? An Introduction to Extraterrestrial Evolution, Science, and Morality,” in Michael Ruse, The Darwinian Paradigm: Essays on Its History, Philosophy and Religious Implications (London: Routledge, 1989), 209–45. 29. Even in the limited venue of the Earth evolution has brought forth creatures with markedly different capacities for color vision; cats, for instance, see fewer colors than humans and some species of mantis shrimp detect more; see John Bradshaw, Cat Sense: How the New Feline Science Can Make You a Better Friend to Your Pet (New York: Basic books, 2013), 104–105 and Justin Marshall and Johannes Oberwinkler, “Ultraviolet vision: The colourful world of the mantis shrimp,” in Nature 401, (28 October 1999), 873–74. Thanks to my colleague Tyrone Gnade for the latter reference. 30. Alien “moralities” might systematically include obligations and permissions we would regard as reprehensible, and they might involve concepts and categorizations we find unintelligible. It might be more reasonable to see them as analogues of, not instances of, morality. 31. Institutes of the Christian Religion, Ford Lewis Battles, trans. and ed (Philadelphia: Westminster Press, 1960), XIII.1. 32. Some influential Christian theologies construe the demands of divine justice and of divine love as distinct, and potentially conflicting, principles in God, and explain the incarnation and death of Jesus Christ as the means God employs to satisfy the two simultaneously. This idea is often ascribed to St. Anselm of Canterbury’s (1033–1109) in his Cur Deus Homo? Even if this account succeeds in its own terms, which seems unlikely, it should be rejected: God is love and whatever there is to say about God’s justice must be understood in terms of it. 33. Luke 6.29; see also Matthew 5.39. 34. The biblical story of the Fall is told in mythological terms. But we need not believe in an original human couple, naked in a garden, eating from magic trees and conversing with serpents to confess that it points to the literal truth of the human condition, what is true of all of us most of the time: in thought or action we deny God’s role as God and seek to be as gods, sufficient unto ourselves. The human world is in a perpetual state of fallenness in which God humbly, patiently works toward our rescue, persuading but not compelling. 35. Frans de Waal, Primates and Philosophers, cited above, and his Good Natured: The Origin of Right and Wrong in Humans and Other Animals (Cambridge, MA: Harvard University Press, 1997). Here I follow what I take to be the scientific consensus in favor of the innateness of these aspects of human beings, but we should note, given the difficulties of research on neonate behavior, the possibility that what appears innate actually develops very early in life due to interaction with the environment. See Jesse Prinz, Beyond Human Nature: How Culture and Experience Shape the Human Mind (New York: W.W. Norton and Company, 2014). 36. The remainder of the Genesis narrative, and much of the Old Testament, depicts the escalating violence and tragedy that ensues. 37. This hypothesis is similar to one suggested by James K. A. Smith. However, while he ascribes “moral immaturity” to prelapsarian humans, I suspect that it was their epistemic situation, tied to the risks that God took in creating, and reaching down to, the least of persons, that sets the stage for the free choice to try to escape being creatures. I see no reason to suppose that the morally mature are less, rather than more, likely to project themselves into a divine role. See James K. A. Smith, “What Stands on the Fall?” in Evolution and the Fall, William T. Cavanaugh and James K. A. Smith, eds. (Grand Rapids: Eerdmans, 2017), 61–63. 38. Harry Frankfurt applies this distinction to human beings in Taking Ourselves Seriously and Getting It Right (Redwood City, CA: Stanford University Press, 2006), 30–1.
Morality
237
39. This hypothetical humbling “view from the bottom,” i.e., to conceive of humans as as far down into creation as God can reach, could not for Christian faith be the last word on us. God, being God, makes the least in merit the first in honor. It is as one of us that God makes a home in creation. 40. On the role of sensory religious ritual see Harvey Whitehouse, Modes of Religiosity: A Cognitive Theory of Religious Transmission (Walnut Creek, CA: Altamira, 2004).
EIGHT Religion
EXPLAINING RELIGION Contemporary science offers a compelling account of the universe. It explains how the galaxies, stars, and planets arose naturally from the initial singularity. It credibly promises to account for how there came to be living things on the Earth and it advances a powerful general explanation of how life on this planet diversified into a host of extraordinarily complex, highly adapted forms. It extends this evolutionary explanation to the human species and offers an explanation of our origins that cannot reasonably be denied. Further, science makes a convincing case that mind and meaning are features of the natural world. It reveals that whatever freedom and responsibility we possess must be consistent with being subject to the causal laws of nature. It plausibly portrays our capacity for moral agency as an adaptation to social life. Science systematically erodes the illusion that humans are mysterious beings that elude natural explanation and transcend the material world. If there is a last stand for the traditional human self-image, safe above the rising waters of scientific explanation, it is in religion. However, the naturalizing project has turned to religion, seeking to explain it too as a natural phenomenon. Not even the sacred is sacred. We saw in the preceding chapter that the scientific explanation of morality “puts it in its place,” portraying it not as a grasp of transcendent norms but as merely human. Contrary as this is to traditional views of our moral nature, it is on reflection a discovery Christian faith can accept with equanimity. The authority and objectivity of morality do not accord with what many have assumed, or at least hoped, but what it actually possesses suffices to constrain rational self-interest, which is all we should expect of it. What is right and wrong is not, on the account of 239
240
Chapter 8
morality as a product of natural selection, subjective or relative in ways that much matter to how we live our lives. Indeed, Christians can, and in fact should, welcome the scientific account of the origin of our moral nature and the constraints it places on the nature of moral truth. In naturalizing morality, science deflates some of its pretensions, but it does not support a radical denial of moral truth or leave us without good reasons to be moral. Science does not debunk morality, undermining its claim to rationality. In contrast, it does threaten to debunk religion, including Christianity, as irrational. This chapter sketches a current scientific strategy for explaining human religiosity, demonstrates its potential to undermine religious belief, and asks whether Christians can reasonably and honestly accept this account of the nature and origin of religion, or if we have at last reached a juncture where Christianity and scientific naturalism must part ways. THE COGNITIVE THEORY In keeping with this book’s general strategy, this chapter focuses on what is arguably the most radically naturalistic explanation available, the cognitive theory of religion. This hypothesis is not in the least hospitable to the attempt to soften the blow science delivers to traditional ideas of human nature. The cognitive theory explains human religiosity as an accident, a mere byproduct of the interaction of mental mechanisms evolved for other purposes. Religion is a vast, ancient and diverse phenomenon. We cannot reasonably seek a single explanation for the manifold thoughts, feelings, and forms of individual and social behavior that constitute it. The explanatory focus of the cognitive theory is narrow: the tendency of human beings to find plausible the existence of invisible agents, non-human persons who take an interest in certain aspects of human life, and whose actions explain certain kinds of event. Proponents of the cognitive theory identify religious beliefs as beliefs about these unseen agents, their desires, beliefs, intentions and actions. Rituals and other religious activities are attempts to interact with and influence the spirits, gods, demons, ancestors and other beings populating the thousands of past and present human religions. The cognitive theory relies on the computational theory of the mind and the modularity thesis, the idea that the mind is, or at least includes, a collection of evolved, specialized information processing systems. A mental module is dedicated to specific computational tasks. It receives a particular type of data, input from the sense organs or from other mental modules, processes that information in a particular way, and then outputs a result that becomes input to other modules, or efferent signals in the systems that control bodily behavior. Some, but relatively little, of the
Religion
241
output of the specialized information processing systems enters conscious awareness as a sensation, idea, or emotion. Introspection affords at most minimal access to the modular mental machinery where the contents of consciousness originate. The mind’s computational modules are adaptations, crafted by natural selection to track and respond to the natural and social environment of our ancestors. In contrast to other hypotheses, the cognitive theory portrays human beings’ disposition for religious experience and belief not as an adaptation, but as the accidental result of the interaction of various mental modules adapted for other functions. Pascal Boyer, a principal advocate of this deflationary theory, writes: Religious believers and sceptics generally agree that religion is a dramatic phenomenon that requires a dramatic explanation, either as a spectacular revelation of truth or as a fundamental error of reasoning. Cognitive science and neuroscience suggests a less dramatic but perhaps more empirically grounded picture of religion as a probable, although by no means inevitable by-product of the normal operation of human cognition. 1
Cognitive theorists accord a central role in making the existence of gods or spirits plausible to an agency detection mechanism. The mind contains a module that takes sensory information as input and outputs to consciousness the idea that some agent, human or otherwise, lurks in the vicinity. The agency detector is hair-triggered, much more likely to be activated when no agent is present—a false positive—than to fail to activate when an agent is present—a false negative. Walking in the woods at night, one is much more likely to interpret a noise as due to an unseen, possibly threatening, person or animal than to the wind in the trees. False positives can be inconvenient, but a false negative can be fatal. On the assumption that natural selection can produce a mechanism that is reliable, but not infallible, it is no surprise that we are equipped with what Justin Barrett describes as a “hypersensitive agency detector.” 2 As such, its principal function is the detection of predators. We can, of course, evaluate evidence and arrive at a conclusion that overrides the initial automatic response: after the momentary startle response I realize the sound was just the wind. However, when I already have reasons to believe that there are invisible agents in the neighborhood, say because all the people I trust say so, this experience reinforces my general belief. But the theory must account for why they believe it. Why do people initially believe the unobserved entities are there? The threshold for accepting unseen agents as explanations of what we observe is low. The hair-triggered agency detection system accounts for my being primed to jump to the conclusion, at least momentarily, that someone unseen is in the vicinity, but this suspicion, typically, is soon either confirmed or discarded. In the absence of further evidence, I do not
242
Chapter 8
become convinced that someone is there. In contrast, humans persist in believing in gods, ghosts, spirits, still present ancestors, etc., in the absence of empirical confirmation. They remain unseen, yet they are an enduring element of the way humans make sense of the world. For cognitive theorists, part of an explanation lies in the mind’s general thirst for explanation. Human beings have, in Paul Bloom’s words, “a terrible eye for randomness” 3 We are inept at reasoning about the probability of singular events and we find it unsatisfying that a personally significant pattern is a mere matter of chance. We are disposed to see purpose or intention whether it is there or not. Studies of young children suggest that this is innate. Human beings approach the world implicitly assuming what psychologist Deborah Keleman labels a “promiscuous teleology,” spontaneously explaining what we find in nature as the product of intelligent design. 4 This tendency is pronounced with respect to events that are personally important, especially misfortune. As with agency detection, here natural selection favors false positives over false negatives: better to have the false belief that my house burning down a second time was no coincidence, when it is, than to believe it was, when there is an arsonist out to get me. The unobserved beings of religion are attractive explanations for significant events for which no natural explanation is readily available. Whatever the difficulties of rationally justifying belief in supernatural agents, the mind’s unconscious information processing mechanisms make such beliefs intuitively plausible. Human beings naturally find it obvious that they share their world with supernatural beings because they evolved to find it intelligible, not because they are irrational. But the religiously relevant intelligibility is typically of particular, personally important events, not of the cosmos as a whole. 5 In this regard, the theistic religions, with their belief in a Creator, are atypical. Another cognitive system, the theory of mind module, plays an important role in the attempt to account for religiosity as a side effect of the operation of human minds. When sensory information activates the agency detector, it in turn activates the theory of mind module, an information processor dedicated to making inferences about what goes on in the minds of other agents, human or otherwise. 6 If my first thought on hearing an unexpected noise in the woods is that a predator is there, my second is that it wants to do me harm. We do not need to see other agents—animal, human, or divine—to make inferences about their mental states. We spend a great deal of time speculating about and trying to explain the mental states of minds that are not present. The theory of mind module embodies the assumptions of our folk psychology, the innate disposition spontaneously to infer beliefs and desires to make sense of observed behavior, and to predict future behavior on the basis of inferred desires and beliefs. 7 Superficially, the unseen agents of the world’s religions are wildly diverse, yet on closer examination they fall into an unexpectedly narrow
Religion
243
range of possibilities. The mind’s information processing mechanisms embody implicit assumptions about the world. There is, for instance, a cognitive system for dealing with living things, and the assumptions of folk biology are built into it. When sensory input triggers this mechanism, and something is identified as a biological organism, a variety of inferences ensue. Cognitive anthropologist Scott Atran writes: In every human society, people think about plants and animals in the same special ways. These special ways of thinking, which can be described as “folk biology,” are fundamentally different from the ways humans ordinarily think about other things in the world, such as stones, stars, tools or even people. 8
There is, Atran continues, “a faculty of the human mind that is innately and uniquely attuned to perceiving and conceptually organizing living kinds.” 9 When a child sees a cat she already knows, in the sense that this assumption is hard-wired into the mental module dedicated to thinking about animals, that it belongs to a species nested within a taxonomic hierarchy, and that each member of that species possesses a hidden causal nature, an essence that accounts for the typical appearance and behavior of animals of that kind. She expects that when a cat gives birth, it will be to more cats, not to animals of other kinds. This innate biological understanding overrides superficial similarities and differences; the child groups a small black puppy with a large white dog, not with a small black kitten. The child takes for granted that the same individual creature can take on a variety of appearances; she has no difficulty believing that the frog was once a tadpole. Other mental modules are dedicated to identifying and reasoning about inanimate material objects (folk mechanics, folk physics), artifacts, tools, and as we saw above, the minds of agents (folk psychology.) Taken together, the evolved modules of the human mind comprise a finegrained structure of implicit assumptions about the world and its contents: an innate folk ontology. Characteristic of the concepts and categories of our folk ontology, in contrast to other categories and concepts, is the wealth of default inferences they sustain and the large number of expectations they generate. These default inferences manifest our innate assumptions about the world and its contents. However, these ontological expectations can be violated, either in reality, as when scientific taxonomy refutes the natural assumption that whales are fish, or in the imagination, when we hear a story about a dog that gave birth to kittens. The cognitive theorist points to the fate of ideas that violate the expectations embodied in our implicit folk ontology. Ideas that are counterintuitive insofar as they violate these categorical expectations capture our attention and tend to be remembered better and longer than those that do not. A dog that gives birth to kittens is attention grabbing and memorable in ways many other strange ideas are not, say, a dog with a green tail. In
244
Chapter 8
contrast to what is merely unusual, ontologically counterintuitive ideas are more likely to be transmitted from person to person and thus to become socially entrenched. The tale of the dog that had kittens is more likely to be told and retold than the story about a green-tailed canine. However, to be readily remembered, shared, and preserved an idea must be only minimally counterintuitive. It violates just one, or at most a few, of the default expectations of an ontological category. In contrast, prodigiously counterintuitive ideas are too hard to remember and impossible to reason about. Something that violates one of the assumptions of a category holds our attention but leaves the rest of its default characteristics in place, so we still draw a variety of inferences about it. When many default assumptions are violated, the natural inclination to make inferences is stymied. Rather than an arresting idea to dwell on, there is a forgettable list of strange properties. As Justin Barrett points out, “a dog that was made in a factory, gives birth to chickens, can talk to people, is invisible, can read minds, and can walk through walls” is too cumbersome for a good story. 10 Ideas of things that are minimally counterintuitive invite, rather than discourage, further thought. The concept of a dog that is invisible, but otherwise normal, leaves in place an array of default assumptions about dogs, so we automatically make inferences that follow from its being a dog but compatible with invisibility. We wonder whether other dogs will detect it by smell alone, ask whether its invisibility will facilitate catching cats, or speculate on how people will react to barking that seems to come out of thin air. The minimally counterintuitive concepts that successfully compete for space in human minds and eventually secure a place in cultural lore are those that have what Barrett calls “good inferential potential:” they activate a variety of mental modules and elicit explanations, predictions, and interesting stories. 11 Contrasting the idea of an invisible person, rich with inferential potential, with the equally counterintuitive idea of an invisible tree, Barrett notes that concepts that activate the theory of mind module are especially good candidates to become widespread. Humans are endowed with highly developed mental machinery specialized for thinking about the minds of other persons. Our ancestors got to be someone’s ancestors because they were expert at quickly making reliable inferences, on the basis of meager data, about the minds of other human beings, and about the minds of the animals they hunted, or which hunted them. Minimally counterintuitive stories of disembodied agents, whether taken as fact or fiction, activate that cognitive machinery and maximally engage our interest. There is great diversity in the particulars, yet across cultures, minimally counterintuitive ideas, ideas of invisible persons that have minds similar to human minds, are common to human religiosity. The cognitive theory also accounts for the close connection of religion to morality. Our moral sentiments, judgments, and behavioral dispositions do not originate with the gods, spirits, ancestors, or other bodiless
Religion
245
agents of religion. They arise from innate mental mechanisms that are adaptations to social life. However, supernatural persons are conceived as having an interest in the moral quality of human action. Categorizing something as a person activates mental modules for social interaction including those evolved for moral thought and feeling. Divine beings, disembodied and unseen, are well situated to know what humans are up to. A human being cannot be sure that one is not nearby, observing without being observed. Even where a religion makes no theological claim about divine omniscience, the default assumption is that the gods, spirits, or ancestors, not being limited to the perspective of a body, have potentially unlimited access to morally relevant information. 12 When someone contemplates a morally forbidden act, but believes that the gods see her even when no human can, normal moral responses are elicited. In the absence of reasons to think otherwise, the default inference will be that the unseen person believes that the act is wrong, does not want her to do it, and will be angry if she does it. Feelings of guilt and fear of retribution ensue. We saw earlier that part of what makes belief in supernatural agents plausible is the strong drive to explain what might otherwise be thought a matter of chance. Supernatural agents provoked by moral infractions plausibly account for otherwise inexplicable misfortunes. Unexpected good fortune can be explained in an intuitively satisfying way as the result of a supernatural agent rewarding good behavior. Beyond clearly moral considerations lies the general idea that bad befalls people because they have offended unseen but powerful agents, and that otherwise unexplained good fortune results from having pleased them. The spirits, demons, gods, ancestors, and so on of religions are persons, yet they lack bodies. Human beings at all times and places readily believe there are minds without bodies. Cognitive theorists explain this as a result of the fact that while we have one cognitive system activated by observed bodies, and dedicated to reasoning about them, a distinct mental module is specialized for reasoning about the minds associated with those bodies. When we perceive a dead body, the biological organism module registers this and supplies the proper inferences: this organism will no longer move, breathe, eat, and so on. However, as an accident of human cognitive architecture, the theory of mind module does not get the news. Because of this encapsulation it continues making inferences about the mind of the deceased organism, producing conscious beliefs about the dead creature’s psychological states. Young children, told a story about a mouse eaten by an alligator, respond correctly to such questions as, “Will the mouse still eat, breathe, run?” Children know that biological functions cease at death. Yet they respond affirmatively to questions like “Now that the mouse is dead, can it still think and feel?” 13 Human beings are “natural born dualists” who find intuitively plausible the idea of a bodiless entity having beliefs, desires, and intentions. 14 Iron-
246
Chapter 8
ically, thanks to this contingent feature of its neural wiring, the human brain finds highly implausible the thought that it is just a brain, and not something immaterial that transcends the body. This facilitates belief in the supernatural agents of whatever religious culture one is born into, and it helps account for the almost universal belief in life after death and its connection to religion. The same mental mechanisms involved in making us natural dualists make us naturally religious. Pascal Boyer reports, “the connection between notions of supernatural agents and representations about death may take different forms in different human groups, but there is always some connection.” 15 Religions often include some account of what becomes of the soul after death, whether it goes to heaven, hell, moves on to occupy another body, human or otherwise, or simply persists in the neighborhood, now one of the ancestors. Religion is essentially practical, not theoretical. People do not just believe that the gods exist as the causes of otherwise inexplicable events; they seek to interact with them. Individuals pray, enlisting their aid. Sacrifices, sometimes costly, are made to gain favor. Corporate religious ritual, devised to influence them, characterize human cultures. Making anthropological and sociological sense of the multiplicity of forms human religiosity takes as a social reality leads far beyond the narrow, but foundational, project of the cognitive theory. Its focus is the crucial question, “Why do humans so naturally believe in supernatural beings?” In this it is analogous to the evolutionary theory of morality, which leaves still to be explored a vast realm of historically shaped and culturally articulated meanings, while seeking to explain the central fact that human beings have altruistic tendencies. 16 For the most part, humans confidently believe that divine beings exist because trusted fellow human beings say they do, and because human brains happen to be structured in ways that makes these reports seem eminently reasonable. 17 Intelligent creatures with a different evolutionary history, lacking a trip-wired agency detection module, without encapsulated mental modules that produce incommensurable output pertaining to biological and psychological functioning, and without a powerful tendency to prefer an explanation for which there is little evidence to no explanation, would lack the religiosity so deeply entrenched in human nature. A DEBUNKING EXPLANATION The cognitive theory is an intriguing hypothesis widely accepted by many researchers. 18 However, it is not a well-confirmed scientific theory, something we cannot reasonably reject. 19 Arguably, the general evolutionary account of the origin of the altruistic dispositions underlying morality discussed in the preceding chapter approaches this status, but the cognitive theory of religion so far falls short of it. Nonetheless, it is rea-
Religion
247
sonable to regard it as the best scientific attempt to explain why rational human beings so readily believe in the unseen agents that populate religions. 20 Suppose, then, that this explanation of the nearly ubiquitous human disposition to believe in supernatural beings turns out to be correct: would it render religious belief in general, and Christian belief in particular, unreasonable? Were the cognitive theory to become a well-confirmed part of the scientific picture of human nature would we at last encounter a real conflict between faith and science? Scientific inquiry can make beliefs unreasonable by showing that they are false or improbable. It can also make a belief unreasonable by revealing its causes. An explanation debunks a belief if it accounts for it in a way that, while not bringing forth evidence that it is false, nonetheless makes it unreasonable by delineating its causes. The fact that would make the belief true, if it is true, plays no role in the explanation of why it is believed. For example, when we discover that someone’s belief that the Red Sox will win the World Series results from wishful thinking, the belief is debunked. It could be true, but a belief caused by the desire for it to be true is not caused by a reliable process, one that tracks the truth, causing a belief if, but only if, it is probably true. Unfortunately, what one wants to be true bears no systematic connection to what is true. When we discover that this is the cause of a belief, we no longer take it seriously. When we realize that our own beliefs are products of wishful thinking we abandon them. Does explaining the disposition to religious belief as a byproduct of our cognitive architecture have a similar debunking effect? The bare fact that a belief has a cause does not undermine its rationality. All our beliefs, rational or not, have causal histories. Some mental mechanisms reliably produce true beliefs and others do not. I believe that the Red Sox won the World Series in 2013. My believing this is, by some complicated route, causally connected with the events of that October that make my belief true. If those events had not occurred, if the Red Sox had not played, or played and lost, then in all probability I would not believe that they won the series in 2013. Human knowledge is a matter of beliefs having reliable causal connections to the world that makes them true. 21 The debunking potential of the cognitive theory is obvious. Minds like ours would more or less inevitably wind up believing in supernatural beings, whether or not such beings happen to exist, whether or not the beliefs happen to be true. Belief in supernatural agents arises as a predictable feature of human nature, but the cognitive mechanisms that produce them do not track the truth. 22 Science can explain, e.g., why human beings believe that dogs exist, but this explanation depends on the assumption that there are dogs that causally interact with human minds. Science also purports to explain why human beings believe that gods exist, but this explanation does not depend on the assumption that there are gods
248
Chapter 8
that causally interact with human beings. If the explanation succeeds, it does so whether or not the religious beliefs are true. 23 How serious is this debunking challenge? In the preceding chapter, we saw that an analogous challenge, originating in the evolutionary explanation of our moral psychology, does not render moral belief, and the action it motivates, unreasonable. Arguably, what is debunked is the idea that the moral facts are ultimately objective and that they are a source of reasons to act independent of our interests. But the debunking force is limited, for in supplying us with our innate moral sentiments evolution leaves us with reasons to act morally. Perhaps the scientific threat to the reasonableness of religious belief and behavior can also be contained. Justin Barrett writes, “ . . . identifying that a belief has a natural cognitive basis does not bear upon whether it is true, but may justify holding such a belief to be true until sufficient reasons to the contrary arise.” 24 Beliefs we are innately disposed to have are, even when we cannot bring forth evidence for them, “innocent until proved guilty.” 25 Barrett’s point is perfectly sensible. The demand that every belief must be justified by appeal to other beliefs in order to be reasonable is itself unreasonable. We cannot go on forever producing reasons to believe what we believe. There can be no infinite regress of reasons for reasons for reasons . . . If there is anything we reasonably believe, there must be something it is reasonable to believe even though we cannot supply reasons to justify believing it. Some beliefs must be, as epistemologists put it, properly basic. What types of belief are good candidates for being reasonable without explicit reasons? If someone believes something that on reflection strikes her as patently obvious, and she knows of no good reason to doubt it, she has a good candidate for a properly basic belief. Most human beings regard the existence of some kind of divine being as obvious. When they give the matter serious thought, they persist in this conviction even if they cannot come up with good reasons for the belief. Assuming that they know of no good reason to suspect that the religious belief is false, why not regard it as a plausible candidate for something that it is reasonable to believe without justification? Why not agree with Barrett that religious belief is innocent until proved guilty? We should. But doing so does not overcome the debunking challenge. Finding evidence that a belief is false is not the only way for it to lose its innocence. Discovering that those who believe it would do so whether or not it happens to be true is also evidence of guilt. It is not proof that the belief is false, but that is not the issue here. What is at issue is whether it is reasonable. At face value, successful scientific explanation of a belief that makes no reference to what would make it true indicates that it is not reasonable, not a belief that one would have just if it is true. It is revealed as something we would believe whether or not it happens to be true.
Religion
249
DEBUNKING DEFLECTED This debunking potential is narrowly focused. Whatever force it has, it applies only to instances of the belief caused by the suspect mechanism, not to the belief as such. Consider an explanation of how someone acquired religious belief. There is a slave who, as he reaches adulthood, discovers that his master, an agnostic who could not care less whether God exists, subjects his slaves to Christian teaching from childhood. His cynical intent is to make them compliant workers who accept their unhappy lot as God’s will. The slave now realizes that his belief in God is explained in a way that fails to connect it to a reality that makes it true; raised as he was, he would believe in God whether or not God exists. He might now conclude that his belief in God is unreasonable, and that he must abandon it, but he avoids this disheartening conclusion. He realizes that other people share his religious belief, people never in the evil slave master’s control. He seeks them out, asks them why they believe, and they direct his attention to evidence that after careful examination he finds adequate. He continues to believe in God, and he does so reasonably, even though he continues to accept the debunking explanation of how he initially acquired his belief. An explanation that debunks a belief leaves open the possibility of its being reacquired by other routes that make it reasonable. A debunking explanation does not prove that a belief is false, nor does it prove that it is in every case held unreasonably. It challenges its reasonableness, but this challenge might be overcome. A useful analogy lies in conspiracy theories, which are both perennially popular and widely dismissed as irrational. The innate readiness of the human mind to accept explanations of untoward events that involve the intentions of unseen agents figures in the cognitive theory of religion. A reasonable conjecture is that is also explains the common disposition to fix on conspiracy explanations on the basis of minimal evidence. Someone is most likely worse off if enemies are plotting against him and he is unaware of it, than if they are not but he thinks they are. Recognizing this evolved structure of human cognition—the preference for false positives over false negatives—we are skeptical of such explanations when others promote them, and we are reasonably cautious when it comes to offering them ourselves. Some theorists go so far as to contend that while we should acknowledge the possibility, it is never reasonable in practice to accept conspiracy as an explanation. 26 This is an understandable reaction to the proliferation of bizarre conspiracy explanations, but what seems more reasonable is simply to recognize the attraction such explanations can have and proceed with extra caution, accepting them only in the rare cases where there is strong evidence. 27 In particular, any inclination to arrive at a conspiracy conclusion simply because it seems obvious should be dismissed and whatever evidence there might be in its favor should be weighed at a
250
Chapter 8
discount. In light of natural human inclinations, no conspiracy theory deserves the benefit of the doubt. It should be treated as guilty until proved innocent. The cognitive theory suggests the need for similar care with respect to the manifold claims of religion, including those of one’s own religion. Any religious claim should initially be regarded as probably false and should be accepted only on the basis of cautiously evaluated evidence. There is one simple, though not necessarily easy, solution to the debunking challenge: the threat can be neutralized by finding good reasons for religious beliefs, reasons that can be identified independent of the operation of the cognitive mechanisms the theory posits. The reasonable response to the realization that human beings are innately disposed toward certain kinds of belief, for which we actually have little or no evidence, is not to commit ourselves to avoid them altogether. It is to nurture a degree of suspicion toward our inclination to such beliefs, and to accept them only on the basis of evidence carefully vetted and reasoning rigorously examined. 28 It is in particular to reject out of hand claims that it is simply obvious that such and such is the case, and to insist on real evidence. The human mind is replete with cognitive machinery that makes various things seem entirely plausible, irrespective of their truth or falsity. Science can reveal that what strikes us as obvious is what most warrants critical examination. It is common practice to label religious beliefs matters of faith and to assume that this somehow insulates them from demands for rational justification. The contrary view, advocated in Chapter Two, is that at least so far as Christianity is concerned, faith should be understood not as subrational belief, but as trust in the God who resurrected Jesus Christ. Trust can be either reasonable or unreasonable. Faith does not relieve us of the obligation to proportion belief to the quality of the evidence. The cognitive theory of religion, or any plausible scientific explanation of human religiosity, does not debunk the Christian faith, but it constrains any tendency to be overly sanguine about the quality of the justification for the beliefs presupposed by faith in Jesus Christ. Realizing that simply in virtue of being human we are inclined to believe in divine beings on the basis of meager evidence, we see how ill-considered it is to regard our theological convictions as more securely established than the well-confirmed theories of the sciences. We see, too, how inappropriate it is to condemn those who do not accept the claims of Christianity as intellectually dishonest or unreasonable. A proper Christian intellectual humility confesses that the reasons for Christian belief are good enough, not that they are better, or even as strong as, the reasons we have for many other things we believe. When it comes to our beliefs, there is no correlation of quality of evidence and importance. Perhaps, in light of this, it is especially worthwhile to remove spurious grounds for regarding the Christian
Religion
251
faith as unreasonable, such as the idea that it is at odds with what contemporary science shows us about the world and our place in it. DEBUNKING AND DESIGN The cognitive theory of religion, if true, does not debunk religious belief as such, but it does show that it is not properly basic. If we identify entirely natural mechanisms that make it seem simply obvious that there are supernatural agents, then we cannot reasonably believe in them just because their existence seems so obvious. Religious belief might be reasonable, but only if it is grounded in explicit evidence. Some thinkers believe this conclusion can be avoided, that we can accept both the cognitive theory of religion and religious beliefs as properly basic, ones that can be reasonably held even in the absence of evidence. The debunking threat arises because the scientific explanation posits causes of the belief in divine beings that would do their work whether or not such beings exist, and thus whether or not the beliefs are true. Possibly, the belief formation process the cognitive theory describes only appears to be unreliable, and in fact the belief is the effect of a reliable causal mechanism designed by God. Any natural explanation can, in principle, be integrated into a more comprehensive supernatural explanation, one in which the scientifically explained phenomenon was intended by God. What science portrays as a mere accidental byproduct of human evolution might in fact be a product of divine design. What on the surface looks like the interaction of a hodgepodge of disparate evolved mental modular mechanisms could in fact constitute the divinely designed process that reliably disposes human beings to believe in supernatural beings. Paul Bloom writes, “If there is an omnipotent God, then he could have orchestrated the universe so that belief in him could have emerged in any fashion whatever.” 29 The belief in God to which our cognitive architecture disposes us could be reliably connected to the divine reality that makes it true. Innate human religiosity could be God’s design, either because the creation is deterministic, and this apparently accidental outcome is the inevitable outcome of the laws and initial conditions, or because God intervened in human evolution, seeing to it that we acquired mental mechanisms that would not have appeared in the natural course of events. Justin Barrett, contemplating the origin of religion, has a version of the latter in mind when he suggests that God might have “orchestrated mutations and selection to produce the sort of organisms we are—evolution through ‘supernatural selection.’” 30 Science does not debunk belief in God merely by discovering how God caused it. In response to the charge that those who believe in God would believe whether or not the belief is true, those who adopt this strategy, whether in its determinist or interventionist version, are entitled
252
Chapter 8
to say that if God did not exist, then they would not believe that God exists, since if God did not exist there would be no world and no evolutionary processes giving rise to creatures with beliefs about God or anything else. 31 Thus their belief tracks the truth of the matter, as rational belief must. Those who invoke the cognitive theory—or any scientific explanation of the origins of religious belief—to challenge the rationality of belief in God need not agree, but it is not incumbent upon theists to abandon their belief that God created the world and designed its human inhabitants in order cogently to respond to the debunker’s claim. When the reasonableness of a belief is challenged, one does not beg the question by refusing to grant the assumption that one’s belief is false. If human beings spontaneously believe in a divine reality because this is how God intended them to be, then our disposition to believe is not an accident, a fluke that bears no causal connection to a reality that makes such beliefs true. Some Christian traditions explicitly embrace this view. Calvin, probably its most well-known advocate, taught that God has implanted a sensus Divinitatis in human beings. This innate sense of divinity is a disposition, triggered by ordinary experience of the world, to believe that it is the creation of a wise and good God and to feel oneself obligated to worship and obey. 32 Belief in God is reasonable, even when no reasons for it can be articulated, because it is the product of a mechanism divinely designed to produce true beliefs. Some beliefs are, at face value, properly basic, and scientific investigation into their causes does not undermine their reasonableness. To recycle an example, someone has a belief about the past: that he had cornflakes for breakfast earlier today. Asked what he had for breakfast, he pauses for a moment and he finds himself with this belief about the past. He might have had evidence, e.g., dishes in the sink, a stray flake in the beard, a leaked NSA surveillance video, but in this instance he has none. Nonetheless, in the absence of reasons to doubt what he seems to remember, he reasonably thinks that he really does remember. We have some scientific knowledge of the neural mechanisms that cause human beings to have beliefs about the past, and while it indicates that they are not always as reliable as we might have supposed, they do reliably cause true beliefs about the past. The causal explanation that identifies the cognitive mechanisms that provide us with beliefs about the past does not threaten to debunk them. Having acquired some scientific knowledge of the causes of memory beliefs, we remain reasonably confident that what we seem to remember is true. 33 It is possible that what the cognitive theory of religion uncovers are the workings of a divinely designed sensus Divinitatis, one that disposes humans to properly basic, and thus reasonable, beliefs about God. Yet invoking this possibility does not make for a credible response to the debunking challenge. We should reject the orchestrated creation that Barrett proposes. To see this, return to the story of the slave. Suppose that
Religion
253
instead of seeking independent reasons for his religious beliefs, he realizes that what he has learned about the causes of his belief can be subsumed into a more comprehensive religious account. He decides that the master’s evil deeds were the means God employed to engender faith in him and his fellow slaves. Pondering the biblical story of the deliverance of the people of Israel from bondage in Egypt, when God manipulated the mind of Pharaoh to achieve desired ends, he judges that this is just the kind of thing the God in whom he believes would do. The slave’s enlarged account of how and why he was caused to acquire belief in God might be true, but it does not deliver him from the debunking threat. An adequate response to the threat exposure of the master’s machinations poses calls for more than the possibility the slave envisages. He can avoid the debunking of his belief in God only if he has good reasons to believe that this is what actually happened, and those reasons must not be tainted by the master’s program of indoctrination. To incorporate the slave master’s scheme into God’s greater plan, and thereby neutralize the challenge to the reasonableness of his belief in God, he must appeal to what he knows independently of anything the master taught him. For while the slave’s comprehensive theistic explanation could be true, it could instead be true that the master is very clever, and exercised the foresight to prime his victims with the disposition to arrive at precisely these beliefs about his being God’s instrument. Thus, even when the slaves mature and realize what has been going on, rather than regarding their beliefs as debunked, they continue as docile workers. If the slave has nothing to go on other than what he has been taught by the master, he has no reason to think one of these possibilities is more likely than the other, and he cannot reasonably cling to his belief. 34 If the slave master instigated the slave’s conviction that the Exodus story is true it has no power to ward off debunking. Faced with the explanation that appears to undermine his belief, he cannot reasonably defend it by assuming that it is true. As in the first case, his only hope for reasonable belief in the God he was raised to believe in is to find evidence not connected to his master’s machinations. Similarly, if we accept the cognitive theory of religion, we may reasonably continue to believe, and go on to devise an account that locates the mental machinery it describes in God’s creation, but only if we acquire reasons to believe that the cognitive theory does not explain. In the event of a well-confirmed scientific explanation of religious belief, the charge that belief in God is unreasonable can be answered, but not by appeal to its being properly basic.
254
Chapter 8
RELIGION WITHOUT DESIGN Even if appeal to a divinely designed sensus Divinitatis could rebut the debunking charge inherent in the scientific explanation of religiosity, it would be a strategy we have good reasons to avoid. Divine design is fundamentally at odds with God’s purpose of creating persons truly distinct from, and capable of communion with, their Creator. God eschewed design and made use neither of deterministic natural processes nor miracles to bring us into existence. We must seek to integrate the cognitive theory into a larger account that portrays the world as God’s creation without invoking design. Human beings have an innate tendency toward belief in divine beings, but not because this is what God specifically intended. It is, as secular proponents of the cognitive theory tell us, merely a byproduct of human evolution. Therefore, we require an alternative account of the role of natural human religiosity, one that does not portray it as designed. 35 The cognitive theory of religion does not show that Christian belief is unreasonable, but it does highlight the challenge of showing that it is reasonable. It might seem, then, that Christians have no reason to cheer on the project of naturalizing religious belief. However, this is too quick a conclusion. We should reject the attempt to integrate the scientific explanation of human religiosity into a grand scheme of divine design, but there is another, better, way to understand human religiosity and God’s interest in it. Abandoning the idea that our religious instincts are part of a divine design plan, we can proceed to an account of their role in God’s dealings with human beings, one that makes it reasonable to welcome rather than to fear the incursion of science into the domain of religion. RELIGION AND MORALITY From the perspective of Christian faith, we can acknowledge human religiosity as in significant ways analogous to morality. Both are at bottom features of human nature, explicable without appeal to the miraculous or to divine beings. Neither on its face indicates any trace of human contact with anything that transcends the natural world. Both are to all appearances the product of the long and contingent evolutionary history that resulted in Homo sapiens. That history could have followed a different course; if it had, the personal creatures—if any—that would have existed instead of us might have had quite different, or no, moral nature, and no, or a very different, religious sensibility. However, the naturalistic explanation of morality and of religion differ in a significant way: morality is an adaptation while religiosity, although it arises from the interaction of mental mechanisms that are adaptations, is not itself an adaptation. There was selection, e.g., for the desire
Religion
255
to retaliate against humans who cheat us, but no selection for the desire to make sacrifices to placate unseen spirits. The explanation of the former presupposes cheaters; the explanation of the latter posits no angry spirits. The evolutionary explanation of morality undermines certain ways of construing the nature of moral truth, giving us good reasons to doubt the vision of moral knowledge as a grasp of something beyond humanity. But it does not warrant systemic doubt about there being moral truth, even if it points to its being more mundane than some have envisaged. It reveals the superficiality of our moral categories, projected upon the social world, but it does not imply that the actions we have evolved to categorize as, e.g., cheating, do not occur. Natural selection has crafted human minds to detect and respond to free riders. To ascertain that a human being, in virtue of neural circuitry that came about as an adaptation to life as a hunter-gatherer, is inclined toward a cognitive, emotional, and behavioral response to a particular kind of social interaction which she describes as unfair, is not to discover that her moral judgment is mistaken or irrational. It may be firmly rooted in desires she has no reason to disown and good reasons to nurture. Learning the origins of our moral minds gives us some worthwhile critical distance on our moral responses and it teaches us that moral reasons cannot be finally disentangled from our own desires, but it does not debunk morality. 36 The causal explanation of how we came to have moral beliefs reliably links them to the social reality in virtue of which they are often true. Natural selection tracks something in reality that makes a difference to fitness. In contrast, the cognitive explanation of the causes of religious belief reveals no causal connection of belief in supernatural agents to any such beings that actually exist. Science, in the form of the cognitive theory of religion, challenges the rationality of religious belief in ways that evolutionary theory does not challenge the rationality of morality. GOD AND RELIGION Despite the significantly different implications of naturalistic explanations of morality and of religion, there is a perspective from which a deeper similarity comes into view. Contrary to what many suppose, morality has no necessary, originating connection to God. Morality is a naturally occurring feature of the human persons to which the Creator condescends, employing our moral sentiments and judgments when the behavior they prompt corresponds to what God cares about: the beloved persons who benefit from altruistic acts. Rejecting the possibility of design, God utilizes indeterministic means to create persons. Once they exist, God sets about self-revelation, inviting them to know, trust, and love their maker. For this task, God makes use of what is ready to hand, whatever aspects of evolved human nature help elicit a positive response.
256
Chapter 8
An important dimension of the positive response God seeks in personal creatures is for them to share God’s attitude to humans, to love them in a way that mirrors the way God loves them. Nurturing our innate dispositions to altruism—our moral nature—is far from the whole of this, but it constitutes an essential core. We cannot love one another as God loves us unless, at the very least, we act with minimal moral decency toward one another. On the other hand, when the actions that our moral instincts motivate are contrary to the well-being of the persons God loves the divine attitude toward them is critical. Human morality is under divine judgment, not because it falls short of some sort of ultimately objectively correct morality—there is no such thing—but because it is at odds with God’s aims. I suggest that we see human religiosity as given an analogous role in God’s purposeful interaction with human beings. 37 Our natural religiosity, like our natural morality, has no necessary connection to God, but God finds aspects of it of use in the task of bringing humanity into fellowship with our Creator. From the perspective of inveterately religious creatures like ourselves, God’s relation to religiosity can seem confusingly ambiguous. Our Creator appears to be both for and against religion. Christianity is one religion among others. There is no reason to deny that it fits the definition of religion advanced by cognitive theorists. Christians believe that there is an unseen agent who has powers far greater than those of human beings, who knows what human beings are doing and cares about certain aspects of their behavior, particularly behavior that affects human well-being, and whose actions are to some degree intelligible in terms of our familiar psychology of belief and desire. We pray to this being, hoping to exert influence, and we engage in communal rituals intended as acts of worship. Irrespective of whether Christian beliefs can be rationally justified, we know that to all appearances Christians acquire their beliefs about the object of their worship, and reason religiously, in ways essentially similar to those of adherents of other faiths. 38 Similarly, leaving aside the question of whether they are veridical, the religious experiences that Christians report do not appear to involve neural structures or activity essentially different than those involved in the experience of other religious persons. Yet the God revealed to human beings in the Bible, as a human being in Jesus Christ, is strangely ambivalent about religion GOD AND THE GODS Whatever its paradoxical flavor, the idea that God is both for and against religion, that it is placed under judgment while used it in pursuit of divine aims has deep roots in biblical faith. The modern project of justifying belief in an unseen deity is foreign to the authors of the Bible; they
Religion
257
take it for granted that human beings believe in gods. Scripture bears witness to the one true God, the Yahweh of Israel, but other gods lurk in the background. A trajectory from polytheism to henotheism to monotheism is discernible in the biblical narrative. It can be disconcerting to the modern reader of Scripture that initially the existence of various gods is assumed, and that Yahweh at first asserts not that no other gods exist, but that the beneficiaries of the divine self-revelation owe Yahweh their sole allegiance. The Decalogue begins: “I am the Lord your God, who brought you out of the land of Egypt, out of the house of slavery; you shall have no other gods before me” (Exodus 20:2). God continues, prohibiting the worship of idols, “ . . . for I the Lord your God am a jealous God . . . ” (20:5). God accepts human belief in a plethora of gods, accommodating to the innate, and at this point culturally entrenched, religious dispositions of the elect people. However, the movement out of polytheism is soon underway. Yahweh, the God that called Abraham and liberated Israel, is not comparable to other gods. This is the Deity that defeated and humiliated the powerful gods of mighty Egypt. However, once the people are installed in the promised land, Yahweh’s demand for unique allegiance seems unreasonable. Impressive as Yahweh is as a warrior-god, surely other gods must be negotiated with if human needs are to be supplied, if, e.g., people, their animals, and their lands are to be fertile. Thus we find the persistent impulse of God’s people to hedge their bets, to make offerings to the local gods, particularly to those in control of the weather in a land vulnerable to drought. 39 The opposition to this natural religious tendency does not immediately take the form of denying the existence of competing gods, but of undercutting their relevance. Against the temptations of idolatry the prophets insist that Yahweh is supreme among the gods, and can be relied on to provide all Israel’s needs. The demand that this Deity alone be worshipped is reasonable only on the assumption that the people need nothing from other gods. The local gods are useless, not worthy of worship. In fact, these lesser gods worship Yahweh, the supreme Deity, attending the high God in the heavenly court. In due course, the absolute supremacy of God above all gods is articulated in the elegant Genesis account of the creation of heaven and Earth, clearly critiquing the creation mythology of the dominant Mesopotamian culture. The world humans inhabit is God’s good and trustworthy creation. There is no need to appease lesser gods to ensure its proper functioning. In God’s creation, the rains come in their season; there is no need to deal with the storm god Baal to prevent drought. No gods must be placated to guarantee human well-being. The old rituals devised for this purpose are transformed into celebrations of, and thanksgiving for, the Creator’s gracious care of this world and its inhabitants. Condescending to natural human religiosity, God does not immediately expose the nothingness of the gods. It appears that the people of
258
Chapter 8
Israel need first to be assured that the gods are useless before they can be persuaded that they do not exist. When it comes, the announcement is dramatic. A remarkable text, Psalm 82, describes the demise of the gods: God has taken his place in the divine council; In the midst of the gods he holds judgment.
God’s judgment on the gods arises from concern for the welfare of human beings: they have shown partiality to the wicked. They have withheld justice from the weak and the orphan. They have not maintained the right of the lowly and destitute; they have not rescued them from the hand of the wicked. God is just, fully committed to downtrodden persons. But the gods have no knowledge and no understanding; they are in the dark (82:2–5). God pronounces judgment on them: I say, “You are gods, children of the Most High, all of you; nevertheless, you shall die like mortals, and fall like any prince (82:7).
Immortality, the distinctive characteristic of the divine, is taken away; the gods are no longer gods. The henotheistic belief in a single high God, like the polytheistic belief that preceded it, is cast aside, having served God’s purposes. Worship of the gods can now be exposed for what it is: bowing down to what is no god at all. In texts that must have been stunningly irreligious in the ancient world, God’s prophets mock the idols and those who revere them. The prophetic voice recorded in Isaiah 44 describes a craftsman, taking a block of wood, some of which he uses to make a fire to cook his food and to keep warm, and painstakingly fashioning an idol from the remainder. Having made it from this mundane material with his own hands, ludicrously he now bows and prays, “Save me, for you are my god!” (44:12–17). Sounding like an ancient Richard Dawkins, Jeremiah lampoons the pious worshippers of his day: “Their idols are like scarecrows in a cucumber field, and they cannot speak; they have to be carried, for they cannot walk” (10:5). God has made his use of humans’ natural inclination to worship and serve a variety of gods, but it can now be gleefully dispensed with. RELIGION AND CREATION Perhaps, there is nothing remarkable in the idea of a Deity who demands to be the preferred object of religious devotion, and who opposes the worship of competing gods. Nor is it so amazing if this Deity eventually proclaims all competing divine beings non-existent. Yet the biblical narrative, beginning with the story of creation, suggests a deeper and more systemic divine objection to human religiosity. This is difficult to make sense of on the assumption that human religiosity is a product of divine design, a feature of human nature the Creator always intended. If God’s
Religion
259
attitude is broadly negative, even—paradoxically—involving the final rejection of religious practices once called for, this coheres better with the idea that religion is not God’s design, but something in us found to be of use, but of strictly limited use, in God’s attempts to save wayward creatures. The primeval history found in the beginning of Genesis is not, of course, an historical record of actual events. Its aim is to teach us how to see the world through the lens of God’s self-revelation in the real history of Israel. Taken this way, it provides insight into how God sees human religiosity. Since the late nineteenth century it has been known that the biblical account of creation is not sui generis but draws on mythic material from the surrounding cultures, especially those of Mesopotamia. What is significant about the biblical story of origins is not that it is utterly original—it isn’t—but that it deploys the concepts and categories of the Ancient Near East to challenge and subvert the prevailing religious world view. The Bible accepts the cosmology of a flat Earth spread out under the metal dome of heaven in which the sun, moon, and stars are affixed, but it denies the celestial entities their customary divine status. It refers to them not by their proper names, which derive from names of gods, but merely as “lights,” created by the God of Israel and placed in the heavens to make possible the ordering of day and night, and of the seasons, and to provide illumination for the Earth’s inhabitants. They are not divine beings but mere creatures, made by God for the benefit of human beings. As a chronological ordering of events the narrative is incongruous: there is light, the ordering of day and night, and vegetation before the sun comes into existence, but this drives home the point that the sun is not a great, ordering deity, on whom all this ultimately depends, but just one of God’s creatures. 40 Nor, in God’s creation, is there any need to bargain with gods for the fertility of crops or livestock. The Earth gives rise to plants and animals that in the natural course of events reproduce after their kind. The Genesis creation story says nothing about human beings serving God or offering sacrifices. J. Richard Middleton, pointing out the significance of this omission in Israel’s cultural context, writes: Since the primary cultic means of securing divine blessing and fertility in ancient Mesopotamia would have been the sacrificial system—the provision of food and drink for the gods, as part of the imposed servitude of humanity—it is significant that in Genesis 1 it is God who graciously provides food for both humans and animals. 41
The role of human beings in Ancient Near Eastern culture is “to bear the yoke of the gods,” serving them in cultic practices. 42 This sharply contrasts with the role God gives human beings as such, which is to rule over and care for the rest of creation. Edenic humans have a vocation—to till the garden and watch over it—but this is not arduous toil, in contrast
260
Chapter 8
to what they endure after they are evicted and they must feed themselves from the cursed ground. In Mesopotamian religion, humans must feed the gods as well as themselves. The Babylonian Astrahasis epic describes humans as originally created to alleviate the gods of work, but later overpopulating the Earth and becoming a noisy nuisance that the gods deal with by means of disease, drought, famine, and eventually a great flood. 43 The God of biblical faith, in contrast, instructs human creatures: “be fruitful, and multiply, and fill the earth” (Genesis 1:28). Of singular importance is the Genesis assertion that God created human beings in the divine image. Middleton writes, “the Mesopotamian ideology of sacred kingship (which includes the notion of the king as the image of a god) provides an appropriate background for understanding the biblical imago Dei” We noted in chapter 5 that contemporary scholarship favors interpreting the image not as a claim about human nature or capabilities, but in light of the practice of describing a local king as the viceroy of the distant high king, and of regarding the kings themselves as divine, or quasi-divine, images of a deity. In the religious mythology of Mesopotamian civilizations, humans are created to relieve the lesser gods of the arduous tasks of building and maintaining irrigation canals, growing food, building temples, and thus sustaining the elaborate system of cultic sacrifice by which the gods were fed. This mythic explanation of human origins explained and legitimized the actual hierarchical social structure, in which the mass of people labored for the benefit of the ruling royal and priestly classes. In this context, the biblical description of human beings as such, both male and female, as the image of God was a radical critique of religion and the political regimes it justified and maintained. In a world where a human elite claims to represent the divine and rule over lesser humans, this conception of the imago has “revolutionary political potential.” 44 The Creator gives human beings as such, not princes and priests, and not lesser gods, the authority to rule in creation (Genesis 1:28). We do not have to construe this as a critique of religion as such, but what cannot be denied is that it functions as a profound attack on the salient religious mentality of the dominant surrounding culture. One might suppose that God created humans always intending them to worship correctly, and that the creation account is intended to steer the people of Israel away from the false religions of Babylonia and Assyria toward the true religion, but if we do, then with Matthew Myer Boulton we must ask, “Where is the temple in Eden? If God created human beings to offer God thanks and praise, then how is it that the first man and woman receive no such instruction, nor offer any such thing?” 45 At face value, God created humans to care for creation for the purpose of using and enjoy it and communing with its Creator; religious practice has no obvious place in this vocation. Prior to the Fall there are no rituals, no offerings, no altars; all this appears only after humans rebel and damage the
Religion
261
initially intimate relation to their Creator who is empirically—though not yet overtly—present. 46 The account of the aftermath of the first religious dispute, which eventuates in the first murder, concludes with the enigmatic statement, “At that time people began to invoke the name of the Lord” (Genesis 4.26), perhaps a reference to the beginning of religion as a human institution. SACRIFICE Soon after God expels Eve and Adam from Eden, their sons, Abel and Cain, bring offerings to God (Genesis 4). There is no mention of God commanding this, and no explanation is offered of why they do it. It appears that as soon as humans are evicted from Eden, they naturally and spontaneously offer a portion of what they have produced to God. The brothers seem to regard religious sacrifice simply as what one ought to do, as though this is, as our scientific account has it, an innate disposition. It is easy to suspect that Cain, the farmer, feels he needs to give God something to ensure rain and fertility for his crops. If so, humankind’s first religious act manifests a lack of trust in God. Also absent from the story is any explanation of why God accepts Abel’s sacrifice but rejects Cain’s. 47 Eventually, as the biblical narrative develops, the difference between proper and improper forms of worship, and the condemnation of the worship of other gods, become crucial matters in the history of Israel. Even if, as it appears, God does not institute religion, from its inception it is recruited to the divine purposes. In this sense, God accepts religion, but the use of it is critical, not an uncritical endorsement. Abel’s offering of animals to God foreshadows the later religion of Israel, in which the ritual sacrifice of animals occupies a central role. Regulations for a complex system of animal and other sacrifices are spelled out in detail, yet almost nothing by way of explanation is offered for why God wants people to engage in this practice. That good relations with the Deity call for gift-giving is, as the cognitive theory of religion tells us, something human beings find intuitively obvious. The unseen divine agents, differing from human beings in significant ways, are nonetheless similar to humans in any number of other ways, in particular, they might have desires that humans can satisfy, and satisfying those desires might induce them to do things we want. Belief in gods activates the mental machinery evolved for human social interaction, and we find intuitively plausible the prospect of reciprocally beneficial exchanges with them. Like us, they are angered by bad behavior but they can be placated with gifts. Because they are more powerful than human beings, it is dangerous to offend them, so if offerings can make amends, it is wise to do so. (Here we should recall that the disposition to believe in gods is
262
Chapter 8
due in part to our mechanisms for detecting predators, not beings that love us.) Gifts might also recruit the superhuman powers of the gods to our benefit, for example to cure the sick, or to ensure fertility or a good crop. It is natural for human beings to believe that the gods (ancestors, spirits, etc.) are there, and can be bargained with, as we trade what we have and they want, e.g., roast pig, for what they can supply and we want, e.g., victory in the upcoming battle. Even if the unseen god cannot actually eat the offered meat—which remains conveniently available to sustain the priests—he receives it in some way peculiar to a god, by way of the smoke and the delightful smell, perhaps in this way consuming its life or soul. A few Old Testament texts allude to this type of explanation. After the Flood, Noah makes burnt offerings of animals and God, pleased with the odor, promises not to repeat the destruction of humankind (Genesis 8:20–21). And in Leviticus instructions on the proper methods for sacrificing bulls, sheep or goats, and birds conclude with the description, “a burnt offering, an offering by fire of pleasing odor to the Lord (Leviticus 1:9, 13, 17). 48 Whatever its rationale, for Israel the sacrificial activity continued, when possible, until the destruction of the Second Temple in A.D. 70. However, there are countervailing biblical voices. God, having made use of the natural human inclination to religious sacrifice, appears to repudiate it. In Psalm 50 God speaks about sacrifices, at first positively: Gather to me my faithful ones, Who made a covenant with me by sacrifice! (50:5) Hear, O my people, and I will speak, O Israel, I will testify against you I am God, your God. (50:7) Not for your sacrifices do I rebuke you Your burnt offerings are continually before me (50:8)
Whatever God has against the people, it is not the ritual offerings they have made. Yet the psalm continues: I will not accept a bull from your house, Or goats from your folds. (50:9) If I were hungry, I would not tell you For the world and all that is in it is mine. Do I eat the flesh of bulls, Or drink the blood of goats? (50:12–13)
In a similar vein, Isaiah delivers God’s denunciation of human religious efforts: What to me is the multitude of your sacrifices? I have had enough of burnt offerings of rams and the fat of fed beasts; I do not delight in the blood of bulls, or of lambs, or of goats.
Religion
263
When you come to appear before me who asked this from your hands? Trample my courts no more; bringing offerings is futile! (1:11–12)
In what might be the most overtly anti-religious assertion in the Old Testament, the prophet Micah asks: With what shall I come before the Lord, And bow myself before God on high? Shall I come before him with burnt offerings, with calves a year old? Will the Lord be pleased with thousands of rams, with ten thousands of rivers of oil? Shall I give my firstborn for my transgression, the fruit of my body for the sin of my soul? He has told you o mortal, what is good; and what the Lord requires of you, but to do justice, and to love kindness, and to walk humbly with your God (Micah 6:6–8).
Ritual sacrifice supplies God with nothing wanted or needed, and God emphatically repudiates the thought, so deeply rooted in the human mind, that deals can be made with God. God makes a covenant with human beings, but this is ultimately a matter of what God freely gives, not a reciprocal exchange in which humans offer something to appease God. Insofar as the desire to negotiate with the divine and thereby accrue favor is integral to human religiosity, we should read these texts as God rejecting religion. God disowns the cultic practices once commanded. I suggest that this is best explained on the view that God, finding human beings infected with a dangerous but deep and initially ineradicable impulse to try to relate to God in this way, made use of it, regulating and channeling it toward faith, but aims finally to discard it. In the end, what God requires is not humans’ sacrifices but participation in God’s justice: religion is not an end in itself but to draw Israel into God’s cause of knowing, and loving, human creatures. For a people surrounded by religious cultures seeking to propitiate gods by sacrificing not only animals but sometimes children, God devises a religion in which animal sacrifices are made, but rather than being gifts to God—really payments to secure his favor—God symbolically gives the life of the animal to the person who, having sinned, has lost his connection to God and God’s people and is now for all practical purposes dead, beyond the community, a non-person. God neither punishes the sinner nor exacts a payment that sets things right. God brings him back to life. 49 God’s subversive use of the human impulse to propitiate the divine by means of costly sacrifice is vividly illustrated in the binding of Isaac (Genesis 22). Abraham mutely follows God’s instructions to sacrifice his son, but at the last moment God rescinds the command, rejects this supreme act of religious devotion, the sacrifice of the beloved son, and supplies the sacrificial victim in place of Isaac. What is the meaning of an offering if God, not the human individual, supplies the offering? It makes
264
Chapter 8
no sense to imagine that one enters into a mutually beneficial exchange with God when God gives what the one who makes the sacrifices would, in the natural course of events, give. Using the forms under which human beings try to give things to God, God vividly demonstrates who the giver is. The natural religious mentality, insofar as it involves the assumption that we can make ourselves right with God by giving something, is judged and rejected. Just when it looks as though Abraham is about to give up what he values above all—his beloved son and the promised future he represents—God intervenes to give, rather than to receive. CHRISTIAN FAITH AND RELIGION Christians confess that the God revealed through Israel finally offers God’s own life as a sacrifice for sinful yet beloved humankind, letting us do our worst and, in so doing, saving us from ourselves. The Christian faith that proclaims this is one among the world’s many religions, but it embodies the same critical attitude toward religion found in the Old Testament. In the aftermath of its long cultural preeminence, it is easy to forget how dangerously irreligious Christianity was in its early days. 50 Despite the persistent propensity to construe sin moralistically, Christian theology contends that the root human problem is not morally bad behavior, but unwillingness to trust God and the ensuing hopeless quest for self-sufficiency and self-justification. The Edenic temptation is not to moral wickedness; it is to be like gods, judging between good and evil. The aim of making this judgment is, of course, to ascertain what is good and to do it, to decide for oneself. It is to deny the reality of what humans are, not gods, but mere creatures, made of the dust of the ground, yet called to partnership with their Creator. Refusing that vocation, ironically humans find themselves not much like gods at all, but vulnerable creatures cast into a perilous world haunted by gods whose approval and help they can only hope to earn. And thus human religiosity becomes the epitome of human sinfulness, the place where we are most emphatically proceeding on our own, trying to be rightly related to the divine by our own efforts. Historically, this comes into sharp focus in the theology of Martin Luther. Contrasting salvation by grace, through faith alone, not by the ubiquitous human attempt to justify oneself, he writes, “all the religions and forms of worship under heaven have been thought up by men to obtain righteousness in the sight of God.” 51 This applies to Christianity as a human institution no less than to other religions: The self-righteous, who do not have faith, do many things. They fast, they pray, the make the sign of the cross. Yet in their hearts they envision God not as he is but as an angry judge who must be placated by their works. All this religious activity is reasonable: of course our correct
Religion
265
belief and right action will make us right with God. But proceeding with what seems reasonable they dismiss the gospel of God’s grace as foolishness, and thus “they depose God from his throne and set themselves up in his place.” 52
Turning away from God, fallen humans turn inward and become—in Luther’s startling image—curved in upon themselves (homo incurvatus in se), self-centered, self-satisfied, self-justifying. 53 “Man,” Luther contends, “is so turned in upon himself that he uses not only physical but spiritual goods for his own purposes and in all things seeks only himself.” 54 Sinful human beings imagine that they are justified before God by getting their religion right, but in reality it is precisely here that they are most desperately closed to God and trapped in sin. Luther condemned late-medieval Christendom as a corruption of authentic Christianity, a false religiosity of works—vain attempts to appease God—in opposition to the good news of salvation by grace, through faith, alone. Arguably, as far as he went, he did not push the critique of our religious dispositions to its logical—even if paradoxical— conclusion. Did Luther grant the authentic Christian religion an exemption from the general condemnation of religion as sin at its most entrenched and dangerous? 55 Perhaps, but in any event, four centuries later Karl Barth emphatically reaches this conclusion and in the most uncompromising terms describes even Christianity—in fact especially Christianity—as subject to God’s judgment on religion. Barth rejects the notion that human beings can escape religion, either by being thoroughly secular, or by arriving at some kind of post-religious Christianity. 56 “As men living in the world, and being what we are, we cannot hope to escape the possibility of religion.” 57 But he also accepts that religion is always idolatry, self-righteousness and unbelief, and that Christianity is no exception. “On the contrary, it is our business as Christians to apply this judgment first and most acutely to ourselves . . . ” 58 We must, in fact, confess that Christianity is religion at its worst, for it is here that there is explicit witness to God’s revelation that salvation is by grace alone, and thus we rely on our religious works not unknowingly but “with a high hand.” 59 Karl Barth’s theological claims correspond to experience. There is no reason to believe that, whatever overt theology of grace we profess, persons of Christian faith are essentially different than the rest of the human species with respect to our deep religious inclinations. Thought, feeling, and action betray the ineradicable suspicion that God can be bargained with, that our good behavior and other “sacrifices” obligate God on our behalf, that our correct theology—including, absurdly, the avowed belief that God loves us unconditionally—puts us in a privileged position. We inevitably draw lines that separate insiders from outsiders, the faithful and the faithless, the good and the bad, as though all such lines run
266
Chapter 8
between us and those beyond God’s grace, rather than through each of us. As Christians we partake of universal forms of human religiosity: we bow, pray, sing, worship, perform rituals said to be prescribed by God, and we preach and teach, claiming to proclaim and imbibe the divine word. Barth would not have us forget that in all this we are engaging in the paradigmatic activity of sinful humankind. It is closer to the truth to confess that we Christians, like everyone else, are saved ultimately not because of, but despite, our religion. Yet Barth would also insist that Christianity is the one true religion. Using Luther’s famous slogan, he insists that we can call it the true religion only in exactly the same sense that we can call ourselves justified sinners. Christianity is true not because here religious humans do better than they do anywhere else, but only because it is here that God has acted, is acting, and will act, to save, disrupting human life, adapting our religiosity to God’s own purposes. The biblical authors witness to a God who makes room for human beings, giving us space to be ourselves even when this means letting us do our worst, letting us ignore, reject, and rebel to the limit, and in the end using what is worst in us to subvert our rebellion and save us. In the story’s climax God falls into the hands of angry sinners who condemn the incarnate God as religiously and politically dangerous and put him to death. This was the all too likely fate for a God who so recklessly loves creatures who everywhere and always turn away, aspiring to be not beloved creatures in communion with God, coworkers in creation, but like gods, going it alone. Where human rejection of God is at its worst, precisely there God acts to make reconciliation. It is, as Matthew Myer Boulton points out, “ . . . only fitting that God’s healing activity should occur in the fatal wound itself . . . ” 60 God speaks the decisive good word to humankind, the refusal to accept our rejection and betrayal as the last word, by way of religion, where that rejection makes it natural home. It is as religious beings, in particular as adherents of the Christian religion, that we know ourselves as sinners, in need of God’s saving grace. On this theological account natural human religiosity makes its appearance not as a gift from God, but as a natural proclivity subject to divine judgment. Yet God condescends, commandeers it and turns it to creation’s ultimate ends. In light of this, persons of Christian faith have no reason to look askance on the cognitive theory of religion. That our religious instincts have a humble, even embarrassing origin is a discovery we can with due humility embrace. 61 As with our moral nature, so with our religiosity: the possibility of a scientific explanation that puts it in its natural place is a promise, not a threat. The idea that it, like morality, is no more than a natural feature of being human that God can utilize for our benefit coheres with a proper Christian ambivalence toward religion.
Religion
267
The Christian faith itself demands that we be highly critical, even suspicious, of our religious attitudes, beliefs, and practices. What to our innate religiosity seems obviously right may be at odds with God’s selfrevelation. We may be wired to conceive God on the model of an unseen predator that can be rendered harmless by our sacrifices and manipulated by our rituals, but God has something very different to say. We saw above that facing up to the debunking potential of the cognitive theory of religion calls for a self-critical attitude toward our most basic religious beliefs. The fact that it seems obvious that God exists is not a good reason to believe it. The cognitive architecture of human minds explains this psychological phenomenon without assuming the belief is true. Reasonable belief that God exists is possible only on the basis of careful attention to explicit evidence. Similarly, within the Christian religion, what seems obviously correct stands in need of ongoing biblical and theological scrutiny. Science and the Christian faith agree in counseling a critical attitude to our natural pieties. We can embrace the scientific project that unearths the absurd roots of our religious instincts, but we cannot regard our own religious beliefs and practices with complacency. RELIGION AND THE SOUL The cognitive theory explains religion’s connection with death and what follows it. The architecture of human cognition that makes plausible the idea of minds without bodies, and thus enables belief in unseen divine agents, also makes for ready belief that human minds continue to function after the death of the body. Disembodied minds, whether humans after death or divine agents, are naturally plausible because the mind’s information processing modules do not efficiently communicate with one another. One mental module processes sensory information and arrives at the conclusion that a biological organism is dead, and this information proceeds to conscious awareness, but the theory of mind module does not get the word; it continues to output beliefs about the mind of the deceased. 62 This manifestation of encapsulation is, presumably, a contingent feature of the human mind. If the neural wiring of our material minds differed slightly differently we might find gods, souls and postmortem life highly implausible. 63 On the contemporary scene, belief in the immortal human soul and belief in God are for many so closely intertwined as to be inseparable, and religion is understood as offering the means to a happy post-mortem life. Science has to a degree eroded this, but the assumption that human beings are, by nature, immortal, because the soul, bearer of what is essential to the person, has a disembodied life after the body dies, and that we can anticipate reward or fear punishment, remains widespread. This convergence of a dualist view of human nature and religion, sustained by
268
Chapter 8
innate cognitive mechanisms and culturally reinforced and elaborated, is taken for granted even among Christians. In light of this it is worth recalling that the creation story in Genesis denies the natural immortality of human beings. The primal human couple is expelled from Eden not, at least not explicitly, as a punishment, but to prevent their eating from the Tree of Life and living forever (3:22–24). Eve and Adam, having eaten from the Tree of Knowledge, are now like gods, knowing good and evil— a sarcastic description?—and this is God’s reason for sending them into exile. The combination of this knowledge and immortality is, in God’s eyes, to be avoided, perhaps for the well-being of the human beings themselves. Denied access to the Tree of Life, humans die. 64 In keeping with this, and in contrast to the religions of some of its powerful neighbors, ancient Israel singularly lacks interest in what becomes of humans after death. The Old Testament shares the idea, common in the ancient world, that the dead descend to the underworld and there continue to exist. They are imprisoned in Sheol, the gloomy realm of the shadowy, lethargic, gibbering, and perhaps only semi-conscious dead. John C. Cooper writes, “the afterlife, if it can be called that, is hopelessly pale and dull in comparison with the shalom of a full earthly life.” 65 In Sheol, the insubstantial residue of the living human being is cut off from the air and light of the human community and at the farthest remove from God in heaven. We may call this insubstantial existence life after death, since we equate life after death with continued existence after death, but this would sound odd to the ancients; for them, what follows dying is death, neither non-existence nor genuine life. The Old Testament’s attitude to Sheol exemplifies its hostility to the religious inclinations of its neighbors. Commerce with the human dead and the gods and spirits, seeking their counsel or aid by divination, was a pervasive feature of ancient religion, but prohibited in Israel. Isaiah condemns those who “consult the ghosts and familiar spirits that chirp and mutter . . . who consult their gods, the dead on behalf of the living” (8:19). Communication with the spirit realm is condemned as disloyalty to Yahweh, whose provision for the people is sufficient; they need not look to a divine underworld for help (Leviticus 19:31, 20:6; Deuteronomy 18:11). 66 The people of God looked ahead in vicarious hope, not to a personal post-mortem life, but to future life for Israel as a people and their descendants in particular. N. T. Wright reminds us, “for the vast majority in ancient Israel the great and solid hope, built upon the character of the covenant god, was for Yahweh’s blessing of justice, prosperity and peace upon the nation and land, and eventually upon the whole earth.” 67 However, the idea of post-mortem personal salvation is not entirely absent from the Old Testament. A handful of texts at least hint that God will deliver people as individuals from Sheol, e.g., “Your dead shall live, their corpses shall rise, O dwellers in the dust, awake and sing for joy!” (Isaiah 26:19). Crucially, this is not a glimpse of a blessed existence in a shadowy
Religion
269
underworld, nor does it refer to disembodied existence. They point to the resurrection of the embodied human individual, to complete deliverance from death and restoration to full, physical, human life in God’s good creation. Unambiguous expressions of faith that God will resurrect human beings are rare in the Old Testament, but Jewish belief in resurrection of the dead became widespread during the four centuries between the last Old Testament prophets and the time of Christ. 68 This resurrection hope was sharply at odds with the ideas of post-mortem existence found throughout the ancient world. It was not an aspiration to a pleasant version of Sheol, translation to the heavenly abode of the gods (on the model of Enoch in Genesis 5:28 or Elijah in 2 Kings 2:11), or the disembodiment of the soul envisaged by Greek philosophers. It was hope for a fully embodied human life in God’s restored material creation. This hope emerges not from reflection upon human nature, but from experience of the character of God witnessed in Israel’s history. The God of Israel is faithful and has the last word on what lies in store for them. Even when persecution results in death, Israel’s God can be trusted to vindicate the martyred. God is the Creator and Lord of all, and not even death can finally defeat God and nullify the covenant promises. Human religious imagination generates notions of an escape from this world to something better, but faith in the God who resurrects the dead to authentic human life in the redeemed future of creation challenges that natural religiosity. In the Gospels, Jesus’s friends and disciples believe in the resurrection of the body, as do the Pharisees, in contrast to the more conservative Sadducees, who mock the idea. 69 As the Christian faith arises out of the witness to the resurrection of Jesus, the New Testament writers take pains to stress that the resurrected Jesus is no ghost or spirit but flesh and blood. That the same God who resurrected Jesus in the first century will in due course resurrect those who put their trust in Christ remains an essential Christian confession. Christianity has no good use for the widespread conviction that we possess immaterial, immortal souls. In the natural course of events, the death of the body is the end of human life and existence. Christian hope is not in natural immortality but that the faithful God who miraculously resurrected Jesus will likewise resurrect us, bringing about what would never occur in the natural course of events. However, before the immaterial, immortal soul can be disposed of an underlying problem must be dealt with. The problem is, at bottom, not theological but metaphysical. The Christian theological confession is the resurrection of the body, a future event that will take place for everyone at once. The Christian creeds are silent about what becomes of the human being between death and this general resurrection. Some Christians believe that in this interim state the dead are simply unconscious, experiencing nothing until they are resurrected. Others believe that the dead are in some manner aware and find themselves in God’s presence. Both views
270
Chapter 8
assume that the human beings, though not yet resurrected, exist. If we continue to exist after we die but before we are resurrected, then we must be something other than, or at least more than, the body. Sooner or later, the body stops existing. If something continues to exist it must be the immaterial soul (or mind) of the person. Thus, while a dualist theory of what humans are is not, in itself, a component of Christian theology, on this account it is an unavoidable presupposition of the essential doctrine of bodily resurrection. Some of the reasons to accept a materialist understanding of human nature were presented in chapter 5. There I also pointed out the inadequacy of various biblical and theological reasons offered on behalf of dualism. However, I deferred the crucial issue of whether belief in life after death demands an immaterial soul until our final chapter. We now turn to this and ask whether we have at last arrived at a genuine contradiction between scientific naturalism and the Christian faith. Scientific naturalism contends that human beings are material things for whom death, in the natural course of events, is the end. Christianity affirms that though we shall die, we shall live. The conviction that these cannot be reconciled is common. The principal task before us is to make the case that we can faithfully embrace both materialism and resurrection. Belief in the soul and its post-mortem existence is a religious inclination from which the Christian faith frees us. Our religiosity, like our morality, is among creation’s unintended consequences which God, patiently resourceful in our deliverance, puts to use. That deliverance triumphs over the grave, as we will see in our next, and final, chapter. NOTES 1. “Religious thought and behaviour as by-products of brain function,” in Trends in Cognitive Science 7 (2003), 123. 2. Justin Barrett, Why Would Anyone Believe in God? (Walnut Creek, CA: Altamira, 2004), 32–4. 3. Paul Bloom, “Religion Is Natural,” Developmental Science 10(2007), 150. 4. “Are Children ‘Intuitive Theists?’ Reasoning about Purpose and Design in Nature” in Psychological Science 15(2004), 295–301. 5. Pascal Boyer, Religion Explained: The Evolutionary Origins of Religious Thought (New York: Basic Books 2001), 13. 6. This, rather than introspection, may well be the main source of knowledge of one’s own mind. See Peter Carruthers, The Opacity of Mind: An Integrative Theory of SelfKnowledge (Oxford, UK: Oxford University Press, 2011). 7. See Bertram F. Malle, How the Mind Explains Behavior: Folk Explanations, Meaning, and Social Interaction (Cambridge, MA: MIT Press, 2004). 8. Scott Atran, “Folk Biology and the Anthropology of Science: Cognitive Universals and Cultural Particulars,” in Behavioral and Brain Sciences 21(1998), 547–69. 9. Atran, “Folk Biology and the Anthropology of science…” 549. 10. Barrett, Why Would Anyone Believe In God? 23. 11. Barrett, Why Would Anyone Believe in God? 25–26. 12. Pascal Boyer, Religion Explained, 156.
Religion
271
13. Jesse M. Bering and David F. Bjorklund, “The Natural Emergence of Reasoning about the Afterlife as a Developmental Regularity” in Developmental Psychology 40 (March 2004), 217–33. 14. See Paul Bloom, Descartes’ Baby: How the Science of Child Development Explains What Makes Us Human (New York: Basic Books, 2004). 15. Pascal Boyer, Religion Explained, 204. 16. Roger Scruton, in, e.g., The Face of God (New York: Continuum, 2012), 1–49, stresses the limitations of evolutionary explanations of morality and religion. 17. In some ways analogously, a child raised in contemporary society takes as given the existence of germs, unseen but dangerous entities. He believes this simply because older people warn him about them and instruct him in rituals devised to keep them at bay, like hand washing. 18. Evolutionary psychologist Jesse Bering reports, “…all but a handful of scholars in this area regard religion as an accidental byproduct of our mental evolution.” The Belief Instinct: The Psychology of Souls, Destiny, and the Meaning of Life (New York: Norton and Company, 2011), 6. 19. Critics argue that the theory depends on controversial and unconfirmed hypotheses about the nature, and even the existence, of the mental modules, and that, even if the modules the theory posits exist, it is not confirmed that they produce the religious beliefs they allegedly explain. See, e.g., Russell Powell and Steve Clarke, “Religion as an Evolutionary Byproduct: A Critique of the Standard Model,” in British Journal for the Philosophy of Science 63(2012), 457–86. For a trenchant analysis of the limits of the modularity thesis, especially in explaining the human ability to cope with varied and changing local environments, see Kim Sterelny, The Evolved Apprentice (Cambridge, MA: MIT Press, 2012). 20. Competing explanations portray religiosity as an adaptation. Typically they invoke group selection: religion is a universal feature of human societies because it promotes social cohesion. See, e.g., David Sloan Wilson, Darwin’s Cathedral: Evolution, Religion, and the Nature of Selection (Chicago: University of Chicago Press, 2002). Whether selection at various group levels, in contrast to selection at the level of the individual organism or the gene, is more than a rare phenomenon, is controversial among evolutionary theorists. The cognitive theory’s unique value lies in its promise plausibly to explain why individual human beings were initially prone to believe in entities for which there is at best minimal evidence. It may well be true that if most everyone in a society has particular beliefs and motivations, then that society will enjoy enhanced cohesion, and this will enhance its members’ inclusive fitness. Once this is the case, and perhaps seen to be the case by the members of this society, individual selection can come into play: at the limit, individuals who lack the belief are denied the ability to reproduce by being killed. But we still need to explain why, originally, individuals began to believe something that, prior to becoming socially entrenched, lacks justification and demands costly behavior. 21. In chapter 5, I disputed the anti-naturalistic contention that the mere fact that there is a causal explanation of a belief renders it irrational, but some causal explanations do undermine the reasonableness of belief. 22. Exchanging the cognitive theory for an adaptationist explanation would not solve this problem. Nor would giving up on evolutionary explanation altogether and retreating to earlier explanatory attempts that invoke the social function of religion (e.g., Durkheim, Marx) or individual psychology (Feuerbach, Freud). None of these explanations require religious beliefs to be true and thus all threaten to debunk the beliefs they explain. 23. In a way, this debunking mirrors the reasoning we saw in chapter 5, whereby C. S. Lewis contended that knowledge of what naturalism implies about the causes of naturalistic belief undermines that belief, revealing it as irrational. 24. Justin Barrett, Cognitive Science, Religion and Theology: From Human Minds to Divine Minds (Templeton Press, 2011), 108. 25. Barrett, Cognitive Science, Religion, and Theology, 108
272
Chapter 8
26. For example, Brian L. Keely, “Of Conspiracy Theories” Journal of Philosophy 96 (1999), 109–26. Whatever plausibility this has seems to depend on the practice of labeling an explanation a conspiracy theory only if one believes it is false and absurd; conspiracy explanations one accepts do not count as conspiracy theories. “Conspiracy theory” is in practice short for “unsubstantiated conspiracy theory.” Those who believe that John Wilkes Booth belonged to a group seeking to kill not only Abraham Lincoln, but also the Vice President and the Secretary of State, dismiss those who think that Lee Harvey Oswald did not act alone on November 22, 1963, as “conspiracy buffs.” And those who take for granted that it was not a coincidence that four airplane hijackings occurred on the morning of September 11, 2001, but believe that this was the work of a conspiratorial group, dismiss the idea that the United States government was behind the attacks as a conspiracy theory. Many explanatory hypotheses are unreasonable, including many conspiracy theories, but not simply because they involve conspiracies. 27. The reasonable attitude here is analogous to a reasonable attitude to claims about miracles, in contrast to David Hume’s well-known contention in Sec. 10 of An Enquiry Concerning Human Understanding, Eric Steinberg, ed. (Indianapolis, IN: Hackett Publishing Co., 1993) that it is never rational to accept reports of the miraculous. 28. J. L. Mackie, in The Miracle of Theism: Arguments for and against the Existence of God (Oxford, UK: Clarendon Press, 1982), 7–8, writes, “if we can show that a certain belief would be almost universally held, even it if it were groundless . . . there is no clear onus of proof on either side.” This esteemed philosopher, who was an atheist, seems to me overly generous to theism. I take it that a heavy burden of justification falls to those of us who persist in belief in the face of debunking explanations. For a Christian’s perspective, see Merold Westphal, Suspicion and Faith: The Religious Uses of Modern Atheism (New York: Fordham University Press, 1998). 29. Paul Bloom, “Religious Belief as an Evolutionary Accident” in Jeffrey Schloss and Michael Murray, eds. The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origin of Religion (Oxford, UK: Oxford University Press, 2009), 126. 30. Justin Barrett, Why Would Anyone Believe in God? 123 31. See, e.g., Michael J. Murray, “Scientific Explanations of Religion and the Justification of Religious Belief,” in The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origin of Religion, Jeffrey Schloss and Michael Murray, eds. (Oxford: Oxford University Press, 2009), 168–78. 32. Institutes of the Christian Religion, Ford Lewis Battles, trans. and ed. (Philadelphia: Westminster Press, 1960), I, xi, 8. For a close examination of the interpretive issues see Paul Helm, “Natural Theology and the Sensus Divinitatis” in his John Calvin’s Ideas (Oxford, UK: Oxford University Press, 2004). 33. Scientific inquiry itself could not, of course, get far without relying on memory, so any wholesale discrediting of its reliability would be self-defeating. The same is true of perception and reasoning. The interesting results are those in which scientific investigation reveals the complex profiles of reliability and unreliability in the operation of our faculties; on memory see, e.g., Daniel L. Schacter, The Seven Sins of Memory: How the Mind Forgets and Remembers (New York: Mariner Books, 2002). 34. The imagined case of the indoctrinated slave, while extreme, is similar to that of anyone raised in a religious home. A normal human child believes whatever she is taught about religious matters and many other things; she is not unreasonable or intellectually defective in believing that there is a God if this is what her parents believe. When she reaches adulthood, she will have any number of beliefs that she cannot justify; they are simply things she has “always” believed. In the absence of reasons to doubt them, it is reasonable for her to continue to believe them. This intellectual conservatism, applied to beliefs acquired as a matter of childhood credulity, renders many beliefs reasonable even in the absence of justification. However, as soon as she comes upon reasons to doubt them, she can no longer reasonably believe without explicit justification. Likewise, once she encounters a plausible debunking explanation
Religion
273
of how she and her forebears acquired a belief, she can no longer reasonably continue with it without justification. 35. A virtue of this account is that it makes the vast diversity of human religiosity unproblematic. However, some explanation is required of why, if the innate mechanisms that incline to religious belief were designed by God, they do not reliably produce theistic belief. Why would God, intending to equip human beings with an innate disposition to religious belief, endow us with one which is not only generic, but readily triggered to produce polytheism or animism? Why not simply belief in the one God? Calvin held that the original endowment was theistic, but among the noetic effects of the Fall was the degradation of the sensus Divinitatis. In sinful human beings a defective sense of divinity remains; corrupted by human pride and arrogance, it has become “a perpetual factory of idols” (Institutes, I. 11. 8). Whether this is psychologically, anthropologically, or historically plausible is, at best, arguable. We may ask, for example, whether there is evidence of devolution from theistic to animistic and polytheistic religion in the human past. 36. Recall, though, that fictionalism rejects this happy conclusion and contends that the evolutionary explanation implies that our moral beliefs are systematically false. In contrast to the cognitive explanation of religious belief, which is consistent with the existence of the unseen beings in which the religious believe, the fictionalist says that the implication is that moral beliefs are false, not merely unreasonable. 37. We may note the analogy with government, where humans purport to know good and evil and force others to comply. The people of Israel want a king, like other nations (1 Samuel 8:1–22). God warns them of what vesting such godlike power in human hands entails, but their desire does not abate. God condescends, acquiescing to the establishment of a monarchy, one subject to God’s critical judgment yet used to bring about Israel’s, and the world’s, salvation in Christ, “Son of David.” Christ, Israel’s true king, is condemned as a seditious blasphemer and put to death, a joint effort of the religious and political powers that enthrall and subjugate human beings. 38. For example, the disparity between the “theologically correct” conceptions of deities and the concepts the faithful actually employ in daily life appears to be the same for Christians and other believers. What the Christian theist officially believes as a result of religious instruction, e.g., God is infinite and omnipresent, can be trumped by implicit anthropomorphic ideas shaped by the innate information processing that gives rise to religion in general. See Justin Barrett, Born Believers: The Science of Children’s Religious Belief (New York: Free Press, 2012), 57–166. For more discussion of the experimental results see, Justin L. Barrett and Frank C. Keil, “Conceptualizing a Nonnatural Entity: Anthropomorphism in God Concepts” Cognitive Psychology 31(1996), 219–47. 39. See, e.g., Tikva Frymer-Kensky, In the Wake of the Goddesses: Women, Culture, and the Biblical Transformation of Pagan Myth (New York: Free Press, 1992), 83–99. 40. Leon R. Kass, The Beginning of Wisdom: Reading Genesis (Chicago: University of Chicago Press, 2006), 30–31. 41. J. Richard Middleton, The Liberating Image: The Imago Dei in Genesis 1 (Grand Rapids, MI: Brazos Press, 2005), 211. 42. Claus Westermann, Genesis, translated by David E. Green (Edinburgh: T. & T. Clark, 1988), 11. 43. The account of the Flood in Genesis clearly derives from Mesopotamian traditions. See, e.g., Peter Enns, Inspiration and Incarnation: Evangelicals and the Old Testament (Grand Rapids: Baker Academic, 2005). However, the biblical explanation of the legendary deluge is radically different: it is God’s attempt to save a remnant from extinction threatened by human violence, not to exterminate humanity. 44. Jürgen Moltmann, God in Creation: A New Theology of Creation and the Spirit of God, translated by Margaret Kohl (Minneapolis: Fortress Press, 1993), 219. 45. Matthew Myer Boulton, God Against Religion: Rethinking Christian Theology Through Worship (Grand Rapids, MI: Eerdmans, 2008), 64.
274
Chapter 8
46. The story describes Eve and Adam hearing God speak and walk in the garden, but it gives no indication that they see God. 47. God’s response to human religion is not easy to predict. Yoram Hazony, in The Philosophy of Hebrew Scripture (Cambridge, UK: Cambridge University Press, 2012), 107–110, points out how surprising it is that God rejects Cain’s offering, despite the fact that it is he, not his brother, who submits to the divine pronouncement (Genesis 3:17–19), making his living by agricultural toil. Abel, seemingly ignoring the curse, turns to sheepherding, yet it is his sacrifice that God accepts. I take this as an indication that the “curse” is to be read not as punishment for the Fall, but among its natural consequences, ones God approves of human efforts to mitigate. 48. Modern readers automatically read this as metaphorical, but we need not suppose ancient readers did. 49. See Colin Gunton, The One, the Three and the Many: God, Creation and the Culture of Modernity (Cambridge, UK: Cambridge University Press, 1993), 225–27. 50. For a vivid portrayal consult Larry W. Hurtado, Destroyer of the Gods: Early Christian Distinctiveness in the Roman World (Waco, TX: Baylor University Press, 2016). 51. Luther’s Works, Jaroslav Pelikan and Helmut Lehmann, trans. and eds. Vol. 26 (Saint Louis, MO: Concordia Publishing House, 1963), 231. 52. Luther’s Works, Vol. 26. 229 53. Luther’s Works, Vol., 25. Hilton G. Oswald, ed. (St. Louis, MO: Concordia Publishing House, 1972), 291 54. Luther’s Works, Vol. 25. 345. 55. Matthew Myer Boulton, in God Against Religion, 162–64, suggests that this is so, and that it was left to Barth to come to grips with the matter. 56. Barth construes as essentially religious all projects that aim at achieving the good, including moral programs and political ideologies: even if they do not include belief in unseen divine agents, they are nonetheless human attempts to secure themselves in the right. Epistle to the Romans, Edwyn C. Hoskyns, trans. (Oxford, UK: Oxford University Press, 6th ed. 1968), 241–54. 57. Barth, Epistle to the Romans, 230. 58. Karl Barth, Church Dogmatics, I.2. G. T. Thomson and Harold Knight, trans. (Edinburgh: T. & T. Clark, 1956), 326–27. 59. Barth, Church Dogmatics, I.2, 337. 60. Boulton, God Against Religion, 165. 61. In the preceding chapter, I speculated about how we might conceive the Fall and original sin in terms of the actual contingencies of human history: we are born into a world in which God is not empirically present. A further conjecture along these lines is that the explicit empirical presence of God renders the disposition to take seriously the existence of unseen agencies harmless. In a redeemed world, in which we see God with us as and as one of us, the output of these interacting mental mechanisms acquires no individual credence or social currency. In the absence of God appearing to us incarnate, the output of our cognitive architecture runs amok. 62. In many cultures, the connection is direct: the unseen agents with whom religious ritual deal are post-mortem human beings, the spirits of the ancestors. 63. This might explain why there are people, such as your author, who find in themselves no inclination to dualism and, perhaps correspondingly, do not find it at all obvious that any sort of divine beings exist, even in the face of systematic childhood training in favor of both. Our cognitive architecture might be abnormal in this regard. 64. In The Garden of Eden and the Hope of Immortality (Minneapolis: Fortress, 1993), James Barr contends that the story should be construed primarily as an account of why and how humans lost their chance for immortality. 65. John C. Cooper, Body, Soul, and Life Everlasting: Biblical Anthropology and the Monism Dualism Debate (Grand Rapids: Eerdmans, 1989), 41. 66. In a well-known episode, King Saul, rejected by God and desperate for help, consults a medium who calls up the recently deceased prophet Samuel. The prophet—
Religion
275
referred to by the witch as an elohim, i.e., a “divine being” (NRSV) or a “god” (RSV)— denounces Saul and reiterates God’s judgment on the king (1 Samuel 28:6–20). Christian materialists deny there are disembodied deceased humans, but what about other putatively disembodied persons? Scripture depicts angels as visible, and thus as material. Whether these messengers are human beings, as sometimes appears to be the case, or non-human material persons from elsewhere in the universe, is a further issue. Demons, on the other hand, are considered immaterial. Some biblical accounts of demonic possession may describe mental illnesses, perhaps of unfamiliar kinds. In other cases, my speculation, drawing on the computational theory of mind, is that they are errant, sub-personal fragments of “code” that infest the human mind and cause aberrant behavior, and that can replicate in other minds. This phenomenon may require cultural conditions that occur in some eras and locales but not in others. Thanks to Col. Fred Chesbro and Dr. Robert Halgren for raising this issue. 67. N. T. Wright, The Resurrection of the Son of God (Minneapolis: Fortress, 2003), 102. 68. Wright, The Resurrection of the Son of God, 129. 69. John’s Gospel records Martha speaking to Jesus about her dead brother, Lazarus: “I know that he will rise again in the resurrection on the last day” (John 11:24). All three synoptic Gospels relate the Sadducee’s attempt to discredit belief in resurrection by asking Jesus who a woman, married sequentially to seven brothers, will be married to when they are all resurrected (Matthew 22:23, Mark 12:18, Luke 20:27).
NINE Resurrection
THE LIFE TO COME We turn, at last, to last things. Foremost among eschatological matters is the question of what becomes of human beings after death. Christians confess that God never forsakes us: though we die yet we will live again. Some versions of Christianity dispense with this hope, but it is motivated by the indispensable claim that God is committed to us unconditionally, always anticipating a shared future. Nothing, not even death, separates us from God’s love. This universe exists for the purpose of there being creatures called into an everlasting relationship of love and trust with the Creator. If death is the end, then God’s creative aims are thwarted. Many Christians, and many others, take for granted that life after death is possible only if dualism is true, and thus that scientific naturalism and Christianity cannot be reconciled. Today, despite the formidable challenges to dualism, many persons of faith cling to it, at least in some attenuated form, on the assumption that to give it up is to abandon hope for God’s ultimate victory over death. I argued in chapter 5 that materialism is not just consistent with Christianity, but that we have theological reasons to oppose dualism, for it portrays humans as transcending this material world and risks an implicit rejection of our status as creatures. But the case for the consistency of Christian faith with materialism, and thus with scientific naturalism, is incomplete until the issue of life after death is dealt with. This chapter offers an account of the Christian hope in everlasting life with God that is consistent with a materialist conception of human beings. The chapter also offers necessarily speculative responses to problems that ensue from the idea of everlasting life for material persons.
277
278
Chapter 9
DUALIST RESURRECTION A dualistic conception of the human person plays a familiar role in efforts to explain how the Christian teaching of life beyond death could be true. The body dies, but that is not the end of the person’s existence. The immaterial soul or mind, the essential part of the person, the immaterial substance that bears her mental properties, continues to exist. The person continues to exist, even if in an unnatural, truncated way. On most accounts, the disembodied person remains conscious and immediately finds herself in the presence of God, in heaven. After a period of disembodied existence God miraculously supplies a body for her, and she continues to exist forever as an embodied human being. To be miraculously re-embodied in this way is what it is to be resurrected. All the disembodied dead are resurrected together; it is the general resurrection. The body with which God supplies the temporarily disembodied soul is, on the model of the post-resurrection appearances of Jesus, recognizably similar to the body that died yet qualitatively changed so as to be immortal. On this account, death is not non-existence, but the loss of one’s body and one’s connection to the physical world. This understanding of the resurrection has been held for centuries by Christians and its popularity continues largely unabated. This account offers a clear explanation of how Christian claims about life after death could be true. For many centuries almost everyone, Christian or not, took dualism in one form or another as the obvious truth, so it is no surprise that Christians constructed an explanation of the possibility of life after death that relies on it. Shorn of dualism, the Christian belief in the life after death becomes problematic in ways it has not been since the earliest centuries of Christian faith. If we are nothing more than these bodies that will in due course be destroyed is it even possible that we will live after we die? Natural possibility is not at issue here. The resurrection Christians hope for, whether they are dualists or materialists, would not occur in the natural course of events. 1 Christian dualists, who believe that humans by nature possess immortal souls, and thus that post-mortem life is naturally possible, believe that souls will be re-embodied, and thus resurrected, only when God miraculously re-embodies them. In contrast, the materialist Christian, with the secular naturalist, knows that in the natural course of events death ends human existence. In contrast to the secular naturalist he trusts in God who can make things happen that would not have occurred in the natural course of events. He asserts that life after death is neither a natural possibility nor an absolute impossibility. God miraculously resurrects the dead. The dualist reasonably believes that God has the capability to re-embody the dead. The principal challenge for the dualist lies in the defense of dualism itself, not the possibility of resurrection. The Christian materi-
Resurrection
279
alist faces a different challenge. He must show that the resurrection is, in an absolute sense, possible, despite the fact that there is no immaterial soul and that human existence ceases at death. The secular materialist and the Christian dualist may well agree that this cannot be done and that Christian faith cannot be reconciled with materialism. MATERIALIST RESURRECTION The Christian who denies that human beings possess an immaterial mind or soul offers an account of how the resurrection is possible distinctly different than that of his fellow believers. A human person is a functioning human body, and this includes a brain that has various psychological properties. Among these are rationality and self-consciousness, characteristics in virtue of which he is a person. When he dies this body becomes inoperative, no longer capable of bearing mental properties. The materialist defines death as what occurs when his brain becomes permanently incapable of sustaining personhood. Perhaps, for a time, the body remains biologically alive, and probably, for a longer time, a corpse continues to exist, but the person stops existing when higher brain function ceases. Eventually this body decays and its microscopic components go their separate ways, most to become parts of other things. In the natural course of events, the person whose existence death terminates will never exist again. Nonetheless, in the future God will act, intervening in the natural order to bring about something that would never happen naturally. God will miraculously recreate the body, and in so doing recreate the person that died. Unlike the dualist, for whom resurrection is the reembodiment of an existing soul, the physicalist conceives of resurrection as akin to God’s original creation from nothing: God brings the dead to life from the nothingness of non-existence. The resurrected body will differ qualitatively from the body that died, but it will be essentially similar to, and recognizable as, the person who died. It will be the same body, numerically identical to the body that died, decomposed, and stopped existing. The Christian physicalist affirms the same creeds as the dualist—“We look for the resurrection of the dead, and the life of the world to come” (Nicene Creed) and “I believe in . . . the resurrection of the body, and the life everlasting” (Apostles’ Creed)—but he construes the import of this crucial confession differently. The materialist regards death as non-existence, while the dualist sees it as disembodied existence. Correlatively, the materialist describes the interim state as a gap in the person’s existence, while the dualist regards it as a gap in embodiment but continued existence. The dualist conceives of the resurrection as God’s miraculous re-embodiment of the dead, but the materialist conceives it as God’s miraculously recreation of those who no longer exist. Both confess that we
280
Chapter 9
have everlasting life, but the dualist believes, and the materialist denies, that everlasting life means being alive at all times after the death of the body. The dualist may speak of those who die as going to heaven, meaning that they exist in an interim state in the presence of God, but if the materialist uses this language, he refers to the new heavens and the new Earth, i.e., the redeemed future of this physical world, which restored humanity will inhabit after the general resurrection. 2 The materialist finds the dualist’s commitment to the natural immortality of human beings theologically problematic, seeing it as at odds with the view that we live after death only by grace, through faith in Jesus who God miraculously resurrected. He is suspicious of dualism as a pagan philosophical intrusion into the Christian faith. The dualist’s principal objection is not theological but metaphysical. He accepts that God the author of nature’s laws is not hindered by natural possibility and can do anything that is in an absolute sense possible, anything that is not logically or metaphysically impossible. But he worries that the materialist conception of the resurrection is an absolute impossibility. 3 THE EXISTENTIAL GAP AND REASSEMBLY This fundamental objection to the materialist view of the resurrection is that not even God, who can do anything possible, can cause a person who no longer exists to exist anew. Once a material object—which on the materialist account is all a human person is—stops existing, nothing that exists in the future could possibly be it, the same—numerically identical— thing existing in the future. The physicalist account of the resurrection involves an existential gap: the person exists, stops existing, but then exists again. But, says the objection, this categorically precludes the future person being the person who died. Whatever God could cause to exist after a material thing is destroyed can be at best a copy of the original, not numerically identical to it, no matter how qualitatively similar. This may seem plausible with respect to ordinary physical objects. You send your toaster to the shop for cleaning and servicing. On its way there it falls out of the delivery van into the street where it is run over by a truck and smashed to pieces. A storm washes the pieces into the sewer and out to sea. Some parts sink into the mud; others are swallowed by fish. They wind up miles apart. Later, you inquire at the shop about the whereabouts of your toaster. A resolute technician, averse to letting you find out what has befallen it, collects parts exactly similar to the parts that once comprised your toaster, assembles them into a toaster, one qualitatively almost exactly similar to the original, and gives it to you. You now believe that you have your old toaster back, but you do not. Your old toaster no longer exists. There are two toasters with different properties. One was manufactured at a factory near Shanghai in 2006 and was given
Resurrection
281
to your wife for her birthday; the other has existed for only a few days and has never been near either your wife or China. The toaster you now have, although it is almost exactly similar to, and for you indistinguishable from, the toaster you sent off, is just a copy. The toaster did not survive. The same goes, the objection continues, for anyone alive as a result of the so-called resurrection the materialist describes. Someone might look like, think like, and act just like the person killed in the twenty-first century, electrocuted while trying to retrieve a burning bagel from a toaster. He might sincerely believe that he is that person and seem to remember being that person, right up to the ill-fated attempt to recover the bagel, but he can’t possibly be him. The original person did not make it; he stopped existing for good back in the twenty-first century, when the surge of electricity from the toaster stopped his heart, cut off the flow of oxygenated blood to his brain, and rendered it inoperative. A human person can survive various vicissitudes, but non-existence is not one of them. He might find comfort in the fact that a perfect simulacrum of him could miraculously exist in the future, but he cannot reasonably anticipate existing in that future. He cannot reasonably hope to be resurrected. He can reasonably believe that only if he believes that he has an immaterial soul that will continue to exist after the body is gone. This kind of example is meant to show that continuous existence is necessary for personal identity through time. A person existing at a later time cannot possibly be the same person as someone who existed at an earlier time unless she exists at all the intervening times. Physical objects cannot go in and out of existence so, if a human being is no more than a physical object, a body, there can be no gaps in her existence. However, this is a mistake, because reassembly can make it possible to exist on both sides of an existential gap. If the later thing is made of the same parts, arranged in exactly the same way as in the original, then it is the same thing, no mere qualitatively similar copy, but numerically identical to the object that was destroyed. The toaster is sent in for repairs but this time it arrives safely at the shop for servicing where it is completely disassembled into an array of small parts. For some time, the parts are in various locations, remote from one another. During this period the toaster does not exist. 4 The parts, tested and cleaned, are reunited and reassembled and your toaster is returned to you. It would be unreasonable to insist that it is just a copy of the toaster you wanted serviced and to demand the original toaster. You have the original toaster. A complex material object can have gaps in its existence, at least if it has the same parts, configured the same way, before and after the gap. These conditions can be relaxed. So long as the object has mostly, even if not all, the same parts and is reassembled in essentially, even if not exactly, the same way, it is the same thing.
282
Chapter 9
If a human being is, as materialists contend, a complex physical object, then she could survive an existential gap by the same means. Some early Christians embraced this possibility, among them dualists who believed that the resurrected person must have the body he had in this life, and that in cases in which that body is destroyed, this is possible only by reassembly of its microscopic components. St. Augustine assured Christians: Far be it from us to fear that the omnipotence of the Creator cannot, for the resuscitation and reanimation of our bodies, recall all the portions which have been consumed by beasts or fire, or have been dissolved into dust or ashes, or have decomposed into water or evaporated into air. 5
God, being omnipotent and omniscient, can keep track of all the parts of a particular body and see to it that, at the resurrection, they are miraculously reunited and reassembled so as to make up the very same body. God can do this directly, or by endowing the soul, which desires its body so as to again be a complete person, with the power to find and reunite the constituent parts of its body. The reassembly scenario shows that the resurrection of a material person is, in an absolute sense, possible: God could do it. It gives rise to interesting complications. Microscopic parts of one body can be parts of other bodies at different times. This prompts what Augustine describes as “that question which seems the most difficult of all—To whom, in the resurrection, will belong the flesh of a dead man which has become the flesh of a living man?” 6 Augustine’s concern was that if microscopic parts belong to both the cannibal and his dinner, it will be impossible to resurrect both. This problem was especially pressing for Augustine, who with other early Christians, believed that the resurrected body must consist of every bit of matter present prior to death. 7 Even if, as we may reasonably suppose, successful reassembly of a complex object does not require retaining every part, the problem persists so long as we assume that the earlier and later body must have some proportion of common parts. If a body at one time and a body at another time can be the same body only if, say, they have at least half of their parts in common, it is conceivable that 51 percent of the atoms that comprise a cannibal’s body once comprised 51 percent of the body of someone he ate, in which case God cannot resurrect both. The solution, of course, is that God can see to it that no bit of matter needed to resurrect one individual winds up constituting another. The Creator who has the wherewithal to keep track of every elementary component of innumerable dead bodies dispersed in nature can just as easily ensure that whatever human flesh is eaten or otherwise assimilated by another is not “assimilated to the substance of the eater,” but evacuated, or evaporated into the air. 8 God can, either by miraculous intervention, or by designing the laws of nature from the
Resurrection
283
beginning to prevent it, ensure that no, or not too many, elementary parts of one body end up in another. THE RESURRECTION OF THE BODY As bizarre as this story is, it defeats the claim that it is impossible for a human body to exist both before and after non-existence. God need only see to it that enough of the original parts are restored to it. However, my aim here in The Material Image is not to establish the bare logical coherence of the Christian faith, including its assertions about resurrection, with scientific naturalism, but to make the case for an intellectually satisfying fit. As a Christian materialist I hope to show that the resurrection is possible without resorting to scenarios that we know are at best highly implausible, if not false. 9 If we can do better than showing that resurrection is possible by invoking a possible, but most likely non-actual, scenario, then we should do so. We can agree with Augustine that God could prevent atoms or molecules that comprise one human body from being incorporated into another, but we have no reason to doubt that part of the cannibal’s victim does become part of the cannibal. We know, then, that it is possible for a material object to exist, stop existing, and then exist again: continuity of existence is not necessary for identity. However, we also know that if your toaster goes to the shop where it is disassembled and its parts dispersed, and exactly similar, but numerically distinct, parts are assembled, so that a toaster very similar to the one you sent in comes into existence, what you get back is not the original toaster. Similarly, over a lengthy period, a toaster’s parts are one by one removed, discarded, and replaced, so that eventually, what exists is a very similar toaster with all new parts, but it is not the same toaster. Having mostly original parts is necessary for being the same toaster. In general, material objects can survive gaps in their existence, but they cannot survive a total replacement of parts. However, we also know that when it comes to identity through time, there is something special about biological organisms. All, or almost all, the microscopic constituents of a human body at any given time sooner or later disperse far and wide into the natural environment, where they become parts of other things, some of them living things that humans eat. We need not worry about cannibals to realize that, in all likelihood, any human body at any moment contains many atoms that were once parts of other human bodies. The scenario of God at the last day summoning some subset of them to reconstruct one’s body seems gratuitous in light of the fact that having any of the same parts is not, after all, necessary for bodily identity. A human being has the same—qualitatively different yet numerically identical—body at age eight and at age eighty, even if the child’s body and the old person’s body have not even a single atom in
284
Chapter 9
common. 10 We should hope to explain how God can resurrect her without the need to reassemble particles that at one time or another were parts of her, but also parts of many other things. Neither continuous existence, nor sameness of elementary parts, is necessary for a human body that exists at one time to be numerically identical to one that exists at another time. This is good news for the Christian materialist, whose account of the resurrection assumes that neither of these conditions is satisfied. However, even if neither continuous existence nor sameness of parts is a necessary condition for personal identity, it might be that identity requires that at least one of them be satisfied. Suppose that Ray exists continuously from 1990 to 2020. Because of the natural processes of cellular replacement, his body in 2020 shares no microscopic components with his body of thirty years earlier. Despite this, the body that exists in 2020 is the body that existed in 1990. Sameness of parts is not necessary for the identity through time of a human body, in contrast to most things, such as toasters, for which it is necessary. However, in 2030 Ray engages in a practice familiar from science fiction. He is teleported from Earth to Mars. He enters a matter transporter in one location, disappears, and soon thereafter reappears at a distant location. How might this technology work? One possibility is that he is scanned, disassembled and his elementary particles are shot through space to Mars. Instructions for reassembly, based on the scan, are also transmitted to the target location, where they direct the reassembly of the transported particles. The transmission is not instantaneous, so there is a temporal gap in Ray’s existence. But the same particles, arranged in the same way, constitute his body immediately before, and immediately after, teleportation, and there is no question that the man who arrives on Mars is the man who entered the transporter on Earth. Teleportation might work differently and, if it did, it would be a mundane analogue of resurrection as the materialist conceives it. Ray’s body is scanned and the configuration of the particles that comprise it recorded, but all that is transmitted to Mars are instructions for assembling new particles. On Earth, Ray’s body is destroyed and discarded, and what appears at the target location is an exactly similar body, made of materials available there. Now there is neither sameness of parts nor continuous existence. Does Ray still exist, or is the Martian who calls himself Ray an imposter, a mere copy of the original? If we rely on the judgment that we make about ordinary physical objects like toasters, then we must acknowledge that this kind of teleportation is not a way to travel but to die. The implication, ominous for the materialist account of the resurrection, which dispenses with both bodily continuity and sameness of parts, is that while neither identity of parts, nor continuous existence, is individually necessary for the identity of a human person through time, what is necessary is that there be one or the
Resurrection
285
other. Lacking the original parts, the latter body, discontinuous with the one that was disassembled, is just a copy, not identical to it. Ray never makes it to Mars, and the dead are not resurrected. The best anyone, even God, can do is make some qualitatively exactly similar, yet numerically non-identical, person, a copy of the person who died, not the same person. If the materialist accepts this, he might have no alternative but to retreat to the reassembly theory, settling for this ancient view as the best we can do, contending that it is at least not as implausible as dualism. Rather than resort to this we should more closely consider the judgment that in the second teleportation story the person who steps out of the transporter on Mars is not Ray. Instructions are transmitted for making a Martian copy of the person that was scanned and destroyed on Earth. The reassembly of something preserves its identity but here there is no reassembly, only copying. And, we assume, a copy of a thing is not identical to it. With this assumption in the foreground, it is easy to judge that identity is not preserved. There can be, at most, something qualitatively exactly similar to an original, but not numerically identical to it, not the thing itself. This is plainly true of toasters and other inanimate objects. An existential gap is survivable, but not unless the original parts are reunited. For such things, a copy is not the original. However, in the background lie facts about biological life that complicate matters of identity. We noted earlier that sameness of parts is not necessary for the identity of a biological organism through time. An organism, such as a human body, might exist at two different times, with no microscopic components in common. Your current body need not contain a single atom with which you were born. Organisms exist from one time to another not by retaining all their microscopic parts, but by systematically discarding and replacing them. This process is a kind of copying. An organism is an assemblage of smaller organisms, its constituent cells, and the overall organism persists through time because its cells make copies of themselves, by mitosis and cytokinesis. In a sense, a biological organism that exists at a time is a copy of an earlier biological organism, a copy of an earlier version of itself. For living things, the fact that Y is a copy of X does not imply that Y is not identical to X. Biological identity through time does not occur despite copying, but because of it. Were cellular copying to stop, the organism would soon stop existing. It would be a serious mistake, one indicating something amiss in one’s concept of biological entities, to claim that a forty year old human being does not have the same body he had forty years ago, on the grounds that his body contains no forty year old cells, but only copies of copies of copies . . . of the cells of four decades earlier. He might have a forty-yearold toaster, but only because its original parts remain together, but he has a forty-year-old body thanks to the patterned self-replication and selfdestruction of its constituent cells. 11
286
Chapter 9
With this in mind, we can label copies that are not identical to the things they are copies of mere copies, but recognize that not all copies are mere copies, and in particular, that while a biological organism is a copy of a biological organism, it is not be a mere copy of it, but a later version if it and thus numerically identical to it. When a biological organism dies and is destroyed, and God later miraculously creates something exactly similar to it, not by reassembling its parts but out of all new materials, the fact that it is a copy of the organism that died does not imply that it is not the same organism. Being a copy would be nothing new for it: when it died, it was a copy of an earlier organism to which it was identical, and in fact to which it was less similar in many important ways, than it is to the resurrected organism. 12 The crucial question is whether miraculous divine action of the sort envisaged in resurrection could be copying of the kind that preserves biological identity, or if even God can create only a mere copy of the person who died. There is no compelling reason to believe that an organism miraculously created by God as a copy of a deceased organism must be a mere copy of it and thus not the same—numerically identical—organism. To see this, let us begin with natural copying, move to a copying process that is artificial rather than natural, and conclude with a process which is supernatural. Organism A consists at a time of some number of cells; each either dies without dividing, or divides and later dies. Generation by generation, each daughter cell either dies without dividing, or divides. For the sake of simplicity assume the organism does not grow; the overall number of cells remains constant. This natural process continues until a later time, when descendants of descendants of the cells that comprised organism A now comprise organism B. B contains no cells that were parts of A. B exists in virtue of a natural copying process and is a copy of organism A. But it is not a mere copy. Despite being a copy of something with which it shares no parts, it is the same organism. A and B are numerically identical. Next consider a possibility intermediate between the natural course of events and divine intervention: human intervention. There is a defect in the copying mechanism in the cells that comprise organism A. Left to its own devices, mitosis fails, there are no new cells, and the organism does not survive. The natural physical processes of cellular replication, ineffective in this individual, would be sufficient for the survival of the organism, but they are not necessary. The artificial can replace the natural. A medical intervention can ensure that something functionally equivalent to natural mitosis occurs, so that nuclei divide, cellular replication occurs, and later organism B exists. The intervention is accomplished by skilled human microbiologists who introduce a chemical that enables the cells to replicate properly. Or they introduce self-replicating nanobots programmed to manipulate the molecules in the cells to produce the desired
Resurrection
287
result. A copying process that begins with organism A eventuates with organism B, which is the same organism, a copy, yet no mere copy. Biological life naturally persists through time by way of immanent causation: B has the characteristics it has in virtue of the characteristics A had, which, in the natural course of events, explains why B exists and has the characteristics it has. No external agency enters into the causal process. The organism reaches the future by its own power. However, this kind of causal history is not necessary for the organism to exist from one time to another. What is necessary is a reliable causal connection of the organism’s earlier to its later characteristics, even if it is not the natural connection. There is, we might say, a sequence of causes and effects that transmits information about the configuration of A that causes the configuration of organism B. The necessary causal connection can involve external agencies, including the actions of persons who intervene with the aim of continuing the organism’s existence. In place of the normal sequence of causes, there can be a “detour” through the mind of an intentional agent. The organism at an earlier time has a particular characteristic, this causes the biologist who wants to keep the organism alive to have a reason to perform a particular action, and this in turn causes her to perform that action, with the result that organism B has that characteristic too. Thanks to the intervention of an intelligent agent, an organism could exist through time by artificial copying. The individual who exits the transporter on Mars really is Ray, in virtue of being a copy of him, created by reliable human technology. Whatever human science could, in principle, accomplish by technological means, God could bring about directly. And while any humanly possible process of copying at best reliably copies the original, God can achieve the result infallibly. By one means or another, the biological continuity that suffices for the identity of the organism through time, is sustained. So far, we have assumed a continuous copying process, whether natural, artificial, or supernatural, but now we introduce an existential gap. In the natural course of events, an organism stops existing at death and does not exist again, but we can imagine another science fiction story in which it exists again by artificial means. Organism A suffers from a defect that can be remedied only if it is disassembled and its constituent cells treated in isolation from one another. The treatment takes considerable time, and the various components must be worked on in widely separated labs. The exotic process even involves some cells being inserted temporarily into other organisms. For much of this extended period, the cells are frozen and inert. No replication occurs. Eventually, the far-flung cells are collected and artificially induced to replicate. They are discarded and after many generations the descendant cells are reassembled into an organism essentially similar to A. None of its cells were parts of A. This organism, B, has no cells, and—we will again suppose—no more elementary parts in common with A. While all this was going on, the organism
288
Chapter 9
did not exist, but now it exists again; organism B is numerically identical to organism A. It is a copy, but not a mere copy. The artificial copying preserves identity, and it does so despite a gap in the organism’s existence. Once again, there is no reason to doubt that what humans might bring about by means that reliably causally connect A to B God can do so miraculously. Mary dies at an advanced age and ceases to exist; eventually, her body also ceases to exist. Later, the very same human body, with all new microscopic components, exists in virtue of divine intervention that makes the right sort of copy in the right way. The body of young Mary and the body of resurrected Mary are the same body, thanks to the reliable causal connection between them. THE RESURRECTION OF PERSONS A human being’s body can go out of existence and miraculously exist again, even with all new matter. However, this does not guarantee that the same person exists again. The requirements for the identity of a person through time, and across an existential gap, differ from what bodily identity requires. The brain of the resurrected body could be configured in such a way as to have a drastically different mind, or no mind at all, with the result that the person who died no longer exists, although the body she once had does. But the conditions necessary for sameness of body and sameness of person can be satisfied together, naturally, artificially, or supernaturally. On the materialist assumption that the mental properties of a human person supervene on the neural state of her brain, by recreating the brain God can at the same time ensure the psychological continuity required for personal identity. God need only see to it that that the later body is in the appropriate neural states. Suppose, for example, that early in life, Sarah visited the zoo where she encountered Bonzo the bonobo, who bit her when she poked her finger between the bars of his cage. This vivid childhood experience produced a long-lived memory trace in her brain. At various times later in her life, it seemed to her that she remembered this encounter and these experiences were actual memories; they were reliably causally connected to the event. God knows that Sarah remembered this event, wants her to exist again, even though she died, and has a reason to preserve psychological continuity, which includes her memory. God recreates the body that dies, and sees to it that it is equipped with a brain that reliably produces the experience of seeming to remember being bit by Bonzo. This experience of seeming to remember the event is reliably causally connected to the event, and thus resurrected Sarah really does remember being bitten by Bonzo. Similar accounts can be given of other aspects of psychological continuity. Much of the psychology of the
Resurrection
289
resurrected person is an effect of the psychological characteristics she acquired in the course of her life. The causal connections that make bodily, and with it personal, identity possible are a hybrid of natural reliable, and then supernatural infallible, causal sequences. The detour into the mind of God does not in the least make the causal connection less reliable. 13 Post-resurrection Sarah remembers being young Sarah at the zoo, and this implies that this is the same person, existing at different times. God has preserved psychological continuity by means of creating a body that is a copy, but no mere copy, of the human person. God has resurrected a human person in the fullness of her humanity, as Jesus was resurrected. God’s grand intention to share the everlasting Triune life with created persons is not defeated. Death does not have the last word. The fictional example considered above, in which a human being survives the existential gap caused by being teleported to Mars, more closely resembles resurrection, in that whoever operates the transporter need not do anything while the information makes its way to the target location. The transmission, the period between the body being disassembled and the copy made, could be delayed for any amount of time. What is necessary is the continued existence of something that preserves the information needed to recreate the person. In the case of supernatural resurrection, the same is true: the crucial information needed to recreate the person exists continuously in the mind of God. THE TRANSITIVITY PARADOX Have we shown that the Christian dualist’s belief that the materialist account of the resurrection involves an absolute impossibility is unfounded? Christian hope in the coming resurrection, and assurance of its possibility, resides not in metaphysical convictions about human immortality, but in the resurrection of Jesus Christ as a historical event. The biblical accounts of his post-resurrection appearances make it hard to deny that his resurrected body was the same—numerically identical— body as the body that was crucified, buried, and resurrected. This is not to say that Christ’s resurrected body was qualitatively exactly similar to the body that died. The resurrected body was no longer mortal, no longer subject to the power of sin and death. It was what St. Paul called a “spiritual body.” The accounts of Christ’s post-resurrection appearances also indicate that in some hard to grasp way he was free with respect to space and time. Yet the resurrected body bore the wounds of the crucifixion. If it was not the crucified body, but a replacement, the wounds were simulacra, not made by Roman nails and a Roman spear, but copies. Those who deny that it was the original body face embarrassing questions: was the tomb not empty, because it contained an abandoned corpse? Was the crucified body disposed of in some unknown way? For these reasons,
290
Chapter 9
Christians, whether dualists or materialists, have traditionally held that the resurrected will have the same bodies they had from birth to death. In light of this, perhaps we can conclude that Christian dualists who adhere to this traditional view can have no objection to the materialist conception of the resurrection as such. Their objection to materialist resurrection must revert to the basic issue of whether there is more to human persons than their bodies. If, as materialism asserts, the human person just is the functioning body, and if the bodies we have now will be resurrected, not replaced, then human persons, though mere material things, will be resurrected. A dualist might be tempted to describe the resurrection of Jesus as a special case, because when he was resurrected his body, dead only for a short time, still existed. However, many of the bodies of the dead have long since gone out of existence, yet they will be recreated at the general resurrection. The Christian dualist, unless she denies the identity of pre-death and post-resurrection bodies, agrees with the materialist that human bodies can die, go out of existence, yet exist again. She simply denies that this means that the human person dies, stops existing, yet will exist again. She shares with the materialist the need for an account of how human bodies can exist, not exist, and exist again. Christian dualists should acknowledge that if what Christian materialists say about human nature is true, then what we say about the resurrection could be true as well. However, to show that this is true, the materialist must overcome the paradox of transitivity. It appears that God can resurrect human beings, with no need for continuous physical existence or the same physical parts, let alone an immaterial mind or soul. However, it is also true that what God can do once God can do more than once. William dies and does not exist for many years, but in the eschaton, God miraculously resurrects him by bringing into existence a material object, a human body, which stands in the appropriate causal relation to him, one that suffices for bodily and personal identity. Surely God could create two such bodies rather than one. 14 William dies, and later, long after his body no longer exists God creates two persons. Let’s call one Bill and one Billy. Both have what, according to the materialist account of resurrection, it takes to be William. Identity is transitive—if William is the same person as Bill and William is the same person as Billy, then Bill is the same person as Billy—so the man who died would be numerically identical to two resurrected persons. Bill and Billy are the same person, not just exactly similar, but numerically identical. This is absurd: at any given time Bill and Billy must be in different places, where they will have different experiences, acquire different beliefs and desires, make different choices, and so on, diverging, physically and psychologically, as time goes by. Billy, say, eats pizza but Bill never does, which implies that it is both true and false of William that he will eat pizza after he is resurrected. Therefore, while there could be two future persons initially exactly similar to William, both of whom
Resurrection
291
believe that they are William, neither can actually be him. After death, William is not recreated; he is replaced by two mere copies, imposters. A dualist might tolerate this untoward result, on the view that there is still just one immaterial soul that God has supplied with two different bodies. The resurrected person is the whole consisting of an immaterial soul and two bodies, rather than the standard issue one body. The situation is very strange, but not impossible. If an immaterial soul can interact with one body then, for all we know, it can interact with more than one. A person is in two different places at once, but only in the sense that his immaterial mind interacts with bodies that are in two places. It would be true that this person ate pizza with his Billy body but not with his Bill body, a situation no more disconcerting than someone eating pizza with his left, but not with his right, hand. But the materialist cannot so easily avoid contradiction. A body, which is all a human being is, cannot be in two places at once. Thus, the physicalist account of resurrection runs afoul of the paradox of transitivity. If the causal relations that the materialist has in view were really sufficient for identity, then Bill and Billy would be the same person; they cannot be, so the materialist is mistaken. A common but inadequate response is to point out that even though God could do this, there is no reason to believe that God would, and good reason to believe that God would not. 15 God loves William and wants him to exist even though he has died, so his Creator recreates him. Sometimes, when we value something, we regard two as better than one, but what is true of, say, $20 bills or cats, is not true of all things and it is clearly not true of human individuals. God would not manifest greater love for William by creating two copies of him, even if both somehow could be William. There is no reason to think that God would bring about a duplex resurrection, but this does not reach the heart of the problem. The mere possibility seems to lead to absurdity. There can be multiple copies of something, but at most one can be identical to it. The rest must be mere copies, not identical to the original embodied person. William dies and no longer exists. At the resurrection God creates something that stands in the relation to William that normally would suffice for identity. The newly created person believes that he is William and he seems to remember being William. He is happily living his eschatological life, now calling himself Billy, thinking how blessed he is to have been resurrected, when he walks around a corner and comes face to face with the other person, Bill, who was created by God at the same time he was created, and who stands in an exactly similar relation to William. Lest we say he has just run into himself, we must acknowledge that he discovers that he is not, after all, William or any other resurrected person. He realizes that William remains non-existent and that two mere copies now stand in for him. Meeting Bill, Billy reasons that had God not created that other copy, then he would have been not a mere copy of William, but William resur-
292
Chapter 9
rected. Bill is entitled to the same reasoning. This leads to the conclusion that whether a person that God creates at the resurrection is identical to someone who died depends not just on the facts about him and his relation to the deceased person, but on the non-existence of competitors, on God refraining from making more than one copy. But this conclusion seems absurd. Whether I am identical to some person who will exist after I die depends entirely on the facts about that person and his relation to me, not on facts about anyone else. 16 My survival is a matter of what happens to me, not to someone else. The transitivity problem arises because the relation that makes for identity through time—psychological continuity for persons and biological continuity for their bodies—is not guaranteed to be unique. God could create two bodies, both reliably causally connected to the person who died in a way that suffices for personal identity and identity of body. Neither psychological continuity nor biological copying guarantees uniqueness. The same appears to be true of any reasonable conception of what it is to be the same person, or the same body, at different times. Once we acknowledge that identity through time is a matter of some kind of causal relation connecting past and future versions of the human being, there is no guarantee of uniqueness. No matter how the materialist envisions God doing it, he cannot rule out the possibility of God doing it more than once, and thus cannot avoid the conclusion that whether the person God creates at the resurrection is William depends on other persons not existing. In contrast, if the dead person and the resurrected person are identical in virtue of a non-physical soul or mind that exists continuously from death to resurrection, the transitivity issue is moot. 17 God simply permits the soul to exist during the interim state, and then, at the resurrection, provides a body for it. There is no re-creation of persons, only re-embodiment, so the issue of how many God creates does not arise. Thus, in light of the paradox of transitivity, the dualist may contend that whatever the evidence against the immaterial soul, it must exist if the Christian claims about resurrection can be true. 18 SUPERFICIALITY At this point, the Christian dualist and the secular materialist reach common ground: the Christian faith cannot give up dualism and thus cannot be reconciled with what science clearly tells us about the kind of beings we are. But the Christian materialist rejects the conviction they share. He points out, first, that while the account of the miraculous resurrection of material persons makes the transitivity paradox vivid, it challenges any materialist conception of the identity of persons through time. The physicalist cannot exclude the possibility of a human being undergoing fission
Resurrection
293
with no divine intervention. An example devised by Derek Parfit illustrates this. 19 A human being’s body is threatened by some fatal condition, but his brain is healthy. He is cloned, and the cloned cell is grown into an adult human body, but gene expression is suppressed, so no brain develops. The person’s brain is now removed from his body and installed in his cloned body. The patient survives the procedure by means of his brain being transplanted into a new, but genetically exactly similar, body. There seems no good reason to doubt that the story involves just one person, not two. Next, we modify the story, relying on the fact that a human being can survive the loss of half his brain. The condition that affected his original body now also infects the left hemisphere of his brain. To save him, the healthy right brain hemisphere is transplanted into the cloned body. Again, there seems no reason to doubt that the patient survives this procedure, drastic as it is. The third case relies on the possible, even if unlikely, supposition that in some individuals the two brain hemispheres have equal capabilities. Two clones are made from the original body. The left hemisphere is transplanted into one, the right hemisphere into the other. Parfit names the resulting persons Lefty and Righty. It appears that a human being has undergone fission. But has the patient survived? Transplanting his whole brain would suffice for personal identity, as would transplanting either hemisphere alone. When the hemispheres are separated and transplanted into different bodies, the transitivity paradox emerges. We have good reasons to declare Lefty the survivor, and equally good reasons to declare Righty the survivor, but if we say that both survive, we arrive at the conclusion that Lefty and Righty are the same person. Parfit’s example of partial brain transplantation is just one of various exotic scenarios in which person fission occurs. The resurrection of the body, as the materialist Christian conceives it, is the supernatural analogue, where it is possible for God to bring about the fission of a human person and thus present us with the same paradox of identity. Parfit draws profound conclusions from the possibility of fission. There are, he points out, four possibilities for the patient to consider as he deliberates about the procedure: 1. 2. 3. 4.
He does not survive He survives as Lefty He survives as Righty He survives as both Lefty and Righty
Each possibility prompts significant objections. (1) seems unreasonable because it counts a double success as a failure. Neither (2) nor (3) are attractive since any reason to conclude that he survives as one is just as good a reason to conclude that he survives as the other. He could embrace (4), but this demands a radical revision of our concept of what a
294
Chapter 9
person is. The proper course, Parfit contends, is to recognize that in such a case we know all the relevant facts and that no further fact about personal identity remains hidden. The problem is not how to discover the truth about personal identity in this situation, but how to describe what we already know. How should we deploy the concept of being the same person in this unprecedented situation? The facts about personal identity are in this instance indeterminate. There is no objective fact of the matter whether, say, Righty is, or is not, the person whose brain was removed and bisected. Nor is there an objective fact of the matter as to whether fission is survivable. Although there is no fact about personal identity to discover, there is a practical issue. We need to choose the best description of a situation that can be described in various ways. Parfit believes that the best way to describe a case of fission is (1): the patient does not survive; Righty and Lefty are new persons. Fission is fatal, but Parfit mitigates this by demoting identity. We cannot survive fission, but what we reasonably care about—what ought to matter to us—is preserved when fission occurs. What the patient cares about when he thinks that he cares about survival is not survival—being identical to someone in the future— but the psychological continuity that does exist between him and Lefty, and between him and Righty. The transitivity problem is solved since, in contrast to identity, psychological continuity is not a transitive relation— Righty and Lefty can be psychologically continuous with the same person without being psychologically continuous with one another. Leaving aside what is for Derek Parfit the crucial issue about what reasonably matters to human beings, I want to focus on the idea that fission is a possibility in which there is no objective fact of the matter as to whether a particular person still exists. Our concept of what it is to be a person works well enough in normal circumstances, but it breaks down when we apply it to the possibility of fission, whether the result of advanced human technology or divine intervention. 20 God has, and human beings might someday acquire, the ability to create situations in which there is no fact of the matter as to whether future person B is the same person as currently existing person A. Even our best account of personal identity in some conceivable circumstances provides no determinate answer, not because it is there and we cannot know it, but because there is nothing further to know. The transitivity paradox helps us see that there are possible circumstances where the concept same person is subject to indeterminacy. If God does no more than create B, who stands in an identity preserving relation to person A, then A is the same person as B. But if God creates B and C, both standing in that relation to person A, it is indeterminate whether A is B. It is not false that A is identical, but neither is it true that A is identical to B. This seems absurd, but only in virtue of the implicit assumption that a human person is the kind of thing the existence of which is necessarily determinate. If there need not always be a definite yes or no answer to
Resurrection
295
such questions as, “Does person A still exist?” or “Is person A the same person as B?” then it is not problematic to discover possible circumstances, such as the existence of a competitor, where the identity of persons becomes indeterminate. We have plenty of perfectly good concepts that are vague in the sense that on occasion it is neither true nor false that they apply. While there are, e.g., instances in which we reasonably say that someone is rich, and instances in which we say that someone is not rich, there are also instances where there is no matter of fact whether a given individual is rich. We might, for various practical reasons, draw a sharp line, declaring those who have a certain amount of money rich and those with one cent less not rich, but we do not imagine that this cuts reality at its joints. It is obvious that taking one cent from a rich person does not really result in her no longer being rich. We reasonably understand this, and many similar examples, as a matter not of reality being vague or indeterminate, but of the looseness of fit of our concepts and words to reality. “Vagueness,” David Lewis said, “is semantic indecision.” Lewis, an American who was a fan of things Australian, writes: The only intelligible account of vagueness locates it in our thought and language. The reason it’s vague where the outback begins is not that there’s this thing, the outback, with imprecise borders: rather, there are many things, with different borders, and nobody has been fool enough to try to enforce a choice of one of them as the official referent of the word “outback.” 21
On this widely-held view, our concepts are vague, but only because we lack agreed upon rules for applying them in every possible situation to a reality that in itself is not vague. Indeed, many regard the idea that existence or identity might be vague as unintelligible or incoherent. 22 While it makes sense to think the statement, “Marvin is rich,” might be neither true nor false, they deny that sometimes there is simply no fact of the matter whether Marvin exists, or whether Marvin is identical to some future person, for instance someone that God will recreate at the general resurrection. That, they fear, is not vague thought and talk, but an inadmissible vagueness in reality; an intolerable vagueness in our reality. However, there are contexts in which existence and identity is at issue, yet indeterminacy is unremarkable. Generally, we have ways to ascertain whether, e.g., a toaster that exists at one time is the same toaster that exists at another time. There are possible odd circumstances in which this is not the case: on my way home with my new toaster in 1970, a mishap necessitates the replacement of one minor part. As the years go by, I gradually replace various small parts. I keep careful records, so when in 1990 I replace a minor component, I note the transition from mostly original, to mostly new, parts. And when in 2000 I replace one more part, I see that I now own a toaster with no original parts. In 2000, did I have the
296
Chapter 9
same toaster I had in 1970; did I have a 30–year old toaster? That is implausible, but if the 1970 toaster is not numerically identical to the 2000 toaster, when did the original toaster cease to exist? Surely it was not destroyed by the mishap on the day I bought it: a toaster can survive the replacement of one minor part. Nor in 1990 did the transition from, say, 51 percent original parts to 49 percent original parts destroy the old toaster and bring a new one into existence. Nor, when the last original component was replaced in 2000 did the original toaster cease to exist to be replaced by a new one. No mysterious further fact about the identity of the toaster lurks here. We have all the relevant information, but it does not tell us what to say about identity. The lawyers who wrote the warranty might have come up with a criterion for the identity of the toaster, one that serves the manufacturer’s needs, but they did not solve the metaphysical puzzle of its identity. Objective reality, independent of human interests, holds no solution. We can accept this as a possible case in which there is no fact of the matter about the identity of toasters without doubting that in ordinary circumstances we make determinately true or false judgments about their identity, just as, despite any number of indeterminate cases we know that some of us are definitely rich and others not rich. Acknowledging this does not lead to any unsettling, perhaps logically incoherent, conclusion about reality itself being indeterminate. Reality’s determinateness lies with its fundamental microscopic components. We have concepts that usefully refer to complex material objects, things made by assembling nature’s fundamental components, but there need not always be a matter of fact as to whether what is there counts as such an object and, if it does, whether it counts as identical to another one. At face value, humans fare no better than toasters on this issue. Any attempt to shield human beings from the risk of ontological indeterminacy conflicts with evident facts about us. A human ovum is fertilized and processes of cellular replication lead to an embryo, a fetus, and in due course an infant. Initially, this living, human entity lacks the capacity for conscious rationality and thus is not a person, although it is a potential person. No actual person is composed of just a few cells. Eventually, a human person exists, a human body possessing, at least dispositionally, the capacity for rational self-consciousness. But there is no moment when the fetus becomes a person, no sharp line that separates what is not yet from what already is a human person. We come into existence gradually. For someone now alive born in December 1980, it is a plain fact that she did not yet exist in December 1970, and that she did exist in December 1990 (assuming no existential gaps), but perhaps it is indeterminate whether she existed in, say, June 1980. One might specify, with whatever precision becomes available, criteria for human personhood, and on this basis narrow the period when we do not know what to say about the person’s existence, but indeterminacy cannot be completely banished. At
Resurrection
297
the far end of life, for some, death comes suddenly, and something approximating a distinct border between existence and non-existence can be identified. In other cases, human beings depart existence as gradually as they began it, the brain losing its capacities piecemeal. That the existence of human persons can be vague in this way follows inevitably from the fact that we exist at any moment only as a result of the activity of a great many interacting small components whose functionality can emerge, or degrade, gradually over time. Moving on from actual human life to mere possibilities, contemplating technologies the future might hold, such as cloning, nanotechnology, artificial replacement of brain circuitry, human-machine hybrids, and so forth, there are conceivable situations in which there is no matter of fact whether a person still exists or, if she does, whether she remains a human person. Some of these possibilities, e.g., teleportation mishaps, resemble the imagined situation in which God creates two or more copies of a dead human being, where any copy on its own would be the dead person resurrected, but there is no objective fact about who these persons are, or whether the original person lives again. Due either to conceivable human machinations or to possible acts of God, circumstances could come about in which human identity is indeterminate. This is a manifestation of the ontological superficiality of human beings we have seen in other contexts. The physical world contains large things made of smaller things, and smaller things are made of still smaller things, and so on. We are among the world’s middle-sized objects. Complex macroscopic objects, whether assembled naturally like human bodies or artfully like toasters cannot always escape vulnerability to what David Lewis described as semantic indecision. The basic, simple constituents of the world can be arranged so that statements like, “That’s a toaster” and “That’s the toaster that electrocuted Sally,” are true or false, but there are also ways they can be arranged such that our ordinary means of sorting and counting go awry, and there is no determinate fact as to whether such statements are true or false. One well-known possibility: after years of systematically, gradually replacing old toaster components with new, we learn that someone else has been taking and refurbishing the old parts, and has just reassembled them as they were originally assembled. Continuity of existence and sameness of parts, usually found together, in this case diverge, leaving us not knowing what to say, probably because there is no fact of the matter about the identity of the toasters in this circumstance. The indeterminacy lies on the surface of reality; it does not penetrate to the depths, to the simple constituents. Human beings, like toasters, exist in the shallows where there is not always an objective fact about existence and identity. Perhaps there are physical things that are not made of other things. If so, they are the basic constituents of the world, the things that, as it were, come directly from the hand of the Creator. All other things exist only when these basic constituents are arranged in certain ways. One of dual-
298
Chapter 9
ism’s attractions is that it locates the human person, at least her essential, immaterial part, as a basic constituent of creation, along with the most elementary physical constituents. The human soul, being non-physical, has no parts and, like the simplest physical things, is brought into being by God directly. Unlike the body, it is not assembled by natural processes. In the beginning, God acted directly to create the elementary particles, strings, branes, or whatever, and he acts in a similarly direct way to create human souls. By this route, the unsettling vagueness of human existence and identity is banished. God can create a human soul and affix it at a precise moment to an embryonic brain, bringing a human being into existence and rendering the statement that she exists determinately true. There can be a parallel moment when the immortal soul leaves its body. No matter what weird things befall the bodies with which souls interact, there is always a determinate matter of fact about who exists and about who is who, for human beings are fundamental parts of God’s creation, not mere assemblages of other things. If ontological matters such as this had a bearing on the value or importance of human persons, then the demise of dualism would be bad news. Creation follows a route from its initial state, when there were only the most basic physical items, to a world that contains, among other things, human persons, in virtue of the causal laws authored by the Creator. This causal route is, however, indeterministic. It is a matter of chance that human persons are among the ways the elementary constituents of the material world came to be configured. The world was not programmed to bring human beings into existence. We are superficial in this regard too, existing only in virtue of complex arrangements of more basic components, arrangements that are far from inevitable, but the outcome of a highly contingent history. The elementary physical constituents were called into being from nothing, and contingently assembled, to make things like us possible. In light of this, it is not surprising that there are conceivable situations in which the universe provides no answer to the question whether a particular human being already exists, or still exists, or whether she is identical to someone existing at another time. Such possible indeterminacy is inevitable for complex material objects. It does not preclude reasonable judgments about a human person at one time being the same human person at a different time. I know that I am the person who wrote the first chapter of this book despite the fact that I can conceive of an outlandish, yet theoretically possible, scenario involving a malfunctioning teleporter in which there is no fact of the matter as to whether I am the person who materializes on Mars. I believe that the God who resurrected Jesus Christ will resurrect me, despite the fact that it is possible for God to create multiple copies of me and thus bring about a situation in which there is no fact of the matter as to whether I exist. God promises to resurrect the dead, so I am confident that there is a determinate fact of the matter:
Resurrection
299
there is a future human person to whom I am identical. The objection to materialist resurrection that invokes the transitivity paradox, like other objections to scientific naturalism, relies on failure to recognize our ontological superficiality. DUALISM’S LAST STAND If the foregoing is sound, Christians are entitled to embrace the full, traditional idea of the bodily resurrection of human persons, even though we do not exist between death and resurrection. This is a happy outcome, for science offers no comfort for the view that we possess non-physical minds or souls that could make possible continuous existence from death to resurrection. However, while many thoughtful Christians are aware of the scientific case for materialism, they see no way to defend belief in resurrection without positing an immaterial component of the human person that survives death. Thus, some seek ways to believe what science tells us about the brain without abandoning dualism, but it is not easy to find a reasonable way to do this. The fundamental difficulty such an account faces lies in the conjunction of two facts. One is that the existence of a normally functioning human brain suffices for the existence of a human person. The other is that an existential gap is avoided only if something exists between death and resurrection the existence of which is sufficient for the existence of the person. After death, no functioning brain exists. So if there is no gap in the person’s existence, there must be something other than the brain which suffices for the existence of the human person. Thus, there must be two different things, one the brain and one something else, both of which are sufficient for the person’s existence. In standard dualism “soul” and “mind” are used interchangeably to refer to one thing, the immaterial substance that bears a person’s mental properties and is her essential part. However, there is a popular idea that there is a mysterious immaterial thing, the soul, distinct from the mind, which apparently has no mental, or any other known, properties and has no role other than to ensure the person’s ongoing disembodied existence. Yet the mere fact that some part of the person, material or immaterial, spans the period from the death of the body to the resurrection does not mean that the person exists during that period. If it did, then preserving a dead person’s left big toe in a jar of formalin until the last day would maintain him in existence and there would be no existential gap. 23 Supposing the thing that exists during this period to be non-physical does not help unless it is the person, not just something that was once a part of the person. A gap in the person’s existence is avoided only if something exists between death and resurrection which has the characteristics necessary for the person’s existence. Anything that has these characteristics
300
Chapter 9
ipso facto is the person. It must be the person, even if truncated because it lacks a body. It need not have all the mental properties the person once had, and will have again, but it must have whatever mental properties are necessary for being the person. The problem is that these properties belong to the brain. Conceivably, science reveals that the brain possesses some mental properties, but does not reveal that there is a non-physical substance which also has mental properties and it is these, not those that belong to the brain, which make something a person. This entity survives the demise of the brain and, so long as it exists, the person exists. How might psychological properties be distributed between the person’s brain and this immaterial thing? Some find plausible that while we possess a good general idea of how brains are capable of cognitive activities that involve computation, we have no idea how they can have consciousness, subjective experiences such as having a sensation of blue. Perhaps, then, conscious experience occurs in an immaterial substance, which is the essential, immortal part of the human person. However, as we noted in Chapter Five, this appears to involve an illicit inference from the fact that we do not know how something can have a property to the conclusion that we do not know that it has that property. Despite our lack of understanding how material things can be conscious, we do not doubt that some are, e.g., cats. This proposal also needs to deal with the powerful evidence from science, as well as from ordinary experience—recall the effects of gin, or a hard blow to the head, on human subjectivity—that the brain is the seat of consciousness no less than other aspects of mind. This strategy also needs to explain how something which is conscious, but which lacks cognitive capacities, can be a person. To ensure that the immaterial thing that exists after death is a person, various cognitive capabilities that plainly belong to the brain will have to be ascribed to it. The dualist may, of course, do so, but at the cost of dismissing a great deal of scientific knowledge of what human brains do. An alternative but related strategy holds that material things, such as the brains of human beings and some animals, are conscious, but that the self-reflective thought in virtue of which one is a person is rooted in an immaterial substance. This implies that self-consciousness is not what arises naturally when consciousness and significant cognitive capacity are combined, but that it is a very different, sui generis, property. 24 Possibly, while material things can be conscious, and material things can engage in cognition, no material thing can have both features, but the proponent of this view owes us a reason to believe so. A similar idea is that the brain carries out routine mental tasks, such as mathematical calculation, but that the immaterial mind engages in more challenging activities that call for creativity and deliberate, conscious reasoning. This casts the brain as something akin to a calculator or computer that the person, i.e., the non-physical mind, uses. In Paul
Resurrection
301
Bloom’s words, it is conceived as “a cognitive prosthesis, added to the soul to increase its computing power.” 25 Such maneuvers, like the attempt to have the immaterial soul bear consciousness, fail in light of the scientific evidence that the activities they assign to an immaterial mind are no less clearly activities of the brain than are other mental activities. An alternative distribution assigns cognition and emotion about certain things, say, moral matters, or one’s relation to God, to the immaterial component, while accepting that the brain attends to less exalted matters. At best, this division of labor is arbitrary: If my immaterial mind thinks about God, and my material brain thinks about dogs, what goes on when I entertain the idea that God created dogs? Is the one mental act distributed over the two entities? At worst it reflects the notion that mind is the “spiritual” part of a human being, the part to which God can connect, and reveals a concomitant denigration of material reality fundamentally incompatible with the Christian view of creation. A more exotic tactic gives up on parceling the mind out between the mortal brain and an immortal non-physical entity simultaneously, and instead posits a transfer of mental activities. Science indicates that each mental property is a property of the brain, but perhaps a non-physical substance can, after the brain’s demise, acquire that property. On this account, the immaterial mind is a kind of backup device, into which the mind is uploaded at death. The brain becomes inoperative, but the individual’s mental life continues without interruption until she is resurrected, when she will be downloaded to a newly created, or re-created, embodied brain. God might have endowed human beings with a nonphysical backup for their minds to make it possible to resurrect them. The objection is that if what it describes is possible, then there is no reason to fear temporary non-existence. If the mind can be uploaded into an immaterial substance, why not simply upload it to a copy of the original brain? No organ will be available for this purpose until the general resurrection, but why doesn’t God upload the information from the neardeath brain to a brain he brings into being on that later occasion? The psychological continuity personal identity requires does not require spatio-temporal continuity. A reliable causal connection is all that is called for, and the continuous existence of God who remembers, and intends to re-create, the mind/brain that dies suffices for psychological continuity. Finally, it is worth noting attempts to preserve continuity in the absence of the body by appeal to an analogy drawn from the computational theory of the mind: the mind is the software and the brain is the hardware. After the hardware is destroyed, the software continues to exist, and can be run again on future hardware. There are two senses in which this can be true. The word processing program I am now using is, on the one hand, a complex algorithm, a set of instructions for a machine to do various things under certain conditions. As an abstract object, it exists outside space and time, in whatever manner such things can be said to
302
Chapter 9
exist. So, it will exist after this computer is destroyed, just as it existed before there were any computers or any programmers. But human persons are not abstract objects, and the existence of the software does not suffice for the existence of the person. If materialism is true, a person is a concrete physical particular. If dualism is true, a person is a whole that has a concrete non-physical particular as a proper and essential part along with a body (a concrete physical particular) as an inessential part. It is also true that the software can continue to exist without the hardware in the sense that some concrete thing can exist which is a representation of it. That representation, a concrete particular, can be ink on paper, magnetized spots on a disk, or neural states in the brain of someone with a remarkably capacious memory. Or it can be a state of an immaterial mind, such as God’s. Persisting representations can be used to program a new computer to do precisely what the computer that was destroyed once did. Thus, John Polkinghorne can write that information about the way a brain is structured so as to be a particular human mind continues to exist in the mind of God after that person no longer exists, and could later be used to structure a brain to function in the way the brain of the deceased once functioned. This is plainly correct, but Polkinghorne mistakenly describes this as the mind of the person existing after he dies. 26 It does not involve the continuous existence of the person. It is on the contrary what occurs when there is a gap in the person’s existence. The error flows from misconstruing the analogy. If something like computationalism is true, then the human mind is what exists when the software is executed by the brain. It is neither the software nor the hardware as such. The human person exists when a particular dynamic pattern is imposed on appropriate matter; the person is not the pattern. The idea that the mind exists without the brain, because software exists without hardware, is akin to thinking that a cake exists after it is eaten because the recipe continues to exist. The recipe is just the instructions for making various ingredients into a cake. For those who have died, being remembered by God is necessary for their future existence, but it is not sufficient for their current existence. No doubt there are other, increasingly exotic, strategies dualists might devise to try to fit the fact that prior to death a human mind is a functioning brain with that mind’s continued existence after that brain no longer exists, but no plausible prospects appear. We see here a clear example of the difficulty faced by those who seek to be honest about contemporary science while denying its naturalistic implications. Insofar as such strategies introduce non-physical entities whose only purpose is to avoid an existential gap between death and a resurrection which is allegedly impossible without them, they appear ad hoc. In contrast, the account of the possibility of resurrection on offer in this chapter relies only on the material human person known to ordinary experience and to science, and on
Resurrection
303
an account of what suffices for the identity of a human person through time that we have good reason to believe in any case. THE ALLEGED TEDIUM Christians refer to the resurrection as the blessed hope, but some critics contend that even if possible, it is not desirable, not to be reasonably hoped for. We could not, they say, rationally hope to live again, or at least to live forever, even if we could rationally believe that there is a God who gives life to the dead. Given the current state of humankind, our destructive impulses, our vain desire to be as gods, knowing good and evil, it is, all things considered, better that human life comes to its natural end. Fallen human life is still good, but not so good that there cannot be too much of it. We may judge that even if death is bad, endless life in our current condition would be worse. We can concur with God’s judgment in Genesis, Adam and Eve are denied access to the Tree of Life lest they eat from it and live forever.” 27 No objection to resurrection lies in this, since the promised life to come is the life of the redeemed, where the damage humans have inflicted on themselves and on one another is undone. Other issues arise because human beings, even when resurrected, remain finite. A common objection is that the life of a finite creature extended for infinite duration would eventually lead to utter boredom, the alleged tedium of endless existence. A finite mind, no matter its capabilities, can have only some finite number of distinct experiences. 28 Eventually, no new experiences are possible, and whatever one’s subjective state, it will be one she has already been in any number of times. 29 Bernard Williams, contending for the undesirability of everlasting life, writes, “Nothing less will do for eternity than something that makes boredom unthinkable.” Williams asks rhetorically, “What could that be? Something that could be guaranteed to be at every moment utterly absorbing?” 30 Williams, who was an atheist, believed there is no such thing, but it is remarkable that his seminal essay on the subject makes no reference to the Christian claim, which is, of course, that God is precisely what makes boredom unthinkable in the life to come. It is totally implausible to claim that an infinite Deity lacks the capacity to keep finite persons enthralled, fresh, and forever anticipating the future with joy. One might deny the possibility or reality of the beatific vision, but it cannot seriously be described as boring. For Christians, discussion of resurrected life with no reference to God is misconceived. One might as well discuss the pros and cons of marriage without mentioning that it involves acquiring a spouse. In Christian perspective the duration of life after death is secondary to its quality. The point of being resurrected is not to achieve life without end, but to share in the everlasting life of God. Gregory of Nyssa
304
Chapter 9
wrote that to be the friend of God is the perfection of human life, and we can add that that perfection is endless because God is everlasting and infinite. 31 Communion with our Creator will not become stale. The internal life of the infinite Triune God, into which creatures are called, is never repeated, ever new. That is a life in which boredom is unthinkable. 32 The conviction that creatures limited in the ways we are will receive the gift of everlasting life raises questions beyond the prospect of endless tedium. Personal identity can be maintained over arbitrarily long periods of time only with robust memory connections, but does the finite memory capacity of human brains imply that most of what resurrected humans experience must be forgotten? If resurrected persons do not undergo qualitative psychological change, they face a torturous eternity of stasis, but if they change at all, however gradually, in due course they will have changed drastically. This, with the worry about the finitude of memory, points to the possibility that even as bodily identity continues forever, over the ages to come persons slowly fade out of existence as new persons gradually replace them. It is not clear how to deal specifically with these and similar challenges. One consideration is that we do not suppose that our relation to time will be the same as now. We may be, as the resurrected Jesus appears to have been, freed from some of the constraints of space and time. I conjecture that God is not bound by the temporal nature of creation, not in virtue of being timeless, but because divine temporality is the reality which our world images. The human beings God retrieves from the nonexistence of death to share the Triune life may experience a changed relationship to time that obviates the problems that might otherwise arise from sheer endless duration. But we can only speculate. Another consideration also comes from Jesus, our model of resurrected humanity but more generally of human life in full fellowship with God. Jesus, God incarnate, is a finite creature, a normal human being. He was endowed with no superhuman, let alone infinite, capacities. The Gospels record Jesus doing things, and knowing things, that are not naturally possible for human beings. The incarnate God does miracles, not because of intrinsic capabilities he possesses—all those he gave up to become one of us—but only because the God he loves, trusts, and obeys acts through him. And Jesus makes the astounding promise that the rest of us, joined in faith to him, and thus to God, will do even greater things (John 14:12). God the Father, together with the Spirit, makes unlimited power available to the incarnate Son. We saw that what makes possible the future existence of human persons and their human bodies, despite the end of existence death entails, is God intervening in the causal processes necessary for personal identity over time. The necessary causal connections are detoured through the mind of God. We may suppose that God will also be available to those who are resurrected, providing for them whatever is needed to sustain them in joyful everlasting life.
Resurrection
305
THE COMING KINGDOM Because God is God, God is not all there is. The divine community of Father, Son, and Spirit calls into being things other than God, and this includes created persons. Created, but not designed, they have their own lives, not merely imagined by, nor necessarily emanating from, God. Distinctness of persons is the necessary ground of love freely given. Yet God is endlessly resourceful in being available to them, condescending to whatever they have turned out to be, patiently but relentlessly calling them into communion with their Creator. God does not monopolize creatures’ attention. Calling them to fellowship with their maker, the Creator calls them to fellowship with one another as well. God’s aim is for created persons to love one another. The paradigmatic form of loving fellowship with the Triune God is life in a community of human persons bound in love to one another. Relations of mutual acceptance and care among human persons mirrors the life within God and between God and creatures. The Christian hope, grounded in confidence that human perversity cannot in the end defeat God’s aims, is that all humanity will at last come into communion with God. Salvation, we trust, is not only for those of us who in this life have been privileged to hear and proclaim the good news. Christians are often unnecessarily pessimistic about this. Natural selection has endowed human beings with a deeply rooted sense of justice, one that demands that the good be rewarded and that the wicked receive retribution. The Christian faith scandalizes this sensibility. God’s justice is an implacable commitment to God’s loving purposes for creatures. This does not imply that the wicked are not wicked, or that God does not care what human beings do. The God who loves human beings hates everything that degrades and damages them, including what they do to themselves and to one another. Some see in this a demand for divine retribution, and the unrepentant wicked at last being cast into hell, or at least annihilated. More in keeping with the revelation in Jesus Christ is confidence that God is committed at all costs to healing, forgiveness, and reconciliation. 33 Here, as in creation, God’s means are appropriate to God’s ends. God does not invite creatures freely to enter into a relationship with their Creator while threatening perdition for the wrong choice; coercion, even (perhaps especially) at the hands of God, is incompatible with freedom. We are now in a realm of conjecture, but it seems more likely that while God devises no punishment for evildoers or for the faithless, God is not willing to give up on anyone, however obdurate. God maintains the post-mortem recalcitrant in existence for as long it takes, never willing to use force, but always ready to persuade. We may call this hell, but confident in God’s power of persuasion and unwillingness to be complicit in any creature’s self-destruction, the hope is for hell being harrowed and, in the fullness of time, emptied. At the same time, resurrected sinners, which includes all of us but for Jesus himself, are
306
Chapter 9
called to make reconciliation with one another, seeking forgiveness, and perhaps no less difficult, forgiving as we have been forgiven. I imagine that this, even if joyfully entered into, will in many ways be far from easy. What lies beyond is the new beginning. World without end. Amen. God’s creative project has been long delayed but not forsaken. In the Genesis creation story, what God intends for a material image is neither a life devoted to religious contemplation nor a life of indolence. God calls humans to meaningful work, shared with one another, and shared with the Creator. Humankind is, as Philip Hefner often reminded us, the created co-creator. 34 God rests after the six days of creation, not because the work of creation is complete, but to make room for the material creatures who are the image of their God to join in the unending creative endeavor. At first assigned the task of tending a safe place, suitable for the immature, but then going out into the wild, partnering with the Creator to explore and transform a treasured but still inchoate world. “Be fruitful and multiply, and fill the earth and subdue it; and have dominion over the fish of the sea and over the birds of the air and over every living thing that moves upon the earth” (Genesis 1:28). The tragedy of human history makes it all too clear how unfit we have made ourselves for this task. Our partnership with God has been largely on hold, but the Christian hope is that resurrected humanity will at long last set out toward its destiny. God wills to be with humankind and thus, as Karl Barth says, humankind is “not enclosed in the circle of its intrinsic possibilities.” 35 The projects in which the Creator enlists resurrected humans may be ones it would be preposterous for humans, or any other creatures, to consider unaided, e.g., to maintain the universe itself in existence, like the mythic human pair in Eden, called to tend the garden, working to make it flourish rather than head toward entropic demise. Being who he is, God does not demand our undivided attention, but calls us to love one another. When human beings answer the call to trust God and join in the work of reconciliation, we do not do so alone. When resurrected humanity takes up its share in the Creator’s life and work, it will probably not do so alone. This is a universe that, in due course, gives rise to personal creatures. But it does so as a result of laws that are indeterministic. The creation was not pre-programmed to produce any specific species of person; the general outcome was guaranteed but the specifics left to chance. This is what we should expect of a God whose aim is not control but to give life to truly distinct persons, persons with their own lives that they can choose to join to the life of God. As noted in chapter 4, a reasonable expectation is that a universe so constructed did not produce persons just once. In a chancy universe, once is too close to never. Personal life may in relative terms be exceptionally rare, but from the perspective of Christian faith, we should assume that in absolute terms it has come forth many times, and that humans are not the only persons God has created. Our future relations with such fellow creatures,
Resurrection
307
and their possible relations with one another and with their Creator can be no more than matters for speculation. 36 But we should not be surprised if what awaits us is coming to know and love a great host of persons who with us share the call of being God’s image. NOTES 1. Strictly speaking, it is possible, but so wildly improbable as to be immediately dismissed. All the atoms that comprised Napoleon shortly before he died might by chance find their way to one location and reassemble as they were previously assembled. We are entitled to say that this physically possible event would never happen in the natural course of events. Perhaps it could happen artificially, by means of some technology possessed by future humans or other creatures. However, this would not be resurrection. It would be the revivification of a mortal life, not resurrection to what St. Paul in 1 Corinthians 15 calls a “spiritual body,” i.e., a body freed from the power of sin and death by the Holy Spirit. 2. The Christian materialist is committed to the implication that Jesus did not exist between his death and resurrection. God’s existence is necessary, not contingent. Jesus is fully God, as he is fully human, so the at least surprising further implication is that there is a temporal discontinuity in the existence of God incarnate. I believe that the conviction that Jesus Christ is fully human, and thus a mortal, material being, as he is fully God, properly shapes any views we might have about the specifics of how God’s necessary existence—which I affirm in chapter 3—applies to the persons of the Trinity. The death of Christ means that loss, death and non-existence are somehow taken into and forever present in God. Thanks to Taelor Lamansky for helpful discussions of this issue. 3. We should pay some heed to what dualists take as biblical support for their view. The materialist account implies that humans do not exist in an ‘interim state’ between death and the general resurrection. This equation of death with non-existence, in contrast to disembodied existence, encounters biblically grounded objections analogous to those discussed with respect to materialism in chapter 5. They should be dealt with in similar ways, as we distinguish theoretically committed from everyday language, the text’s actual intent, and what it teaches as abstracted from the cultural suppositions of the authors. For example, when the Gospels report that Jesus says to the believing thief on the cross, “Today you will be with me in paradise” (Luke 23:43) this should be understood in terms of the thief’s subjective perspective—his next experience after death will be resurrected life with Jesus—not his objective relation to time. St. Paul writes, “to be absent from the body is to be present with the Lord” (2 Corinthians 5:8). Here, as elsewhere in his writings (e.g., 2 Corinthians 12:2–4), we most likely see Paul’s personal metaphysical assumptions in play, ones we can leave in the background with no loss of essential meaning. Here it is important to keep in view that in the one place where he addresses the resurrection in theoretically serious terms (1 Corinthians 15) he does so without reliance on these assumptions. In Jesus’ parable of the rich man and Lazarus (Luke 16:19–31) the characters are embodied, not disembodied souls awaiting resurrection. The parable’s intent is not to portray the human condition between death and resurrection, but the obduracy of resistance to Jesus in this life. A decisive New Testament text is Romans 4:17, where St. Paul proclaims the God who “gives life to the dead and calls into being that which does not exist.” We should see God’s re-creation of the dead as akin to the original creation out of nothing, creation from the non-existence of death. 4. Someone might say that it continues to exist while its parts are temporarily scattered, but this seems implausible. If all the pieces are close by, piled in a heap on the workbench, someone might say, “There’s the toaster” and contend that it still exists, albeit in a nonfunctioning, disassembled way. The plausibility diminishes as
308
Chapter 9
the parts are further separated, as they become incorporated into other things, or as the chances of their being reassembled decline. If the company were suddenly to stop operations, so it is never reassembled, the truth would be that the toaster stopped existing at the time it was disassembled. 5. The City of God, Marcus Dods, trans. (New York: Random House, 1950), Bk. 22, Ch. 20. 6. City of God, Bk. 22, Ch. 12 and Ch. 20. 7. Augustine’s conviction that the resurrected body must contain every bit of matter present in the mortal body appears to be grounded not in a metaphysical theory about what is necessary for the identity of a body through time, but in what to the modern reader is an unreasonably literal application of Jesus’s words, “Not a hair of your head shall perish” (Luke 21.18). In Augustine’s view (City of God, Bk. 22, Ch. 12), even the matter that comprised hair and nails cut from the body in the course of one’s earthly life will be restored to it at the resurrection; fortunately, that matter will be redistributed over the body as a whole, and we need not worry about having excessively long hair, toenails, and fingernails in the next life! Not all early Christians agreed. Origin, e.g., knew that there is a constant flow of microscopic parts through biological organisms; see the discussion of his views in Caroline Walker Bynum, The Resurrection of the Body in Western Christianity, 200–1336 (New York: Columbia University Press, 1995), 63–71. 8. City of God, Bk. 22. Ch. 12 and Ch. 20. This appears to have the odd implication that no matter how many victims the cannibal feasts on, he cannot gain weight; all their parts just pass through him. 9. The philosopher Peter van Inwagen offers a proof that resurrection is possible by contriving an even less plausible scenario: when a human being dies God surreptitiously removes the corpse to an undisclosed location and replaces it with a simulacrum. The body is preserved until the Last Judgment, when it will be restored to life. “The Possibility of Resurrection,” in International Journal for the Philosophy of Religion 9 (1978), 114–21. 10. That there actually is not a single atom in common is, I assume, improbable, given the disparity between the number of atoms and the number of body-sized volumes. But I will make this simplifying assumption here and in what follows. 11. After mitosis the daughter cells are exactly similar to one another. Cells die either because they receive a signal that instructs them to die, or because they do not receive a signal that instructs them not to die; this depends on the type of cell. Whether a cell receives a signal depends on its location in the body, i.e., its proximity to other cells. It’s also worth noting that the microscopic components of certain types of cell are replaced in the course of their existence. Thanks to my colleagues, biologists Elizabeth Heeg and Tyrone Gnade, for giving me a sense of how this works. 12. We assume that human beings who die in adulthood will not be resurrected as infants. 13. Although God is affected by what occurs in creation, ultimately God transcends the realm of natural cause and effect. God’s beliefs and desires do not cause God’s choices. The immaterial God possesses the radical free will some mistakenly ascribe to human agents. Therefore, there is a gap in the causal chain that connects the mind of young Sarah to the mind of resurrected Sarah. This does not make the process unreliable. If God knows that a particular act is necessary to achieve a goal, we can be confident that this is precisely what God will do. The detour through the mind of a metaphysically free Deity introduces a gap in causation, but not a gap in reliability. 14. God could make any number of such bodies, but two is all it takes to make trouble. 15. John Hick, Death and Eternal Life (New York: Harper and Row, 1976), 292. 16. Not everyone insists that personal identity cannot depend on such extrinsic facts about persons. See, e.g., Robert Nozick, Philosophical Explanations (Cambridge, MA: Harvard University Press, 1981), 29–114, and Stephen T. Davis, Risen Indeed: Making Sense of the Resurrection, (Grand Rapids, MI: Eerdmans, 1993), 132–146.
Resurrection
309
17. The transitivity paradox appears to overturn the traditional dualist’s conviction that the bodies of resurrected persons are the bodies that died. It looks as though she has to regard the resurrection of Jesus as a special case. But it does not threaten her primary conviction that resurrected persons are persons who existed before death. The immaterial soul, necessary and sufficient for the person’s existence, never goes out of existence and in never recreated. 18. We need not share the assumption that identity of soul is necessary or sufficient for personal identity: we can conceive of a soul undergoing drastic psychological change that breaks the psychological continuity necessary for identity, and we can conceive of psychological continuity sustaining personal identity across a multiplicity of souls. 19. Derek Parfit, Reasons and Persons (Oxford, UK: Oxford University Press, 1984), 245–80. 20. No doubt there are other conceivable circumstances, e.g., the fusion of two persons into one entity, in which we have no objective grounds to apply the concept one way rather than another. 21. David Lewis, On the Plurality of Worlds (Oxford, UK: Basil Blackwell, 1986), 212. I should note that Lewis does not invoke the vagueness of identity to cope with Parfit’s fission scenario. Instead, he identifies the original patient, Righty, and Lefty not as whole persons, but as temporal parts of persons, person stages. Thus, the original patient and Righty are stages of a person, the original patient and Righty are stages of a person, but Lefty and Righty are not stages of the same person. Now the odd implication is that prior to the procedure, the original patient is a stage of two different, overlapping persons. See David Lewis, “Survival and Identity,” in Amélie Oksenberg Rorty, ed. The Identities of Persons (Berkeley: University of California Press, 1978), 17–40. 22. An essentially simple reductio ad absurdum argument for the incoherence of vague identity originates with Nathan Salmon and Gareth Evans. See Salmon’s “Identity Facts” in Metaphysics, Mathematics, and Meaning, Philosophical Papers, Volume I (Oxford, UK: Oxford University Press, 2005), 166–91 and Evans’s “Can There Be Vague Objects?” Analysis 38(1978), 208. Suppose that it is indefinite whether x = y. This implies that x has the property of being something which is neither identical to, nor not identical to, y. But y does not have this property. It is definitely the case that y = y. Therefore, y has a property that x lacks. By the principle of the non-identity of discernibles, this implies that x ≠ y. Therefore, it is false that it is indefinite whether x = y. For attempts to interpret and refute the argument see, e.g., Terrence Parsons, Indeterminate Identity (Oxford, UK: Oxford University Press, 2000) and Peter van Inwagen, Material Beings (Ithaca, NY: Cornell University Press, 1990), 244–70. 23. Or, as some early Christians, as well as Jews and Muslims, proposed, an indestructible bone found at the base of the spinal cord. See Caroline Walker Bynum, The Resurrection of the Body in Western Christianity, 200–1336 (New York: Columbia University Press, 1995), p. 54. 24. J. R. Lucas’s Gödelian argument against computationalism, discussed in chapter 5, is akin to both this and the preceding strategy: whatever cognitive powers might be generated by the brain’s computation, the human mind is infinitely self-reflective and no mere computer. 25. Descartes’ Baby: How the Science of Child Development Explains What Makes Us Human (New York: Basic Books, 2004), 201. 26. “Eschatological Credibility: Emergent and Teleological Processes,” in Ted Peters, Robert John Russell, and Michael Welker, eds. Resurrection: Theological and Scientific Assessments (Grand Rapids: Eerdmans, 2002), 51–53. This approach resembles that of many Christian dualists, from Thomas Aquinas to the present day, who accept Aristotle’s account of the soul as the form of the body and maintain that the soul/form continues in existence after the body is destroyed. See, e.g., J. P. Moreland and Scott B. Rae, Body and Soul: Human Nature and the Crisis in Ethics (Downers Grove, IL: InterVar-
310
Chapter 9
sity Press, 2000). So far as I can see, this is subject to the same objection as the idea that the soul is software. Abstract objects lack causal efficacy, but human minds do not. 27. In the Old Testament, while premature death, or death uncomforted by hope for one’s progeny, is evil, death in itself is not necessarily portrayed as an unmitigated evil. 28. This is avoidable: the minds of the resurrected need never become actually infinite. They need only forever increase in capacity as fast as novel experience requires. I make no use of this loophole because, while God could do this for resurrected persons, it is not obvious that they would remain recognizably human. 29. Of course, since many experiences will not be available for the blessed, e.g., that of being eaten by wild ferrets, believing that one has been damned to hell, etc., the resurrected human being has at most some proper subset of all possible experiences. 30. “The Makropulos Case: Reflections on the Tedium of Immortality” in Bernard Williams, Problems of the Self: Philosophical Papers 1956–1972 (Cambridge, UK: Cambridge University Press, 1973), 95. 31. Gregory of Nyssa, The Life of Moses, Abraham J. Malherbe and Everett Ferguson, trans. (New York: Paulist Press, 1978), 137. 32. We know already that the boundaries of the human mind are in a way porous. Our cognition, as Andy Clark argues, is not limited to the brain, but is carried out beyond its boundaries by means of our cultural and technological “scaffolding.” See his Supersizing the Mind: Embodiment, Action, and Cognitive Extension (Oxford, UK: Oxford University Press, 2008). 33. Also casting doubt on divine retribution is the fact that faith requires belief, and belief cannot be chosen (chapter 2). Further, whatever free choices humans make, they do not exercise libertarian free will and thus are not the ultimate origin of their actions (chapter 6). 34. See, e.g., The Human Factor: Evolution, Culture and Religion (Minneapolis: Fortress Press, 1993). 35. Karl Barth, Church Dogmatics, III, 2. Harold Knight, G. W. Bromiley, J. K. S. Reid, and R. H. Fuller, trans. (Edinburgh: T. & T. Clark, 1960), 165. 36. For speculations, see Steven J. Dick, ed., Many Worlds: The New Universe, Extraterrestrial Life, and the Theological Implications (Philadelphia: Templeton Foundation Press, 2000).
Conclusion
The universe looks the way we should expect it to look if it is God’s creation, at least if that God is the God of Christian faith. For some, this will sound like post hoc rationalization, a belated attempt to make discredited faith cohere with science. For others, it will sound like craven capitulation to the cultural success of science and surrender to the naturalistic presuppositions of its proponents. To both the only response is to focus on the central Christian confession: there is a God who is a community of loving persons—Father, Son, and Spirit—and this Triune God creates persons truly distinct from their God and calls them to serve as the divine image in creation. We ask what a world created by this God— not some generic, religiously plausible deity devoted to design, command and control—should look like and reasonably conclude that it is the world as science portrays it. Daniel Dennett once said that modern science is incompatible with “all but the most pastel versions of religious faith.” 1 On the contrary, the consonance of faith and science comes into view only when we look to the vivid character we encounter in the particularities of biblical revelation: the God of Israel, made human flesh in Jesus Christ. In the long view, Western civilization has been shaped by two great accounts of the world and the place of humans in it. One originates with the ancient Greeks and, despite its non-theistic origins, has enjoyed a very long association with Christianity, which arrived on the scene in its heyday and made creative use of its materials in the articulation of the faith. The other—the scientific—originates in the late-medieval Christian culture, but only after long reflection on what after centuries was still an only partially successful attempt to integrate the Greek legacy into the Christian faith. To this day, many Christians remain loyal to a human self-image inherited from the ancient account and decry what science now reveals about our nature. Others, dismissive of Christianity—or at least of a caricature of it—are happy to agree that it is incompatible with the implications of science. Nonetheless, when we step away from the contingencies of history, the reasonable judgment is that it is the scientific account, to which Christian faith gave birth, that coheres with that faith. That is what I have tried to show in this book. This is not, however, an exercise in Christian apologetics, except indirectly, as it dispels some misconceived objections to the faith. The universe looks the way we should expect it to look if Christianity is true, but 311
312
Conclusion
this is evidence for Christianity only if we should expect the world to look some other way if Christianity were false. 2 Of course, as a Christian, I believe this: if Christianity were not true, observed reality would have been different. God would not have been revealed to the people of Israel, in the life, death, resurrection, and ascension of Jesus, or in the ongoing life of the Christian Church. But this is a matter not of science but of human history and its explanation. So far as science is concerned, what should the world look like on the assumption that there is no God? The matter seems largely imponderable. The bare supposition that there is no creator seems to tell us only what not to expect; it appears to sustain no positive expectations. 3 If there is no God then we should be very surprised if, say, homo sapiens popped into existence spontaneously, rather than as the result of some natural process, but the same is true if the Creator Christian faith confesses created humanity as God’s material image. The idea that the world science portrays appears not to have been created depends on assumptions about the Creator’s ends and means that on reflection Christians have no good reason to accept. It is, I trust, clear that what I bring together in this book is one of various possible ways to articulate the Christian faith, and some of the many possible ways to solve the philosophical problems that arise as we seek to locate what matters for human beings in the world science describes. I do not imagine that anyone agrees with everything on offer, either theological or philosophical. The point is that we can fashion a more or less coherent account that makes it reasonable for someone with a robust Christian faith enthusiastically to accept the naturalistic implications of contemporary science. Elements of that account are, no doubt, often counterintuitive and sometimes disturbing to my fellow believers: science has rational authority to trump the claims of faith, God does not intend or foreknow all future events, there is no immaterial soul, the specifics of creation were not designed, morality lacks final objectivity, and both it and religion are merely human realities to which God condescends. As disconcerting as all this may be, it leaves the Christian gospel intact. Much that is familiar is revised or denied, but everything essential remains: we are the created image of the God who in Jesus Christ delivers us. The incursion of science into domains once thought sacred and safely beyond its naturalizing reach predictably occasions anxiety, evasion, and denial. But this is not the response of resolute faith. The scientific genius of James Clerk Maxwell was already evident when he began his studies at Cambridge University in 1850. There, the young man who was to become the greatest physicist between Isaac Newton and Albert Einstein earnestly embraced the project of thinking through his deep Christian faith in light of the advance of science. To a friend he confided:
Conclusion
313
Now my great plan . . . is to let nothing be willfully left unexamined. Nothing is to be holy ground consecrated to Stationary Faith. I assert the Right of Trespass on any plot of Holy Ground which any man has set apart . . . .Now I am convinced that no one but a Christian can actually purge his land of these holy spots . . . .Christianity - that is, the religion of the Bible - is the only scheme or form of belief which disavows any possessions on such a tenure. Here alone all is free. You may fly to the ends of the world and find no God but the Author of Salvation. You may search the Scriptures and not find a text to stop you in your explorations . . . .But a Candle is coming to drive out all Ghosts and Bugbears. Let us follow the light. 4
Since Maxwell wrote, the scientific challenge to traditional beliefs has only intensified, but his confidence in God is no less justified and his advice no less needed. In our day, many lament the “desacralization” of nature that James Clerk Maxwell celebrated: “the once-sacred cosmos, in which the world of nature had been suffused with supernatural presence,” was supplanted by modernity. 5 The loss of a “sacramental ontology” accounts for the ills of modern, and post-modern, culture. 6 God’s immanence in creation, they say, is no longer evident, a tragedy inflicted in part by modern science. The prescribed remedy is the revival of ancient and medieval philosophies. If this book’s claims are correct, our pressing need is not for a resuscitation of pre-scientific Platonic or Thomistic metaphysics, but to look more deeply into the desacralized, demythologized, objectified, mechanistic universe of modern science and discover a world illuminated by the Christian Gospel, which is holy, dedicated to the Creator’s aim of fellowship with the creature. I began this book by quoting Carl Sagan, who rhetorically asked why we have not seen in the grand scientific picture of the universe proof of God’s greatness. I answered that we cannot do so until we abandon the attempt to separate science from its naturalistic implications and learn to see the world as scientific naturalism renders it as just the kind of world God would create. Sagan goes on to describe what he calls humankind’s “great demotion.” At the start, science deprived us of our imagined position at the center of the universe, and over the next few centuries went on to undermine our illusions of uniqueness and transcendence. 7 Modern science has been a course in humility, systematically prying away the pretension that we are privileged in our nature or origins. Far from transcending the physical world and its causal mechanisms, we are material beings and, among them we are superficial, the unlikely contingent product of blind evolution. Christian faith, mindful of fallen humanity’s engrained impulse to deny its status and to imagine itself godlike, welcomes the power of science to restore us to our place as creatures. But this is far from the last word. To concur with science is not to denigrate human beings, but to replace false notions of what is great about us with
314
Conclusion
the truth: the glory of human beings lies not in what we are, but in the one who creates us, loves us, and calls us to share in God’s everlasting life and work. Friedrich Nietzsche, rightly celebrating the demise of what he mistook for the true God, but which was a mere human artifact, wrote: at hearing the news that “the old god is dead,” we philosophers and “free spirits” feel illuminated by a new dawn; our heart overflows with gratitude, amazement, forebodings, expectation—finally the horizon seems clear again, even if not bright; finally our ships may set out again, set out to face any danger; every daring of the lover of knowledge is allowed again; the sea, our sea, lies open again; maybe there has never been such an open sea. 8
We must correct the greatest atheist. The future is bright, without fear or foreboding: we sail not alone but with God. Yet he was right about the promise of an open future. Who knows what wonders of divine love await us in the fullness of renewed time? NOTES 1. “Back from the Drawing Board,” in Bo Dahlbom, ed. Dennett and His Critics (Cambridge, MA: Blackwell, 1993), 203. 2. There is, of course, the pressing question whether we should expect a world created by God to contain the apparently gratuitous evil this world contains. We should note that insofar as evil is evidence against belief in God, this is not a conflict of faith and science. If science plays a role here it is most likely one of mitigation: it is reasonable to suspect that a world designed to produce things like us, i.e., beings capable of being personally related to their Creator, is one in which evil of various kinds is an unavoidable possibility. Whether taking such risks is consistent with the loving nature Christian theology ascribes to God depends on the value of the ends for which they are taken, the probability of the evil actually occurring, and what the risktaker does when things go badly. The Christian sees the problem as improperly framed if it implicitly asks, “Why doesn’t a good, omnipotent God do something about the evil in the world?” The Christian claims that God, being perfectly good, wise, and omnipotent deals with the possibility and actuality of evil in the best possible way, and that God’s entrance into human affairs, centrally as the incarnate Christ, is that way. For those of Christian faith, the problem of evil is the problem of defending this claim. Here, as with the problem of faith and science, attention to the specifics of Christian faith, not generic belief in God, is essential. 3. Perhaps there is just the metaphysical expectation, discussed in chapter 3: if there is no God then we should be surprised to find that this universe, or anything else, exists. 4. Letter to Lewis Campbell, 7 March 1852, in Lewis Campbell and William Garnet, eds. The Life of James Clerk Maxwell (New York: Johnson Reprints, 1969), 178. 5. Hans Boersma, Heavenly Participation: The Weaving of a Sacramental Tapestry (Grand Rapids, MI: Eerdmans, 2011), 56. 6. Boersma, Heavenly Participation, 124. 7. Pale Blue Dot: A Vision of the Human Future in Space (New York: Random House, 1994), 26–39. 8. The Gay Science, Walter Kaufmann, trans. (New York: Vintage Books, 1974), 280.
Bibliography
Adams, Robert M. 1987. “Must God Create the Best?” In his The Virtue of Faith and Other Essays in Philosophical Theology, 51–64. Oxford, UK: Oxford University Press. Ainslie, George. 2001. Breakdown of Will. Cambridge, UK: Cambridge University Press. Anselm. 1998. Basic Writings: Proslogium, Monologium, Gaunilo's In Behalf of the Fool, Cur Deus Homo. 2nd revised edition. N. S. Deane, trans. Chicago: Open Court Publishing. Aquinas, Thomas. 2018. Summa Theologiae Prima Pars, 1–49. Fr. Laurence Shapcote, trans. Green Bay, WI: Aquinas Institute. Ariely, Dan. 2010. Predictably Irrational: The Hidden Forces that Shape Our Decisions. Revised edition. New York: Harper. Ashton, John F, ed. 2000. In Six Days: Why 50 Scientists Choose to Believe in Creation. Green Forest, AR: Master Books. Asimov, Isaac. 1976. Only A Trillion. New York: Grosset and Dunlap. Atran, Scott. 1998. “Folk Biology and the Anthropology of Science: Cognitive Universals and Cultural Particulars.” Behavioral and Brain Sciences 21:547–569. Augustine. 1950. The City of God. Marcus Dods, trans. New York: Random House. Axelrod, Robert. 1984. The Evolution of Cooperation. New York: Basic Books. Barkow, Jerome H., Leda Cosmides, and John Tooby, eds. 1992. The Adapted Mind: Evolutionary Psychology and the Generation of Culture. Oxford, UK: Oxford University Press. Barr, James. 1993. The Garden of Eden and the Hope of Immortality. Minneapolis: Fortress Press. Barrett, H. Clark, Leda Cosmides, and John Tooby. 2007. “The Hominid Entry into the Cognitive Niche.” In The Evolution of Mind: Fundamental Questions and Controversies. Steven W. Gangstead and Jeffry A. Simpson, eds. New York: Guilford Press. Barrett, Justin L. 2004. Why Would Anyone Believe in God? Walnut Creek, CA: Altamira Press. ———. 2011. Cognitive Science, Religion and Theology: From Human Minds to Divine Minds. West Conshohocken, PA: Templeton Press. ———. 2012. Born Believers: The Science of Children’s Religious Belief. New York: Free Press. Barrett, Justin L., and Frank C. Keil. 1996. “Conceptualizing a Nonnatural Entity: Anthropomorphism in God Concepts.” Cognitive Psychology 31:219–247. Barth, Karl. 2004 [1940] Church Dogmatics, II.1. T. H. Parker, W. D. Johnston, Harold Knight, and J. L. M. Haired, trans. New York: Continuum. ———. 1956. Church Dogmatics, I.2. G. T. Thomson and Harold Knight, trans. Edinburgh: T. & T. Clark. ———. 1960. Church Dogmatics, III.2. 1960. Harold Knight, G. W. Bromiley, J. K. S. Reid, and R. H. Fuller, trans. Edinburgh: T. & T. Clark. ———. 1968. Epistle to the Romans. 6th edition. Edwyn C. Hoskyns, trans. Oxford, UK: Oxford University Press. Behe, Michael. 2008. The Edge of Evolution: The Search for the Limits of Darwinism. New York: Free Press. Bering, Jesse M. 2011. The Belief Instinct: The Psychology of Souls, Destiny, and the Meaning of Life. New York: Norton.
315
316
Bibliography
Bering, Jesse M., and David F. Bjorklund. 2004. “The Natural Emergence of Reasoning about the Afterlife as a Developmental Regularity.” Developmental Psychology 40:217–233. Berkouwer, G. C. 1962. Man: The Image of God. Derk W. Jellema, trans. Grand Rapids, MI: Eerdmans. Bishop, John. 2007. Believing By Faith: An Essay in the Epistemology and Ethics of Religious Belief. New York: Oxford University Press. Bloom, Paul. 2004. Descartes’ Baby: How the Science of Child Development Explains What Makes Us Human. New York: Basic Books. ———. 2007. “Religion Is Natural.” Developmental Science. 10:147–150. ———. 2009. “Religious Belief as an Evolutionary Accident.” In The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origin of Religion, edited by Jeffrey Schloss and Michael Murray, 118–27. Oxford, UK: Oxford University Press. Boer, Harry R. 1990. An Ember Still Glowing: Humankind as the Image of God. Grand Rapids: Eerdmans. Boethius. 1999. The Consolation of Philosophy. Revised edition. Victor Watts, trans. New York: Penguin. Book of Common Prayer. 1979. New York: Seabury Press. Boulton, Matthew Myer. 2008. God Against Religion: Rethinking Christian Theology Through Worship. Grand Rapids, MI: Eerdmans. Boyer, Pascal. 2001. Religion Explained: The Evolutionary Origins of Religious Thought. New York: Basic Books. ———. 2003. “Religious thought and behaviour as by-products of brain function.” Trends in Cognitive Science 7:123. Bradshaw, John. 2013. Cat Sense: How the New Feline Science Can Make You a Better Friend to Your Pet. New York: Basic Books. Brooke, John Hadley. 1991. Science and Religion: Some Historical Perspectives. Cambridge, UK: Cambridge University Press. Brown, Warren S. 2010. “Nonreductive Human Uniqueness: Immaterial, Biological, or Psychosocial.” In Human Identity at the Intersection of Science, Technology and Religion, edited by Nancey C. Murphy and Christopher C. Knight, 81–93. Burlington, VT: Ashgate. Brown, Warren S., Nancey Murphy, and H. Newton Malony, eds. 1998. Whatever Happened to the Soul? Scientific and Theological Portraits of Human Nature. Minneapolis: Fortress Press. Brown, Warren S., and Brad D. Strawn. 2012. The Physical Nature of Christian Life: Neuroscience, Psychology and the Church. Cambridge, UK: Cambridge University Press. Buss, David. 2013. Evolutionary Psychology: The New Science of the Mind. 4th revised edition. New York: Pearson Education. Bynum, Caroline Walker. 1995. The Resurrection of the Body in Western Christianity, 200–1336. New York: Columbia University Press. Calvin, John. 1960. Institutes of the Christian Religion. Ford Lewis Battles, trans and ed. Philadelphia: Westminster Press. Carpenter, Humphrey. 1979. The Inklings: C. S. Lewis, J. R. R. Tolkien, Charles Williams, and their Friends. Boston: Houghton Mifflin. Carroll, Sean. 2019. Something Deeply Hidden: Quantum Worlds and the Emergence of Spacetime. New York: Dutton. Carruthers, Peter. 2011. The Opacity of Mind: An Integrative Theory of Self-Knowledge. Oxford, UK: Oxford University Press. Chater, Nick. 2018. The Mind is Flat: The Remarkable Shallowness of the Improvising Brain. New Haven, CT: Yale University Press. Chisholm, Roderick. 1989. “Human Freedom and the Self.” In his On Metaphysics. Minneapolis: University of Minnesota Press. 5–15. Churchland, Patricia. 2011. Braintrust: What Neuroscience Tells Us About Morality. Princeton, NJ: Princeton University Press.
Bibliography
317
———. 2013. Touching a Nerve: The Self As Brain. New York: W.W. Norton and Company. ———.2019. Conscience: The Origins of Moral Intuition. New York: W.W. Norton and Company. Churchland, Paul. 2007. “On the Reality (and Diversity) of Objective Colors: How Color-Qualia Space is a Map of Reflectance-Profile Space.” In his Neurophilosophy at Work, 198–231. Cambridge, UK: Cambridge University Press. ———. 2007. “Rules, Know-How, and the Future of Moral Cognition.” In his Neurophilosophy at Work, 61–79. Cambridge, UK: Cambridge University Press. Clark, Andy. 2008. Supersizing the Mind: Embodiment, Action, and Cognitive Extension. Oxford, UK: Oxford University Press. ———. 2013. Mindware: An Introduction to the Philosophy of Cognitive Science. 2nd edition. Oxford, UK: Oxford University Press. ———. 2015. Surfing Uncertainty: Prediction, Action, and the Embodied Mind. New York: Oxford University Press. Cooper, John C. 1989. Body, Soul, and Life Everlasting: Biblical Anthropology and the Monism Dualism Debate. Grand Rapids, MI: Eerdmans. Cosmides, Leda, and John Tooby. 1995. “Cognitive Adaptations for Social Exchange.” In The Adapted Mind: Evolutionary Psychology and the Generation of Culture, edited by Jerome H. Barkow, Leda Cosmides and John Tooby, 163–228. Oxford, UK: Oxford University Press. Craig, Edward. 1987. The Mind of God and the Works of Man. Oxford, UK: Oxford University Press. Crane, Tim. 2003. The Mechanical Mind: A Philosophical Introduction to Minds, Machines, and Mental Representations. 2nd edition. London: Routledge. Cross, Richard. 1999. Duns Scotus. New York: Oxford University Press. Davidson, Donald. 1989. “Actions, Reasons, and Causes.” In his Essays on Actions and Events, 3–20. Oxford, UK: Oxford University Press. Davies, Paul. 2010. The Eerie Silence: Renewing Our Search for Alien Intelligence. Boston: Houghton Mifflin Harcourt. Davis, Stephen T. 1993. Risen Indeed: Making Sense of the Resurrection. Grand Rapids, MI: Eerdmans. Dawkins, Richard. 1976. The Selfish Gene. New York: Oxford University Press. ———. 2006. The God Delusion. New York: Houghton Mifflin. ———. 2007 Sunday Times Oxford Literary Festival. Audio file at http://richarddawkins.net. De Waal, Frans. 1997. Good Natured: The Origin of Right and Wrong in Humans and Other Animals. Cambridge, MA: Harvard University Press. ———. 2006. Primates and Philosophers: How Morality Evolved. Princeton, NJ: Princeton University Press. Dembski, William. 1998. The Design Inference: Eliminating Chance Through Small Probabilities. New York: Cambridge University Press. ———. 2007. No Free Lunch: Why Specified Complexity Cannot be Purchased without Intelligence. 2nd edition. Lanham, MD: Rowman & Littlefield. Dennett, Daniel. 1984. Foreword to Ruth Garrett Millikan, Language, Thought, and Other Biological Categories: New Foundations for Realism. Cambridge, MA: MIT Press. ———. 1993. “Back from the Drawing Board.” In Dennett and His Critics, edited by Bo Dahlbom, 203–235. Cambridge, UK: Blackwell. ———. 1995. Darwin’s Dangerous Idea: Evolution and the Meanings of Life. New York: Simon and Schuster. ———. 1998. “Self-Portrait.” In his Brainchildren: Essays on Designing Minds, 355–66. Cambridge, MA: MIT Press. ———. 2017. From Bacteria to Bach and Back: The Evolution of Minds. New York: W. W. Norton and Company. Descartes, René. 1993. Meditations on First Philosophy. 3rd edition. Donald A. Cress, trans. Indianapolis, IN: Hackett Publishing Company.
318
Bibliography
Dewey, John. 1910. “The Influence of Darwinism on Philosophy.” In his The Influence of Darwin on Philosophy and Other Essays in Contemporary Thought, 1–19. New York: Henry Holt and Company. Dick, Steven J., ed. 2000. Many Worlds: The New Universe, Extraterrestrial Life and the Theological Implications. Philadelphia: Templeton Foundation Press. Diggins, Frank. 2003. “The True History of the Discovery of Penicillin by Alexander Fleming.” Biomedical Science. 246–249. Drees, Willem B. 1996. Religion, Science and Naturalism. Cambridge, UK: Cambridge University Press. Einstein, Albert. 1936. “Physik und Realität.” Journal of the Franklin Institute 221:313–47. Enns, Peter. 2003. Inspiration and Incarnation: Evangelicals and the Old Testament. Grand Rapids, MI: Baker Academic. Evans, Gareth. 1978. “Can There Be Vague Objects?” Analysis 38:208. Flanagan, Owen. 2002. The Problem of the Soul: Two Visions of Mind and How to Reconcile Them. New York: Basic Books. Fodor, Jerry. 1992. The Big Idea: Can There Be A Science of Mind? Times Literary Supplement. 3 July. 5–8. Frankfurt, Harry G. 1988. “Alternate Possibilities and Moral Responsibility.” In his The Importance of What We Care About: Philosophical Essays, 1–10. Cambridge, UK: Cambridge University Press. ———. 1999. “On God’s Creation.” In his Necessity, Volition, and Love, 117–28. Cambridge, UK: Cambridge University Press. ———. 2006. Taking Ourselves Seriously and Getting It Right. Redwood City, CA: Stanford University Press. Frankish, Keith, ed. 2017. Illusionism as a Theory of Consciousness. Exeter, UK. Academic Imprints. Frost, David, and Fred Bauer. 1997. Billy Graham: Personal Thoughts of a Public Man. Colorado Springs, CO: David C. Cook. Kingsley, Charles. 1885. The Water Babies: A Fairy Tale for a Land-Baby. London: Macmillan. Frymer-Kensky, Tikva. 1992. In the Wake of the Goddesses: Women, Culture, and the Biblical Transformation of Pagan Myth. New York: Free Press. Gazzaniga, Michael S. 2009. Human: The Science Behind What Makes Us Unique. New York: Harper Collins eBooks. Ghiselin, Michael T. 1969. The Triumph of the Darwinian Method. Berkeley: University of California Press. Giberson, Karl W. 2008. Saving Darwin: How to be a Christian and Believe in Evolution. New York: Harper Collins. Gigerenzer, Gerd. 2010. Rationality for Mortals: How People Cope with Uncertainty. Oxford, UK: Oxford University Press. Gilkey, Langdon. 1959. Maker of Heaven and Earth: The Christian Doctrine of Creation in the Light of Modern Knowledge. New York: Doubleday. ———. 1961. “Cosmology, Ontology, and the Travail of Biblical Language.” Journal of Religion 41:194–205. Gödel, Kurt. 1992. On Formally Undecidable Propositions of Principia Mathematica and Related Systems. New York: Dover. Goetz, Stewart, and Charles Taliaferro. 2008. Naturalism. Grand Rapids, MI: Eerdmans. Goldman, Alvin I. 2012. Reliabilism and Contemporary Epistemology: Essays. Oxford, UK: Oxford University Press. Gopnik, Alison, and Andrew N. Meltzoff. 1997. Words, Thoughts and Theories. Cambridge, MA: MIT Press. Gould, Stephen J. 1989. Wonderful Life: The Burgess Shale and the Nature of History. New York: W.W. Norton and Company. ———. 1999. Rocks of Ages: Science and Religion in the Fullness of Life. New York: Ballantine Publishing Group.
Bibliography
319
Graziano, Michael. 2013. Consciousness and the Social Brain. Oxford, UK: Oxford University Press. ———. 2019. Rethinking Consciousness: A Scientific Theory of Subjective Experience. New York: W. W. Norton and Company. Gregory of Nyssa. 1978. The Life of Moses. Abraham J. Malherbe and Everett Ferguson, trans. New York: Paulist Press. Green, Joel B. 2008. Body, Soul, and Human Life: The Nature of Humanity in the Bible. Grand Rapids, MI: Baker Academic. Gunton, Colin. 1992. Christ and Creation. Grand Rapids, MI: Eerdmans. ———. 1993. The One, the Three and the Many: God, Creation and the Culture of Modernity. Cambridge, UK: Cambridge University Press. ———. 1998. The Triune Creator: A Historical and Systematic Study. Grand Rapids, MI: Eerdmans. Haldane, J. B. S. 1928. Possible Worlds and Other Papers. New York: Harpers. Hardwick, Charley D. 1996. Events of Grace: Naturalism, Existentialism and Theology. Cambridge, UK: Cambridge University Press. Hare, John. 2004. “Is There an Evolutionary Foundation for Human Morality?” In Evolution and Ethics: Human Morality in Biological and Religious Perspective, edited by Philip Clayton and Jeffrey Schloss, 187–203. Grand Rapids, MI: Eerdmans. Harris, Sam. 2004. The End of Faith: Religion, Terror and the Future of Reason. New York: W. W. Norton and Company. ———. 2011. The Moral Landscape: How Science Can Determine Human Values. New York: Free Press. Hasker, William. 2013. Metaphysics and the Tri-Personal God. Oxford, UK: Oxford University Press. Hazony, Yoram. 2012. The Philosophy of Hebrew Scripture. Cambridge, UK: Cambridge University Press. Hefner, Philip. 1993. The Human Factor: Evolution, Culture and Religion. Minneapolis: Fortress Press. Heidegger, Martin. 1962. Being and Time. 7th edition. John Macquarrie and Edward Robinson, trans. New York: Harper and Row. Helm, Paul. 2004. “Natural Theology and the Sensus Divinitatis.” In his John Calvin’s Ideas. Oxford, UK: Oxford University Press. 209–45. Herrmann, Robert L., and John Marks Templeton. 1989. The God Who Would Be Known: Revelations of the Divine in Contemporary Science. Philadelphia: Templeton Foundation Press. ———. 1994. Is God the Only Reality? Science Points to a Deeper Meaning of the Universe. New York: Continuum. Hick, John. 1976. Death and Eternal Life. New York: Harper and Row. Hodge, Charles. 1874. What Is Darwinism? New York: Scribner Armstrong and Company. Hume, David. 1993. An Enquiry Concerning Human Understanding. 2nd edition. Eric Steinberg, ed. Indianapolis, IN: Hackett Publishing Company. ———. 2000. Treatise of Human Nature. David Fate Norton and Mary L. Norton, eds. Oxford, UK: Oxford University Press. Hunter, Cornelius G. 2001. Darwin’s God: Evolution and the Problem of Evil. Grand Rapids, MI: Brazos Books. Hurtado, Larry W. 2016. Destroyer of the Gods: Early Christian Distinctiveness in the Roman World. Waco, TX: Baylor University Press. Hutchins, Edwin. 1995. Cognition in the Wild. Cambridge, MA: MIT Press. Jackson, Frank. 2003. “Mind and Illusion.” In Minds and Persons, edited by Anthony O’Hear, 261–263. Cambridge, UK: Cambridge University Press. James, William. 1897. “The Will to Believe.” In The Will to Believe and Other Essays in Popular Philosophy. New York: Longmans, Green, and Company. Jenson, Robert W. 1997. Systematic Theology Volume I: The Triune God. New York: Oxford University Press.
320
Bibliography
John Paul II. 1996. “Message to the Pontifical Academy of Sciences: On Evolution.” 22 October. Johnson, Phillip. 1995. Reason in the Balance: The Case against Naturalism in Science, Law and Education. Downers Grove, IL: InterVarsity Press. ———. 1997. An Easy-To-Understand Guide for Defeating Darwinism by Opening Minds. Downers Grove, IL: InterVarsity Press. Johnston, Mark. 2009. Saving God: Religion After Idolatry. Princeton, NJ: Princeton University Press. Jónsson, Gunnlaugur A. 1988. The Image of God: Genesis 1:26–28 in a Century of Old Testament Research. Translated by Lorraine Svendsen. Stockholm: Almquist and Wiksell International. Joyce, Richard. 2007. The Myth of Morality. Cambridge, UK: Cambridge University Press. Julian of Norwich. n.d. Revelations of Divine Love. Grace Warrack, trans. Grand Rapids, MI: Christian Classics Ethereal Library. Kahneman, Daniel. 2010. Thinking, Fast and Slow. New York: Farrar, Straus and Giroux. Kahneman, Daniel, Paul Slovic, and Amos Tversky, eds. 1982. Judgment under Uncertainty: Heuristics and Risks. Cambridge, UK: Cambridge University Press. Kane, Robert. 1998. The Significance of Free Will. Oxford, UK: Oxford University Press. Kass, Leon R. 2006. The Beginning of Wisdom: Reading Genesis. Chicago: University of Chicago Press. Kaufmann, Stuart. 1996. At Home in the Universe: The Search for the Laws of Self-Organization and Complexity. New York: Oxford University Press. Keely, Brian L. “Of Conspiracy Theories.” Journal of Philosophy 96:109–126 Keleman, Deborah. 2004. “Are Children ‘Intuitive Theists?’ Reasoning about Purpose and Design in Nature.” Psychological Science 15:295–301 Kim, Jaegwon. 2004. Physicalism, Or Something Near Enough. Princeton, NJ: Princeton University Press. Kitcher, Philip. 2014. The Ethical Project. Cambridge, MA: Harvard University Press. Kornblith, Hilary. 2002. Knowledge and Its Place in Nature. Oxford, UK: Clarendon. Kripke, Saul. 1980. Naming and Necessity. Princeton, NJ: Princeton University Press. Lacey, Alan. 1995. Naturalism. Oxford Companion to Philosophy. Ted Honderich, ed. Oxford, UK: Oxford University Press. 604–606. Levin, Michael. 1998. “How Philosophical Errors Impede Freedom.” Journal of Libertarian Studies. 14:125–34. Lewis, C. S. 1947. Miracles: A Preliminary Study. New York: Macmillan. Lewis, David. 1978. “Survival and Identity.” In The Identities of Persons, edited by Amélie Oksenberg Rorty, 17–40. Berkeley: University of California Press. ———. 1986. On the Plurality of Worlds. Oxford, UK: Basil Blackwell. Libet, Benjamin. 2004. Mind Time: The Temporal Factor of Consciousness. Cambridge, MA: Harvard University Press. Liebenberg, Louis. 1990. The Art of Tracking: The Origin of Science. Claremont, South Africa: David Philip. Lipton, Peter. 2004. Inference to the Best Explanation. 2nd edition. London: Routledge. Livingstone, David N. 1997. Darwin’s Forgotten Defenders: The Encounter between Evangelical Theology and Evolutionary Thought. Grand Rapids, MI: Eerdmans. Locke, John. 1979. An Essay Concerning Human Understanding. Revised edition. Peter H. Nidditch, ed. Oxford, UK: Clarendon. Lucas, J. R. 1961. “Minds, Machines, and Gödel.” Philosophy. 36:112–27. Luther, Martin. 1963. Luther’s Works. Volume 26. Translated and edited by Jaroslav Pelikan and Helmut Lehmann. Saint Louis, MO: Concordia Publishing House. ———. 1972. Luther’s Works. Volume 25. Hilton G. Oswald, ed. St. Louis, MO: Concordia Publishing House. Lucas, J. R. 1961. “Minds, Machines, and Gödel.” Philosophy 36. 112–27. Mackie, J. L. 1977. Ethics: Inventing Right and Wrong. New York: Penguin.
Bibliography
321
———. 1982. The Miracle of Theism: Arguments for and against the Existence of God. Oxford, UK: Clarendon Press. Malle, Bertram F. 2006. How the Mind Explains Behavior: Folk Explanation, Meaning, and Social Interaction. Cambridge, MA: MIT Press. Marcus, Gary. 2008. Kludge: The Haphazard Construction of the Human Mind. Boston: Houghton Mifflin. Marr, David. 2010. Vision: A Computational Investigation into the Human Representation and Processing of Visual Information. Cambridge, MA: MIT Press. Marshall, Justin, and Johannes Oberwinkler. 1999. “Ultraviolet vision: The colourful world of the mantis shrimp.” Nature 401:873–74. Maxwell, James Clerk. 1969. Letter to Lewis Campbell, 7 March 1852. In The Life of James Clerk Maxwell by Lewis Campbell and William Garnet. New York: Johnson Reprints. 178. McCosh, James. 1890. The Method of Divine Government: Physical and Moral. 8th edition. New York: Robert Carter Brothers. McDowell, John. 1998. “Two Sorts of Naturalism.” In his Mind, Value and Reality. Cambridge, MA: Harvard University Press. 176–97. McGrath, Alister. 2015. The Big Question: Why We Can’t Stop Talking about Science, Faith and God. New York: St. Martin’s Press. Mead, James K. 2007. Biblical Theology: Issues, Methods, and Themes. Louisville, KY: Westminster John Knox Press. Mercier, Hugo, and Dan Sperber. 2017. The Enigma of Reason. Cambridge, MA: Harvard University Press. Middleton, J. Richard. 2005. The Liberating Image: The Imago Dei in Genesis 1. Grand Rapids, MI: Brazos Press. Miller, William Ian. 1997. The Anatomy of Disgust. Cambridge, MA: Harvard University Press. Moltmann, Jürgen. 1993. God in Creation: A New Theology of Creation and the Spirit of God. Minneapolis: Fortress Press. Moreland, J. P., and Scott B. Rae. 2000. Body and Soul: Human Nature and the Crisis in Ethics. Downers Grove, IL: InterVarsity Press. Murphy, Nancey. 2009. Introduction and Overview. In Downward Causation and the Neurobiology of Free Will, edited by Nancey Murphy, George Ellis, and Timothy O’Connor, 1–28. Berlin: Springer. Murphy, Nancey, and Warren S. Brown. 2007. Did My Neurons Make Me Do It? Philosophical and Neurobiological Perspectives on Moral Responsibility and Free Will. Oxford, UK: Oxford University Press. Murray, Michael J. 2009. “Scientific Explanations of Religion and the Justification of Religious Belief.” In The Believing Primate: Scientific, Philosophical, and Theological Reflections on the Origin of Religion, edited by Jeffrey Schloss and Michael Murray, 168–78. Oxford, UK: Oxford University Press. Nagel, Thomas. 1979. “What is it Like to be a Bat?” In his Mortal Questions, 165–80. Cambridge, UK: Cambridge University Press. ———. 1986. The View from Nowhere. New York: Oxford University Press. ———. 2012. Mind and Cosmos. Oxford, UK: Oxford University Press. National Center for Science Education. 1995. Reports of the National Center for Science Education. 17:31–32 Neander, Karen. 2017. A Mark of the Mental: In Defense of Informational Teleosemantics. Cambridge, MA: MIT Press. Nielsen, Kai. 1996. Naturalism Without Foundations. New York: Prometheus Books. Nietzsche, Friedrich. 1966. Beyond Good and Evil: Prelude to a Philosophy of the Future. Translated by Walter Kaufmann. New York: Vintage Books. ———. 1974. The Gay Science. Walter Kaufmann, trans. New York: Vintage Books. Noë, Alva. 2015. Strange Tools: Art and Human Nature. New York: Hill and Wang. Nozick, Robert. 1981. Philosophical Explanations. Cambridge, MA: Harvard University Press.
322
Bibliography
Nussbaum, Martha. 2001. Upheavals of Thought: The Intelligence of Emotions. Cambridge, UK: Cambridge University Press. O’Connor, Timothy. 2008. Theism and Ultimate Explanation: The Necessary Shape of Contingency. Malden, MA: Blackwell. Papineau, David. 2002. Thinking About Consciousness. Oxford, UK: Oxford University Press. Parfit, Derek. 1984. Reasons and Persons. Oxford, UK: Oxford University Press. ———. 2013. On What Matters, Volume 1. Oxford, UK: Oxford University Press. Parsons, Terrence. 2000. Indeterminate Identity. Oxford, UK: Oxford University Press. Peacocke, Arthur. 1985. “A Christian ‘Materialism?’” In How We Know. Michael Shafto, ed. San Francisco: Harper and Row. 146–68 ———. 1993. Theology for a Scientific Age: Being and Becoming—Natural, Human, and Divine. Minneapolis: Fortress Press. ———. 2000. “The Challenge and Stimulus of the Epic of Evolution to Theology.” In Many Worlds: The New Universe, Extraterrestrial Life and the Theological Implications, edited by Steven J. Dick, 89–118. Philadelphia: Templeton Foundation Press. ———. 2004. “Biological Evolution and Christian Theology—Yesterday and Today.” In his Evolution: The Disguised Friend of Faith? Selected Essays, 22–49. Philadelphia: Templeton Foundation Press. ______. 2004. “Chance, Potentiality and God.” In his Evolution: The Disguised Friend of Faith? Selected Essays, 50–64. Philadelphia: Templeton Foundation Press. ______. 2007. All That Is: A Naturalistic Faith for the Twenty-First Century. Minneapolis: Fortress Press. Pereboom, Derk. 2014. Free Will, Agency, and Meaning in Life. Oxford, UK: Oxford University Press. Perry, John. 2001. Knowledge, Possibility, and Consciousness. Cambridge, MA: MIT Press. Peters, Ted. 1992. God−The World’s Future. Minneapolis: Fortress Press. Peters, Ted, ed. 1989. Cosmos and Creation: Theology and Science in Consonance. Nashville, TN: Abingdon Press. Peters, Ted and Martinez Hewlett. 2003. Evolution from Creation to New Creation: Conflict, Conversation, and Convergence. Nashville, TN: Abingdon Press. Petzold, Charles. 2008. The Annotated Turing: A Guided Tour Through Alan Turing’s Historic Paper on Computability and the Turing Machine. Indianapolis, IN: John Wiley and Sons. Pinker, Steven. 1997. How the Mind Works. New York: W. W. Norton and Company. ———. 2002. The Blank Slate: The Modern Denial of Human Nature. New York: Viking. Pizarro, David, Yoel Inbar, and Chelsea Helion. 2011. “On Disgust and Moral Judgment.” Emotion Review. 3:267–68 Plantinga, Alvin. 2001. “Methodological Naturalism?” In Intelligent Design Creationism and Its Critics: Philosophical, Theological, and Scientific Perspectives, edited by Robert T. Pennock, 339–62. Cambridge, MA: MIT Press. ———. 2001. “When Faith and Reason Clash.” In Intelligent Design Creationism and Its Critics: Philosophical, Theological, and Scientific Perspectives, edited by Robert T. Pennock, 113–46. Cambridge, MA: MIT Press. ———. 2002. “Introduction: The Evolutionary Argument Against Naturalism.” In Naturalism Defeated? Essays on Plantinga’s Evolutionary Argument against Naturalism, edited by James Beilby, 1–15. Ithaca, NY: Cornell University Press. ———. 2002. “Reply to Beilby’s Cohorts.” In Naturalism Defeated? Essays on Plantinga’s Evolutionary Argument against Naturalism. James Beilby, ed. Ithaca, NY: Cornell University Press. 258–65 ———. 2011. Where the Conflict Really Lies: Science, Religion, and Naturalism. Oxford, UK: Oxford University Press. Plantinga, Alvin, and Michael Toole, 2008. Knowledge of God. Malden, MA: Blackwell. Plato. Phaedo. 2009. David Gallop, trans. Oxford, UK: Oxford University Press. Pohier, Jacques. 1985. God-In Fragments. London: SCM Press.
Bibliography
323
Polkinghorne, John. 2002. “Eschatological Credibility: Emergent and Teleological Processes.” In Resurrection: Theological and Scientific Assessments, edited by Ted Peters, Robert John Russell and Michael Welker, 51–55. Grand Rapids, MI: Eerdmans. ———. 2004. Science and the Trinity: The Christian Encounter with Reality. New Haven, CT: Yale University Press. Pollard, William G. 1958. Chance and Providence: How God Acts in a World Governed by Scientific Law. New York: Charles Scribner’s Sons. Powell, Russell, and Steve Clarke. 2012. “Religion as an Evolutionary Byproduct: A Critique of the Standard Model.” British Journal for the Philosophy of Science 63:457–86. Prinz, Jesse. 2014. Beyond Human Nature: How Culture and Experience Shape the Human Mind. New York: W.W. Norton and Company. Provine, William. 1987. Review of Edward J. Larson, Trial and Error: The American Controversy over Creation and Evolution. In Academe. January/February 1987. 51–52. Putnam, Hilary. 1981. Reason, Truth and History. Cambridge, UK: Cambridge University Press. Pyysiäinen, Ilkkla. 2004. Magic, Miracles, and Religion: A Scientist’s Perspective. Walnut Creek, CA: AltaMira Press. Quine, W. V. O. 1960. Word and Object. Cambridge, MA: MIT Press. ———. 1995. “Naturalism; Or, Living within One’s Means.” Dialectica. 49:251–61. Ratzsch, Del. 2000. Science and Its Limits: The Natural Sciences in Christian Perspective. 2nd ed. Downers Grove, IL: InterVarsity Press. Reppert, Victor. 2003. C. S. Lewis’s Dangerous Idea: In Defense of the Reason Argument. Downers Grove, IL: InterVarsity Press. Ridley, Matt. 2010. The Rational Optimist: How Prosperity Evolves. New York: Harper. Rosenberg, Alex. 2011. The Atheist’s Guide to Reality: Enjoying Life without Illusions. New York: W.W. Norton and Company. ———. 2019. How History Gets Things Wrong: The Neuroscience of Our Addiction to Stories. Cambridge, MA: MIT Press. Roughgarden, Joan. 2006. Evolution and Christian Faith: Reflections of an Evolutionary Biologist: Washington, DC: Island Press. Rouse, Joseph. 2018. “Naturecultural Inferentialism.” In From Rules to Meanings: New Essays on Inferentialism, edited by Ondřej Beran, Vojtěch Kolman, and Ladislav Koreň, 239–48. New York: Routledge. Rubin, Paul H. 2002. Darwinian Politics: The Evolutionary Origin of Freedom. Piscataway, NJ: Rutgers University Press. Ruse, Michael. 1989. “Is Rape Wrong on Andromeda? An Introduction to Extraterrestrial Evolution, Science, and Morality.” In his The Darwinian Paradigm: Essays on Its History, Philosophy and Religious Implications, 209–45. London: Routledge. ———. 2003. Darwin and Design: Does Evolution Have A Purpose? Cambridge, MA: Harvard University Press. Rushdie, Salman. 1991. “Imaginary Homelands.” In his Imaginary Homelands: Essays and Criticism 1981–1991. London: Granta Books. Russell, Bertrand. 1935. Religion and Science. London: Oxford University Press. Russell, Robert J. 1989. “Cosmology, Creation, and Contingency.” In Cosmos as Creation: Theology and Science in Consonance, edited by Ted Peters, 177–209. Nashville: Abingdon Press. ———. 2008. “Special Providence and Genetic Mutation: A New Defense of Theistic Evolution.” In his Cosmology from Alpha to Omega: The Creative Mutual Interaction of Theology and Science, 212–25. Minneapolis: Fortress Press. Sagan, Carl. 1995. Pale Blue Dot: A Vision of the Human Future in Space. New York: Random House. Salmon, Nathan. 2005. “Identity Facts.” In his Metaphysics, Mathematics, and Meaning, Philosophical Papers, Volume I. Oxford, UK: Oxford University Press. Salmon, Wesley C. 1984. Scientific Explanation and the Causal Structure of the World. Princeton, NJ: Princeton University Press.
324
Bibliography
Sandel, Michael J. 2007. The Case Against Perfection: Ethics in the Age of Genetic Engineering. Cambridge, MA: Harvard University Press. Sanders, John. 1998. The God Who Risks: A Theology of Providence. Downers Grove, IL: InterVarsity Press. Santayana, George. 1955. Scepticism and Animal Faith: Introduction to a System of Philosophy. New York: Dover Publications. Sarkar, Sahotra. 2007. Doubting Darwin? Creationist Designs on Evolution. Malden, MA: Blackwell. Schacter, Daniel L. 2002. The Seven Sins of Memory: How the Mind Forgets and Remembers. New York: Mariner Books. Schopf, J. William, editor. 2002. Life’s Origin: The Beginning of Biological Evolution. Berkeley: University of California Press. Scruton, Roger. 1994. Modern Philosophy: An Introduction and Survey. New York: Penguin Books. ———. 2012. The Face of God. New York: Continuum. Searle, John. 1983. Minds, Brains and Science. Cambridge, MA: Harvard University Press. Sellars, Wilfrid. 1963. “Philosophy and the Scientific Image of Man.” In his Science, Perception and Reality, 1–40. London: Routledge and Kegan Paul. ———. 1997. Empiricism and the Philosophy of Mind. Cambridge, MA: Harvard University Press. Sesardic, Nevin. 1995. “Recent Work in Human Altruism and Evolution.” Ethics 106: 328–57. Skinner, B. F. 1971. Beyond Freedom and Dignity. New York: Knopf. Smith, James K. A. 2017. “What Stands on the Fall?” In Evolution and the Fall, edited by William T. Cavanaugh and James K. A. Smith, 61–63. Grand Rapids, MI: Eerdmans. Smolin, Lee. 2013. Time Reborn: From the Crisis in Physics to the Future of the Universe. Boston: Houghton Mifflin Harcourt. Stanovich, Keith E. 2005. The Robot’s Rebellion: Finding Meaning in the Age of Darwin. Chicago: University of Chicago Press. Stenger, Victor. 2014. God and the Multiverse: Humanity’s Expanding View of the Cosmos. Amherst, NY: Prometheus Books. Sterelny, Kim. 2003. Thought in a Hostile World: The Evolution of Human Cognition. Oxford, UK: Blackwell. ———. 2012. The Evolved Apprentice. Cambridge, MA: MIT Press. Strawson, Galen. 1994. “The Impossibility of Moral Responsibility.” Philosophical Studies 75:5–24. Thorson, Walter D. 2014. The Woodpecker’s Purpose: A Critique of Intelligent Design. Wenham, MA: Gordon College Center for Faith and Inquiry. Trivers, Robert. 2002. “Reciprocal Altruism.” In his Natural Selection and Social Theory: Selected Papers of Robert Triver, 3–55. Oxford, UK: Oxford University Press. Tye, Michael. 2000. Consciousness, Color, and Content. Cambridge, MA: MIT Press. van Inwagen, Peter. 1978. “The Possibility of Resurrection.” International Journal for the Philosophy of Religion. 9:114–121. ———. 1983. An Essay on Free Will. Oxford, UK: Oxford University Press. ———. 1988. “And Yet They Are Not Three Gods But One God.” In Philosophy and the Christian Faith, edited by Thomas V. Morris, 241–78. Notre Dame: IN: University of Notre Dame Press. 241–78. ———. 1990. Material Beings. Ithaca, NY: Cornell University Press. ———. 2008. Metaphysics. 3rd edition. Boulder, CO: Westview Press. van Leeuween, Mary Stewart. 2001. “Of Hoggamus and Hogwash: Evolutionary Psychology and Gender Relations.” Journal of Psychology and Theology 30:101–111. von Rad, Gerhard. Genesis: A Commentary. John H. Marks, trans. Philadelphia: Westminster Press. Wacome, Donald. 2004. “Reductionism’s Demise: Cold Comfort.” Zygon 39:321–37. Ward, Keith. 1996. God, Chance and Necessity. Oxford, UK: Oneworld.
Bibliography
325
Ward, Peter, and Don Brownlee. 2003. Rare Earth: Why Complex Life is Uncommon in the Universe. New York: Springer. Webb, Stephen. 2002. If the Universe Is Teeming with Aliens…Where Is Everybody? Fifty Solutions to the Fermi Paradox and the Problem of Extraterrestrial Life. New York: Copernicus Books. Wegner, Daniel M. 2002. The Illusion of Conscious Will. Cambridge, MA: MIT Press. Wegner, Daniel M. and Kurt Gray. 2016. The Mind Club: Who Thinks, What Feels, and Why It Matters. New York: Viking Press. Westermann, Claus. 1988. Genesis. David E. Green, trans. Edinburgh: T. & T. Clark. Westphal, Merold. 1998. Suspicion and Faith: The Religious Uses of Modern Atheism. New York: Fordham University Press. Whitehouse, Harvey. 2004. Modes of Religiosity: A Cognitive Theory of Religious Transmission. Walnut Creek, CA: Altamira Press. Williams, Bernard. 1973. “Deciding to Believe.” In his Problems of the Self: Philosophical Papers 1956–1972, 136–51. Cambridge, UK: Cambridge University Press. ———. 1973. “The Makropulos Case: Reflections on the Tedium of Immortality.” In his Problems of the Self: Philosophical Papers 1956–1972, 82–100. Cambridge, UK: Cambridge University Press. ———. 1981. “Internal and External Reasons.” In his Moral Luck. Cambridge, UK: Cambridge University Press. 101–113. Williams, Margery. 1975. The Velveteen Rabbit: Or How Toys Become Real. New York: Avon Books. Williams, Rowan. 1990. The Wound of Knowledge: Christian Spirituality from the New Testament to Saint John of the Cross. 2nd edition. Cambridge, MA: Cowley Publications. ———. “On Being Creatures.” 2000. In his On Christian Theology, 61–78. Malden, MA: Wiley-Blackwell. Wilson, David Sloan. 2002. Darwin’s Cathedral: Evolution, Religion, and the Nature of Selection. Chicago: University of Chicago Press. Wilson, E. O. 1975. Sociobiology: The New Synthesis. Cambridge, MA: Harvard University Press. Wilson, Timothy D. 2004. Strangers to Ourselves: Discovering the Adaptive Unconscious. Cambridge, MA: Harvard University Press. Wolff, Hans Walter. 1996. Anthropology of the Old Testament. Margaret Kohl, trans. Mifflintown, PA: Sigler Press. Wright, N. T. 1996. Jesus and the Victory of God. Minneapolis: Fortress Press. ———. 2003. The Resurrection of the Son of God. Minneapolis: Fortress Press. Zizioulas, John D. 1993. Being As Communion: Studies in Personhood and the Church. Crestwood, NY: St. Valdimir’s Seminary Press.
Index
abductive reasoning, 75 agency detection, 241–242, 246 agent causation, 17, 71, 175–176, 178, 180–181, 184, 188 altruism, 196, 202–203, 204, 205, 208, 211–212, 233n3, 234n6, 256; indiscriminate, 200; reciprocal, 197, 198–199; trans-kin, 197 amoralism, 217 angels, 274n66 antecedent (prior) probability, 34–35, 36–37, 46n10, 84, 86, 89, 92, 93, 96, 103, 107, 110, 123, 124, 202 anthropic selection, 68 Aquinas, Saint Thomas, 69, 119n36 Asimov, Isaac, 105 Atran, Scott, 243 Augustine, Saint, 282, 283, 308n7 Barrett, Justin, 241, 244, 248, 251, 252 Barth, Karl, 108, 265–266, 274n56, 306 biblical interpretation, 20n22, 32–33, 45n8–45n9, 116n5, 129–132, 307n3 Bloom, Paul, 251 Bohr, Niels, 72 Boulton, Matthew Myer, 260, 266 Boyer, Pascal, 241, 246 Boyle, Robert, 32 Calvin, John, 107, 224, 252, 273n35 categorical imperatives, 211, 214, 217, 218 causal exclusion, 160n39 Chinese room, 151–152, 153 Chisholm, Roderick, 178 choosing to believe, 41–43, 310n33 Christian morality, 225 Churchland, Patricia, 170 cognitive theory of religion, 240–246, 247, 249–255, 262, 266, 267
cold fusion, 52 color vision, 154–156, 208–209, 223, 236n29 compatibilism, 11, 16, 164–170, 179, 183, 192n20 computational theory of mind, 11, 129, 136, 151–153, 240–241, 301–302 consciousness, 128–129, 134, 146–148, 300 continuing creation, 88 Cooper, John C., 268 cosmological fine-tuning, 68, 69, 119n43 Craig, Edward, 40–41 created co-creator, 306 creation science, 64 creationism, 46n14, 86 Darwin, Charles, 32, 64, 68, 86, 90, 92, 103, 109 Darwinism, 5, 10, 16, 83, 83–84, 88, 89–90, 91, 92, 101, 103, 106, 109, 110, 112 Dawkins, Richard, 13, 20n30, 76, 90, 104, 202, 258 debunking, 17, 69, 211–216, 226, 240, 246–251, 253, 255 deism, 8, 88, 89 demons, 274n66 Dennett, Daniel, 84, 123, 154, 311 desacralization, 313 Descartes, René, 38, 40, 158n20 design, xii, 13, 16, 68, 91–93, 98, 100–103, 106, 110, 124, 142, 184, 188, 251, 258, 311 determinism, 55, 61, 68, 71, 91, 92, 100, 101, 116n17, 174, 182, 184, 185–186, 222, 251, 254 Dewey, John, 88 disgust, 209–211, 235n20–235n21 327
328
Index
divine condescension, 17, 224, 228, 255, 257, 266, 312 divine conservation, 117n25 divine foreknowledge, xi, 61, 101, 103, 186, 191n16 divine intervention: in history, 15, 63, 65, 67, 83; in nature, 15, 63, 65 divine justice, xi, 94, 224–225, 228, 236n32, 263, 305 divine temporality, xi, 9, 61, 80n16, 191n16, 304 downward causation, 17, 174–175, 190n2 Drake equation, 119n41 dualism, 16, 83, 125, 126–130, 134, 142, 158n16, 163, 176, 277, 278, 280, 299–302 Einstein, Albert, 27, 135 emergence, 11, 16, 145–148, 157, 174, 175 empirical content, 13, 29–30, 36, 131 encapsulation, 141, 245–246 epiphenomenalism, 144, 160n38–160n39 error theory, 212, 217 ethical naturalism, 235n19 Everett, Hugh, 36 existential gap, 279, 280–284, 285, 287, 288, 289, 299, 302 extraterrestrial persons, 104–106, 119n41, 306 faith, 9, 28, 29–31, 53, 67, 76–77, 250, 310n33 Fall, the, 9, 17, 95, 194, 226–229, 236n34, 236n37, 260, 273n35, 274n47 falsifiability, 29–30, 46n13 Fermi paradox, 105 Fields, W. C., 11 first philosophy, 15, 38 fission of persons, 292–293, 294 Fodor, Jerry, 128, 129 folk-psychology, 80n11, 242, 243 foundationalism, 38–39 functionalism, 143, 145, 160n35 Gage, Phineas, 127
general resurrection, 269, 278, 280, 290, 295, 301, 307n3 Giberson, Karl, 110 Gilkey, Langdon, 67, 96 Gödel, Kurt, 136–137, 159n27, 309n24 Gould, Stephen J., 29, 118n32 Graham, Billy, 114 Gregory of Nyssa, 303 Gunton, Colin, 60, 96, 117n20 heaven, xi, 7, 246, 278, 280 Hefner, Philip, 306 hell, xi, 222, 246, 305 Hodge, Charles, 89 Hume, David, 51, 235n23, 272n27 imago Dei, xi, 2, 14, 28, 107, 125, 134, 156, 260 incarnation, 1, 3, 8–9, 19n6, 95, 118n33, 189, 230, 231, 266, 304, 307n2 incompatibilism, 165, 168–178, 187–189 indeterminacy: causal, 91, 102, 173–174, 175; ontological, 294–298 intelligent design, 64, 68, 80n20 intentionality, 151–156 interim state, 269, 279, 292, 307n3 internalism, 46n16 irreducibility, 16, 144–145, 146, 157 irreducible complexity, 80n20 Jenson, Robert, 7 John Paul II, Pope, 114 Johnson, Phillip, 88 Kabbalah, 103 Keleman, Deborah, 242 kenosis, 118n33 Lemaître, Georges, 36 Levin, Michael, 92 Lewis, C. S., 21n32, 135–136, 138, 159n22, 271n23 Lewis, David, 295, 297 libertarian free will, 71, 163, 175, 188, 310n33 Lucas, J. R., 137–138, 141, 309n24 Luther, Martin, 264–265, 266 Lyell, Charles, 64
Index macroevolution, 83 Marcus, Gary, 142 materialism, 21n31, 115, 123–127, 129, 133, 135–144, 145, 157, 163, 174, 277, 299, 300, 302; eliminative, 144; nonreductive, 11, 143, 144, 145 matter-first, 149 Maxwell, James Clerk, 312–313 McCosh, James, 109, 114 McGrath, Alister, 12 mere copy, 284, 286–287, 288, 289, 291 methodological naturalism, 5, 19n15, 64 microevolution, 83 Middleton, J. Richard, 259, 260 mind-first, 148, 150 minimally counterintuitive concepts, 243–244 modal intuitions, 134 modal properties, 74 modularity of mind, 141, 190n2, 240–243, 246, 267, 271n19 moral fictionalism, 218, 273n36 moral relativism, 221; individual (subjectivism), 17, 207–208; cultural (social), 17, 207–208; species, 207, 208, 210, 221 moral sentiments (emotions), 194, 200, 213–214, 217, 219, 220, 227, 244, 248, 255 multiple realizability, 80n12, 143 multiple universes, 68, 69, 70, 81n26 Murphy, Nancey, 174, 191n10 Nagel, Thomas, 44, 128 National Association of Biology Teachers, 110 naturalizing project, 11, 84, 156, 194, 239, 312 nephes, 130, 132 Nielsen, Kai, 4 Nietzsche, Fredrich, xii, 314 Nussbaum, Martha, 97 objective prescriptivity, 212, 214, 217 objectivity: scientific, 34; of morality, 17, 68, 218, 222, 223, 239, 312 omniscience, xi, 91, 97, 101, 182, 245, 282
329
open theism, 190n1 panentheism, 59, 60 pantheism, 59, 60 Parfit, Derek, 234n12, 293, 294 Paul, Saint, 130, 133, 289, 307n1, 307n3 Peacocke, Arthur, 23, 44, 55, 59, 117n20, 118n27, 144 perfect being theology, 94 personhood, 105, 107–108, 128, 134, 148, 157, 231, 234n10, 279 Peters, Ted, 4, 80n16, 118n35 physicalism, 21n31, 127, 134; nonreductive. See materialism Pinker, Steven, 203 Plantinga, Alvin, 4, 88, 110–111, 112 Polkinghorne, John, 19n1, 302 prisoner’s dilemma, 197–198 problem of evil, 314n2 proper basicality, 248, 251–253 Provine, William, 88 psychological continuity, 288, 292, 294, 301, 309n18 quantum mechanics, 17, 46n13, 66, 79n10, 92, 93, 101–102, 110, 117n18, 164, 172–174 Quine, W. V. O., 78 reductionism, 16, 144–145, 174 reliabilism, 38, 46n16 Rosenberg, Alex, 18 Roughgarden, Joan, 114 Russell, Bertrand, 25, 88 Russell, Robert John, 72, 79n10, 110, 120n60 Ryle, Gilbert, 142 sacrifice, 260, 261–264 Sagan, Carl, 3–4, 313 Sandel, Michael, 99 Sanders, John, 101 scientism, 25 Scruton, Roger, 154 Searle, John, 151–153 secondary causation, 50, 55, 56, 60–62, 87–89, 90, 92, 109, 113–114, 177 selfish gene, 202–203, 204, 213, 215, 234n13
330
Index
semantic indecision, 295, 297 sensus Divinitatis, 252, 254, 273n35 Sesardic, Nevin, 196 sheol, 268–269 Shields, Charles Woodruff, 106 sin, 140, 222, 227–228, 230, 233, 264, 265; original, 17, 227, 230, 232, 274n61 Skinner, B. F., 139 special creation, 78, 83, 111, 113, 132, 228 spiritual beings, 16, 123, 133, 289, 301, 307n1 Strawson, Galen, 177 superficiality: of moral reality, 2, 206–211, 222, 255; ontological, 2, 16, 80n16, 145, 148–151, 154–156, 160n39, 209, 210, 292, 294–297, 298, 313 supervenience, 58, 288
theistic evolution, 68, 80n21, 91, 93, 103, 120n60 theory of mind, 190n3, 242, 244, 245, 267 tit-for-tat, 198–199 token identity, 160n39 top-down causation. See downward causation transitivity paradox, 289–292, 293, 294, 299, 309n17 Turing, Alan, 136 van Inwagen, Peter, 95, 308n9 view from nowhere, 44, 210 vitalism, 31, 132 Ward, Keith, 59 Wegener, Alfred, 36 Williams, Bernard, 303 Williams, Marjory, 156 Williams, Rowan, 9, 14
teleology, 103, 119n36, 242 zombies, 134, 147
About the Author
Donald H. Wacome is professor of philosophy at Northwestern College. He is an active lay member of Church of the Savior (Episcopalian) in Orange City, Iowa.
331