From Quantum Mechanics to the Eucharistic Meal: John Polkinghorne’s ‘Bottom-up’ Vision of Science and Theology

From Quantum Mechanics to the Eucharistic Meal: John Polkinghorne’s ‘Bottom-up’ Vision of Science and Theology

Print Friendly, PDF & Email

Introduction

Rev Dr. John Polkinghorne KBE FRS needs no extended introduction to Metanexus Online readers. [1] In brief, he has had two successive careers, first as an elementary particle physicist at Edinburgh and then Cambridge (1956-1979), and then as an Anglican priest (1981- ), during which he has served in a variety of ecclesial positions (as curate, vicar, and chaplain) as well as president of Queen’s College (1989-1996). Since entering the priesthood, Polkinghorne has worked tirelessly at the frontiers of the science and theology conversation, having authored, co-authored or edited over twenty five books on various related topics. His most recently book, Science and the Trinity: The Christian Encounter with Reality (2004), serves as the occasion for this review and assessment of his work. [2]

I first met Polkinghorne only last fall, being privileged to work with him at a symposium on pneumatology sponsored by the Templeton Foundation. While I had read some of his books before, I so appreciated the clarity with which he spoke to difficult matters related to science and theology that I was led to revisit his oeuvre. I had long been convinced that the major challenges for Christian theology in the twenty-first century remain in two areas: in the engagement with the sciences and in the encounter between religions. Although my own training in the sciences is rather limited (my background and training is in systematic and constructive theology), yet going back and (re)reading Polkinghorne’s work has helped me to refocus the issues much more specifically.

In what follows, I will use Science and the Trinity as a springboard to explore Polkinghorne’s contributions to the dialogue between science and theology over the last twenty plus years. Our review and assessment will move from considerations of theological method through some of the “thick” details of Polkinghorne’s theology to an analysis of his eschatological vision. My objective in this exercise is twofold: to understand and further appreciate Polkinghorne as a scientist-theologian, and to map some of the present trajectories anticipating future conversations between science and theology.

Methodological Considerations: A “Bottom-up, Eucharistic-Assisted Logic”

Having begun as a scientist and come to theology only later in his career, Polkinghorne has written his fair share of books attempting to integrate theology with the current scientific conception of the world. [3] In Science and the Trinity, however, Polkinghorne’s goal is to allow theology to shape the agenda, rather than science. This reversal was already signaled almost a decade ago when Polkinghorne contrasted the more assimilationist strategy of fellow scientist-theologians Ian Barbour and Arthur Peacocke with his own approach to find consonance between science and theology in a way which preserved “the autonomy of theology in its dialogue with scientific culture.” [4] In this case, meaning in theology would be governed by Christian praxis and self-understanding, rather than imported from the outside. From the perspective of theological method, two features of Science and the Trinity are particularly noteworthy with regard to theology setting the stage for the conversation.

First, the terms of the science-and-theology relationship are established via theological categories. In contrast to the now classic four models approach – conflict, independence, dialogue, integration – attempting to relate science to religion, [5] Polkinghorne proposes instead four theological frameworks within which theologians have sought to engage the sciences. The Deistic model is most akin to traditional natural theology in acknowledging that scientific data point to some sort of creator of the universe, but is theologically “thin” in terms of acknowledging not much more from this same data as did the earlier natural theology: that this creator “wound up” the universe at the beginning and has been absent from it since. The Theistic model goes further than the deistic one in drawing inspiration from the Bible and even the life of Jesus, but stops short of adopting the dogmatic pronouncements of conciliar Christianity in articulating a theology for a scientific age.

The Revisionary model engages with the broad scope of traditional Christian theology, but does so seeking somewhat radical revisions of orthodox theological claims – e.g., the doctrines of the incarnation, virgin birth, and resurrection of Jesus, just to name some christological touchstones – in light of scientific discoveries. Finally, the Developmental model, Polkinghorne’s preferred approach, seeks to articulate the science-and-religion relationship less in terms of radical revision and more in terms of “a continuously unfolding exploration” (26). [6] The emphasis here is on continuity rather than discontinuity between past theological formulations and present affirmations.

How does Polkinghorne seek to ensure such continuity between his own scientifically-informed theological reflections and those of historic Christian orthodoxy? This leads to the second feature of Science and the Trinity whereby theology’s setting the agenda is clearly seen: a “bottom-up” or abductive theological method beginning with the basic data of Christian faith experience. As a scientist, Polkinghorne had long ago been trained to allow empirical data to give rise to scientific theory. While theory itself certainly shapes what scientists observe to begin with, yet theories are also revised according to empirical observations. Similarly, then, Polkinghorne has sought to reflect theologically “from the bottom-up” by beginning with the empirical data of Christian faith. [7]

What does such a bottom-up approach mean methodologically? Three interlocking methodological movements can be seen in Science and the Trinity. Initially, Polkinghorne takes the testimony of Scripture seriously as empirical data for theological reflection. His understanding of Scripture, however, is quite nuanced. While there is an evidential role for scripture, there are also controlling factors limiting or constraining the interpretation of Scripture such as its various genres, unedifying material which needs revaluation based on the developmental character of the scriptural “database,” and theological commitments regarding the divinely inspired (not dictated) character of the biblical text. Hence Polkinghorne acknowledges the polysemous character of the biblical witness, and insists that we need a “flexible hermeneutic” (35). In any case, Christian theology in a scientific age cannot proceed by ignoring the testimony of Scripture. [8]

Secondly, and building on the first, the experiences and testimonies of the earliest Christians serve as the fundamental empirical data for Christian theological reflection. Of course, these experiences have been preserved first in the apostolic tradition and then in the New Testament writings. What Polkinghorne calls “trinitarian thinking” (99-103) is precisely the apostolic witness that “arose primarily as a response to the insistent complexity of human encounter with the reality of God experienced within the growing life of the Church” (99-100). Not without reason, then, Polkinghorne has attempted to take seriously the apostolic portraits of and testimonies to the person and work of Jesus Christ. [9]

Finally, the ongoing experiences of all Christians expand the empirical database for theological reflection. By this, Polkinghorne the Anglican priest is referring to what he calls the “liturgy-assisted logic” of Christian theology which presumes a sacramental understanding of the human encounter with God in and through the eucharistic experience. While this has been a long-standing motif in Polkinghorne’s writings, [10] it finds its most developed expression to date in chapter five of Science and the Trinity, titled “The Eucharist: Liturgy-assisted Logic.” I am less interested here in the content of Polkinghorne’s theology of the Eucharist, than in its methodological implications. [11] On this latter front, the relationship between the eucharistic and liturgical experience and theological reflection is analogous to the relationship between empirical experimentation and scientific theory: “In each case, the cost of illumination is the willingness to have one’s everyday habits of thought revised and expanded under the influence of the reality encountered” (141). Hence Polkinghorne’s “liturgy-assisted logic” is part and parcel of his “bottom-up” approach to theology that admits “faith seeking understanding receives its impetus from religious experience” (118). His can therefore be rightly understood as a sacramental vision of science and theology: “Sacramental theology is as complex and sophisticated, and ultimately as powerfully insightful, as the considerations that support a fundamental theory in science” (141). [12]

Clearly Polkinghorne’s “bottom-up approach” puts him squarely in the Anglican tradition with regard to his theological method. Early on in his theological career he had already affirmed what has historically been affirmed as the “Anglican triad” of reason, tradition, and Scripture as the foundational resources for theological reflection. [13] Reason as a mode of discovery, best represented in scientific inquiry, cannot be rejected, even if its powers have to be recognized as limited. As we shall see momentarily, Polkinghorne’s commitment to historic Christian orthodoxy means that he cannot put aside (at least) the creedal confessions handed down by the church’s tradition. At the same time, as we have already seen that for Polkinghorne, the church’s tradition is not static but is fluid, constituted by the ongoing experiences and reflections of her members. [14] This experiential dimension, however, undergirds not only the church’s tradition, but also the reasoning modes of scientific investigation whereby scientists make tacit judgments as they engage personally with the objects of inquiry within a truth-seeking community. [15] In this sense, then, I suggest that Polkinghorne’s theological method is actually quite compatible with the approach also known as the Wesleyan “quadrilateral” of Scripture, tradition, reason, and experience. [16]

Before moving on, we need to ask what keeps this experiential, communal, and scriptural approach from lapsing into epistemic subjectivism or relativism. If Polkinghorne grants that scientists and theologians both formulate their theories/theologies from specific vantage points, are not truth claims relative to those perspectives? Certainly Polkinghorne insists that both scientists and theologians need to hold their truth claims in ways that allow for their revisability. At the same time, however, for this scientist-theologian, there is only “one world” (the title of an earlier Polkinghorne volume), not two. Science and theology provide complementary perspectives on this one world. Hence, there is a “unity of knowledge” which emerges over time as scientists and theologians engage this one world both practically and theoretically. [17] The question now becomes this: to what degree does Polkinghorne hold the theological claims of the Christian tradition fallibilistically and, hence, as revisable?

Theological Commitments: Polkinghorne and Theological Thickness

The extent of revision required for theology in a scientific age can be discerned in chapters three and four of Science and the Trinity regarding a theology of nature and the nature of divinity. In these chapters, Polkinghorne attempts to accomplish three broad objectives: 1) to rehabilitate, through discussions of the theology of nature and the theology of scripture (chapter two), the ancient Christian conviction regarding the Books of Nature and of Scripture as revelatory of the divine; 2) to correlate our present scientifically-informed understanding of the universe with a trinitarian vision of God on the one hand, and to view the natural world through a trinitarian theological perspective on the other; and 3) to rearticulate the traditional doctrine of God in light of our current scientific knowledge of the natural world understood as the result of the creative work of God. I will briefly summarize the content of these two chapters, paying special attention to the trinitarian character of Polkinghorne’s theological vision.

There are “seven scientifically disclosed features of our universe” that Polkinghorne sees both as “vestiges of the Trinity” on the one hand, and as illuminated by trinitarian thinking on the other (61-62). [18] First, the universe is deeply intelligible, rationally transparent, and beautiful; this evidences both the Father’s gift of the divine image and of the Holy Spirit of truth to human beings who are thereby enabled to recognize and explore this rational beauty of the universe that we inhabit. Second, the universe manifests an unpredictable evolutionary history that includes many tragic dead-ends even while also seemingly being guided by a so-called “Anthropic Principle” that has allowed for the emergence of complex human life; this reveals not only that the evolutionary course of natural history is an improvisation of God and creation, but also that God is a “fellow-sufferer” (73) who enters into the creation’s processes and through death on the cross experiences solidarity with the groanings of the created order. Third, we live in a relational universe (confirmed, for example, through entanglement effects discovered by quantum mechanics); this correlates well with the inner-trinitarian life of God understood by the ancients in terms of the notion of perichoresis (literally, mutual envelopment). Fourth, we have come to realize, through quantum experimentation, for example, that behind the appearances of our everyday experience are hidden realities which belie our commonsense conceptions of the world. Yet on this point, the trinitarian connection seems rather forced. Polkinghorne suggests that just as we must engage the quantum realm on its own terms (rather than on the terms of the Newtonian physics which govern our everyday engagements with the world), so also must we engage the divine on God’s own terms as triune (rather than on our own anthropologically derived preconceptions). In this case, Polkinghorne appears to be drawing an epistemological analogy rather than a trinitarian correlation.

Fifth, following from two (above), the course of the universe is an open process, given that the unpredictability of quantum events and chaotic systems seem to be a matters of ontological principle rather than of epistemological deficiencies; while Polkinghorne proceeds to reiterate the point about divine action as “an unfolding improvisation and not the performance of an already written score” (81), the trinitarian conclusion he draws is that creation is the work of the Word and Spirit as the “two hands of” divine order and contingency. Sixth, and further explicating two (above), we live in an information-generating universe that has allowed and sustained the emergence of complex realities; God is thus understood to be working behind-the-scenes (so to speak) as the deus absonditus and as the “hidden” Spirit who interacts with the world through “the input of information within its open history” (84). Finally, scientific cosmology predicts a universe of eventual futility: either a Big Crunch if the universe begins to collapse back upon itself, or a gradual whimpering out if the universe continues to expand forever; the trinitarian response to this anticipation of the end of cosmic history is the promise of the resu rection as seen in the life of Jesus (on which more in the next section).

Polkinghorne then turns from theology of nature to the nature of God. Here, Polkinghorne seeks a “theological thickness” which continues to allow theology (and the theological and dogmatic tradition) to set the agenda, but which at the same time remains open to the insights of contemporary science. The most important claims have to do with the relationship between God and the world articulated in terms of kenosis. Rather than adopting the more prevalent (in science and theology circles) panentheistic model to affirm God as both transcendent and immanent to the world, Polkinghorne opts instead to see creaturely “freedom” secured through the divine self-limitation driven by divine love. In particular, God’s self-limitation pertains also to the creation of time so that God “exists” in complementary modalities of both eternality and of temporality. This divine kenosis into the temporal process means that, in contrast to classical theism which affirmed God’s simple foreknowledge of future events, God possesses what Polkinghorne calls a “current omniscience, temporally indexed” (108). [19] Similarly, in contrast to the medieval doctrine of divine simplicity, God is viewed instead as internally complex, albeit without compromising the divine unity. These theological moves, Polkinghorne suggests, follows from the trinitarian thinking informed by God’s self-revelation in the incarnation: the God revealed as acting in the world in the form of the Son’s life being poured out unto death must describe the essential nature of God in Godself.

One might add that a more robust trinitarian theological vision would be not only incarnational but “pentecostal,” in terms not so much of the modern Pentecostal movement (although this may be argued as an expansion of the theme) but of the Spirit’s outpouring on all flesh on the fiftieth day after Jesus’ ascension (Acts 2). Although entitled Science and the Trinity, the Spirit receives surprisingly little mention in this volume. Building further on the theology of nature and the doctrine of God articulated in these pages, does not the narrative of the Day of Pentecost (and after) lend itself to our saying more about the Spirit than that the Spirit is the Deus absonditus who works hiddenly and quietly in the world, or that the Spirit symbolizes the improvisational God who “dances” with the contingencies of a free creation? Let me briefly pursue this line of thinking with regard to Polkinghorne’s continually developing theory of divine action.

As a scientist theologian, Polkinghorne has long wrestled with the topic of God’s action in the world. [20] The predominant models for conceiving divine action have been unsatisfactory. Interventionism or supernaturalism is dismissed either because God is reduced to being just another actor in the world or because God’s action in the world would be inconsistently intermittent; God acting only as creator of the world is too deistic, leaving little if any room for ongoing divine action; the Thomistic doctrine of God as primary cause and creatures as secondary causes results in the bifurcation of one world into two ontological realms, the theological and the scientific; and process thought’s doctrine of divine persuasive power is unable to sustain the eschatological promises of God as revealed in Scripture. Polkinghorne’s own proposal over the last decade plus has been to advocate a kind of “top-down” or holistic model of divine causality, albeit through God’s inputting of “pure” or “active” (as opposed to “energetic”) information at the level of non-linear or chaotic systems that are finely-tuned and extremely sensitive to initial conditions and are intrinsically open to the future. [21] Such input of information does not violate the law of conservation of energy, and also avoids the criticism of the god-of-the-gaps since the “gaps” are ontological rather than epistemological. [22] The analogies of other top-down causal inputs of information are the laws of nature, the interaction between the human mind and the processes of the brain, and quantum events. Yet, conceived in this way, of course, divine action would also be imperceptible to empirical inquiry.

More recently, however, Polkinghorne has reconsidered his claim that God acts in the world only through the input of pure information. Drawing from kenotic theory – that the second person of the Trinity emptied himself in the incarnation (cf. Phil. 2:5-8) – Polkinghorne suggests four levels of divine kenosis, the last of which has implications for reconceiving divine action in general and of the causal joint between God and the world in particular. [23] The kenosis of omnipotence, of simple eternity, and of omniscience leads, in part, to the previous claims regarding divine dipolarity and creaturely freedom, both of which combine to suggest the view that God knows the future according to its modal status, rather than as actual. The kenosis of causal status, on the other hand, leads Polkinghorne to suggest that divine action should not be limited to the input of pure information, but would be analogous to creaturely activities which involve “a mixture of energetic and informational causalities”; in this case, God’s activity in the world allows for “divine special providence to act as a cause among causes.” [24] Polkinghorne’s chief (only) example of such divine action is connected to the kenotic framework of thought itself: the incarnation of God in Jesus Christ.

My question concerns whether or not Polkinghorne’s present account of divine action is sufficiently trinitarian. In this new model, the role of the Spirit remains substantially what it was before: associated with the input of pure information and, hence, remaining empirically veiled. But what if we were to take not only the incarnation but also Pentecost seriously? If in the incarnation the Son takes on human flesh, at Pentecost the Spirit is “poured out on all flesh” (Acts 2:17). This more robustly trinitarian framework would not require abandonment of Polkinghorne’s complementary model that includes both “energetic transactions” and “active information” inputs. Rather, it seems to me, divine action explicated pneumatologically could be explored at various levels: that of the indeterminacy of quantum events; that of chaos/dynamical systems which are extremely sensitive to initial conditions; that of “top-down” or whole-part causation of wider (and wider) environments on their sub-systems; that of the input of pure information; and that of neurobiological, neuropsychological, and psychosociological processes. My point is that once the possibility of divine action in energetic terms is granted by way of taking the incarnation seriously, then a theologically thick account of divine action must be pneumatologically informed as well. [25]

Eschatological Anticipations: Continuity and/or Discontinuity?

Given what we have already seen in Polkinghorne’s theology of nature with regard to scientific cosmology’s predictions about the end of the world, it should not be surprising that a thick trinitarian theology would take up the eschatological question. The penultimate chapter of Science and the Trinity turns to this topic. Whereas in previous work Polkinghorne has defended the need for a fairly traditional notion of eschatological hope, [26] here focus is on the nature of life after death. If the universe is to make sense and human experience is to find final fulfillment, then creation must be redeemed from its transcience, decay, and brokenness in ways that yet retain some sort of continuity with this present world.

This continuity-in-discontinuity must also exemplify the eschatological experiences of human beings. Characteristic conditions include, Polkinghorne suggests: embodiment (albeit with transformed spiritual bodies); temporality (albeit no longer subject to death and decay); and processive experience open to the ongoing transformative graciousness of God (even in judgment, surely in purgation, definitely in the dynamical perfection of unending life with God). [27] Perhaps most important is how we account for the continuity of personal selfhood in the afterlife. Polkinghorne retrieves the ancient Aristotelian and Thomistic idea of the soul as the form or pattern of the body, but subjects it to significant revisions insofar as the human soul is an “information-bearing pattern” (161) that is dynamically shaped by its embodiment, social relations, and environmental locatedness. Not intrinsically immortal, the only hope for life after death is the faithfulness of God with regard to the promise of the resurrection from the dead. [28] The final resurrection will be a triple vindication: of God, of Jesus Christ, and of human hopes. [29]

What is at stake for Polkinghorne? Ultimately, eschatology concerns both the credibility of Christian belief and human hope beyond this life. If this is the case, then two other sets of questions arise for Polkinghorne’s scientifically informed theological vision: that related to his views regarding the omniscience of God, and the world’s religious traditions. Let me briefly explicate on the implications of each of these topics for Christian eschatology.

I have already introduced (in the preceding section) Polkinghorne’s understanding of divine kenosis into time (leading to divine temporality), and divine kenosis of knowledge (leading to a revision of the traditional doctrine of divine omniscience). Put bluntly, “God does not yet know the unformed future, simply because it is not yet there to be known” (54). Unlike process theology which asserts that God is necessarily limited by temporality, Polkinghorne’s kenosis theology insists that God has “chosen to possess only a current omniscience, temporally indexed” (108, emphasis orig.). World history is continuously unfolding, not a “fixed score” (67-68), and biblical prophecy reflects “a consonance of understanding, rather than confirmation of prediction” (53), such that later events are correlated retrospectively with previous writings. God is therefore mutable (107), continuously adjusting his plans in light of developments in the world. [30] It is clear, however, that Polkinghorne has been moved to this view by his work as a physicist: the unpredictability of quantum events and of chaotic systems are ontological features of the world’s openness to the future (79-80). Hence, nuancing Polkinghorne’s doctrine of divine omniscience, “if God’s creation is intrinsically temporal, surely the Creator must know it in its temporality. In other words, God will not simply know that events are successive but God will know them according to their nature, that is to say, in their succession” (104, italics orig.). [31]

While Polkinghorne is driven by scientific and theodicy considerations to revise the classical doctrine of God’s exhaustive and definite foreknowledge of future events, a small group of North American evangelical theologians have recently proposed a similar “open theistic” view of divine omniscience largely through a retrieval and reinterpretation of the biblical text. [32] This proposal has generated intense controversy and polemical literature, especially among members of the Evangelical Theological Society, as open theism is seen to be a plain denial of various biblical claims regarding God’s knowledge of the future. [33] Further, the open view is also understood by its detractors to undermine the sovereignty of God, resulting in God being recreated in the image of humankind. Most importantly for our purposes, there is the question about how God’s promises regarding eschatological redemption can be accomplished if the historical details are continuously being decided by free creatures, often choosing against the will of God. [34]

Now Polkinghorne has anticipated this question for his own view of divine current knowledge. In response, he asserts that, “God will not be caught out by the movements of history into the future, in the way that human beings are so often caught out…. This does not negate ultimate divine sovereignty, for we may suppose that God can bring about determinate ends through contingent paths” (108). Here, Polkinghorne’s position can actually be further bolstered through the reformulations of the open theistic position by one of its chief North American advocates, Gregory A. Boyd. In response to the criticism that the open view God is incapable of guaranteeing the eschatological promises of Scripture, Boyd has made two key moves that may be of aid to Polkinghorne. First, rather than just saying that God knows future events in their successive nature, Boyd is more specific: God knows past events as past, present events as factual, some future events as necessary (as predetermined by natural law or by God’s sovereign decision), and other future events as contingent (due to creaturely freedom). This last class of events is known by God as possibilities or probabilities, rather than as actualities. [35] The important point here is, as Polkinghorne also suggests, the nature of the future as being truly open. Second, Boyd suggests that an infinitely intelligent God is not limited like human beings in terms of the number of possibilities that need to be attended to, and hence can give undivided attention to each possibility which actualizes as though that were the only possibility to which God needed to respond. [36] Put together, might such an account buttress Polkinghorne’s eschatological confidence? “If even the omnipotent God cannot act to change the past, it does not seem any more conceivable that the omniscient God can know with certainty the unformed future”; so even if “God may not fix his will on delivering checkmate by promoting that pawn on that square…, he will certainly win the game.” [37]

Yet there are at least two further issues here for Polkinghorne. The first is the question raised by North American evangelical critics of open theism: how biblical is Polkinghorne theological vision given the scriptural testimony to God’s foreknowledge of future contingents? The second has to do with whether or not this reformulation of the classical doctrine of divine omniscience is so radical that Polkinghorne becomes a revisionary rather than developmental theologian, to use Polkinghorne’s own terms. If the former, then the lines between Polkinghorne and Peacocke (who is labeled a revisionist by Polkinghorne), for example, become blurred. In this case, for all of Polkinghorne’s desire to remain faithful to historical Christian teachings, has his own “bottom-up” thinking led him to cross the line of orthodoxy? [38]

This reference to Polkinghorne’s “bottom-up” approach to theology raises the second set of questions relative to his eschatology: that concerning the diversity and plurality of the world’s religious traditions. These topics are related because most of the important truth claims made by the religions that appear to be conflicting usually concern either transcendental or eschatological matters. [39] Now Polkinghorne himself has repeatedly acknowledged the difficult theological questions raised by religious pluralism, albeit not usually in the context of discussing eschatology. Throughout, he has consistently maintained a kind of inclusivistic position with regard to the salvation of those in other faiths, even while registering his perplexity with regard to articulating a coherent theology of religions. [40]

In the concluding chapter of Science and the Trinity, Polkinghorne defends the particularity of his Christian commitment in even while recognizing that, “The problems presented by religious diversity are serious” (175). There are certainly similarities underlying the bewildering diversity of religions, but conflicting concepts such as resurrection, reincarnation, or release from illusion in the religions “do not seem to be culturally different ways of expressing the same idea” (175). The way forward is neither to deny truth in other traditions nor to smooth over the differences. [41] Rather, we cannot but “hope that dialogue between the faith traditions, only just beginning to take place with due seriousness and probably needing centuries rather than years for its full development, will help to resolve some of these perplexities” (176).

Besides the value of the interreligious dialogue for dealing with the perennial human questions concerned with life after death, I suggest that it is important for Polkinghorne’s theological project for at least three reasons. First, the promise of a “bottom-up” approach to theology remains unfulfilled so long as the challenges raised by the diversity of religions are not confronted head on. To be sure, Polkinghorne attempts to stay true to the scriptural and dogmatic tradition of Christianity as much as possible. Nevertheless if Polkinghorne is to be truly a “bottom-up” thinker, then, as Ann Pederson and Lou Ann Trost have suggested, he needs to pay attention not just the empirical deliverances of the hard sciences but also to the experiences of all human beings. [42]

Second, the fully trinitarian theology Polkinghorne is attempting to articulate will need to emphasize not only the historical particularity of the Word incarnate but also the universal outpouring of the Spirit on all flesh. I have elsewhere argued at length that a pneumatological approach to theology of religions has remained largely unexplored, and should be pursued. [43] Elsewhere, Polkinghorne himself has wondered about the possibility of understanding the testimony of all the world’s religious traditions “as a sign of the Spirit’s veiled working within the manifold cultural contexts of humanity. Indeed, the concept of the Spirit, immanently present to creation, seems to offer Christian theology its most promising resource for a respectful approach to other faiths…. Thus it is possible for the Christian to acknowledge the authenticity of others’ experience, without denying those unique aspects of Christian understanding of the divine nature that have to be held as non-negotiable by the believer.” [44] Indeed, Polkinghorne is correct to say that we are only at the very beginning of the interreligious dialogue on matters of ultimate concern.

Finally, however, the interreligious dialogue is important also for invigorating the science-and-theology conversation. Polkinghorne himself has noted that science may serve as a meeting point for the interreligious dialogue. [45] I wholeheartedly agree. [46] More than that, might not the interreligious dialogue also serve to sharpen scientific theorizing about empirical data precisely because of the multiplicity of perspectives that are invited to the science-and-theology discussion? In fact, how might the question regarding the nature of time itself be discussed if representatives from the eastern religious and philosophical traditions were present at the science-and-religion dialogue table? [47] And, of course, to get any clarity at all about the nature of time is to do the same with regard to the nature of eternity – in which case, the results of the science-and-interreligious dialogue may also have something to say about Christian eschatology.

Conclusion

Wherever the conversation goes from here, the work of John Polkinghorne can be seen as a stimulus to the ongoing encounter between science and religion. On the one hand, Polkinghorne’s “bottom-up” approach demonstrates how any theology necessarily begins “on the ground” where human beings encounter God. On the other hand, the challenges of theologizing unavoidably from below have made much more explicit the costs involved in theological revision and the dialogue with other faiths. Further, Polkinghorne’s insistence on theological thickness is a welcome breath of fresh air in the science-and-theology dialogue, even as his vision of theology reveals how the scientific advances of our time inevitably implicate the kinds of claims that theologians can and should make. Last but not least, Polkinghorne’s perennial engagement with eschatological matters not only does not shy away from the most controversial issues for theology in a scientific age, but also points to how religion and theology do indeed have something to say about human hope which cannot be gained through the sciences alone. In all of these matters, theologians who wish to engage the sciences in their work and scientists who are also theologically motivated can do much worse than to follow the lead of John Polkinghorne, FRS. May the conversation continue…

Endnotes

1 See http://www.polkinghorne.org/; Polkinghorne was elected as a Fellow to the Royal Society in 1954, and then appointed Knight Commander of the Order of the British Empire in 1997.

2 John Polkinghorne, Science and the Trinity: The Christian Encounter with Reality (New Haven and London: Yale University Press, 2004). Unless otherwise noted, all references to this volume will be by page numbers inserted parenthetically in the main text, or notes, as the case may be.

3 By my count, Polkinghorne has written at least twelve books of this genre over the last twenty plus years: The Way the World Is: The Christian Perspective of a Scientist (Grand Rapids: Eerdmans, 1983); One World: The Interaction of Science and Theology (London: SPCK, 1986; reprint, Princeton, NJ: Princeton University Press, 1987); Science and Creation: The Search for Understanding (Boston: Shambhala/New Science Library, 1989); Science and Providence: God’s Interaction with the World (Boston: Shambhala/New Science Library, 1989); Reason and Reality: The Relationship between Science and Theology (Philadelphia Trinity Press International, 1991); Quarks, Chaos and Christianity: Questions to Science and Religion (London: SPCK, 1994; reprint, New York: Crossroad, 1998); Serious Talk: Science and Religion in Dialogue (Valley Forge: Trinity Press International, 1995); Scientists as Theologians: A Comparison of the Writings of Ian Barbour, Arthur Peacocke and John Polkinghorne (London: SPCK, 1996); Belief in God in an Age of Science (New Haven and London: Yale University Press, 1998); Science and Theology: An Introduction (London: SPCK, and Minneapolis: Fortress Press, 1998); Traffic in Truth: Exchanges between Science and Theology (Norwich, Eng.: Canterbury Press, 2000; reprint, Minneapolis: Fortress Press, 2002); and Faith, Science and Understanding (New Haven and London: Yale University Press, 2000).

4 Polkinghorne, Scientists as Theologians, 85. To be fair, Polkinghorne admitted that Peacocke’s “assimilationism” is much more theologically robust than Barbour’s process theism, even if not quite manifesting the “theological thickness” of historic Christian orthodoxy.

5 See Ian G. Barbour, Religion in an Age of Science (San Francisco: Harper & Row, 1990), ch. 1; cf. also Richard F. Carlson, ed., Science and Christianity: Four Views (Downers Grove: InterVarsity Press, 2000), and Niels Nenrik Gregersen and J. Wentzel van Huyssteen, eds., Rethinking Theology and Science: Six Models for the Current Dialogue (Grand Rapids and Cambridge: William B. Eerdmans Publishing Company, 1998).

6 Not surprisingly, Polkinghorne also suggests that Peacocke represents the revisionary approach, and Barbour the theistic model. Polkinghorne’s example of the deistic perspective is the work of physicist Paul Davies; see Davies, The Mind of God (New York: Simon and Schuster, 1992).

7 Notice then the subtitle of Polkinghorne’s Gifford Lectures given in 1993-1994: Polkinghorne, The Faith of a Physicist: Reflections of a Bottom-Up Thinker (Princeton, NJ: Princeton University Press, 1994).

8 For more on Polkinghorne’s views regarding the nature and interpretation of Scripture, see Reason and Reality, ch. 5.

9 Most extensively in his Gifford Lectures, The Faith of a Physicist, chs. 5-7, on Jesus, the crucifixion and resurrection, and the confession of Jesus’ divinity, respectively. Elsewhere, Polkinghorne has suggested that the apostolic experience serves a kind of revelatory function that bears “analogy with the role played by observations and experiments in science” (Faith, Science and Understanding, 52).

10 See One World, 98; Science and Providence, 92-94; and Faith of a Physicist, 158-60.

11 For the record, the some of the central features of Polkinghorne’s theology of the Eucharist are that: a) it is celebrated in the context of external threat and internal betrayal – hence it is closed in some respects (not anything goes), but open in others; b) it is a communal experience and activity – the Eucharist conducts the church and her members rather than vice-versa; c) through it we commune with one another and with God; d) the bread and wine are not merely natural goods, but “the products of human labor” – hence the Eucharist “unites nature and human culture” (133); and e) it is about the present and future worlds intersecting as we meet the broken and yet resurrected Christ (see Science and the Trinity, ch. 5).

12 Polkinghorne’s long-standing commitment to these historic sources for theology and to orthodoxy itself have led some to see him (rightly, in my opinion) as a apologist for a scientific age; see Paul Avis, “Apologist from the World of Science: John Polkinghorne FRS,” Scottish Journal of Theology 43 (1990): 485-502.

13 Polkinghorne, Science and Creation, 96. For more on the “triad,” see Ross Thompson, Is There an Anglican Way? Scripture, Church, and Reason: New Approaches to an Old Triad (London: Darton, Longman, and Todd, 1997).

14 I would call this, following Dale Irvin, the church as a “traditioning” reality; see Dale T. Irvin, Christian Histories, Christian Traditioning: Rendering Accounts (Maryknoll: Orbis, 1998).

15 On this point, Polkinghorne draws on the scientific method of Michael Polanyi; see Science and the Trinity, 58; Reason and Reality, ch. 4; and Belief in God in an Age of Science, ch. 5.

16 See Donald A. D. Thorsen, The Wesleyan Quadrilateral: Scripture, Tradition, Reason and Experience as a Model of Evangelical Theology (Grand Rapids: Zondervan, 1990), and W. Steven Gunter, et al., Wesley and the Quadrilateral: Renewing the Conversation (Nashville: Abingdon, 1997). My own theological method is both “Anglican” and “Wesleyan”; see Yong, Spirit-Word-Community: Theological Hermeneutics in Trinitarian Perspective (Burlington, Vt.: Ashgate, 2002).

17 Polkinghorne, Traffic in Truth, 4, and passim.

18 Much of the scientific material presented in this chapter has appeared in previous Polkinghorne volumes, but the trinitarian perspective is most succinctly and systematically laid out in this chapter.

19 I return to this matter of divine omniscience in part III.

20 For an introduction to the discussion on divine action in the science-and-religion conversation, see Christopher Southgate, et al., God, Humanity, and the Cosmos: A Textbook in Science and Religion (Harrisburg, Penn.: Trinity Press International, 1999), ch. 7, and Wesley J. Wildman, “The Divine Action Project,” Theology and Science 2:1 (2004): 31-75.

21 Polkinghorne writes: “The word ‘information’ is being used…to represent the influence that brings about the formation of a structured pattern of future dynamical behaviour. This is not the same as the registration or transmission of bits of information in the sense used by telephone engineers or, more formally, by the mathematical theory of communication. A much closer analogue is provided by the ‘guiding wave’ of Bohm’s version of quantum theory. The latter encodes information about the whole environment (it is holistic), and it influences the motion of a quantum entity by directional preferences but not by the transfer of energy (it is active in a non-energetic way). For information in the sense of the telephone engineer, there is a necessary cost in energy input, since the signal has to rise above the level of the noise of the background. For the Bohmian guiding wave there is no such energy tariff; the wave remains effective however greatly it is attenuated. I believe, therefore, that it is possible to maintain a clear distinction between energetic causality and ‘informational’ causality…” (Belief in God in an Age of Science, 66-67; cf. also Faith, Science and Understanding, 124-25). For further details, see Polkinghorne, Science and Providence,”chs. 2-4; Quarks, Chaos and Christianity, ch. 5; Science and Theology, ch. 5; “The Metaphysics of Divine Action,” in Robert John Russell, Nancey Murphy, and Arthur R. Peacocke, eds., Chaos and Complexity: Scientific Perspectives on Divine Action (Vatican City State: Vatican Observatory, and Berkeley, Calif.: The Center for Theology and the Natural Sciences, 1995), 147-56; and “The Laws of Nature and the Laws of Physics”, in Robert John Russell, Nancey Murphy, and C. J. Isham, eds., Quantum Cosmology and the Laws of Nature: Scientific Perspectives on Divine Action (Vatican City State: Vatican Observatory, and Berkeley, Calif.: The Center for Theology and the Natural Sciences, 1996), 429-40.

22 At various places, Polkinghorne has argued, against Bohm and others, that “epistemology models ontology” – e.g., Belief in God in an Age of Science, 52-53; Science and Theology, 30-31; and Quantum Theory: A Very Short Introduction (Oxford: Oxford University Press, 2002), 85-86 – and that the unpredictability of quantum phenomena has to do with their ontological character rather than with our epistemic ignorance. Yet others – e.g., Arthur Peacocke, “A Response to Polkinghorne,” Science and Christian Belief 7 (1995): 109-15, esp. 11 – have raised questions precisely about whether or not ontological indeterminism and openness can be legitimately drawn or inferred from epistemological uncertainty. These are issues which require further discussion.

23 Polkinghorne, “Kenotic Creation and Divine Action,” in John Polkinghorne, ed., The Work of Love: Creation as Kenosis (Grand Rapids and Cambridge: Eerdmans, and London: SPCK, 2001), 90-106, esp. 104-5. In dialogue with the work of Jörgen Moltmann, a kenotic theology had already made its way into Polkinghorne’s toolkit by the mid-1980s; see Science and Creation, ch. 4; cf. also Faith, Science, and Understanding, ch. 6.

24 Polkinghorne, “Kenotic Creation and Divine Action,” 101 and 104 respectively.

25 In a fascinating paper, Steven D. Crain, , “Divine Action in a World Chaos: An Evaluation of John Polkinghorne’s Model of Special Divine Action,” Faith and Philosophy 14:1 (1997): 41-61, suggests both that direct divine intervention with regard to saving humans from their sin is theologically and scientifically unobjectionable, and that given God’s transcendence, divine action is metaphysical in nature and hence obviates the need for us to find a “causal joint” between God and the world. While recognizing the need to tread carefully with regard to correlating scientific theories with pneumatological theology, I am convinced that any theologically thick account cannot avoid either metaphysical assumptions (if not arguments) or the plausibility conditions of a scientifically informed worldview. In this sense, my proposals are not meant to somehow uncover the causal joints for “pentecostal” phenomena, but rather to take the biblical narratives seriously in the science-and-th
eology conversation.

26 See Faith of a Physicist, ch. 9, and Quarks, Chaos, and Christianity, ch. 7.

27 There are indications that Polkinghorne is a hopeful universalist, as in his reading of 1 Cor. 15:22 (30-31), his denial of traditional (e.g., Dantean) images of hell (46-47), and his view of divine mercy having no limits, even on the other side of death (158-59). Yet Polkinghorne wishes to also take seriously human freedom to reject God, and in that sense, allows for the possibility of hell understood as “the dreary town, lost down a crack in the floor of heaven, of C. S. Lewis’s The Great Divorce” (159).

28 But it is here at this point of the “intermediate state” where another contemporary Thomistic view of the soul, by Terence Nichols, differs from Polkinghorne’s. While also conceiving the soul in terms of “information,” Nichols sees no other way to account for biblical data like that of the appearance of Samuel’s ghost, or the promises in the Christian Testament that to be absent from the body is to be present with God, than to argue that God graciously endows the soul with immortality, even if such immortality awaits the final resurrection of the body. See Terence L. Nichols, The Sacred Cosmos: Christian Faith and the Challenge of Naturalism (Grand Rapids: Brazos Press, 2003), ch. 7.

29 See Polkinghorne, Searching for Truth: Lenten Meditations on Science and Faith (New York: Crossroad, 1996), 155.

30 With regard to relativity theory, God’s time is suggested to be that of “the frame that is at rest with respect to the cosmic background radiation” (110). For a similar argument regarding the divine temporal frame after creation as measured according to the cosmic time of the General Theory of Relativity, see William Lane Craig, “God and Real Time,” Religious Studies 26 (1990): 335-47.

31 For previous articulations of this view of omniscience, see Polkinghorne, Quarks, Chaos, and Christianity, 73-78; Belief in God in an Age of Science, 73-74; and Faith, Science, and Understanding, 150-51.

32 The chief spokespersons are Clark H. Pinnock, John Sanders, Richard Rice, William Hasker, and Gregory A. Boyd. I summarize the basic issues in the debate in two articles: “Divine Knowledge and Future Contingents: Weighing the Presuppositional Issues in the Contemporary Debate,” Evangelical Review of Theology 26:3 (2002): 240-64, and “Divine Knowledge and Relation to Time,” in Thomas Jay Oord, ed., Philosophy of Religion: Introductory Essays (Kansas City, Mo.: Beacon Hill Press/Nazarene Publishing House, 2003), 136-52.

33 See, e.g., Bruce A. Ware, “Defining Evangelicalism’s Boundaries Theologically: Is Open Theism Evangelical?” Journal of the Evangelical Theological Society 45:2 (2002): 193-212; Douglas Wilson, ed., Bound Only Once: The Failure of Open Theism (Moscow, Id.: Canon Press, 2001); and John Piper, Justin Taylor, and Paul Kjoss Helseth, eds., Beyond the Bounds: Open Theism and the Undermining of Biblical Christianity (Wheaton, Ill.: Crossway, 2003).

34 This critical question regarding the impossibility of an open theistic eschatology is raised by Bruce A. Ware, Their God is Too Small: Open Theism and the Undermining of Confidence in God (Wheaton, Ill.: Crossway, 2003), esp. ch. 5.

35 For a much clearer articulation of the claim that God knows the future as future and future contingents in terms of possibilities and probabilities, see Gregory A. Boyd, God of the Possible: A Biblical Introduction to the Open View of God (Grand Rapids: Baker, 2000), esp. 15-18.

36 See Boyd, “Neo-Molinism and the Infinite Intelligence of God,” Philosophia Christi, Series 2, 5:1 (2003): 187-204.

37 Polkinghorne, Science and Providence, 79 and 98.

38 This is no mere speculative question, as theologians the stature of Thomas Oden have called the open view of God “heretical”; see Oden, “The Real Reformers are Traditionalists,” Christianity Today 42:2 (9 Feb. 1998): 45.

39 I argue this point in Yong, “The Spirit Bears Witness: Pneumatology, Truth, and the Religions,” Scottish Journal of Theology 57:1 (2004): 14-38, esp. 15-25.

40 Polkinghorne’s inclusivism is best seen in his claim that the monotheistic Semitic faiths “are surely all seeking to speak of the same God, but they do so with very different voices” (Scientists as Theologians, 61). For more on Polkinghorne’s inclusivism, see his Faith of a Physicist, ch. 10, and Science and Theology, 124. For his confession of theological perplexity about the diversity of world’s religions, see Science and Providence, 58, and Serious Talk, 16.

41 So, “I do not believe that progress would come from denying the reality of others’ religious experience or of my own Christian convictions” (Faith, Science, and Understanding, 65).

42 Ann Pederson and Lou Ann Trost, “John Polkinghorne and the Task of Addressing a ‘Messy’ World,” Zygon 35:4 (2000): 977-83. Pederson and Trost are concerned primarily with the experiences of women, the marginalized, and the oppressed. I see no reason to exclude the experiences of those in other faiths.

43 See Yong, Beyond the Impasse: Toward a Pneumatological Theology of Religions (Grand Rapids: Baker Academic, 2003). And beyond even this, there is also the potential for a pneumatological approach to the science-and-theology conversation;

44 John Polkinghorne and Michael Welker, Faith in the Living God: A Dialogue (Minneapolis: Fortress Press, 2001), 74.

45 Polkinghorne, Science and Theology, 125-27.

46 See Yong, “Christian and Buddhist Perspectives on Neuropsychology and the Human Person: Pneuma and Pratityasamutpada,” Zygon: Journal of Religion and Science 40:1 (2005): 143-65; cf. also see Yong, “Discerning the Spirit(s) in the Natural World: Toward a Typology of ‘Spirit’ in the Theology and Science Conversation,” Theology and Science, forthcoming.

47 Polkinghorne himself raises this question in Scientists as Theologians, 62.

Amos Yong (Ph.D., Religious and Theological Studies, Boston University), is Associate Professor of Theology at Bethel University, St. Paul, Minnesota. He has published in scholarly journals such as Zygon: Journal of Religion and Science, and Theology and Science. His has written four books, the latest forthcoming this summer, The Spirit Poured Out on All Flesh: Pentecostalism and the Possibility of Global Theology (Baker Academic).