Anda di halaman 1dari 22

HeyJ XLIX (2008), pp.

582602

DIVINE AGENCY, CONTEMPORARY PHYSICS, AND THE AUTONOMY OF NATURE


WILLIAM E. CARROLL

Thomas Aquinas Fellow in Theology and Science, Blackfriars University of Oxford

Over the past fteen years the Vatican Observatory and the Center for Theology and the Natural Sciences in Berkeley, California sponsored a series of conferences on what they called scientic perspectives on divine action, which resulted in the publication of ve impressive volumes,1 with a sixth, retrospective volume, to appear soon. The scientic subjects of these books ranged from quantum mechanics and quantum cosmology to chaos theory, evolutionary biology, and the neurosciences. What emerged as a major theme in the contributions of many of the scholars in these volumes is how contemporary science points to a kind of metaphysical space which can allow for divine agency in the world.2 Thus, for example, the fascination with quantum mechanics and chaos theory, since each has been viewed as providing a kind of metaphysical indeterminacy needed to provide an arena in which God can act without somehow interfering with the laws of nature. Another feature of many of the essays is the view that contemporary science lends support to a process theology which challenges traditional notions of divine omnipotence, immutability, and a-temporality. Divine agency is crucial for believers since, not only has God caused all that is to be at the very beginning of time, the universe also depends on Gods continuing creative act at every moment of its existence. Furthermore, God is providential: He guides and directs the universe towards the fullling of the purposes He has established. In addition to such a general providence, God acts in particular and special ways, either in history or in individual lives. For many theologians there is a real urgency in seeking to understand divine agency in the context of what the natural sciences tell us about the world; the task is especially important since, according to these theologians, modern science that is, science since the seventeenth century has presented theology with a tremendous challenge: to nd a place for God in a universe increasingly susceptible to explanation in scientic terms. A widely accepted view is that, with the rise of modern science, God has increasingly been pushed to the margins.
r The author 2008. Journal compilation r Trustees for Roman Catholic Purposes Registered 2008. Published by Blackwell Publishing Ltd, 9600 Garsington Road, Oxford OX4 2DQ, UK and 350 Main Street, Malden, MA 02148, USA.

DIVINE AGENCY, CONTEMPORARY PHYSICS

583

A German chemist, Heinrich Caro, writing at the end of the nineteenth century, remarked: Science has conducted God to its frontiers, thanking him for his provisional services.3 What many theologians nd so fascinating with contemporary science is a view of the universe fundamentally different from a causally closed system operating according to the prescriptions of Newtonian mechanics. So long as the universe was seen in deterministic Newtonian terms, it was easy, so such an interpretation suggests, either to see God as interfering with the laws of nature or to limit divine action to an initial act of creation. Even the notion of Gods continual causing of the existence of things was thought to be challenged by the principle of inertia, which seemed to entail the view that the universe was self-sufcient and thus required no appeal outside of itself, once it existed, to account for all motion and change.4 Wolfhart Pannenberg, for example, claims that the acceptance of the principle of inertia in the seventeenth century represents not only a radical break in the history of science5 but also posed a fundamental problem for the Christian doctrine of creation. According to Pannenberg, the principle of inertia lies behind the denial of the radical contingency of the world, a contingency central to the doctrine of creation: The emancipation from the creator God entailed in the principle of inertia did not apply only to natural bodies and beings . . . . Even more serious was the consequence that the system of the natural universe had to be conceived now as an interplay between nite bodies and forces without further need for recourse to God.6 For Pannenberg and others, the deterministic view of the universe, made famous by Pierre Laplace (17491827),7 has been overturned by quantum indeterminism, and, as a result, God could be thought of as acting at the quantum level. The most widely accepted interpretation of quantum mechanics afrms that indeterminism is a universal feature of the microworld,8 and that the probabilistic structure in quantum theory entails [a] radical departure from the philosophical position of classical physics.9 The commitment to ontological indeterminism, associated with many philosophical interpretations of quantum mechanics, has offered a way for some to reafrm Gods acting in nature. In an important sense, Gods action, so conceived, is not an intervention in nature, since nature is fundamentally indeterministic. Robert Russell, Director of the Center for Theology and the Natural Sciences at Berkeley, nds such noninterventionist divine action attractive because he thinks it is a way of maintaining continuing divine agency without sacricing the integrity of nature.10 Writing in the recently published Oxford Handbook of Religion and Science (2006), Russell notes the importance of what he terms NIODA [Non-interventionist objective divine action], the thesis that God acts objectively and directly in and through (mediated by) quantum events to actualize one of several potential outcomes; in short, the

584

WILLIAM E. CARROLL

collapse of the wave function occurs because of divine and natural causality working together even while Gods action remains ontologically different from natural agency.11 The universe which God creates is such that some natural processes at the quantum level are insufciently determined by prior natural events. As a result, Gods action is required to bring about the quantum event. Another contemporary theologian who has written extensively on what he considers to be the positive theological implications of quantum mechanics is Thomas Tracy. Tracy says that we should think of special divine action that is, the production of particular events in the world as the providential determination of otherwise undetermined events. Gods governance at the quantum level consists in activating or actualizing one or another of the quantum entitys innate powers at particular instants. Gods action is pervasive in its effects on the worlds structure, but it remains hidden within that structure. Gods action takes the form of realizing one of several potentials in the quantum system: God does not manipulate subatomic particles as though He were some type of physical force. Although God acts, He does not interfere.12 In the midst of considerable enthusiasm for the theological implications of quantum mechanics, we need to remember that the philosophical interpretations which embrace an ontological indeterminacy are just that philosophical interpretations, which continue to be the subject of lively debate. What is evident, however, is the urgency which many thinkers claim is needed in addressing the question of divine agency and modern science: to nd adequate ways to account for Gods activity in the world without denying the appropriate autonomy of natural processes nor reducing what we mean by Gods activity merely to the subjective religious experience of believers. One scholar who has written extensively on this topic, Philip Clayton, thinks that how to understand divine agency ought to be at the center of contemporary theological reection. As he puts it: how can we attribute events to the causal activity of God when science appears to be able fully to explain what happens in the world? John Polkinghorne, although critical of the appeal to quantum indeterminacy as a way to make room for divine action in the world, does think that contemporary chaos theory offers a fruitful avenue for theological reection. The most obvious thing to say about chaotic systems is that they are intrinsically unpredictable. Their exquisite sensitivity [to slight changes in initial conditions] means that we can never know enough to be able to predict with any long-term reliability how they will behave. Polkinghorne argues that the epistemological limitations which chaos theory presents point to a fundamental feature of the world, what he calls an ontological openness.
I want to say that the physical world is open in its process, that the future is not just a tautologous spelling-out of what was already implicit in the past, but

DIVINE AGENCY, CONTEMPORARY PHYSICS

585

there is genuine novelty, genuine becoming, in the history of the universe . . . . The dead hand of the Laplacean Calculator is relaxed and there is scope for forms of causality other than the energetic transactions of current physical theory. As we shall see there is room for the operation of holistic organizing principles (presently unknown to us, but in principle open to scientic discernment), for human intentionality, and for divine providential interaction.13

For Polkinghorne, chaos theory offers us a metaphysically attractive option of openness, a causal grid from below which delineates an envelope of possibility . . . within which there remains room for manoeuvre.14 This room for manoeuvre, can be seen, according to Polkinghorne, in the act of creation itself, understood as: . . . a kenosis (emptying) of divine omnipotence, which allows for something other than God to exist . . .15 According to Polkinghorne: The act of creation involves a voluntary limitation, not only of divine power in allowing the other to be, but also of divine knowledge in allowing the future to be open . . . . An evolutionary universe is to be understood theologically as one that is allowed by God, within certain limits, to make itself by exploring and realizing its own inherent fruitfulness. The gift of creaturely freedom is costly, for it carries with it the precariousness inherent in the self-restriction of divine control.16 I think that Polkinghorne moves far too easily from claims in epistemology to claims in metaphysics. Various attempts by Polkinghorne, Arthur Peacocke, Nancey Murphy, George Ellis, and others to locate a venue for divine agency in the indeterminism of contemporary physics really amount to the contention that any account of the physical world in the natural sciences is somehow inherently incomplete. In other words, these authors must maintain that the natural sciences cannot in principle provide a complete, coherent scientic account of physical reality.17 Nicholas Saunders, who has written an extensive analysis of the claims of theologians and philosophers who use developments in contemporary science to explain divine action, is very critical of their fascination with quantum mechanics. There is, according to Saunders, no credible account of divine action in the context of contemporary science and the result is a major crisis for Christian theology in its attempt to afrm Gods special providential activity. Saunders shares with those whom he criticizes the assumption that Gods agency needs to be understood in new ways; he differs, however, in that he thinks that there is as yet no adequate account available.18 All these scholars tend to share the distinction between what they call general divine providence (Gods creating and sustaining all of reality) and special divine providence (specic providential acts, envisaged, intended, and somehow brought about in this world by God). They nd the former category, God as cause of existence, to be far less troublesome than the latter category, special acts by God which seem to be inconsistent with the autonomy and integrity of nature. As we have seen, the indeterminism in the quantum realm has proven to be especially

586

WILLIAM E. CARROLL

attractive to those who wish to develop a theory of non-interventionist divine action. In many ways, however, the fundamental problem concerns how to understand divine causality in its broadest sense, and, as we shall see, by examining what it means for God to be cause of existence we can discover a rich understanding of cause, useful for the theological project of understanding divine agency. It is one of the benets of Nicholas Saunders work that it examines in some detail the philosophical and scientic foundations of the arguments of those who are attracted to what is called quantum special divine providence. The debate is highly complex, and I do not want to address the multifaceted questions of how properly to interpret quantum mechanics or chaos theory.19 What I should like to focus on is the concern for metaphysical space which informs the arguments of so many contemporary writers on science and theology, and to show how a return to Thomas Aquinas discussion of divine agency is particularly fruitful, especially his understanding of how God is the complete cause of the whole reality of whatever is and yet in the created world there is a rich array of real secondary causes.20 Indeed, I think that how to understand divine action in a world explained increasingly in terms of quantum mechanics, or chaos theory, or evolutionary biology is really not so different from how to understand divine action in a world explained either in terms of Aristotelian physics or Newtonian mechanics. My point is that there is a philosophical and theological way to discuss divine agency regardless of what scientic conceptions of the world we might have. The concern to afrm both divine agency in the world and also to afrm the integrity of nature so important for contemporary theologians who are attracted to developments in recent science is hardly a new concern. In looking at this concern in a broader historical context, we can take Thomas Aquinas as a guide. Gods creative act, for Thomas, is not an example of divine withdrawal21 but is, rather, the exercise of divine omnipotence. Furthermore, Thomas understanding of Gods action, throughout the entire course of cosmic history, afrms the integrity and relative autonomy of the physical world and the adequacy of the natural sciences themselves to describe this world. For some in the Middle Ages any appeal to the autonomy of nature, that is, any appeal to the discovery of real causes in the natural order, seemed to challenge divine omnipotence. One reaction, made famous by some Muslim thinkers was to protect Gods power and sovereignty by denying that there are real causes in nature. Thus, they would say that when re is burning a piece of paper it is really God who is the true agent of the burning; the re is but an instrument without any causal efcacy. Accordingly, events that occur in the natural world are only occasions in which God acts.22 For these theologians, there is a fundamental incompatibility between the view of God as the on-going cause of all that is and the view that there are autonomous natural processes

DIVINE AGENCY, CONTEMPORARY PHYSICS

587

occurring in the world. They thought that they had to deny the possibility of science (understood as the discovery of real causes in nature) in order to defend Gods omnipotence. A leading critic of this denial of real causes in nature, other than God, was the Muslin philosopher, Averroes. Yet, in his defense of the natural sciences he came to reject the doctrine of creation out of nothing, because he thought that to afrm the kind of divine omnipotence which produces things out of nothing is to deny a regularity and predictability to the natural world. Thus, for Averroes, to defend the intelligibility of nature, that is, of a world in which there really are necessary connections between cause and effect one must deny the doctrine of creation out of nothing.23 Contrary to the positions both of some Muslim theologians and of their opponent, Averroes, Thomas Aquinas argues that a doctrine of creation out of nothing, which afrms the radical dependence of all things upon God as their cause, is fully compatible with the discovery of causes in nature.24 Gods omnipotence does not challenge the possibility of real causality for creatures, including that particular causality, free will, which is characteristic of human beings. Thomas would reject any notion of divine withdrawal from the world so as to leave room, so to speak, for the actions of creatures. Thomas does not think that God allows or permits creatures to behave the way they do.25 Similarly, Thomas would reject a process theology which denies Gods immutability and His omnipotence (as well as His knowledge of the future) so that God would be said to be evolving or changing with the universe and everything in it. For process theologians, the source of conict between science and religion, from the side of religion, is the very doctrine of creation ex nihilo, since, according to them, this doctrine involves a commitment to divine omnipotence which is incompatible with the discovery of any kind of causality or power inherent in nature. Such theologians nd in the thought of Alfred North Whitehead the key to a proper rapprochement between science and theology. Not only does Whitehead highlight the importance of nature as a process of becoming, but he explicitly rejects creation ex nihilo. Whitehead thinks that the kind of extreme voluntarism which sees God as the one supreme reality, omnipotently disposing a wholly derivative world is absolutely incompatible with a true science of nature.26 As David Grifn of the Claremont Graduate School in California, one of the centers of process thought in the United States, observes: to make the very existence of a realm of nite actualities . . . contingent upon a divine decision, is to deny the possibility of any kind of causal nexus within the world. According to Grifn, the God who creates ex nihilo can interrupt not only the law of gravity but also the very principle of causation . . . . In Whiteheads naturalistic theism, by contrast, there are beneath the contingent laws of our particular cosmic epoch, some metaphysical principles, which obtain necessarily, and, therefore, cannot be violated.27 The reasons for rejecting creation ex nihilo by process

588

WILLIAM E. CARROLL

thinkers are quite similar to the ones set forth by Averroes. For David Grifn, the universal web of nite causation cannot be interrupted, even by God. The implication is that the divine causation in the world is always persuasive, never coercive in the sense of wholly determining. Thus, God belongs to the same genus as all other actual entities, thereby exemplifying the same metaphysical categories, and, signicantly, Grifn concludes that the causal relations between God and other entities are not different in kind from the causal relations between [sic] nite entities.28 In defense of the autonomy of the natural order and of the existence of real causal connections in that order, process thinkers are willing to sacrice both divine omnipotence as well as a radical distinction between divine and creaturely causality. Thomas understanding of divine action and the changes which occur in nature would allow us to avoid various attempts to accommodate the contingency afrmed in the natural sciences by re-thinking divine omnipotence, omniscience, and Gods a-temporality. Keith Ward, Regius Professor of Divinity Emeritus at Oxford, is a good example of this latter approach. Ward thinks that the traditional attempt to make God the efcient cause of all things, without compromising the simplicity and unchangeability which are characteristics of the Aristotelian picture of God was an heroic failure, since it could not account for the contingency of the universe. This is so because [t]hat which is wholly necessary can only produce that which is necessary. A contingent universe can only be accounted for if one makes free creativity a characteristic of the First Mover, which entails placing change and contingency within the First Mover itself. According to Ward, Gods omniscience is the capacity to know everything that becomes actual, whenever it does so . . . . The classical hypothesis [of a God who does not change] does not . . . seem compelling.29 In a sense, God must wait to know what is actual since there is an inherent contingency in nature itself, and as actualities change so does Gods knowledge.30 Since modern science, especially evolutionary biology, discloses a fundamental contingency in nature, the classical conception of God, inconsistent with nature so understood, must be jettisoned. So, at least, is the argument of many contemporary theologians who think that in rejecting classical theism they are honoring the insights of science. Furthermore, there is the temptation to think that the view of God-in-time, changing as the world and man change, neither omniscient nor omnipotent in the classical sense, is more consistent with the core of biblical revelation than the God described by Thomas Aquinas and others. In a Thomistic analysis, such a view of God, changing as the world changes, fails to do justice either to God or to creation. Creatures are what they are (including those which are free), precisely because God is present to them as cause. Were God to withdraw, all that exists would cease to be. Creaturely freedom and the integrity of nature, in general, are

DIVINE AGENCY, CONTEMPORARY PHYSICS

589

guaranteed by Gods creative causality. Here is how Thomas expresses this view in the Summa theologiae:
Some have understood God to work in every agent in such a way that no created power has any effect in things, but that God alone is the ultimate cause of everything wrought; for instance, that it is not re that gives heat, but God in the re, and so forth. But this is impossible. First, because the order of cause and effect would be taken away from created things, and this would imply lack of power in the Creator, for it is due to the power of the cause, that it bestows active power on its effect. Secondly, because the active powers which are seen to exist in things, would be bestowed on things to no purpose, if these wrought nothing through them. Indeed, all things created would seem, in a way, to be purposeless, if they lacked an operation proper to them, since the purpose of everything is its operation . . . . We must therefore understand that God works in things in such a manner that things have their proper operation . . . .31

Gods will is so powerful that His causal agency also produces the modality of its effect: the effect is assimilated to Gods will in every way so that not only what happens occurs because God wills it to happen, but it happens in that way which God wills it to happen.32 Gods will transcends and constitutes the whole hierarchy of created causes, both causes which always and necessarily produce their effects and causes which at times fail to produce their effects.
Gods will is to be thought of as existing outside the realm of existents, as a cause from which pours forth everything that exists in all its variant forms. Now what can be and must be are variants of being, so that it is from Gods will itself that things derive whether they must be or may or may not be and the distinction of the two according to the nature of their immediate causes. For He prepares causes that must be for those effects that He wills must be, and causes that might cause but might fail to cause for those effects that He wills might or might not be. And it is because of the nature of these causes that these effects are said to be effects that must be and those effects that need not be, although all depend upon Gods will as primary cause, a cause which transcends the distinction between must and might not. But the same cannot be said of human will or any other cause, since every other cause exists within the realm of must and might not. So of every other cause it must be said either that it can fail to cause, or that its effect must be and cannot not be; Gods will however cannot fail, and yet not all His effects must be, but some can be or not be.33

Since Gods will is the cause of being as such precisely what creation means, His causation does not compete with the causation of creatures, but rather supports and grounds it.34 Since it is characteristic of secondary causes precisely to be causes, Gods causal determination of them is not such as to deny their proper autonomy. God does not need a metaphysical indeterminacy in nature so that His actions would not collide, so to speak, with other causes.35 Gods providence is such that when we speak of an explanation that a natural effect occurs we can refer to a proximate cause. If we ask why wood is heated in the presence of re, we can explain the phenomenon in

590

WILLIAM E. CARROLL

terms of the characteristics of both wood and re. Thus, if a person answers the question of why the wood is heated by saying that God wills it, the person answers appropriately, provided he intends to take the question back to a rst cause; but not appropriately, if he means to exclude all other causes.36 For Thomas, there is no question that there are real causes in the natural order: if effects are not produced by the action of created things, but only by the action of God, it is impossible for the power of any created cause to be manifested through its effects. If no created thing really produced an effect, then no nature of anything would ever be known through its effect, and thus all the knowledge of natural science is taken away from us.37 Thomas thinks that to defend the fact that creatures are real causes, far from challenging divine omnipotence, is a powerful argument for divine omnipotence. As he says, to detract from the perfection of creatures [that is, to deny their power to produce effects] is to detract from the perfection of divine power.38 God is immediately active in all things and, as such, must be within them. In an important sense, God is more intimate to each creature than a creature is to itself.39 God, as the cause of each creatures being, is present at the very center of each creatures being. He is more interior to things than they are to themselves: not as an intrinsic principle entering into their constitution, but as the abiding cause of their existence.40 Thomas draws an analogy from the sun. Just as the air is lighted as long as it is illuminated by the sun, and falls into darkness when the sun does not shine at night, so creatures are caused to be by the creative diffusion of Gods goodness. If God were to withdraw His presence all creatures would fall into non-being.41 Thomas is inuenced in this analysis by the works of Pseudo-Dionysius.42 Simon Tugwell aptly puts it: The fact that things exist and act in their own right is the most telling indication that God is existing and acting in them.43 The source of most of the difculties in grasping an adequate understanding of the relationship between the created order and God is the failure to understand divine transcendence. It is Gods very transcendence, a transcendence beyond any contrast with immanence, which enables God to be intimately present in the world as cause. God is not transcendent in such a way that He is outside or above or beyond the world. God is not different from creatures in the way in which creatures differ from one another. We might say that God differs differently from the created order. Kathryn Tanner, who has written persuasively on this subject, observes: This non-competitive relation between creatures and God is possible, it seems, only if God is the fecund provider of all that the creature is in itself . . . . This relationship of total giver to total gift is possible, in turn, only if God and creatures are on different levels of being, and different planes of causality.44 Rudi te Velde puts it this way: God operates immanently in nature in such a way that He sets nature, so to speak, free in its own operation . . . . Thomas [sees]

DIVINE AGENCY, CONTEMPORARY PHYSICS

591

. . . God as a cause which by its transcending immanence constitutes the causality of nature in its own order.45 Proponents of what has been termed panentheism criticize classical Western theism for understanding the world as being ontologically outside of God, and, thus, as presenting signicant difculties for making sense of Gods action in the world.46 Their concern is to fashion a theology consistent with biblical revelation and the insights of contemporary science and philosophy, but their criticism of classical theism does not do justice to the position of Thomas. If we follow Thomas lead, we can see that there is no need to choose between a robust view of creation as the constant exercise of divine omnipotence and the causes disclosed by the natural sciences. No matter how random one thinks evolutionary change is, for example; no matter how much one thinks that natural selection is the master mechanism of change in the world of living things; the role of God as Creator, as continuing cause of the whole reality of all that is, is not challenged. We need to remember Thomas fundamental point that creation is not a change, and thus there is no possibility of conict between the explanatory domain of the natural sciences the world of change and that of creation.47 In defending the relative autonomy of the natural sciences, Thomas shows us how to distinguish between the being or existence of creatures and the operations they perform. God causes creatures to exist in such a way that they are the real causes of their own operations. For Thomas, God is at work in every operation of nature, but the autonomy of nature is not an indication of some reduction in Gods power or activity; rather, it is an indication of His goodness. It is important to recognize that divine causality and creaturely causality function at fundamentally different levels. In the Summa contra Gentiles, Thomas remarks that the same effect is not attributed to a natural cause and to divine power in such a way that it is partly done by God, and partly by the natural agent; rather, it is wholly done by both, according to a different way, just as the same effect is wholly attributed to the instrument and also wholly to the principal agent.48 It is not the case of partial or co-causes with each contributing a separate element to produce the effect.49 God, as Creator, transcends the order of created causes in such a way that He is their enabling origin. Yet the same God who transcends the created order is also intimately and immanently present within that order as upholding all causes in their causing, including the human will. For Thomas the differing metaphysical levels of primary and secondary causation require us to say that any created effect comes totally and immediately from God as the transcendent primary cause and totally and immediately from the creature as secondary cause.50 In response to the objection that it is superuous for effects to ow from natural causes since they could just as well be directly caused by God alone, Thomas writes that the existence of real secondary causes is not the result of the inadequacy of divine power,

592

WILLIAM E. CARROLL

but of the immensity of Gods goodness. God wills to communicate His likeness to things, not only that they might exist, but also that they might be causes for other things. Indeed all creatures generally attain the divine likeness in these two ways . . . . By this, in fact, the beauty of order in created things is evident.51 Thomas rejects any form of an emanationist scheme which views Gods causality in necessitarian terms. God is a voluntary agent: He acts not by a necessity of His nature, but by His intellect and will.52 Furthermore, God brings things into being by His wisdom, which excludes the views of those who say that all things depend on the simple will of God, without any reason.53 Thomas thinks that it is a mistake to think that justice, goodness, and truth, for example, depend only on Gods will. To make such a claim would be to deny that the will of God proceeds from His wisdom: which denial is blasphemy.54 God does not only give being to things when they rst begin to exist, He also causes being in them so long as they exist. He not only causes the operative powers to exist in things when these things come into being, He always causes these powers in things.55 Thus, if Gods creative act were to cease, every operation would cease; every operation of a thing has God as its ultimate cause. As we have seen, Thomas does not think that such an afrmation of divine omnipotence eliminates the real role of created causes. As is apparent, the analysis of divine causality I have just sketched is a complex topic in metaphysics. In fact, the very notion of cause, either as predicated of God or of things in the world of our experience, which Thomas uses is quite different from cause understood as a temporal relationship. When Thomas speaks of causality he employs a much richer sense of the term than we tend to use today. Whereas contemporary thinkers have come to view causality in terms of a kind of necessary consequentiality between events, Thomas understood causality in terms of metaphysical dependence.56 As part of the philosophy of nature connected to the rise of modern science, two of the four causes of Aristotelian science, the nal and the formal, were considered irrelevant. Furthermore, to the extent that the natural sciences came to be seen as depending exclusively on the language of mathematics, only that which was measurable would fall within their explanatory domains.57 Even the notion of agent or efcient causality underwent a profound change from the Aristotelian sense. It was conceived exclusively in terms of the force or energy that moved the fundamental parts of the universe.58 In the eighteenth century, David Hume called into question even this narrow idea of efcient causality. Since the supposed inuence of a cause upon its effect was not directly evident to sense observation, Hume concluded that the connection between cause and effect was not a feature of the real world, but only a habit of our thinking as we become accustomed to see one thing constantly conjoined to another.59 Causality,

DIVINE AGENCY, CONTEMPORARY PHYSICS

593

thus, becomes not a property of things but of thought; it is no longer an ontological reality in the world outside ourselves, but an epistemological property of the way we think about the world. [Thus,] . . . the hallmark of causality [is to be] . . . found in the epistemological category of predictability rather than the ontological category of dependence.60 One of the consequences of viewing causality exclusively in terms of a physical force is that divine causality, too, comes to be seen in such terms.61 To conceive Gods causality in this way is to make God a kind of competing cause in the world, or, perhaps better put, just one more cause in the world, although considerably more powerful than any other. To view the world as functioning in terms of an ordered regularity of mechanical causes seemed to mean that there was no room for any kind of special divine action.62 As Albert Einstein observed: The more man is imbued with the ordered regularity of all events the rmer becomes his conviction that there is no room left by the side of this ordered regularity for causes of a different nature. For him neither the rule of human nor the rule of divine will exists as an independent cause of natural events.63 Starting from this kind of faulty analysis, several contemporary theologians, as we have seen, have found such room for divine action, what Polkinghorne calls room for divine manoeuvre, in the new scientic view of the world set forth in quantum mechanics and chaos theory. In addition to the narrowing of the notion of causality, at the beginning of the early modern era there was also a signicant transformation in the conception of God, so both divine and agency took on new connotations. For Amos Funkenstein, at the onset of modernity, as God came to be described in unequivocal terms, or even given physical features and functions, [He] eventually became all the easier to discard. As a scientic hypothesis, he was later shown to be superuous; as a being, he was shown to be a mere hypostatization of rational, social, or psychological ideals and images.64 Funkenstein points to what he calls the transparency of God in the seventeenth century. He does not necessarily mean that seventeenth-century thinkers always claimed to know more about God than medieval theologians. To some of them God remained a deus absconditus about whom little can be known. What he means is that what was claimed to be known about God, be it much or little, was expressible in what Descartes calls clear and distinct ideas.65 For Thomas Aquinas, however, as we see in his doctrine of analogy, Gods transcendence is such that our language always falls short when we speak of God. We can only use language about God in a highly qualied and provisional way. In the perceptive observation of Denys Turner, the negative theology of Aquinas does not mean that we fall short of things to say about God, but that everything we say about God falls short of what God is.66 In the early modern era, for many thinkers, language came to be predicated of God in the same unequivocal way that it is predicated of things in the world; the implication of this is that God is, in some sense,

594

WILLIAM E. CARROLL

closer to things in the world, indeed, to such an extent that he becomes a thing himself. In other words, there is a qualitative change in how God is conceived.67 But in order to preserve some sense of Gods transcendence, modern thinkers began to stress that Gods goodness, for example, differed from worldly goodness only in the sense that it was much greater rather than its being the radical source of goodness, as Aquinas would have maintained. Similarly, Gods existence came to be understood to differ from the existence of creatures in the sense that His existence was innitely greater rather than radically other.68 Especially beginning in the seventeenth century, as the scientic elevation of precise, univocal, mechanical language came to inltrate both philosophical and theological thinking, it became necessary to specify what sort of thing God was.69 Yet, for thinkers such as Aquinas, God is not a thing at all. A key philosophical point here is that once being, ranging from the nite to the innite, from creatures to Creator, is conceived in a broad univocal sense, as constituting a single ontological category, God cannot be seen to transcend the ontology of this world, but must somehow be tted in given a function and location with the ontology of this world.70 God becomes a thing, albeit a supreme thing among other things in the world. Such a God is not only liable to appear incredible or unbelievable (a big thing that soon becomes too obviously a projection of ordinary things), but also, as the world becomes more selfexplanatory and self-sufcient, [God becomes] increasingly superuous. In such a world, atheism becomes almost irresistible.71 This revolution in the conception of God has its roots in the fourteenth century, especially in the thought of Duns Scotus. For Thomas Aquinas, being is not something shared by, or common to, God and creatures. Creatures be only insofar as being or existence emerges from and is created by God. Scotus rejected Thomas sense of analogy, especially as applied to being. For him, God and creatures share being, although Gods share is innitely greater than that of creatures. Hans van Balthasar observes that Scotus represents a key turning point that leads to a characteristically modern conception of God. With Scotus, being can be univocally applied to innite and nite being, that is to God and the world.72 And as Jean-Luc Marion notes, with Cajetan, Suarez, and Descartes we nd a univocist drift to any notion of analogy.73 Yet, the argument is that the foundations were really laid by Scotus for a distinctively modern conception of God, if not for a distinctively modern world-view.74

CONCLUSION

In various contemporary accounts of divine action there is a special concern to locate the causal joint, that particular point or way that divine causality can be conceived of as interfacing with the physical

DIVINE AGENCY, CONTEMPORARY PHYSICS

595

world.75 The concern for nding such a causal joint proceeds from assumptions about divine causality which are problematic. For even if we grant that contemporary physics afrms a radical indeterminism in nature, any analysis of Gods action in the world will be impaired if we restrict our notion of cause to the categories of matter, energy, and force. It is important to note, however, that the narrowing of the notion of causality, in the thought of Hume and others, to which I have referred, has occurred in the philosophy of nature, not in the empirical sciences themselves. When scientists adopt such limited or restricted notions of cause, they are operating in a broader arena of analysis than that of the empirical sciences themselves. This broader arena, the philosophy of nature, is a more general science of nature than any of the specialized sciences; it examines topics such as the nature of change and time, the role of mathematics in the investigation of nature, and related questions. One must be careful, however, not to draw too sharp a distinction between the philosophy of nature and the empirical sciences. One must also be careful not to identify the philosophy of nature with either epistemology or metaphysics. The complete dependence of all that is on God does not challenge an appropriate autonomy of natural causation; God is not a competing cause in a world of other causes. In fact, Gods causality is such that He causes creatures to be the kind of causal agents which they are. In an important sense, there would be no autonomy to the natural order were God not causing it to be so. Traditional conceptions of Gods omnipotence need not be abandoned in order to embrace an evolving universe in which real novelty and contingency are characteristic features of nature. Nor ought we to think that divine agency requires a kind of indeterminism in nature in order to have the metaphysical space for God to act without interfering. For Thomas, the natural sciences, philosophy, and theology discover complementary, not competing, truths about nature, human nature, and God. The account he offers of divine agency and the autonomy and integrity of nature is not merely an artifact from the past, but an enduring legacy.

Notes
1 Quantum Cosmology and the Laws of Nature, edited by Robert John Russell, Nancey Murphy, and C.J. Isham (Vatican City: Vatican Observatory Publications, 1993); Chaos and Complexity, edited by Robert John Russell, Nancey Murphy, and Arthur R. Peacocke (Vatican City: Vatican Observatory Publications, 1995); Evolutionary and Molecular Biology, edited by Robert John Russell, William R. Stoeger, S.J., and Francisco J. Ayala (Vatican City: Vatican Observatory Publications, 1998); Neurosciences and the Person, edited by Robert John Russell, Nancey Murphy, Theo C. Meyering, and Michael A. Arbib (Vatican City: Vatican Observatory Publications, 1999); and Quantum Mechanics, edited by Robert John Russell, Philip Clayton, Kirk Wegter-McNelly, and John Polkinghorne (Vatican City: Vatican Observatory Publications, 2001). The subtitle of each is: Scientic Perspectives on Divine Action. For a synopsis and

596

WILLIAM E. CARROLL

analysis of these works, see Wesley J. Wildman, The Divine Action Project, 19882003, Theology and Science 2: 1 (2004), 3175. Some of the contributors to these volumes responded in the October 2004 issue, and Wildman responded in 2005. 2 This would be a correlative to the need for a similar metaphysical space which allows for the causal agency of creatures. 3 From the preface to the English version of Ernst Haeckels The Riddle of the Universe (1900) by Joseph McCabe, ix. 4 This view has been set forth by Hans Blumenberg, The Legitimacy of the Modern Age, translated by Robert Wallace (Cambridge, MA: MIT Press, 1971) and Wolfhart Pannenberg, Toward a Theology of Nature, translated by Ted Peters (Louisville, KY: Westminster/John Knox Press, 1993) and in Metaphysics and the Idea of God, translated by Philip Clayton (Grand Rapids, MI: W. B. Eerdmans, 1990). For a discussion of these claims especially how to understand the principle of inertia see William E. Carroll, The Scientic Revolution and Contemporary Discourse on Faith and Reason, in Faith and Reason, edited by Timothy Smith (South Bend, IN: St. Augustines Press, 2000), 195216. 5 The claim is that the principle of inertia contradicts Aristotles claim that everything that is moved is moved by another and thus the apparent need for a conjoined mover to account for motion. 6 Pannenberg, Metaphysics and the Idea of God, 20. 7 The present state of the system of nature is evidently a consequence of what it was in the preceding moment, and if we conceive of an intelligence that at a given instant comprehends all the relations of the entities of this universe, it could state the respective position, motions, and general affects of all these entities at any time in the past or future. Laplace, Recherches sur ` lintegration des equations differentielles aux differences nies et sur leur application a lanalyse des hasards, 1776, quoted and translated in Charles C. Gillispie, Pierre-Simon Laplace 17491827. A Life in Exact Science (Princeton, N.J.: Princeton University Press, 1997), 26. Laplace was also the author of Exposition du syste`me du monde in which he argued that stability of the solar system did not require divine maintenance. His book was the occasion for a famous, but likely apocryphal, anecdote: to Napoleons query about the absence of God in Laplaces system, Laplace was said to reply that he did not need that hypothesis. 8 P.C.W. Davies, Quantum Mechanics (London: Routledge & Kegan Paul, 1984), 4. 9 Chris J. Isham, Lectures on Quantum Theory: Mathematical and Structural Foundations (London: Imperial College Press, 1995), 131132. 10 In an essay which applies such views to the eld of genetics, Russell adopts the theological view that Gods special action can be considered as objective and non-interventionist if the quantum events underlying genetic mutations are given an indeterminist interpretation philosophically. If it can be shown scientically that quantum mechanics plays a role in genetic mutations, then by extension it can be claimed theologically that Gods action in genetic mutations is a form of objectively special, non-interventionist divine action. Moreover, since genetics plays a key role in biological evolution, we can argue by inference that Gods action plays a key role in biological evolution . . .. Russell, thus, presents a sophisticated form of theistic evolution. Robert J. Russell, Special Providence and Genetic Mutation: A New Defense of Theistic Evolution, in Evolutionary and Molecular Biology: Scientic Perspectives on Divine Action, edited by Robert J. Russell, William R. Stoeger, and Francisco J. Ayala, 191223, at 213 and 206, italics in original (Vatican City: Vatican Observatory Publications, 1998). Russell provides an excellent summary of the views of two theologians, Nancey Murphy and Thomas Tracy, on the general theological signicance of quantum indeterminism 214. Italics in original. 11 Robert J. Russell, Quantum Physics and the Theology of Non-Interventionist Objective Divine Action, in The Oxford Handbook of Religion and Science, edited by Philip Clayton and Zachary Simpson. (Oxford: Oxford University Press, 2006) 586. 12 Particular Providence and the God of the Gaps, in Chaos and Complexity: Scientic Perspectives on Divine Action, 289324. See, as well, Nancy Murphy, Divine Action in the Natural Sciences, Chaos and Complexity: Scientic Perspectives on Divine Action, 324357. Russell in The Oxford Handbook: NIODA allows for Gods special providence, which believers require, since, although God causes all the processes of the ordinary world (general providence), some of those processes genuinely convey special meaning because the choices God makes in causing them, and not the other options available to God, bring them about 592. 13 J. Polkinghorne, The Laws of Nature and the Laws of Physics, in Quantum Cosmology and the Laws of Nature, edited by Robert J. Russell, Nancey Murphy, and C.J. Isham (Vatican City: Vatican Observatory Publications, 1993), 4412.

DIVINE AGENCY, CONTEMPORARY PHYSICS

597

14 How that manoeuvre is executed will depend upon other organizing principles, active in the situation, viewed holistically. A chaotic system faces a future of labyrinthine possibilities, which it will thread its way through according to the indiscernible effects of innitesimal triggers, nudging it this way or that . . . [C]haos theory [is] actually an approximation to a more supple reality, these triggers of vanishingly small energy input become non-energetic items of information input (this way, that way) as proliferating possibilities are negotiated. The way the envelope of possibility is actually traversed will depend upon downward causation by such information input, for whose operation it affords the necessary room for manoeuvre. ibid., 443 Italics in the original. 15 I am suggesting that we need to go further and recognize that the act of creating the other in its freedom involves also a kenosis of the divine omniscience. God continues to know all that can be known, possessing what philosophers call a current omniscience, but God does not possess an absolute omniscience, for God allows the future to be truly open. I do not think that this negates the Christian hope of ultimate eschatological fulllment. God may be held to bring about such determinate purpose even if it is by way of contingent paths. ibid., 4478 On this nal point see D. Bartholomew, God of Chance (London: SCM Press, 1984). The theme of kenosis is explored in a series of essays edited by Polkinghorne in The Work of Love: Creation as Kenosis (London: W.B. Eerdmans, 2001). 16 John Polkinghorne, Chaos Theory and Divine Action, in Religion and Science, edited by W. Mark Richardson and Wesley J. Wildman (New York: Routledge, 1996), 250 and 249. 17 This is a criticism aptly made by Willem B. Drees in Gaps for God? in Chaos and Complexity, op. cit., pp. 223237. That Polkinghorne is particularly susceptible to this criticism can be seen in the following observation he makes in this same volume: For a chaotic system, its strange attractor represents the envelope of possibility within which its future motion will be contained. The innitely variable paths of exploration of this strange attractor are not discriminated from each other by differences of energy. They represent different patterns of behavior, different unfoldings of temporal development. In a conventional interpretation of classical chaos theory, these different patterns of possibility are brought about by sensitive responses to innitesimal disturbances of the system. Our metaphysical proposal replaces these physical nudges by a causal agency operating in the openness represented by the range of possible behaviors contained within the monoenergetic strange attractor. What was previously seen as the limit of predictability now represents a gap within which other forms of causality can be at work. Polkinghorne, The Metaphysics of Divine Action, in Chaos and Complexity. . . , 1534. 18 Nicholas Saunders, Divine Action and Modern Science (Cambridge, U.K.: Cambridge University Press, 2002). Saunders concludes his analysis in this way: Would it be correct to argue on the basis of the foregoing critique that the prospects for supporting anything like the traditional understanding of Gods activity in the world are extremely bleak? Largely the answer to this question must be yes. In fact it is no real exaggeration to state that contemporary theology is in crisis. 215. Italics in the original. The traditional understanding Saunders has in mind refers to special providential acts by God. 19 I am persuaded by those who argue that it is philosophically suspect to argue from the essentially mathematical realm of chaos theory to reach conclusions about determinism or indeterminism in nature. The philosophical issues connected to a proper interpretation of quantum mechanics and chaos theory are extraordinarily complex. Robert J. Russell and Wesley J. Wildman, for example, note the use made by chaos theory in some theological circles: The development of chaos theory has been welcomed by some theologians as powerful evidence that the universe is metaphysically open (i.e., not completely deterministic) at the macro-level. Metaphysical indeterminacy at the quantum level does not even need to be assumed, on this view, for chaos theory makes room for human freedom and divine acts in history that work wholly within natures metaphysical openness and do not violate natural laws . . . . [Such an interpretation is] without justication . . . since it makes little sense to appeal to chaos theory as positive evidence for metaphysical indeterminism when chaos theory is itself so useful for strengthening the hypothesis of metaphysical determinism: it provides a powerful way for determinists to argue that many kinds of apparent randomness in nature should be subsumed under deterministic covering laws. Chaos: A Mathematical Introduction with Philosophical Reections, in Chaos and Complexity . . ., op. cit., 4990, at 84 and 86. 20 Too often, those who examine the distinction Thomas draws between primary and secondary causality, read Aquinas in the light of a Humean understanding of cause. See William A. Wallace, Causality and Scientic Explanation, 2 vols. (Ann Arbor: The University of

598

WILLIAM E. CARROLL

Michigan Press, 1972), and Joseph De Finance, Conoscenza dellessere, translated by M. ` Delmirani (Roma: Editrice Ponticia Universita Gregoriana, 1993), 332423. 21 This is what Polkinghorne calls a kenosis (or emptying) of divine omnipotence. 22 The best known representative of this position in Islam was al-Ghazali (10581111); see The Incoherence of the Philosophers, trans. by Michael E. Marmura (Provo, Utah: Brigham Young University Press, 1997). Maimonides (11351204), an ardent critic, describes the position of the kalam theologians in this way: They [the theologians] assert that when a man moves a pen, it is not the man who moves it; for the motion occurring in the pen is an accident created by God in the pen. Similarly the motion of the hand, which we think of as moving the pen, is an accident created by God in the moving hand. Only God has instituted the habit that the motion of the hand is concomitant with the motion of the pen, without the hand exercising in any respect an inuence on, or being causative in regard to, the motion of the pen. The Guide of the Perplexed I.73; trans. by S. Pines (Chicago: University of Chicago Press, 1963), 202. 23 [al-Ghazalis] assertion [in defense of creation out of nothing] . . . that life can proceed from the lifeless and knowledge from what does not possess knowledge, and that the dignity of the First consists only in its being the principle of the universe, is false. For if life could proceed from the lifeless, then the existent might proceed from the non-existent, and then anything whatever might proceed from anything whatever, and there would be no congruity between causes and effects, either in the genus predicated analogically or in the species. Averroes, Tahafut al-Tahafut, trans. by Simon Van den Bergh (London: Luzac, 1954), 452; also quoted in Barry Kogan, Averroes and the Metaphysics of Causation (Albany, NY: State University of New York Press, 1985), 353. 24 See my essay, Creation and Science in the Middle Ages, New Blackfriars 88 (November, 2007), 678689. 25 Thomas view of divine causality raises the specter of the so-called problem of evil. Thomas is able to respond successfully to objections that his view of Gods causality makes God the source of evil; an exposition of Thomas views on this matter are, however, well beyond the scope of my analysis here. 26 A.N. Whitehead, Adventures of Ideas (New York, NY: Free Press, 1967 [1933]), 166. 27 David Ray Grifn, Religion and Scientic Naturalism: Overcoming the Conicts (Albany, NY: State University of New York Press, 2000). 93. At the heart of the metaphysical principles is creativity, which Whitehead calls the category of the ultimate. Creativity is the twofold power of every actual entity to exert both nal and efcient causation. This idea is embodied in Whiteheads doctrine that every actuality is a momentary event, or actual occasion, which creates itself out of the causal inuences received from prior actual occasions, then exerts inuence upon subsequent occasions. In supernatualistic theism, this twofold creative power belonged essentially to God alone; any creative power possessed by nite events was a wholly contingent gift of God, which could thereby be overridden or canceled out at will, so that God could completely determine what occurs. This is the doctrine that Whitehead rejects as a false metaphysical compliment . . . . According to his naturalistic theism, the ultimate creativity in the universe is necessarily embodied in nite actualities as well as in the divine actuality. This means that power is inherent in the world as well as in God. It means, more precisely, that every one of the worlds units is inherently inuenced by all prior units, that every unit inherently has some power of self-determination, and that every unit inherently has the power to inict itself upon others, for good or for ill. 28 ibid., 96. Ian Barbour notes that what is characteristic of process theology is an afrmation of a God of persuasion rather than compulsion . . . who inuences the world without determining it. Ian Barbour, Religion in an Age of Science (San Francisco: Harper, 1990), 224. 29 Keith Ward, Religion and Creation (Oxford: Clarendon Press, 1996), 202, 188. Italics added. 30 Wards arguments are far more sophisticated than can be adequately set forth here, but for the claim that the Thomistic view of divine agency and the world of change is a great success, rather than an heroic failure, see William E. Carroll, Aquinas on Creation and the Metaphysical Foundations of Science, Sapientia 54 (1999), 6991. 31 Summa theologiae I, q. 105, a. 5. 32 De veritate, q. 23, a. 5. 33 Commentary on Aristotles De Interpretatione, Book I, lectio 14. This translation is found in Timothy McDermott (ed.), Aquinas: Selected Philosophical Writings (Oxford: Oxford University Press, 1993), 282.

DIVINE AGENCY, CONTEMPORARY PHYSICS

599

34 Harm J. M. J. Goris, Free Creatures of an Eternal God: Thomas Aquinas on Gods Infallible Foreknowledge and Irresistible Will (Nijmegen: Stichting Thomasfonds, 1996), 299. Goris notes that the distinction between divine causality and creaturely causality is based on the distinction between divine being and creaturely being: Aquinas distinguishes the being of the Creator from the being of the creature not in terms of necessary being versus contingent being but more radically in terms of being versus non-being, while God causes the either necessary or contingent being of the creature. Likewise divine causation differs from creaturely causation as being differs from non-being. Without Gods causation there is no creaturely causation at all. ibid. For a trenchant criticism of some of the other features of Goris analysis, see the review of his book by Brian Shanley, O.P. in The Thomist (April 1998). 35 In discussing how the human will is free to choose, and yet caused to be so by God, Thomas notes that the autonomy of the will does not require that it be the rst cause of its activity: Not every principle is a rst principle . . . . [A]lthough it is essential to the voluntary act that its principle be within the agent, nevertheless it is not contrary to the nature of a voluntary act that this principle be caused or moved by an extrinsic principle: because it is not essential to the voluntary act that its intrinsic principle be a rst principle. Summa theologiae I-II, q. 6, a. 1, ad 1. If the Thomist solution to the reconcilability of nite free action and divine causal power is to work . . . God cannot be inserted into the worlds causal chains, the divine causal inuence, as ex nihilo, cannot and must not be thought of as univocal with other causes. As in all other things, God is not to be conceived of as a cause in the categorical sense; He does not belong to any categories precisely because He is the cause of them all. John C. Yates, The Timelessness of God (Lanham, Maryland: University Press of America, 1990), 2526. In the Summa theologiae, Thomas writes: God is the rst cause of both natural causes and voluntary agents. And just as His moving natural causes does not prevent their acts from being natural, so also His moving voluntary agents does not prevent them from acting voluntarily, but rather makes it be just that, for He works in each according to its nature. Summa theologiae I, q. 83, a. 1, ad 3. Indeed, every movement either of will or of nature proceeds from God as the First Mover. ibid., ad 3. What is crucial for Thomas, however, is that we recognize that both natural and voluntary movements proceed from an intrinsic principle, but that need not, indeed cannot, be the truly rst principle of action. 36 Summa contra Gentiles III, c. 94. 37 Summa contra Gentiles III, c. 69. 38 ibid. 39 In I Sent., 8, 1, ad 1. 40 In I Sent., 37, 1, 1 ad 1. 41 Summa theologiae I, q. 104, a. 1. 42 Referring to Thomas analogy of the suns illumination of light, Fran ORourke observes: As the sun is naturally luminous, while air is lighted by sharing in the light of the sun although it does not partake of its nature, so also God alone is by his essence Being, while every creature is being through participation since its essence is not identical with its esse. Beings do not share in divine essence but in the illuminative effusion of divine Being which emanates from him . . . . God is present in all things not according to his essence but through a participation of his created likeness . . . . Divine similitude is not just a gift bestowed upon beings, but is their very being itself . . . . Creatures participate in Gods presence but God is not participated. Beings share in the similitude of God while God in no manner resembles them. Fran ORourke, Pseudo-Dionysius and the Metaphysics of Aquinas (London: E.J. Brill, 1992), 257258. 43 Simon Tugwell, Albert and Aquinas: Selected Writings (New York: The Paulist Press, 1988), 213. 44 Kathryn Tanner, Jesus, Humanity and the Trinity (London: Continuum, 2001), 34. For an excellent discussion of the transition between a Thomistic understanding of divine transcendence and a modern sense, especially beginning with Suarez, see William Placher, The Domestication of Transcendence (Louisville, KY: Westminster Press, 1996). 45 Rudi A. Te Velde, Participation and Substantiality in Thomas Aquinas. Studien und Texte zur Geistesgeschichte des Mittelaters, vol. 46 (Leiden: Brill, 1995), 164. 46 Philip Clayton, God and Contemporary Science (Edinburgh: Edinburgh University Press, 1997), 100. 47 For a discussion of Thomas on creation, see Steven E. Baldner and William E. Carroll, Aquinas on Creation (Toronto: Pontical Institute of Mediaeval Studies, 1997) and William E. n Carroll, La Creacio y Las Ciencias Naturales: Actualidad de Santo Tomas de Aquino (Santiago: Editorial Universidad Catolica de Chile, 2003).

600

WILLIAM E. CARROLL

48 Summa contra Gentiles III, c. 70, 8. 49 God and creatures are not two causes collaborating on the same level to produce a joint effect. God causes on the transcendental level and He thereby constitutes the creatures causation on the categorical level. Goris, op. cit., 301. 50 Brian J. Shanley, O.P., Divine Causation and Human Freedom in Aquinas, American Catholic Philosophical Quarterly 72: 1 (1998), 100 and 108. Shanley argues that no real explanation of exactly how Gods causality functions is possible, since God transcends the mundane world of causation. Michael Miller has argued that Bernard Lonergan, following in the tradition of Aquinas, provides a more philosophically satisfying account of divine causation without sacrificing divine transcendence: in Transcendence and Divine Causality, American Catholic Philosophical Quarterly 73:4 (Autumn 1999), 537554. David Burrell observes that the terms primary and secondary [causality] come into play when we are faced with the situation where one thing is by virtue of the other. So each can properly be said to be a cause, yet what makes one secondary is the intrinsic dependence on the one which is primary. This stipulation clearly distinguishes a secondary cause from an instrument, which is not a cause in its own right: it is not the hammer which drives the nails but the carpenter using it. Burrell, Freedom and Causation in Three Traditions (Notre Dame, IN: University of Notre Dame Press, 1993), 97. See also William E. Carroll, Aquinas and the Metaphysical Foundations of Science, Sapientia, op. cit. 51 Summa contra Gentiles III, c. 70. Ian Barbour reects a common criticism (which we have already seen in Keith Wards analysis) of the position of Thomas Aquinas. This position, Barbour thinks, does not fully represent the biblical idea that God has a more active and responsive role in nature and history. Moreover, if all events are predestined in the divine plan, then chance and human freedom are ultimately illusory, though they seem real to us from our limited perspective. Alternatively, if the future is open even in Gods sight, we would have to say that not all secondary causes are instruments of Gods will. In that case, creation is a greater selflimitation in Gods power and in Gods knowledge than classical theism acknowledges . . . . [T]he concepts of primary and secondary causality do not provide a coherent solution to the problem of Gods action in a world of scientic law and human freedom. When Science Meets Religion: Enemies, Strangers, or Partners? (San Francisco: Harper, 2000), 103, 161. I think that Barbour is not adequately attentive to the analogical notion of causality and the profound metaphysical difference between divine and creaturely causality which are characteristic of Thomas thought. 52 Summa contra Gentiles II, c. 23. 53 Summa contra Gentiles II, c. 24. 54 De veritate q. 23, a. 6. 55 Summa contra Gentiles III, c. 67. 56 Il titolo di questo libro richiama una nozione, causa, che suggerisce al lettore contemporaneo contenuti concettuali per qualche aspetto sostanzialmente diversi da quelli che evocava nel lettore medievale. Difatti per i contemporanei il termine causa indica per lo piu la ` ` sola idea di consequenzialita necessaria . . . Per il lettore medievale, invece, accanto allidea di una connessione di fatto, il concetto di causa trasmette quella di un ordinamento metasico. . . . La ` ` ` causa, in questo modo, e superiore alleffetto; e poiche e principio della sua sussistenza in essere, e ` principio anche della sua intelligibilita. Cristina DAncona Costa, Introduzione, in Tommaso DAquino, Commento al Libro delle Cause (Milano: Rusconi, 1986), 7. 57 Mario Bunge points out the important role that empirical science has played in this shift in our understanding of causality: The Aristotelian teaching of causes lasted in the ofcial Western culture until the Renaissance. When modern science was born, formal and nal causes were left aside as standing beyond the reach of experiment; and material causes were taken for granted in connection with all natural happenings . . . Hence, of the four Aristotelian causes only the efcient cause was regarded as worthy of scientic research. Mario Bunge, Causality and Modern Science (New York: Dover, 1979), 32. See William A. Wallace, O.P., Causality and Scientic Explanation, 2 vols. (Ann Arbor: University of Michigan Press, 1972), vol. 2, 246. I do not think that Bunge adequately distinguishes between developments in the empirical sciences and the philosophical reection on such developments. 58 Michael J. Dodds, The Doctrine of Causality in Aquinas and The Book of Causes: One Key to Understanding the Nature of Divine Action, paper delivered at the Thomistic Institute, University of Notre Dame, July 2000. I am grateful to Professor Dodds for his analysis of the narrowing of the notion of causality, which I have used in this section. 59 [U]pon the whole, there appears not, throughout all nature, any one instance of connexion [i.e., between cause and effect] which is conceivable by us. All events seem entirely loose and

DIVINE AGENCY, CONTEMPORARY PHYSICS

601

separate. One event follows another; but we never can observe any tie between them. They seem conjoined, but never connected. And as we can have no idea of any thing which never appeared to our outward sense or inward sentiment, the necessary conclusion seems to be that we have no idea of connexion or power at all, and that these words are absolutely without any meaning, when employed either in philosophical reasonings or common life. David Hume, Enquiries concerning Human Understanding and concerning the Principles of Morals (1777), 3rd edition, edited by L.A. Selby-Bigge and P.H. Nidditch (Oxford: The Clarendon Press, 1975), 74. Italics in the original. Humes analysis of causality has been especially inuential for modern thought. An examination of the philosophical presuppositions which inform his position is beyond the scope of this essay. Hume must deny the rather obvious experience we have of doing things and having things done to us, as well as the insights we have into causal dependence in the natural order. 60 Dodds, op. cit. 61 As Philip Clayton has observed: The present-day crisis in the notion of divine action has resulted as much as anything from a shift in the notion of causality. God and Contemporary Science, op. cit., 189. 62 The scientic world-view seems to leave no room for God to act, since everything that happens is determined by scientic laws. Keith Ward, Divine Action (London: Collins, 1990), l. Langdon Gilkey explains this reluctance of contemporary theologians to speak of divine intervention: Thus contemporary theology does not expect, nor does it speak of, wondrous divine events on the surface of natural and historical life. The causal nexus in space and time which Enlightenment science and philosophy introduced into the Western mind and which was assumed by liberalism is also assumed by modern theologians and scholars; since they participate in the modern world of science both intellectually and existentially, they can scarcely do anything else. Langdon Gilkey, Cosmology, Ontology and the Travail of Biblical Language, in Owen C. Thomas, ed., Gods Activity in the World: the Contemporary Problem (Chico, California: Scholars Press, 1983), 31. 63 Albert Einstein, Out of My Later Years (New York: Wisdom Library, 1950), 32. 64 Amos Funkenstein, Theology and the Scientic Imagination: From the Middle Ages to the Seventeenth Century (Princeton, N.J.: Princeton University Press, 1986), 116. 65 ibid., 25. 66 Denys Turner, Faith, Reason, and the Existence of God (Cambridge, UK: Cambridge University Press, 2004), 236. 67 For a brief discussion of this point as well as an analysis of the history of modern atheism, see: Gavin Hyman, Atheism in Modern History, in The Cambridge Companion to Atheism, edited by Michael Martin, 2746 (Cambridge, U.K.: Cambridge University Press, 2007), 39. 68 ibid. 69 ibid., 39. 70 ibid., 42. 71 ibid., 423. 72 Hans Urs von Balthasar, The Glory of the Lord: A Theological Aesthetics, vol. 5: The Realm of Metaphysics in the Modern Age, translated by Oliver Davies et al. (Edinburgh: T & T Clark, 1991), 16. 73 Jean-Luc Marion, The Essential Incoherence of Descartes Denition of Divinity, in Essays in Descartes Meditations, edited by Amelie Oksenberg Rorty (Berkeley, CA: University of California Press, 1986), 306. 74 The French philosopher, Eric Alliez, writes that with Scotus denial of analogy what can be seen to be constituted . . . is a thought whose moving edges end up leading to that scientic revolution destined to make an epoch of our modernity. Eric Alliez, Capital Times: Tales from the Conquest of Time, trans. by Georges Van De Abbeele (Minneapolis, MN: University of Minnesota Press, 1996), 226. 75 As Philip Clayton explains, If one is to offer a full theory of divine agency; one must include some account of where the causal joint is at which Gods action directly impacts on the world. To do this requires one in turn to get ones hands dirty with the actual scientic data and theories, including the basic features of relativity theory, quantum mechanics and (more recently) chaos theory. Philip Clayton, God and Contemporary Science, 192. The difculties of discovering such a causal joint, however, are evident in the work of Arthur Peacocke who maintains that the continuing action of God with the world-as-a-whole might best be envisaged . . . as analogous to an input of information rather than of energy. The problem with this notion, as Peacocke recognizes, is that in physics any input of information requires some input of matter/energy. Such matter/energy input on Gods part, however, smacks of interference with

602

WILLIAM E. CARROLL

the order of the world. Peacocke concludes that he has located, but not solved, the problem of the causal joint: How can God exert his inuence on, make an input of information into, the world-as-a-whole without an input of matter/energy? This seems to me to be the ultimate level of the causal joint conundrum, for it involves the very nature of the divine being in relation to that of matter/energy and seems to be the right place in which to locate the problem . . . Arthur Peacocke, Theology for a Scientic Age: Being and Becoming- Natural, Divine and Human (Minneapolis, MN: Fortress Press, 1993), 149151, 160161, 164.

Anda mungkin juga menyukai